(Oxford Library of Psychology) Philip Gable, Matthew Miller, Edward Bernat - The Oxford Handbook of EEG Frequency-Oxford University Press (2022)
(Oxford Library of Psychology) Philip Gable, Matthew Miller, Edward Bernat - The Oxford Handbook of EEG Frequency-Oxford University Press (2022)
E E G F R E QU E N C Y
OXFOR D LIBR A RY OF PS YCHOLOG Y
EEG
FREQUENCY
Edited by
PHILIP GABLE, MATTHEW MILLER,
EDWARD BERNAT
Great Clarendon Street, Oxford, OX2 6DP,
United Kingdom
Oxford University Press is a department of the University of Oxford.
It furthers the University’s objective of excellence in research, scholarship,
and education by publishing worldwide. Oxford is a registered trade mark of
Oxford University Press in the UK and in certain other countries
© Oxford University Press 2022
The moral rights of the authors have been asserted
First Edition published in 2022
Impression: 1
All rights reserved. No part of this publication may be reproduced, stored in
a retrieval system, or transmitted, in any form or by any means, without the
prior permission in writing of Oxford University Press, or as expressly permitted
by law, by licence or under terms agreed with the appropriate reprographics
rights organization. Enquiries concerning reproduction outside the scope of the
above should be sent to the Rights Department, Oxford University Press, at the
address above
You must not circulate this work in any other form
and you must impose this same condition on any acquirer
Published in the United States of America by Oxford University Press
198 Madison Avenue, New York, NY 10016, United States of America
British Library Cataloguing in Publication Data
Data available
CIP data is on file at the Library of Congress
ISBN 978–0–19–289834–0
DOI: 10.1093/oxfordhb/9780192898340.001.0001
Printed and bound by
CPI Group (UK) Ltd, Croydon, CR0 4YY
Oxford University Press makes no representation, express or implied, that the
drug dosages in this book are correct. Readers must therefore always check
the product information and clinical procedures with the most up-to-date
published product information and data sheets provided by the manufacturers
and the most recent codes of conduct and safety regulations. The authors and
the publishers do not accept responsibility or legal liability for any errors in the
text or for the misuse or misapplication of material in this work. Except where
otherwise stated, drug dosages and recommendations are for the non-pregnant
adult who is not breast-feeding
Links to third party websites are provided by Oxford in good faith and
for information only. Oxford disclaims any responsibility for the materials
contained in any third party website referenced in this work.
Contents
List of Contributors ix
Foreword xiii
Preface xv
PA RT I
1. Introduction: Methods for Collecting EEG Data for Frequency
Analyses in Humans 3
Philip A. Gable and Matthew W. Miller
2. Logic behind EEG Frequency Analysis: Basic Electricity and
Assumptions 15
Kyle J. Curham and John J. B. Allen
PA RT I I
8. Gamma Activity in Sensory and Cognitive Processing 145
Daniel Strüber and Christoph S. Herrmann
9. Frontal Midline Theta as a Model Specimen of Cortical Theta 178
James F. Cavanagh and Michael X Cohen
10. The Role of Alpha and Beta Oscillations in the Human EEG during
Perception and Memory Processes 202
Sebastian Michelmann, Benjamin Griffiths,
and Simon Hanslmayr
11. Theory and Research on Asymmetric Frontal Cortical Activity as
Assessed by EEG Frequency Analyses 220
Eddie Harmon-Jones, Taylor Popp, and Philip A. Gable
12. Oscillatory Activity in Sensorimotor Function 259
Bernadette C. M. van Wijk
PA RT I I I
13. EEG Frequency Development across Infancy and Childhood 293
Kimberly Cuevas and Martha Ann Bell
14. Developmental Research on Time-Frequency Activity in
Adolescence and Early Adulthood 324
Stephen M. Malone, Jeremy Harper, and William G. Iacono
15. Theta-Beta Power Ratio: An Electrophysiological Signature of
Motivation, Attention and Cognitive Control 352
Dennis J. L. G. Schutter and J. Leon Kenemans
16. Cortical Source Localization in EEG Frequency Analysis 377
Wanze Xie and John E. Richards
17. Frequency Characteristics of Sleep 401
Alpár S. Lázár, Zsolt I. Lázár, and Róbert Bódizs
18. A Review of Oscillatory Brain Dynamics in Schizophrenia 434
Kevin M. Spencer
19. EEG Frequency Techniques for Imaging Control Functions
in Anxiety 464
Jason S. Moser, Courtney Louis, Lilianne Gloe,
Stefanie Russman Block, and Spencer Fix
Contents vii
PA RT I V
20. Bivariate Functional Connectivity Measures for Within-and
Cross-Frequency Coupling of Neuronal Oscillations 495
J. Matias Palva and Satu Palva
21. Multivariate Methods for Functional Connectivity Analysis 514
Selin Aviyente
22. Brain Stimulation Approaches to Investigate EEG Oscillations 532
Florian H. Kasten and Christoph S. Herrmann
23. Parameterizing Neural Field Potential Data 563
Bradley Voytek
Index 579
List of Contributors
Simon Hanslmayr, Professor, Centre for Cognitive Neuroimaging, School for Neuroscience
and Psychology, University of Glasgow, UK
Eddie Harmon-Jones, Professor, School of Psychological Science, University of New
South Wales, Sydney, Australia
Jeremy Harper, Department of Psychiatry and Behavioral Sciences, University of
Minnesota, Minneapolis, MN
Christoph S. Herrmann, Professor, Experimental Psychology Lab, Carl von Ossietzky
University, Oldenburg, Germany
William G. Iacono, Professor, Department of Psychology, University of Minnesota,
Minneapolis, MN
Florian H. Kasten, Experimental Psychology Lab, Carl von Ossietzky University,
Oldenburg, Germany
Andreas Keil, Professor of Psychology, Department of Psychology and Center for the
Study of Emotion & Attention, University of Florida, Gainesville, FL.
J. Leon Kenemans, Professor, Biopsychology and Psychopharmacology, Utrecht
University, Utrecht, Netherlands
Alpár S. Lázár, Associate Professor, School of Health Sciences, University of East
Anglia, Norwich, UK
Zsolt I. Lázár, Assistant Professor, Faculty of Physics, Babeş-Bolyai University, Cluj-
Napoca, Romania
Courtney Louis, Department of Psychology, College of Social Science, Michigan State
University, East Lansing, MI
Stephen M. Malone, Research Assistant Professor and Co-investigator, Minnesota
Center for Twin & Family Research, University of Minnesota, Minneapolis, MN
Ali Mazaheri, Associate Professor, School of Psychology, University of Birmingham,
Edgbaston, Birmingham, UK
Sebastian Michelmann, Princeton Neuroscience Institute, Princeton, NJ
Matthew W. Miller, Associate Professor, School of Kinesiology, Auburn University,
Auburn, AL
Jason Moser, Professor, Department of Psychology College of Social Science, Psychology,
Michigan State University, East Lansing, MI
Belel Ait Oumeziane, Department of Psychological Sciences, Purdue University, West
Lafayette, IN
LIST OF CONTRIBUTORS xi
Steven J. Luck
When Hans Berger conducted his pioneering human EEG recordings in the 1920s, his
first major discovery was the alpha rhythm, a 10-Hz oscillation that grew larger when the
subject’s eyes were closed. A few years later, Lord Adrian (my intellectual great-great-
grandfather) showed that the alpha rhythm also varied according to whether the subject
was focusing intensely or daydreaming. Thus the study of EEG oscillations was born.
But this area of research underwent a protracted childhood, because the scientists
of the mid-twentieth century could not easily see smaller oscillations amidst the cha-
otic twists and turns of the scalp EEG. To pull out specific neural processes from the
complex and noisy EEG, they began to use signal averaging techniques that can isolate
the brain potentials that are triggered by specific events such as the onset of a light (the
event-related potentials or ERPs). However, these techniques assume that the phase of
the signal is constant across trials, and the application of signal averaging to EEG data
eliminates or hopelessly distorts oscillating activity. Indeed, for the first 30 years of my
own research career, I viewed the alpha rhythm as a nuisance that should be suppressed
lest it contaminate my precious ERP waveforms.
All of this began to change in the 1980s and 1990s, partly driven by high-profile studies
of local field potential oscillations in animals and partly driven by the application of
time-frequency analysis methods to human EEG recordings. The brain oscillations that
were obscured by signal averaging could now be visualized and quantified. A new gen-
eration of researchers began studying human EEG oscillations and linking them with
microelectrode data from animals and computational models of brain dynamics.
In science, the introduction of a new approach often leads to a burst of progress
followed by the realization that things are not as simple as they seem. The new wave of
oscillation research followed this common path, with the discovery of important new
phenomena being accompanied by conceptual and methodological challenges. One
such challenge—related to the famous Heisenberg Uncertainty Principle—is that pre-
cision in the time-domain is inversely related to precision in the frequency domain. The
more precisely you determine the frequency of an oscillation, the less you can say about
when the oscillation was present.
A second important challenge is the difficulty of distinguishing between bona fide
oscillations and other kinds of neural events. Part of the genius of the Fourier transform
xiv Foreword
Motivation
The time is ripe for a comprehensive book on the array of historical and cutting-edge
frequency and time-frequency approaches to studying EEG/ERP because of the wide-
spread interest in frequency research. There is a great need for a book organizing the di-
verse and important methods of EEG frequency analyses and interpreting the resultant
measures.
One stream of research comes from traditional (band-based) frequency analyses.
Likewise, understanding the cutting-edge frequency analyses which may not be familiar
to many EEG researchers is increasingly important for investigators applying frequency
analyses. However, there is a major need for a comprehensive handbook on analyses
within this domain. Although research has been rapidly accumulating over the last
decade, there has not been sufficient organization of research on the topic. We believe
this comprehensive handbook is increasingly necessary to help delineate the boundaries
of the area, the major scientific questions that need to be addressed, and the core theor-
etical frameworks that can guide future research and development.
Thus, a specific goal of this book is to bring together these various scientific
perspectives and research approaches within a single reference volume that provides an
integrated, cutting-edge overview of the current state of the field. This volume comprises
contributions from leading researchers within various allied disciplines.
The use of electroencephalography (EEG) to study the human mind has seen tre-
mendous growth across a vast array of disciplines due to increased ease of use and
affordability of the technology. EEG is a non-invasive measure of electrical brain ac-
tivity. Typically, researchers investigate the EEG signal using either time-domain (e.g.,
ERP) analyses or frequency analyses. Several books have examined practicalities of
conducting ERP analyses and interpreting various ERP measures. However, a compre-
hensive book has yet to be developed organizing the numerous ways to process EEG
frequency and interpreting frequency measures linked to cognitive, affective, and motor
processes.
We (editors Philip, Matt, and Ed) felt a great need for a book organizing the diverse
and fascinating methods of EEG frequency analyses and interpreting the resultant
measures. Frequency analyses provide unique assessments of neural functioning,
neural connectivity, and “resting” neural activity studied by EEG researchers. Further,
xvi Preface
frequency-domain measures are reliably associated with cognitive, affective, and motor
processes of great interest to neuroscientists and psychological scientists. For example,
asymmetrical activation of the frontal cortex as measured by the inverse of alpha-band
activity is closely linked with motivation and emotion. In addition, analyses examining
the synchrony of EEG frequencies recorded from different scalp locations allow
researchers to examine brain connectivity without having to incur the costs of magnetic
resonance imaging.
ORIGIN
As EEG frequency researchers, we wanted a resource that introduced the myriad ways
in which EEG frequency analyses are being investigated. This volume began while the
lead editor, Philip, was on sabbatical and seeking to begin a new chapter in his career.
During that time, he visited with Matt Miller and the project quickly developed into a
collaborative project. After developing the project more, Matt and Philip brought Ed
Bernat on board.
As individual editors, we each had areas of expertise in EEG frequency research, but
in developing this volume, we quickly discovered that each of our individual areas of
expertise were far different from each other. Three editors were necessary to even try
to cover the breadth of EEG frequency research being conducted. In addition, we did
not want to create a handbook that focused on EEG frequency research specifically in
emotion, cognition, or clinical applications. Instead, we wanted to provide a survey of
the breadth of work being conducted with EEG frequency research. Try as we might, we
also acknowledge this handbook will inevitably fail to cover everyone’s interest across all
topics. To that end, we hope to receive feedback from readers so that future editions of
this handbook can be expanded to encompass the ever-growing field of EEG frequency
research.
Together, the project has been a long labor, but well worth the time and effort to de-
velop the resource. We have been especially excited to work with leading experts in the
field as they develop chapters for the volume. We are excited for you as the reader to see
what we have gotten to see throughout the editing process: the excitement and develop-
ment of EEG frequency research across a wide range of fields and programs of research.
ORGANIZATION
To aid in reading the handbook, much thought and structuring has been given to the
organization of the chapters. As a whole, the book provides a systematic summary of
EEG frequency analyses and applications. Individual chapters give depth to each type
Preface xvii
LIMITATIONS OF THE
CURRENT VOLUME
As with any book, the current volume is not a complete work on the topic of EEG fre-
quency analyses. In addition, this is not the first book on EEG frequency research. Many
xviii Preface
excellent books have been published this topic. Here we mention what the current book
does not include and refer the reader to additional resources available in other books.
One of the most inspiring resources for us as editors is Steven J. Luck and Emily
S. Kappenman’s Oxford Handbook on Event-Related Potential Components. This book
has been used in our courses and labs, as well as by countless other EEG researchers. It
focuses on the excellent work that has investigated the spectrum of ERP components
derived from EEG research. We were inspired to build a similar handbook that would
cover EEG frequency analyses. The current book does not address ERP research in
much detail. The closest chapters addressing ERP research are Chapters 4 and 5. For
those desiring a more comprehensive volume on ERP analyses, please see Luck and
Kappenman.
Another extraordinary resource is Mike X. Cohen’s Analyzing Neural Time Series
Data: Theory and Practice. This book has served as the primary resource for many EEG
researchers, including us, to learn how to conduct time-frequency analyses, and to teach
our students how to perform the analyses. The book is particularly helpful in guiding
the reader from the mathematical bases of frequency analyses to the implementation
of these analyses in MATLAB. These analyses are referenced throughout our book and
include fast Fourier transforms, complex Morlet wavelet convolution, intertrial phase
clustering, surface Laplacian filtering, phase-and power-based connectivity measures,
and cross-frequency coupling. Chapters in our book that address these “how-to” topics
include Chapters 4, 5, 19, 20, 21, and 23.
Finally, Chapter 1 of this volume describes how to collect EEG data. For readers
who which to have greater detail about implementing and collect EEG from human
participants, we recommend Dickter and Kieffaber’s EEG Methods for the Psychological
Sciences. This book is an excellent resource for researchers beginning to implement EEG.
Pa rt I
Chapter 1
INTRODU C T I ON
Methods for Collecting EEG Data for Frequency
Analyses in Humans
This chapter aims to provide a structure for readers to understand the methodology
behind collecting EEG (electroencephalography) presented in the subsequent chapters.
It is important to first understand the research methods involved in recording EEG fre-
quency before delving into more advanced frequency analyses and the interpretations.
As researchers, we focus this first chapter on a brief introduction to the topic of EEG
methodology and scientific practices. To begin, however, we feel it is important to lay
down definitions for terms used throughout the book.
EEG refers to the recording of electrical brain activity from the human scalp. It is one
of the most common methods for measuring brain functioning in areas of mind, brain,
and behavior science. EEG data contain rhythmic activity or waves that may reflect
neural oscillations, or fluctuations in the excitability of populations of neurons (more
on this later). These rhythmic fluctuations are typically described using two main
descriptors. The first is frequency, which is the speed of the wave, and it is measured
in hertz (Hz), which is the number of wave cycles per second. The second is power,
which is the squared amplitude of the wave. The greater the power of an oscillation, the
greater the energy of that oscillation. All of the chapters in this work discuss frequency
4 PHILIP A. GABLE and MATTHEW W. MILLER
with most referencing power. Sometimes researchers investigate the phase of the wave,
which is the position of the wave measured in radians or degrees. Many of the chapters
included here discuss phase.
The brain produces rhythms in multiple frequencies, which can be isolated from
the raw EEG signal using multiple techniques described by Curhamn & Allen in
Chapter 2, and Voytek in Chapter 23). Different psychological processes are linked to
different frequencies, which are often grouped into bands. The most commonly studied
bands include delta (1–4 Hz), theta (4–8 Hz), alpha (8–12 Hz), beta (13–30 Hz), and
gamma (lower gamma 30–80 Hz; upper gamma 80–150 Hz). While these are not the
only frequency bands, these are the bands most typically associated with processes
of mind and behavior measured by EEG. Importantly, these bands are not defined
without reason, but instead reflect biological changes at the cellular level (see Cohen,
2014 and Buzsaki, 2006 for reviews). However, these bands are not rigid and may vary
depending on individual differences, such as brain development, structure, and chem-
istry. Chapters in the second section of this work note how different frequency bands are
associated with cognitive, motivational, and sensorimotor processes.
These definitions are by no means complete. Individual chapters provide more pre-
cise definitions of terms used. With this initial framework, readers should be able to
venture into subsequent chapters focusing in more detail on these definitions. Because
EEG is a rather complex measure, we focus in more detail on the physiological basis and
scientific methods used to record and process EEG.
EEG is measured because all nerve cells communicate using electrical signals, sending
information throughout the brain and to the rest of the body. Within a neuron, an action
potential is an electrical wave that runs from the axon hillock at the cell body to the axon
terminals. At the axon terminals, the action potential causes neurotransmitter to be
released. This neurotransmitter crosses the synaptic gap and binds to receptors on the
membrane of the postsynaptic cell. Binding to the receptor causes voltage changes by
activating ion channels or second messengers that either excite or inhibit the postsynaptic
neuron. The summation of this voltage change in the membrane of the dendrites and cell
body of the postsynaptic neuron is called a postsynaptic potential. Postsynaptic potentials
tend to occur locally, rather than moving down the axon. This allows postsynaptic
potentials to summate rather than to cancel, resulting in voltage changes that have larger
amplitudes and can be recorded on the cortical surface or at the scalp.
When tens of thousands to millions of neurons are excited or inhibited at the same
time, the voltage change outside the cell (extracellular potential) can be recorded at the
scalp using EEG, which measures the sum of electrical activity from excitatory and in-
hibitory postsynaptic potentials over this collection of neurons. The activity can only be
recorded on the scalp surface because tissue (cerebrospinal fluid, meninges, skull, and
INTRODUCTION 5
skin) between the neurons and scalp conducts the electrical signal. In addition, for elec-
trical activity to be projected to the scalp, cellular alignment must be precisely arranged
in parallel so that their effects cumulate to project the electrical activity to the scalp (see
Curhamn & Allen, Chapter 2; Keil & Thigpen, Chapter 3). Neurons must be arranged
so that the cluster of neurons all have dendrites at one pole and axons departing at the
opposite pole. This arrangement is called an open field and occurs when neurons are
organized in layers. The cortex, cerebellum, and parts of the thalamus tend to have this
open field arrangement of neurons resulting from pyramidal cells.
Due to the electrical basis of the EEG signal, EEG has excellent ability to tell us when
something is happening in the brain. This is called temporal resolution and is one of
the greatest strengths of the EEG signal. The EEG signal measures neural activity at the
accuracy of milliseconds, which allows for the ongoing measurement of psychological
processes as they unfold (Luck, 2014).
However, the EEG signal is limited in its ability to measure where something is
occurring. This is called spatial resolution and is one of the greatest weaknesses of the
EEG signal. Depending on where the source of the EEG signal is generated, the orien-
tation of the open field neurons might not be parallel to the scalp, thus generating EEG
signals in multiple directions. In addition, resistors (e.g., the skull) in the tissue between
the neurons and scalp can cause the EEG signal to spread out. Because of the volume
conduction through the head, as well as the orientation of the pyramidal cells emitting
the signal, the spatial location of the signal is difficult to ascertain. As Keil and Thigpen
(Chapter 3) note, a difference in frequency power between two experimental conditions
could be the result of a different number of neurons activated, the temporal order in
which they were activated, or neurons with different orientations being activated. To
address EEG’s limited spatial resolution, cortical source localization techniques have
been developed and are reviewed by Xie and Richards (Chapter 19). In sum, using an
analogy from Steve Luck’s ERP Boot Camp, the EEG signal is like being able to see every
frame of a movie as it unfolds. However, because of the low spatial resolution, the movie
appears a bit blurry.
The earliest method of EEG measurement was implemented by Hans Berger in the
late 1920s. In his early experiments, he used two sponges soaked in saline connected
to an amplifier (Berger, 1929). While the equipment and processing of EEG signal has
advanced considerably since that time, the basic components remain similar. EEG
6 PHILIP A. GABLE and MATTHEW W. MILLER
electrodes are placed on or near the head, the signal from the electrode is transmitted to
an amplifier, and the signal is digitized and recorded.
In psychological research labs, EEG is usually recorded from 32, 64, 128, or more
electrodes. In other research (e.g., sleep, nonhuman) fewer electrodes (2–8 electrodes)
is more typical. When larger numbers of electrodes are used, electrodes are mounted in
an electrode cap or net. When fewer electrodes are used, electrodes may be positioned
individually on the head using a bonding agent. Electrodes can either be wet electrodes
or dry electrodes. Wet electrodes are made of silver, silver-chloride, or tin, and a con-
ductive gel or liquid is placed inside or around the electrode. Dry electrode systems
use electrodes coated in gold, silver, or nickel, and place electrodes directly on the scalp
without a conductive medium. Wet electrodes generally have higher signal quality, but
dry electrodes may be preferred when high impedance levels are tolerable, or when re-
cording for long periods.
Electrode systems will have active or passive electrodes. Active electrodes con-
tain a small pre-amplification unit directly attached to the conductive metal in
the electrode. This allows the EEG signal picked up at the sensor to be immedi-
ately amplified before additional environmental noise can be introduced. Passive
electrodes do not have amplification at the electrode, and instead carry the signal to
an amplifier about a meter away. Compared to passive electrodes, active electrodes
minimize noise introduced during signal transmission, tolerate high impedance re-
cording, and reduce participant preparation time. Passive electrodes have a lower
profile to benefit transcranial magnetic stimulation over the cap and can be used
inside an MRI bore.
Electrode placement is predominantly based on the 10–20 system (Jasper, 1958).
Electrodes are named using the first letter to refer to the brain region under the electrode
from anterior to posterior (e.g., Fp—frontal pole, F—frontal, C—central, P—parietal,
T—temporal, O—occipital). Numbers following the letter are used to indicate the lateral
position of the electrodes. Ascending odd numbers indicate sites more lateral over the
left hemisphere of the brain, whereas ascending even numbers indicate sites more lateral
over the right hemisphere of the brain. The letter Z is used to designate medial sites. In
addition to the recording electrodes, EEG also requires a ground electrode, which assists
in reducing electrical noise, as well as a reference electrode placed on the head or face.
The raw EEG signal is usually filtered during recording. Signals below 0.1 Hz or above
200 Hz are removed because the frequency bands of interest fall within this range.
A filter at 60 Hz (in North and South America) or 50 Hz (in Europe, Asia, and Africa)
may also be used to further reduce electrical noise from alternating current.
1.6 Artifacts
The quality of the EEG is crucial to EEG frequency analysis. To best record EEG signal
reflecting brain activity, researchers must remove signal that occurs because of anything
other than neural activity. Signal that is not naturally present is called artifact. These
INTRODUCTION 7
can be biological (e.g., muscle movement) or nonbiological (e.g., electrical noise). Much
artifact can by eliminated by taking preventative measures. Artifact that cannot be
prevented should be removed.
Muscle artifact, or electromyography (EMG), is one of the most common types of
artifacts and is usually high in frequency (100–500 Hz). Usually, this falls outside of the
frequency range typically investigated by researchers. However, some muscle artifact
may seep into lower frequencies. Researchers can reduce muscle artifact by instructing
participants to limit their muscle movements. Muscle artifact that does occur can be
removed through visual inspection, filtering, and automatic artifact detection algo-
rithm. It should be noted that some muscle artifact may be related to the experiment
(e.g., sensorimotor studies). In such cases, it may be beneficial to measure EMG at the
site movement is expected (e.g., the hand), then control for it in analyses.
Eye movements are another common type of artifact. The eyeball is polarized which
causes large artifact in the form of voltage changes resulting from moving the eyes. Like
dealing with muscle artifact, eye movement artifact can be removed from signal using
recordings near the eyes called electro-oculograms (EOG). One pair of EOG electrodes
are placed above and below the eye, while another pair is placed just lateral to either eye
on the temple. Eye blinks also create eye movement artifact. It is preferable to correct
blink artifact using an artifact reduction algorithm based on regression, principal com-
ponent analyses, or independent component analyses.
Artifacts occurring in the environment are the result of nonbiological factors. The
most common sources of these artifacts are the result of external electrical noise coming
from compact fluorescent lightbulbs, data hubs, or electrical junctions. Grounding will
aid in reducing these sources of noise, as will electromagnetically shielded rooms.
Once an EEG signal is recorded, the raw data must go through several processing
steps before it is in a format useable in analyses. The raw signal is collected in the
time-domain but must be converted into a frequency-domain representation. One
way this can be accomplished is in the form of a power spectrum, which collapses fre-
quency data across time to map the frequencies present. A frequency analysis can be
conducted over windows that are minutes, seconds, or milliseconds in length; these are
called epochs. Epochs that are seconds or minutes can be analyzed for power spectra
using a Fourier transform, which decomposes a signal into a series of sine and cosine
functions of various frequencies. The function of each frequency begins with its own
phase. A Fourier transformation assumes that the epoch repeats infinitely forward
and infinitely backward in time. A process called windowing is used to prevent artifact
created from the Fourier transform. However, windowing can also introduce artifact
and data loss into the frequency analysis. Overlapping epochs is a way to prevent dis-
continuity, data loss near the ends of the epoch, and to help meet the assumptions of the
Fourier transform in windowing.
8 PHILIP A. GABLE and MATTHEW W. MILLER
One of the most common forms of signal frequency processing is to use a fast
Fourier transform (FFT). An FFT provides the spectrum of frequency power for a
period, which is often averaged across a range of frequencies comprising a band (e.g.,
theta). It also provides a spectrum of phase. The power spectrum reflects the energy
of each frequency determined by the squared amplitude of the wave. The phase spec-
trum reflects the phase in radians or degrees of the sine or cosine wave at each interval
(e.g., 1/T). Most frequency analyses focus exclusively on frequency power. However,
there is increasing interest in examining the phase of frequencies (e.g., see Michelmann,
Griffiths, & Hanslmayr, Chapter 10; Palva & Palva, Chapter 20). Crucially, the FFT has
two limitations: it poorly depicts changes in the frequency spectrum over time, and it
assumes the EEG data are stationary during the period to which the FFT is applied. To
overcome these limitations, time-resolved frequency decomposition techniques are
growing in popularity, particularly wavelet analyses (e.g., complex Morlet wavelets),
which reveal changes in power at various frequencies with excellent temporal precision
(see Aviyente, Chapter 4; Weinberg, Ethridge, Oumeziane, & Foti, Chapter 5).
1 There have been special issues in the International Journal of Psychophysiology and Psychophysiology
devoted to reproducibility and, relatedly, open science in cognitive electrophysiology, and readers are
encouraged to read these special issues (Kappenman & Keil, 2017; Larson & Moser, 2017).
INTRODUCTION 9
Each study should test specific hypotheses because results that confirm hypotheses
are more likely to be true than results based on exploratory analyses (Ioannidis, 2005).
Thus, it is crucial that researchers do not rewrite their hypotheses to fit with their results,
a practice known as HARKing (hypothesizing after results are known), as it exaggerates
the confidence the reader has that the results are true. One method to avoid HARKing
is pre-registering hypotheses using the Open Science Framework (osf.io), aspredicted.
org, or other repositories. However, formulating specific hypotheses for cognitive elec-
trophysiology studies can be difficult, especially if researchers want to frame them stat-
istically. For example, a researcher may be confident in predicting that an experimental
condition will affect EEG activity, but they may struggle to define “EEG activity” as
a dependent variable. In ERP studies, this is less of a concern because the dependent
variables (ERP components) are well-characterized (Kappenman & Luck, 2012); how-
ever, there are fewer well-characterized time-frequency variables. Although researchers
can be vague about defining their time-frequency variables, they should then use stat-
istical analyses with strict corrections for multiple comparisons (see Cohen, 2014). This
reduces the likelihood of making Type I errors, but consequently increases the likeli-
hood of making Type II errors. (Researchers may also consider data-driven region-of-
interest approaches; see Brooks et al., 2017). Therefore, time-frequency analyses will
benefit from having well-defined dependent variables (Indeed, this was an initial im-
petus for this book!).
For example, if a researcher believes an experimental manipulation is likely to influence
EEG activity related to cognitive control, they can have a clearly defined dependent vari-
able of oscillatory activity in the theta frequency bandwidth measured from frontal mid-
line electrodes over a certain time (Cavanagh & Cohen, Chapter 9). In a pre-registration,
researchers should use a more precise definition of their dependent variable than “frontal
10 PHILIP A. GABLE and MATTHEW W. MILLER
midline theta”. For example, they could specify that they will determine the wavelet of 4–
8 Hz that exhibits the greatest peak power between 200 and 600 ms after stimulus onset
at electrode Fz for each participant, and then compute the average power of this wavelet
during the 200–600 ms time epoch at Fz. They could also introduce some flexibility into
the specification of the dependent variable by noting that they will choose a different band-
width, epoch, and/or electrode if the grand average time-frequency plot (averaged across
all conditions) reveals an unexpected time-frequency and/or scalp distribution. In this
example, the average across all conditions avoids biasing the analysis in favor of choosing
a time-frequency window exhibiting differences between conditions. Besides reducing
HARKing and facilitating the specification of dependent variables, pre-registration is cru-
cial for holding researchers accountable to their research design, statistical analyses, and
sample size; however, none of these positive features of pre-registration work if researchers
deviate from their pre- registration without properly noting the deviation (Claesen
et al., 2019).
Another way that researchers can increase the likelihood that study results are reprodu-
cible is by increasing the power of their studies (Ioannidis, 2005). Alarmingly, Button
and colleagues’ (2013) analysis of neuroscience studies found their average power was
very low (8–31%). There are two general ways that researchers can increase the power of
their studies. First, researchers should attempt to maximize the effect they are studying
and minimize its variance (i.e., increase the standardized effect size). This can be done
by using strong experimental manipulations and reliable dependent variables, such
as those discussed in this work. Additionally, researchers should increase the signal
to noise ratio in their studies by optimizing the number of trials and collecting good
EEG data (Cohen, 2017; Luck, 2014).2 Further, when possible, researchers should use
within-subjects designs, which is already the case in many cognitive electrophysi-
ology studies. Second, researchers should collect larger samples, which is particu-
larly important when testing between-subjects effects or within-between subject
interaction effects. Cognitive electrophysiology studies can require a lot of time to
collect and process data, so collecting more participants may seem burdensome, es-
pecially for researchers investigating small–medium effects. For example, a two-tailed
2
A good way for researchers to benefit the field (and their citation count) may be for them to establish
the number of trials required for different time-frequency variables in different paradigms, which has
been done for ERP variables (e.g., Rietdijk et al., 2014). It is worth noting that simply adding more trials
may not increase the signal to noise ratio, since participants may fidget more toward the end of long data
collections, consequently increasing noise.
INTRODUCTION 11
dependent t-test for an effect size of dz =0.35, an alpha =.05, and power =.90 requires
88 participants, according to G*Power 3.1.9.4 (Faul et al., 2009). Indeed, it is likely that
researchers will often find themselves studying small-medium effects. Specifically,
when researchers conduct a priori power calculations to determine their sample sizes,
they should assume that the effect sizes in the extant literature are inflated, due to pub-
lication bias by researchers and journals (i.e., only publishing significant results) (for
a more detailed discussion on sample size calculations in EEG studies, see Larson &
Carbine, 2017). Although it is difficult to collect and process large samples, it is crucial
to the reproducibility of cognitive electrophysiology studies. To reduce the demands
large sample sizes impose, researchers who mentor doctoral students, review for and
sit on the editorial boards of journals, are involved in hiring decisions about faculty and
post-doctoral researchers, and are involved in promotion and tenure decisions should
reconsider expectations about the speed of science and the number of publications (for
further discussion on these issues, see Bradley (2017) and Yeung [2019]). Also, if a re-
searcher is concerned about allocating a lot of time to a study that may not yield signifi-
cant results, they can conduct a sequential analysis where they pause data collection
after a pre-specified sample has been collected and then determine whether to continue
data collection based on if the incremental results are significant (given an adjusted
alpha level) and if the incremental results suggest an effect size that is too small to be of
interest (Lakens, 2014).
3
There is a special issue in the International Journal of Psychophysiology devoted to open science in
human electrophysiology, and readers are encouraged to read it (Clayson, Keil, & Larson, 2022).
12 PHILIP A. GABLE and MATTHEW W. MILLER
The recommendations were made to increase the likelihood that researchers’ ori-
ginal results will be reproducible, but the recommendations also apply to researchers
attempting to reproduce original results. Crucially, researchers attempting to repro-
duce original results should also attempt to replicate and expand an original finding
(Cohen, 2017), preferably increasing the sample size by two and half times the ori-
ginal (Simonsohn, 2015). Specifically, cognitive electrophysiology will benefit if most
studies include an attempt to reproduce an original result and then add a new result
(e.g., by adding a new experimental condition). With results from replication attempts,
more precise estimates about the direction and size of effects can be made. Of course,
it is nearly impossible to reproduce a study methodologically. For example, different
researchers may have different criteria for manually rejecting trials, and different in-
dependent component analyses may yield different components. However, researchers
can still come quite close to a methodological reproduction, especially if methods and
materials are available for them to use. A challenge for cognitive electrophysiologists
is that they often employ different signal processing methods, such as subtracting or
not subtracting the ERP from an epoch of EEG data prior to convolving the data with
a wavelet. To this end, researchers should use the same methods as the original study,
unless a different method is clearly superior. Of course, whether a different method is
clearly superior is debatable; thus it is incumbent upon the researcher conducting the
methodological reproduction to state their case in a compelling way.
1.13 Explore
Some of the most exciting and reproducible effects in cognitive electrophysiology have
been discovered by accident (e.g., Kutas & Federmeier, 2011), meaning that they would
not have occurred if researchers had only conducted confirmatory research testing a
priori hypotheses. Thus, it is imperative that researchers conduct exploratory analyses
in addition to confirmatory analyses. However, results from exploratory analyses should
be clearly labeled as such to avoid misleading readers to having excessive confidence in
the result. Ideally, then, a study will test pre-registered confirmatory hypotheses that
replicate and expand an original result and conduct exploratory analyses. Compelling
results from the exploratory analyses can then serve as a priori hypotheses in future
confirmatory research. Finally, some of the most exciting exploratory research may
come from analyzing old data in new ways. For example, Voytek (Chapter 23) proposes
exciting new analytical methods that researchers can apply; his signal processing scripts
are freely available (https://voyteklab.com/code), and researchers can use these scripts
to analyze their old data or other openly available data.
INTRODUCTION 13
References
Berger, H. (1929). Über das Elektrenkephalogramm des Menschen. Archiv für Psychiatrie und
Nervenkrankheiten, 87, 527–570. https://doi.org/10.1007/BF01797193.
Bradley, M. M. (2017). The science pendulum: from programmatic to incremental—and back?
Psychophysiology, 54, 6–11. doi: 10.1111/psyp.12608
Brooks, J. L., Zoumpoulaki, A., & Bowman, H. (2017). Data-driven region-of-interest selection
without inflating Type I error rate. Psychophysiology, 54, 100–113. doi: 10.1111/psyp.12682
Buzsáki, G. (2006). Rhythms of the brain. Oxford University Press.
Button, K. S., Ioannidis, J. P. A., Mokrysz, C., Nosek, B. A., Flint, J., Robinson, E. S. J., & Munafò,
M. R. (2013). Power failure: Why small sample size undermines the reliability of neurosci-
ence. Nature Reviews Neuroscience, 14, 365–376. doi:10.1038/nrn3475
Claesen, A., Gomes, S., Tuerlinckx, F., & Vanpaemel, W. (2019). Comparing dream to
reality: An assessment of adherence of the first generation of preregistered studies. PsyArXiv
[online]. https://doi.org/10.1098/rsos.211037
Clayson, P. E., Keil, A., & Larson, M. J. (2022). Open science in human electrophysiology.
International Journal of Psychophysiology, 174, 43–46. https://doi.org/10.1016/j.ijpsy
cho.2022.02.002
Cohen, M. X. (2014). Analyzing neural time series data: Theory and practice. MIT Press.
Cohen, M. X. (2017). Rigor and replication in time-frequency analyses of cognitive elec-
trophysiology data. International Journal of Psychophysiology, 111, 80–87. doi: 10.1016/
j.ijpsycho.2016.02.001
Faul, F., Erdfelder, E., Buchner, A., & Lang, A. (2009). Statistical power analyses using G*Power
3.1: Tests for correlation and regression analyses. Behavior Research Methods, 41, 1149–1160.
doi:10.3758/BRM.41.4.1149
Goodman, S. N., Fanelli, D., & Ioannidis, J. P. A. (2016). What does research reproducibility
mean? Science Translational Medicine, 8, 341ps12. doi: 10.1126/scitranslmed.aaf5027
Groppe, D. M. (2017). Combatting the scientific decline effect with confidence (intervals).
Psychophysiology, 54, 139–145. doi: 10.1111/psyp.12616
Ioannidis, J. P. A. (2005). Why most published research findings are false. PLoS Medicine, 2,
e124. doi: 10.1371/journal.pmed.0020124
Jasper, H. H. (1958). The ten- twenty electrode system of the International Federation.
Electroencephalography and Clinical Neurophysiology, 10, 371–375.
Kappenman, E. S., & Keil, A. (2017). Introduction to the special issue on recentering
science: Replication, robustness, and reproducibility in psychophysiology. Psychophysiology,
54, 3–5. doi: 10.1111/psyp.12787
Kappenman, E. S., & Luck, S. J. (Eds.) (2012). The Oxford handbook of event-related potential
components. Oxford University Press.
Kutas, M. & Federmeier, K. D. (2011). Thirty years and counting: Finding meaning in the N400
component of the event-related brain potential (ERP). Annual Review of Psychology, 62, 621–
647. doi: 10.1146/annurev.psych.093008.131123
Lakens, D. (2014). Performing high-powered studies efficiently with sequential analyses.
European Journal of Social Psychology, 44, 701–7 10. doi: 10.10002/ejsp.2023
Larson, M. J. & Carbine, K. A. (2017). Sample size calculations in human electrophysiology
(EEG and ERP) studies: A systematic review and recommendations for increased rigor.
International Journal of Psychophysiology, 111, 33–41. doi: 10.1016/j.ijpsycho.2016.06.015
14 PHILIP A. GABLE and MATTHEW W. MILLER
Larson, M. J. & Moser, J. S. (2017). Rigor and replication: Toward improved best practices in
human electrophysiology research. International Journal of Psychophysiology, 111, 1–4.
doi: 10.1016/j.ijpsycho.2016.12.001
Luck, S. J. (2014). An introduction to the event-related potential technique (2nd ed.). MIT Press.
Open Science Collaboration. (2015). Estimating the reproducibility of psychological science.
Science, 349, aac4716. doi: 10.1126/science.aac4716
Rietdijk, W. J., Franken, I. H., & Thurik, A. R. (2014). Internal consistency of event-related
potentials associated with cognitive control: N2/P3 and ERN/Pe. PLoS One, 17, e102672.
doi: 10.1371/journal.pone.0102672
Simonsohn, U. (2015). Small telescopes: detectability and the evaluation of replication results.
Psychological Science, 26, 559–569. https://doi.org/10.1177/0956797614567341
Yeung, N. (2019). Forcing PhD students to publish is bad for science. Nature Human Behaviour,
3, 1036. doi: 10.1038/s41562-019-0685-4
CHAPTER 2
L O GIC BEHI ND E E G
F REQU ENCY A NA LYSI S
Basic Electricity and Assumptions
2.1 Introduction
The brain is an electrochemical machine. Although nerve cells differ greatly in their
size and morphology, all pass messages using electrical signals, sending information
throughout the brain and to the rest of the body via the spinal cord. Within any neuron,
an action potential is a wave of electrical activity that travels along the nerve membrane.
Action potentials are triggered by the summation of input from other neurons using
chemical neurotransmitters, which create voltage potential changes in the post-synaptic
neuron. Surface-recorded EEG is blind to the activity of single neurons but can non-
invasively measure the electrical activity resulting from summated excitatory and in-
hibitory post-synaptic potentials over millions of these signals in humans. EEG has
excellent temporal resolution, but poor spatial resolution; it can tell us when something
is happening in the brain, but not precisely where it is happening.
EEG gives an incomplete picture of the electrical activity occurring in the brain. The
cortex is the outermost layer of the brain, and the primary generator of the electrical
activity we measure with EEG. Our ability to detect cortical activity largely depends
on the parallel arrangement of cortical pyramidal neurons. When millions of parallel
neurons fire simultaneously, their electrical activity adds together to generate a signal
large enough to detect at the scalp. However, when neural populations fire incoherently,
or when they are not arranged in a parallel formation, the electrical activity does not
add constructively, so there is no observed signal at the scalp. This chapter introduces
basic concepts in electricity in signal processing, with some practical considerations
for data collection and analysis, to help readers understand EEG measurement and
interpretation.
16 KYLE J. CURHAM and JOHN J. B. ALLEN
Every physical thing is made up of atoms, which in turn are made up of fundamental
particles including protons, neutrons, and electrons. Protons have a positive charge,
neutrons have a neutral charge, and electrons have a negative charge. Charges of the
same sign repel, and opposite charges attract. A few simple experiments demonstrate the
existence of electrical charge. For example, rubbing a balloon on a wool sweater makes
the balloon negatively charged as electrons move from the wool to the balloon (Figure
2.1). The degree of attraction or repulsion between two point charges is proportional
to the product of the charges divided by the inverse squared distance between them.
Therefore, when the electron-rich balloon is brought into proximity of the electron-
poor wool, or any neutral surface such as a piece of paper or the wall, the balloon will
be attracted. If you subsequently rub a second balloon in the same way, the two balloons
will repel each other since they are both negatively charged.
Most of the time, atoms have equal numbers of protons and electrons. However,
atoms can become ionized when electrons are removed or added, resulting in a net
charge. Electricity is the phenomenon that describes the behavior and movement of
charge. In general, it doesn’t matter whether it is the electrons or ions that are moving.
The flow of charge can be accomplished either by the transfer of electrons from atom to
atom, or, in the case of electrophysiology, by the diffusion of charged ions across cellular
membranes.
The degree to which electrons are free to move from atom to atom varies by material
type. For example, in metals, the outermost electrons are so loosely bound that they
freely move in the space between atoms at room temperature. Because these unbound
electrons are free to travel from atom to atom, they are called free electrons. The relative
mobility of electrons within a material is known as electrical conductivity. Conductivity
is determined by the types of atoms in a material, and how the atoms are linked to-
gether with one another. Materials with few or no free electrons are called insulators,
and materials with many free electrons are called conductors. The directed motion of
electrons is called electrical current. Just like water flowing through a pipe, electrons
move within the empty space between atoms. Under normal conditions, the motion of
free electrons in a conductor is random, with no particular direction or speed. However,
electrons can be influenced to move in a coordinated fashion through a conductive ma-
terial by supplying a voltage. Voltage is the “pressure” that pushes on free electrons to
cause them to flow. The ability of a current to flow from one location to another depends
on the resistance. In insulating materials, such as glass or rubber, electrons have little
freedom to move from atom to atom. The less freedom electrons have to move from
atom to atom, the greater the resistance to the flow of charge. A conductor’s resistance
generally increases as its length increases or its diameter decreases. It is again useful to
refer to the water analogy: water can flow more easily through a short, wide pipe than a
long, narrow pipe. The international system of units (SI) of current is called the ampere.
LOGIC BEHIND EEG FREQUENCY ANALYSIS 17
(a)
(b)
Figure 2.1 Before rubbing the balloon against the sweater (A), no net accumulation of electrons
exists on the balloon. After rubbing the balloon on the sweater (B), the balloon has accumulated an
excess of electrons, and the resultant negative charge of the balloon and positive charge of the sweater
creates a force of attraction sufficient to keep the balloon from being pulled to the ground by gravity.
Figure credit: K. Ehrmann.
18 KYLE J. CURHAM and JOHN J. B. ALLEN
One ampere is defined as one coulomb of charge (or 6 × 1018 electrons) flowing past a
given point in a conductor in one second. The volt is the unit of pressure, that is, the
amount of electromotive force (EMF) required to push a current of one ampere through a
conductor with a resistance of one ohm, or 1 volt/ampere.
When resistance is high, electrons tend to gather on one side of the insulating material
and the positive ions tend to gather on the other, effectively storing potential energy in
an electric field. At some point, the voltage across the material will exceed a threshold
known as the dielectric constant, at which point current begins to flow. This tendency
for high resistance to result in charge separation is known as capacitance. The amount of
charge stored in the capacitor is directly proportional to the surface area of the dielectric
(the electrical insulator polarized by the electric field) (Table 2.1).
2.3 Circuits
Rseries = R1 + R2 + …+ Rn
LOGIC BEHIND EEG FREQUENCY ANALYSIS 19
(a) R1
+
V1 R2
R3
(b)
+
V1 R1 R2 R3
Figure 2.2 Simple series (left) and parallel (right) circuits. Circuit components labeled with R
indicate resistors. Circuit components labeled with V indicate voltage sources.
Conversely, adding multiple resistors in parallel will decrease the overall resistance:
1 1 1 1
= + + …+
R parallel R1 R2 Rn
As more “pipes” are added, the water has more paths to escape, decreasing the overall
resistance to flow. Capacitors wired in series or parallel follow the same rules, but
reversed:
1 1 1 1
= + + …+
Cseries C1 C2 Cn
C parallel = C1 + C2 + …+ Cn
Most circuits are some complex combination of series and parallel. We can approach
these circuits one piece at a time, deriving a new equivalent circuit at each step
(Figure 2.3).
The next few sections explore examples of equivalent circuit representations, and
we use equivalent circuit representations to learn the voltages and currents at every
point in the original complex circuit. We later show how models of neurons can
be represented as a simple equivalent circuit of capacitors, resistors, and voltage
sources.
20 KYLE J. CURHAM and JOHN J. B. ALLEN
(a) R1
+
V1 R1 R2 R3
(b) R1
+
V1 Req
(c) R1
+
V1 Req
Figure 2.3 Reducing a complex circuit (left) to a simple equivalent circuit (right). In an inter-
mediate step, we combine parallel resistors 2 and 3 (middle). Next, we combine resistor 1 with the
equivalent resistor from the intermediate step.
Circuits come in two basic flavors: direct current (DC) and alternating current (AC).
The difference depends on whether the voltage and current change directions over
time. DC circuits maintain currents flowing in a constant direction within a closed
loop, whereas AC circuits have current that repeatedly reverses direction. A DC electric
source feeds from one terminal to a set of circuit elements and then back to the other
terminal, in a complete circuit. Figures 2.2 and 2.3 are both DC circuits due to their con-
stant voltage power source. Note some resistors are connected in parallel, while others
are connected in series. At each step, we can combine resistors according to the rules in
Section 2.3 to derive a simpler equivalent circuit.
Some circuits may contain both resistors and capacitors. These are known as RC
circuits. In a simple circuit with one resistor and one capacitor in series, the capacitor
must discharge through the resistor. This discharge occurs at an exponential rate
determined by the RC time constant. The time constant indicates the number of seconds
for the capacitor to become 63.2% charged, or, equivalently, the time for current flow
LOGIC BEHIND EEG FREQUENCY ANALYSIS 21
to have slowed by 63.2% from its starting value. This choice of time constant has an in-
tuitive explanation: at any moment in time, the rate of change in voltage is equal to the
voltage divided the time constant. For example, a 1 mF cap and a 1 kΩ resistor yields a
time constant of one second. If the capacitor is charged to 5 volts, the voltage will fall at
a rate of 5 V/s. If the capacitor is charged to 2 volts, the voltage will fall at a rate of 2 V/s.
In contrast to DC signals, some sources of electricity produce AC, where voltages and
currents periodically reverse direction and switch back and forth between positive and
negative polarity. The electricity that comes from an American wall outlet is an example
of AC. The current in North America is 120 VAC and changes direction 60 times per
second. AC circuits can exhibit more interesting behaviors than DC circuits. For ex-
ample, at low frequencies, a capacitor acts like an open circuit, so no current flows in the
dielectric. However, when driven by an AC source, a capacitor will only accumulate a
limited amount of charge before the potential difference changes polarity and the charge
is returned to the source. The higher the frequency, the less charge will accumulate and
the smaller the opposition to the current.
Both resistors and capacitors resist the flow of current when a voltage is applied.
However, unlike in DC circuits, resistance may be frequency dependent. Frequency-
dependent resistance is known as impedance, a complex-valued quantity that can be
broken into two parts: magnitude (the ratio of the voltage amplitude to the current
amplitude) and phase (quantifies how much the current lags the voltage). Alternatively,
we can break impedance down into its real and imaginary parts. Like in DC circuits,
the real part of the impedance acts like resistance, resisting the flow of electric current.
The imaginary part is called reactance, and it quantifies the opposition to a change in
the current of a capacitive circuit element. Ideal capacitors are purely reactive, that is,
they have zero resistance, and the impedance of a resistor is purely real, or resistive.
Note this implies the current in a capacitor always lags the voltage by 90°. Impedance
devices add like resistors in a DC circuit. For a set of components in series, the total im-
pedance is the sum of the component impedances. To obtain the impedance of parallel
circuit components, the inverse total impedance is given by the sum of the inverses of
the component impedances.
Using different combinations of resistors and capacitors, RC circuits can be used to
attenuate some frequencies, while allowing others to propagate through the circuit un-
affected. For example, wiring a resistor in series with a load, and a capacitor in parallel
with the same load, significantly attenuates high-frequency signals. Conversely, wiring
a resistor in parallel with the load, and the capacitor in series, attenuates low-frequency
signals. In Figure 2.4, at low frequencies, the reactance of the capacitor will be very large
compared to the resistance of the resistor. This means that the voltage across the cap-
acitor will be much larger than the voltage across the resistor. At high frequencies the
22 KYLE J. CURHAM and JOHN J. B. ALLEN
reverse is true: the voltage across the resistor is larger than across the capacitor. In other
words, low frequencies pass to the output, and high frequencies are attenuated. This is
known as a low-pass RC filter. Similarly, we can construct a high-pass filter by swapping
the resistor and capacitor. The frequency cutoff for these filters is determined by the
time-constant of the circuit, which is derived from the resistance and capacitance. The
cutoff frequency for an RC circuit is:
1
fc =
2πRC
In the low-pass configuration, frequencies just above the cutoff are attenuated to half
their original amplitude. Conversely, in the high-pass configuration, frequencies just
below the cutoff are attenuated to half amplitude. The amount of attenuation increases
as you move farther beyond the cutoff frequency. Figure 2.4 shows the amplitude roll-off
as a function of frequency for a low-pass filter.
(a)
R1
Vin C1 Vout
—10
—20
—30
0
Phase (deg)
—45
—90
Frequency (Hz)
Figure 2.4 Low-pass RC filter (left). The parallel arm with the capacitor provides a low im-
pedance path for high frequency signals. However, the capacitor saturates for low-frequency
signals, providing a high-impedance path. Low-frequencies signals are thus preferentially
observed at Vout. Frequency response of the low-pass RC filter (right). The signal is not appre-
ciably attenuated below the cutoff frequency of 50 Hz (shown in red). At frequencies just above
50 Hz, the signal magnitude is cut in half. As frequency increases, the amount of attenuation
increases.
LOGIC BEHIND EEG FREQUENCY ANALYSIS 23
We can use a simple circuit model to describe the electrical properties of neurons,
including the initiation and propagation of action potentials. Action potentials are the
result of the diffusion of sodium and potassium ions across neural membranes. The
Hodgkin–Huxley model treats each component of a neuron as an electrical element in
the circuit (Figure 2.5), where current is propagated by the movement of ions across cell
membranes.
The cell membrane is represented by a capacitance (Cm). Cellular membranes are
highly resistive, and act as a dielectric material due to their relatively impermeability.
In the absence of special proteins called ion channels, ions like sodium and potassium
are unable to diffuse across the membrane, effectively turning the membrane into the
dielectric of a capacitor. As ionic currents add or subtract from the charge accumulating
inside the neuron, ions line up along the cell membrane. The differing concentration
of ions on either side of the membrane results in a net voltage potential, represented by
voltage sources (En). To maintain these concentration gradients, neurons have active
sodium-potassium pumps that exchange two sodium ions into the extracellular space
for three potassium ions in the intracellular space.
Sodium and potassium ion channels are represented by electrical conductances (GNA,
GK) that depend on both voltage and time. As the voltage potential increases, the per-
meability of the membrane is selectively modulated, that is, conductance is increased,
for specific ion species. The flux of sodium or potassium ions across the membrane is
represented by ionic currents (Ip). Since the membrane is not perfectly impermeable,
leak channels are also included, represented by another conductance (GL).
The Hodgkin–Huxley circuit model exhibits similar dynamics to real neurons. The
amount of injected current controls the emergence of a stable limit cycle. For a suffi-
ciently large input current, the circuit will exhibit repeating “action potentials” at a min-
imum firing rate. This means that either the neuron is not firing at all (corresponding
to zero frequency) or is firing at the minimum firing rate. Increasing the injected
current beyond the minimum threshold increases the firing rate of the neuron. When
Vm
inside
INa IK IL
GNa GK GL
CM
ENa EK EL
outside
the neuron fires, a series of channel activations occur to produce the action potential.
As the membrane potential approaches threshold, sodium ion channels begin to rap-
idly open, depolarizing the membrane (i.e., discharging the capacitor). The influx of
sodium changes the voltage gradient between the intracellular and extracellular space,
increasing the membrane potential. Once the polarity of the potential changes direction
(when enough sodium ions have crossed the membrane), sodium ion channels begin
to deactivate (decreasing sodium conductance). As the sodium channels close, potas-
sium channels begin to open (increasing potassium conductance), resulting in an efflux
of potassium ions to the extracellular space, restoring the membrane potential to the
resting state following a brief hyperpolarization.
2.7 Filtering
2.8 Analog-to-Digital
Signal Conversion
Most FIR filters are implemented using digital signal processing. Our discussion of elec-
tricity thus far has dealt with analog signals, which are continuous in both time and in
voltage. An economy of representation can be achieved by sampling discrete points in
both the time and voltage domains, a process of creating a digital signal, which has a tem-
poral resolution determined by the sampling rate and a voltage resolution determined
by the resolution of the analog-to-digital converter (Figure 2.6). For example, a 16-bit
converter will allow 1016 of 65,536 discrete voltage values, and a sampling rate of 1,000
Hz will allow one value every millisecond.
With sufficiently large sampling rates, a digital signal can closely approximate the
analog signal it is attempting to represent (Figure 2.7). However, several considerations
LOGIC BEHIND EEG FREQUENCY ANALYSIS 25
0
–0.5 –0.5
–1 –1
–1.5 –1.5
–2 –2
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Seconds Seconds
Figure 2.6 A signal sampled at 20 Hz. Discrete-time sampling (left panel) allows for con-
tinuous y-axis (μV) values, whereas digitally-sampled signals (right panel) must use a limited
number of y-axis values. The three bit converter illustrated here (right panel) allows for 23 =8 dis-
tinct values, providing only a coarse approximation of the signal voltage. The right panel depicts
the discrete sample value (red circle) and the 3-bit digital equivalent (red line), and the discrep-
ancy (dashed vertical black lines).
are essential to ensure signal fidelity when digitizing an analog waveform. In order to
recover all components of a periodic waveform, it is necessary to use a sampling rate at
least twice the highest waveform frequency. This is known as the Nyquist sampling rate,
and it determines whether or not aliasing will occur. Similarly, for a given sampling rate,
0 0 0
—1 —1 —1
—2 —2 —2
0 0.5 1 1.5 2 0 0.5 1 1.5 2 0 0.5 1 1.5 2
2 2 2
1 1 1
5 bits
0 0 0
µV
—1 —1 —1
—2 —2 —2
0 0.5 1 1.5 2 0 0.5 1 1.5 2 0 0.5 1 1.5 2
2 2 2
1 1 1
8 bits
0 0 0
µV
—1 —1 —1
—2 —2 —2
0 0.5 1 1.5 2 0 0.5 1 1.5 2 0 0.5 1 1.5 2
Seconds Seconds Seconds
Figure 2.7 A comparison of a signal (black line) sampled (red line) at three sampling rates (20,
40, 100 Hz) and using three different converter resolutions (4-bit, 5-bit, and 8-bit) that allow for
16, 32, and 128 distinct μV values. Low bit-resolution was used here for illustrative purposes; com-
mercial converters are typically 12-bit (4,096 values) or 16-bit (65,536 values).
26 KYLE J. CURHAM and JOHN J. B. ALLEN
Signal Aliasing
18Hz @ 20 samples/s
Figure 2.8 Signal aliasing due to insufficient sampling rate. Samples from an 18-Hz signal were
sampled at a rate of 20 Hz. Identical samples are obtained for a 2-Hz sine wave sampled at 20 Hz.
These digitized signals are indistinguishable from each other.
the highest frequency signal that can be represented is one-half the sampling rate, and
this is known as the Nyquist frequency. Aliasing occurs when the set of samples obtained
from the analog signal are indistinguishable from samples from a lower-frequency
signal. For example, in Figure 2.8, an 18-Hz signal is sampled at 20 Hz, well below the
Nyquist sampling rate. For this 20-Hz sampling rate, the Nyquist frequency is 10 Hz.
Although the signal is in fact an 18-Hz signal, the sampled signal appears as a 2-Hz sine
wave, so the signal is not well characterized. In general, signals that are x Hz above the
Nyquist frequency will appear as a signal x Hz below the Nyquist frequency. Sampling
above the Nyquist frequency will prevent aliasing but may still not characterize the
signal well in the time domain. As a general guideline, it is recommended to sample at
least 5× the highest frequency of interest to get a good signal.
FIR and IIR digital filters can be used to process digital signals. A simple moving average
is an example of an FIR filter. If we take the average across the last five data samples
and shift the five-sample window forward by one sample at each timestep, the result is
a moving average window. We can denote the value of the filtered signal (x) at the nth
timepoint using summation notation:
2
1
y (n) = ∑ 5 x (n − k )
k = −2
LOGIC BEHIND EEG FREQUENCY ANALYSIS 27
We can take this one step further, using a different weight for each sample in the
moving average window (h) of width M +1:
M /2
y (n) = ∑ h (k ) x (n − k )
k = − M /2
This result is exactly the impulse response when the input signal x is an impulse, that
is, one at the middle timestep and zero at all other timesteps. Note the filter output will
clearly be zero outside the range of the window, demonstrating that this is in fact an FIR
filter. Figure 2.9 shows the output of a moving average filter, given a noisy input signal.
The summation operation described is known as convolution. In general, convolu-
tion indicates the amount of overlap of one function or kernel (h) as it is shifted over
another function (x). In these examples, the convolution is simple to compute. However,
0 .1 .2 .3 .4 .5 .6 .7 .8 .9 1
Time (s)
(b) 1st-Order Filter with 50 Hz Cutoff
Raw
Filtered
Figure 2.9 Moving average FIR filter (top). A window size of 15 ms was convolved with noisy
data to obtain the filtered signal. Low-pass IIR filter with 50 Hz cutoff (bottom).
28 KYLE J. CURHAM and JOHN J. B. ALLEN
In general, signals may have both AC and DC components. For example, you may see
an AC signal that oscillates around a nonzero mean (a DC offset). Arbitrary complex
signals may be approximated by the sum of two or more simpler signals. For example,
we can synthesize a complex signal my summing together multiple sine waves of various
frequencies. Fourier analysis is the reverse process—decomposing a signal into its con-
stituent parts. The Fourier series approximates any complex periodic signal as a finite
weighted sum of sine waves of various frequencies. The more sine waves included in
the summation, the better the Fourier series can approximate the signal. In the limit
that the number of frequencies included in the summation goes to infinity, the Fourier
series converges to the Fourier transform. In this case, we can describe the signal as a
continuous distribution, or spectrum, of frequencies, along with the phases at which
each sine wave begins. The Fourier series can be applied to a wide array of mathem-
atical, physical, and signal processing problems. Fourier analysis is now widely used
across several domains, including audio, images, radar, sonar, X-ray crystallography,
and more.
Rather than analyzing signals as a function of time, Fourier analysis allows us to
study their properties as a function of frequency. Signals that are localized in the time
domain have Fourier transforms that are spread out across the frequency domain, and
vice versa. For example, the Fourier transform of a pure sine wave is a single point in
the frequency domain (Figure 2.10, bottom). Points in the frequency domain may be
characterized by properties such as power and phase, which are of interest for EEG
analyses. The absolute value of a given frequency component of the Fourier series
indicates the “amount” of that frequency present in the original signal. The squared
absolute value is the signal power. The power spectrum describes how signal power
varies as a function of frequency. EEG signals typically follow a 1/f trend, such that
low frequencies have more power compared to high frequencies. However, the power
spectrum may vary as a function of individual differences and task demands. Changes
in power at a frequency may be due to alterations in the slope of the EEG frequency
spectrum, or modulations in frequency-specific oscillatory activity (see Chapters 9, 10,
and 23). Similarly, the phase spectrum of the signal can be extracted from the Fourier
series. The signal phase indicates the amount of “shift” in each of the basis sine waves
(for more info, see Chapter 7).
The Fourier transform is invertible. Given the power and phase spectrum, it is pos-
sible to reconstruct the original time-domain signal. Moreover, it is possible to compute
the Fourier transform of a signal, perform mathematical operations in the frequency
LOGIC BEHIND EEG FREQUENCY ANALYSIS 29
amplitude 0.5
0
—0.5
—1
0 100 200 300 400 500 600 700 800 900 1000
time (ms)
Superposition of Sinusoids
4
amplitude
0
—2
—4
0 100 200 300 400 500 600 700 800 900 1000
time (ms)
Power Spectral Density
0
Power (dB)
–100
–200
–300
–400
0 5 10 15 20 25 30 35 40 45 50
frequency (Hz)
(b) Sinusoid
0 .1 .2 .3 .4 .5 .6 .7 .8 .9 1
Time (s)
Power Spectral Density
0
-50
Power (dB)
-100
-150
-200
-250
-300
-350
0 5 10 15 20 25 30 35 40 45 50
Frequency (Hz)
Figure 2.10 Constructing a complex signal from the superposition of sinusoids (top). The
power spectrum of the signal show distinct peaks at the frequencies of the component sinusoids.
A single sinusoid corresponds to a single peak in the power spectrum (bottom).
domain to alter the power and phase spectrum, and then compute the inverse Fourier
transform to convert back to the time domain. For every mathematical operation
in the time-domain, there is a corresponding equivalent operation in the frequency-
domain. In some cases, operations may be easier to perform in one domain or another.
This makes the Fourier transform very powerful. For example, convolution in the time-
domain is equivalent to multiplication in the frequency domain.
30 KYLE J. CURHAM and JOHN J. B. ALLEN
To sum up: compute the Fourier transform of the time-domain signal and filter
weights, multiply them in the frequency domain, and compute the inverse Fourier
transform to obtain a filtered signal.
2.11 Windowing
The Fourier transform assumes the signal is periodic and includes sine waves of infinite
period. However, EEG recording epochs are finite, spanning just seconds or minutes.
Clearly low frequencies cannot be captured by data segments smaller than the period of
the signal. For example, a segment of 250 ms will contain only a half-cycle of a 2-Hz sine
wave. In contrast, a 50-Hz waveform would complete 10 full cycles within the allotted
window. A general rule guideline to determine an adequate window size is to take 3–5
times the period of the lowest frequency of interest.
Window functions are often used to examine small segments of data from a longer
signal to study transient events that have may have different spectral properties than
other surrounding data segments. Windows are usually constructed such that they are
zero-valued outside of the interval, symmetric around the middle of the interval, and
tapering away to zero at the edge of the window. Several types of window functions are
commonly used in EEG frequency analysis, including Hann, Hamming, and Gaussian
windows. The duration of the window is application specific and is governed by
requirements such time and frequency resolution.
However, we cannot perfectly resolve both time and frequency simultaneously.
Time and frequency are conjugate variables, which are coupled such that knowledge
of one makes knowledge of the other uncertain. Therefore, there is a tradeoff between
time and frequency resolution, such that good time resolution comes at the expense
of frequency resolution, and vice versa. As the window shrinks to zero width, short
tones become clicks, with no discernable frequency. Clicks can be perfectly localized
in time, but the frequency is undefined. As the window gets large compared to the
signal length, it becomes impossible to localize a signal in time, but the frequency
can be easily determined. Time-frequency analyses balance the time-frequency reso-
lution tradeoff.
Frequency analyses are restricted to stationary signals. That is, the spectral
characteristics do not change over time. However, EEG signals change as a function of
state and task demands, which can vary over time. Time-frequency analyses are used
to study the non-stationary characteristics of psychophysiological signals. Rather than
analyzing signals in one domain—time or frequency—signal properties are localized to
points or pixels in a two-dimensional time-frequency plane.
LOGIC BEHIND EEG FREQUENCY ANALYSIS 31
4 Hz
8 Hz
16 Hz
2.13.2 Grounding
Ground electrodes are used for common mode rejection: parts of the signal shared
across all electrodes are removed from the recording. A good ground connection is crit-
ical to reducing the impact of external noise sources. Impedances for all electrodes are
compared to the ground electrode. Therefore, if the ground impedance is high, good
impedances will not be possible on any other electrodes. However, electrical interfer-
ence may persist even after ensuring good electrode impedances and secure grounding.
Therefore, data rejection and correction are often used to handle residual artifacts
during offline processing.
2.13.3 Artifact Rejection
Artifact removal may consist of rejecting entire segments of data, or individual
electrodes from the recording. While the basic strategy is to identify data corrupted by
artifact and exclude it from future analysis, it is not a trivial task. The determination of
whether a signal is sufficiently corrupted by noise to justify removal is time consuming,
subjective, and may vary by researcher. Moreover, significant training is required to
learn to correctly identify different sources of noise. Judgements must also be made as to
whether each artifact should be removed, even if it is correctly identified. In some cases,
a filter may be used to attenuate electronic noise. For example, a 60-Hz artifact from
power-lines may be removed from EEG recordings using a notch filter, which allows
frequencies lower and higher than the specified frequency-band to pass unaffected,
while significantly attenuating the artifact (e.g., between 55–65 Hz). However, filtering
may not be an optimal solution for all artifacts. For example, blink timing is related to
information processing (Stern et al., 1984), and blink rate is correlated to dopamine
release (Taylor et al., 1999), although recent work has called this into question (Dang
et al., 2017). However, many researchers remove data segments with eyeblinks due to
the large artifact they produce in frontal EEG channels. If the phenomenon of interest is
correlated to these physiological artifacts, it is possible to inadvertently remove valuable
data and miss the psychophysiological phenomenon.
Sometimes artifacts may be localized to a single bad channel. In this case, remove
only that channel from the recording and keep the remaining data. Many researchers
examine the standard deviation of the signal to automatically detect bad electrodes. If
it exceeds a threshold much larger than would be expected for a quality EEG signal, the
channel is marked bad. If a channel is marked bad and it is required for a planned EEG
analysis, the missing data can be interpolated. Interpolation uses information from the
surrounding electrodes to guess what the signal would have been if there were a good
electrode at that location. A popular choice is spherical spline interpolation (Perrin
et al., 1989).
34 KYLE J. CURHAM and JOHN J. B. ALLEN
2.13.4 Artifact Correction
If possible, EEG researchers want to retain as much signal as possible and thus avoid
removing noisy signals from data analyses. Several algorithms exist to attempt to
separate signal from noise, thereby sparing valuable data. The three most common
approaches are linear regression, independent component analysis (ICA), and principal
component analysis (PCA).
Linear regression is by far the simplest method to remove ocular artifacts. Ocular
artifacts typically have much higher amplitude than the EEG signals generated by neural
sources. For the purpose of regression, the tiny EEG signals are assumed to be noise rela-
tive to the ocular artifacts. Blinks are assumed to affect electrodes in a linear combination,
such that electrodes closer to ocular channels have a larger weight than electrodes farther
from ocular channels. Once the optimal weights are calculated, a weighted composite of
the ocular artifact is subtracted from each of the EEG electrodes (cf. Gratton et al., 1983),
yielding EEG that should be (mostly) free of the artifact. However, it requires the ocular
artifacts to be linearly dependent and normally distributed, which generally is not the case.
This approach is fast and simple, but sometimes inaccurate, which introduces another
source of residual artifact. Errors in the regression get “subtracted into” the recording.
ICA is a blind source separation technique, unmixing a multivariate signal into a set
of additive components (Delorme et al., 2010). Unlike linear regression, independent
components (ICs) are assumed to be non-Gaussian and statistically independent from
each other. However, ICA can only separate linearly mixed sources, and even when the
sources are not independent, ICA will return maximally independent components.
In many cases, a subset of the ICs will capture artifacts such as ocular and motor
components. Components are determined to be artifact or signal based on their simi-
larity with the topography and time course of known artifacts. This can be determined
based on experimenter ratings or algorithmically using spatial and temporal features
of eye blinks, vertical and horizontal eye movements, and discontinuities in the EEG
time series (Mognon et al., 2011). Artifactual components may also be identified auto-
matically based on a supervised machine learning approach trained from expert ratings
on a large corpus of EEG data (Winkler et al., 2014). ICA is more useful than linear re-
gression because it can extract more than just ocular components, also capturing loose
electrodes and muscle artifacts equally well. Once the bad components are identified,
the signal can be reconstructed without the artifacts.
Other blind source separation techniques similar to ICA can be used in a similar
manner. PCA is like ICA in that it separates the data into a set of linearly independent
components. However, these components are additionally assumed to be mutually or-
thogonal (i.e., uncorrelated with each other).
2.13.5 Referencing
Voltage is always measured between two points. It does not make sense to talk about the
potential at a particular EEG electrode without first defining a reference potential (i.e.,
LOGIC BEHIND EEG FREQUENCY ANALYSIS 35
an arbitrarily chosen “zero”). A common reference is usually chosen for all electrodes, to
which all potentials recorded at each electrode are measured. The reference needs to be
chosen carefully because the electrical activity under the reference site will be reflected
in the activity at every other electrode. If the reference is not neutral, artifacts will be
introduced due to the activity at the reference site. The reference electrodes should be
placed on a presumed electrically neutral area. In many cases, researchers choose an
electrode over the mastoid part of the temporal bone, or the left or right earlobe.
In reality, there are no electrically neutral points to choose as a reference. However,
it is possible to construct a virtual reference that is more likely to be electrically neutral
than any single EEG electrode. For example, choosing the left or right mastoid alone
results in a systematic decrease of EEG amplitude in the electrodes which are closer to
the reference site. The “linked” mastoids reference is obtained by using the average of
both left and right mastoid electrodes as the reference level. This resolves the left/right
asymmetry that would normally occur. Another popular choice is the average reference
when the number of electrodes is large (typically >32). The average reference is obtained
by subtracting the mean potential across all EEG electrodes from each electrode at each
time-point. Theoretically, if we could obtain measurements sampling equally around
a sphere containing the brain, then electric dipoles picked up by the electrodes would
average out to zero, and the average reference would be a truly neutral point (Bertrand
et al., 1985). However, this is a practical impossibility because the ventral surface is in-
accessible for electrode placement. Still, by averaging the electrical activity across the
entire scalp, the “responsibility” is distributed over all electrodes, rather than only one
or two of them.
However, prominent deflections at any given scalp location can differ dramatically
depending on the chosen reference scheme, affecting both the amplitude and latency
of the recorded signal. Not all researchers use the same reference, leading to problems
in the comparability between datasets. Fortunately, it is possible to convert from one
reference to another with a simple mathematical transformation. Note the potential
difference between two electrodes does not depend on the chosen reference. If your ana-
lysis is only concerned with comparisons between two electrodes, then the reference is
irrelevant. A simple analogy is measuring height above sea level. A change in sea level
does not change the shape of the landscape or the relative elevation between any two
points. Thus, if there is a change in sea level, we can readjust to the new reference level by
simply subtracting the change in elevation from every point. Similarly, we can transform
between different references by subtracting the potential at the chosen reference site.
Other transformations can be used to obtain a reference-free representation of EEG
signals. This is usually achieved using the scalp Laplacian, sometimes called the Current
Source Density (CSD). Technically, the scalp Laplacian relates current generators within
an electrical conductor to the second spatial derivative of the potential at each elec-
trode. Approximately, the scalp-Laplacian at a given electrode is obtained subtracting
the potentials of all neighboring sites weighted by their inverse distance. In other words,
the scalp Laplacian measures how extreme the potential at one electrode is compared
to the average of its neighboring electrodes. This intuitive description is complicated by
the fact that scalp electrode montages do not form a flat grid of evenly spaced electrodes.
36 KYLE J. CURHAM and JOHN J. B. ALLEN
Cz AR LM CSD
Open
Closed
Figure 2.12 Topography of alpha power under eyes open (top) and eyes closed (bottom)
conditions as a function of transformation (Cz, average (AR), or linked mastoid (LM) reference
or current source density (CSD) transformation) from a sample of over 2400 resting recordings.
Power values at each site represent natural-log transformed values; thus, negative numbers rep-
resent mean power values less than one. Each transformation is scaled independently, but within
each transformation, eyes open and closed data are plotted on the same scale. Only the CSD
transformation confines occipital alpha to occipital leads, whereas the other three montages show
reflected alpha at frontal regions, visible most clearly by a comparison of frontal leads under eyes
closed compared to eyes open recordings.
Figure modeled after Smith et al., 2017.
Instead, they conform to the nearly spherical geometry of the scalp. However, the
Laplacian can still be efficiently computed using spherical spline interpolation
(Delorme & Makeig, 2004). The scalp Laplacian has the added advantage of addressing
volume conduction. Electrical fields originating from any particular neuronal structure
will influence the electrical potential throughout the brain and surrounding physio-
logical tissue. As a result, EEG electrodes detect a mixture of activity from across the
whole brain, rather than directly under the recording site. Critically, the scalp Laplacian
minimizes the influence of this effect since the Laplacian is close to zero when a signal
is shared similarly across several electrodes (cf. Smith et al., 2017). Figure 2.12 shows a
comparison across referencing schemes .
2.14 Conclusion
A basic understanding of electricity and signal processing is vital to EEG data collection
and interpretation. Investigators need to report clearly how they acquired and analyzed
the data, and consumers need to be able to evaluate the methods and claims critically.
There are many steps to be taken in acquiring and processing EEG signals, from net ap-
plication and impedance checking, to grounding and referencing, to artifact detection
LOGIC BEHIND EEG FREQUENCY ANALYSIS 37
and correction, to data processing and analysis methods in the time and frequency
domains. This chapter provided a basic introduction to these concepts, but more in-
formation can be obtained from several guidelines papers (Picton et al., 2000; Keil
et al., 2014).
Glossary
Aliasing: When an analog signal is insufficiently sampled, the digitized signal may be indis-
tinguishable from samples from a lower-frequency signal.
Ampere: The SI unit of electric current. It is defined as one coulomb of charge per second.
Analog signal: A continuous time-varying signal that may vary in frequency, amplitude,
and phase.
Analog-to-digital converter: System that converts an analog signal into a digital signal.
Capacitance: Ratio of the change in an electric charge to the change in electric potential.
A property that quantified a materials ability to store electric charge.
Conductivity: Property that quantifies how strongly a material conducts electric current.
Conductor: A material that allows the flow of electric current.
Conjugate variables: Pairs of variables that cannot be precisely estimated simultaneously.
Knowledge of one variable necessitates uncertainty in the other variable. Time and fre-
quency are examples of conjugate variables.
Convolution: A mathematical operation that indicates the amount of overlap of one function
or kernel as it is shifted over another function.
Cutoff frequency: Frequency at which the signal power is reduced by half. The amount of at-
tenuation increases as you move farther from the cutoff.
Current: The directed motion of electrons or electric charge.
Digital signal: A discrete sequence of finite values that represent a signal.
Electric field: The force per unit charge at each point in space.
Electricity: Phenomenon that describes the behavior and movement of electric charge.
Electromotive force (EMF): The rate at which energy is drawn from a 1 A current source,
measured in volts.
Finite impulse response (FIR) filter: A filter whose impulse response settles to zero in fi-
nite time.
Free electrons: Any electron that is free to move under the influence of an electric or mag-
netic field.
Impedance: Measures the opposition to an electric current when a voltage is applied.
Impedance is equal to resistance in a DC circuit.
Impulse response: The output of a circuit when presented with a brief input signal.
Insulator: A material that does not allow the flow of electric current.
Ionization: The acquisition or loss of electrons in an atom or a molecule, resulting in a net
positive or negative charge.
Nyquist frequency: Half of the sampling rate of a digital signal. It is the highest frequency that
can be accurately sampled for a given sampling rate.
Nyquist sampling rate: In order to adequately digitize an analog signal, it should be sampled
at least twice the highest frequency of interest.
38 KYLE J. CURHAM and JOHN J. B. ALLEN
References
Bertrand, O., Perrin, F., & Pernier, J. (1985). A theoretical justification of the average reference in
topographic evoked potential studies. Electroencephalography and Clinical Neurophysiology,
62(6), 462–464.
Cohen, M. X. (2014). Analyzing neural time series data: theory and practice. MIT press.
Dang, L. C., Samanez-Larkin, G. R., Castrellon, J. J., Newhouse, P. A., Zald, D. H., Perkins, S.
F., & Ronald, L. (2017). Spontaneous eye blink rate (EBR) is uncorrelated with dopamine
D2 receptor availability and unmodulated by dopamine agonism in healthy adults. ENeuro,
4(5), 1–11.
Delorme, A. & Makeig, S. (2004). EEGLAB: An open source toolbox for analysis of single-
trial EEG dynamics including independent component analysis. Journal of Neuroscience
Methods, 134(1), 9–21. https://doi.org/10.1016/j.jneumeth.2003.10.009
Delorme, A., Sejnowski, T. J., & Makeig, S. (2010). Enhanced detection of artifacts in EEG
data using higher-order statistics and independent component analysis. NeuroImage, 34(4),
1443–1449. https://doi.org/10.1016/j.neuroimage.2006.11.004.Enhanced
Ferree, T. C., Luu, P., Russell, G. S., & Tucker, D. M. (2001). Scalp electrode impedance, infec-
tion risk, and EEG data quality. Clinical Neurophysiology, 112(3), 536–544.
Gratton, G., Coles, M. G., & Donchin, E. (1983). A new method for off-line removal of ocular
artifact. Electroencephalography and Clinical Neurophysiology, 55(4), 468–484.
Keil, A., Debener, S., Gratton, G., Junghöfer, M., Kappenman, E. S., Luck, S. J., Luu, P., Miller,
G. A., & Yee, C. M. (2014). Committee report: Publication guidelines and recommendations
for studies using electroencephalography and magnetoencephalography. Psychophysiology,
51(1), 1–21.
Mognon, A., Jovicich, J., Bruzzone, L., & Buiatti, M. (2011). ADJUST: An automatic EEG arti-
fact detector based on the joint use of spatial and temporal features. Psychophysiology, 48(2),
229–240. https://doi.org/10.1111/j.1469-8986.2010.01061.x
Perrin, F., Pernier, J., Bertrand, O., & Echallier, J. F. (1989). Spherical splines for scalp potential
and current density mapping. Electroencephalography and Clinical Neurophysiology, 72(2),
184–187. https://doi.org/10.1016/0013-4694(89)90180-6
Picton, T. W., Bentin, S., Berg, P., Donchin, E., Hillyard, S. A., Johnson, R., . . . Taylor, M. J.
(2000). Guidelines for using human event-related potentials to study cognition: recording
standards and publication criteria. Psychophysiology, 37(2), 127–152. http://www.ncbi.nlm.
nih.gov/pubmed/10731765
LOGIC BEHIND EEG FREQUENCY ANALYSIS 39
Smith, E. E., Reznik, S. J., Stewart, J. L., & Allen, J. J. (2017). Assessing and conceptualizing
frontal EEG asymmetry: An updated primer on recording, processing, analyzing, and
interpreting frontal alpha asymmetry. International Journal of Psychophysiology, 111, 98–114.
Stern, J. A., Boyer, D., & Schroeder, D. (1994). Blink rate: A possible measure of fatigue. Human
Factors, 36(2), 285–297. https://doi.org/10.1177/001872089403600209
Stern, J. A., Walrath, L. C., & Goldstein, R. (1984). The endogenous eyeblink. Psychophysiology,
21(1), 22–33.
Trujillo, L. T., and Allen, J. B. (2007). Theta EEG dynamics of the error-related negativity.
Clinical Neurophysiology, 118(3), 645–668
Taylor, J. R., Elsworth, J. D., Lawrence, M. S., Sladek, J. R. Jr, Roth, R. H., & Redmond, D. E. Jr.
(1999). Spontaneous blink rates correlate with dopamine levels in the caudate nucleus of
MPTP-treated monkeys. Experimental Neurology, 158(1), 214–220.
Winkler, I., Brandl, S., Horn, F., Waldburger, E., Allefeld, C., & Tangermann, M. (2014). Robust
artifactual independent component classification for BCI practitioners. Journal of Neural
Engineering [online], 11(3), 035013. https://doi.org/10.1088/1741-2560/11/3/035013
CHAPTER 3
Cyclical rhythms are a hallmark property of natural systems, from the movement of
the planets to the sleep and wake cycles of mammals, to neuronal membranes. Such
rhythmic variation, if it occurs repeatedly across time, is called an oscillation. The back-
and forth of tree branches on a windy day, ocean waves, and the rhythmic song of choir
frogs are further examples for natural oscillations. As such, the notion of oscillations
seems fairly straightforward: they occur in complex systems, in which many connected
units interact, and they reflect activity perceived as recurrent, or rhythmic. However, a
woman strolling on an ocean beach will soon realize that although waves break repeat-
edly and somewhat predictably, they also vary in terms of their exact timing, duration,
size, and shape. This seems to be different from the orbit of our planet earth around the
sun, or the vibrations of crystalline matter—rhythms that are regular and fixed enough
for us to measure time on their basis. These latter are typically referred to as “periodic
oscillations”, which are a sequence of recurring events (often taking the shape of a wave
when measured), each of which has identical, constant, duration. Although periodic
oscillations are well defined in the language of mathematics (e.g., the sine function),
the examples at the start of the chapter show that the concept of periodicity may not be
as clear-cut in natural systems. For example, the duration of solar years is not constant
when measured with precision, and even clock-setting crystal oscillators change their
rhythm with temperature. Thus, periodicity may be a continuous dimension ranging
from more periodic (the solar year) to less periodic (the tides) rhythms, rather than
representing a qualitative (yes/no) feature. In fact, various natural systems are likely to
FROM NEURAL OSCILLATIONS TO COGNITIVE PROCESSES 41
oscillate in different fashion, and the same system may display oscillatory behavior at
different levels of complexity when challenged in different ways.
Identifying how a complex system oscillates may go a long way towards its char-
acterization. As a consequence, the natural sciences have developed an impres-
sive toolbox for detecting, quantifying, and understanding oscillatory activity, and
to separate it from non-oscillatory phenomena. This chapter discusses how these
concepts are applied to fundamental problems in cognitive neuroscience, and ask
what the significance is of oscillatory activity for human behavior and cognition? It
also asks how we should define and conceptualize brain oscillations, and how they
can be measured and interpreted. We address these questions in the light of current
research, with a focus on electrophysiological recordings in human beings and ex-
perimental animals.
After over a decade of testing, in 1927 Hans Berger conducted the earliest in vivo
recordings of electric fields from the human brain using his new invention, the electro-
encephalograph (Berger, 1929). Upon visual inspection of these first brain wave (elec-
troencephalography, or EEG) recordings, Berger decided that the most salient feature
of the EEG was its change in frequency content when the participant performed a task
or was stimulated (often by being touched with a glass probe). Most prominently, Berger
identified “first-order waves” oscillating at a rate of 10 cycles/second (i.e., 10 Hertz
(Hz), today referred to as alpha waves) and faster “second-order waves” oscillating at
about 20–30 Hz (today referred to as beta waves). Even without the ability to average
across trials, and in the absence of digitally supported spectral analysis, Berger and his
colleagues (i.e., his family; his wife Ursula von Bülow was his technical assistant and
frequent test participant, along with himself, and his son, Klaus) made a striking ob-
servation: whenever a person was at rest, for example, with their eyes closed, first-order
waves characterized by large-amplitude regular activity at 10 Hz dominated the EEG.
However, touching the person with a glass probe, asking them to open their eyes, or
just addressing them and asking a question led to a dramatic blocking of these alpha
waves, which were replaced by faster and lower-amplitude EEG readings. This effect,
referred to as “alpha blocking” has since been replicated thousands of times, and is still
used today in clinical usage of EEG in neurology and psychiatry to assess the status
of global brain electric states (Başar & Güntekin, 2012). Furthermore, this observa-
tion opened an avenue for experimental studies of EEG-behavior relations, in which
participants are asked to respond to experimenter-defined task demands, and changes
in the EEG are measured as a dependent variable (Klimesch et al., 2005). In addition
to testing hypotheses about brain function, these variables possess practical value for
42 ANDREAS KEIL and NINA THIGPEN
testing hypotheses about behavioral and cognitive processes, their time course, and
about specific characteristics of the participant or the environment, of interest in clinical
and translational studies.
1.2
Frequency (Hz)
30 1.1
20 1
0.9
10
0.8
0 1 2 3 4 5 6 7 8
Time (seconds)
Figure 3.2 Oscillatory changes during a working memory task. Time-frequency decompos-
ition of the EEG activity recorded at sensor Pz over a 9-s trial in which participants memorized
the orientation of two gratings. They then viewed a mask, a task array, and finally were asked to
indicate if the orientation of an item in the task array matched an item in the memory set. Each
task element prompts specific time-frequency dynamics. For example, alpha-band activity (8–
12Hz) is highest during the fixation period before the start of the trial, and during the retention
interval when viewing the mask.
44 ANDREAS KEIL and NINA THIGPEN
A useful approach for a classification of the brain’s oscillatory activity is the widely-
adopted nomenclature first introduced by Robert Galambos (1992). Considering brain
oscillation across several species, Galambos distinguished:
Investigators new to the field are often curious about how different types of oscillations
should be analyzed and interpreted in terms of hypotheses regarding a specific behav-
ioral, cognitive, or neural process. To address these questions, it may be helpful to con-
sider the different types of dependent variables that may be used in studies of oscillatory
brain activity during cognitive processing.
The mathematical foundations of extracting power and phase from time domain neural
signals as described earlier are straightforward and yield estimates of power at a given
frequency, in a given time period, at a specific sensor. What is less clear is the functional
and neurophysiological interpretation of differences in power, especially in population-
level signals, such as local field potentials (LFPs), intra-and extracranial EEG, and MEG.
These signals reflect the synchronous synaptic (dominantly post-synaptic) currents of a
large number (at least several tens of thousands in the case of EEG/MEG) of neurons
which need to be aligned favorably (in parallel) to create measurable extracranial elec-
tric fields (Olejniczak, 2006). Thus, a difference between two experimental conditions,
say in EEG alpha-band power, may reflect differences in the number of neurons
engaged in the respective cognitive process, but it may also represent a combination of
differences in active population size, amount of dendritic engagement at each neuron,
orientation of neurons involved, and temporal synchronization of the post-synaptic
events. Neural mass power in a given frequency band is therefore difficult to map dir-
ectly onto a physiological or cognitive concept, emphasizing the need for research that
links measures of oscillatory brain activity to robust cognitive or behavioral processes,
across different levels of observation (i.e., from single neurons to large populations).
In addition to measuring spectral power, researchers may use approaches that capit-
alize on additional spectral information, often the phase of the signal. Many algorithms
FROM NEURAL OSCILLATIONS TO COGNITIVE PROCESSES 45
exist for quantifying the similarity of the phase across different observations, such as
experimental trials, time points, or channels. The resulting metrics are referred to as
indices of phase-locking or phase coherency. These variables are often used to estimate
the strength of oscillatory interactions across recordings sites (e.g., inter-site phase
locking, often considered a proxy of connectivity), the stability of the temporal pro-
file of oscillations across repeated trials (e.g., phase-locking value), or the interactions
between different frequencies within or across recordings sites (e.g., phase-amplitude
coupling (Chapter 20). Thus, a wide range of hypotheses and, increasingly, formal com-
putational models of neurophysiological processes can be tested using these different
variables, each of which may reflect entirely different facets of oscillatory brain activity.
Later in this chapter we present examples for using these markers across a wide range of
research questions in cognitive neuroscience.
In the many decades since Berger’s and many others’ discoveries, brain oscillations have
been traditionally divided into frequency bands in the 1–4 Hz range (delta), 4–8 Hz
(theta), 8–12 Hz (alpha), 12–30 Hz (beta), and >30 Hz (gamma), with some variability in
the demarcation of these bands. These frequency bands have been consistently observed
at the level of spike trains, local field potentials, and EEG/MEG, making them apparent
in multivariate analyses and meta-analyses across studies and even species (Lopes da
Silva, 1991). The peak frequency in each of these bands is similar across bats, mice, rats,
cats, dogs, horses, dolphins, macaques, and humans, despite the range of brain size and
axon length across these species (Buzsáki et al., 2013). Given this phylogenetic stability, it
has been suggested that the traditional frequency bands (and perhaps their logarithmic
distance from one another) represent fundamental mechanisms underlying specific
neurocomputations and behaviors (Steriade et al., 1990). Klimesch (1999) highlights
the variability and task-dependency of these frequencies, however, with examples for
gamma-like oscillations in the traditional beta band, and overlap between alpha and
beta band oscillations, along with pronounced inter-individual differences. Similarly,
limitations of assigning specific functional roles to oscillations in one of the traditional
bands have been made apparent by phenomena in which oscillations continuously
transition from one traditional band into another during the same task (Donner &
Siegel, 2011).
Some of the stated confusion about these observations may be reflective of
researchers’ desire to equate oscillatory activity in one frequency range with a cat-
egory of cognition of behavior, defined in the language of cognitive psychology, such
46 ANDREAS KEIL and NINA THIGPEN
neurophysiological and behavioral processes, thus providing a second way for action
potentials to temporally organize downstream neural—and ultimately behavioral—
processes (Gütig, 2014). Such a temporal organization may be propagated in space if
neurons act as coupled oscillators, facilitating specific phase relationships among
the units of a functional network of connected neurons. Many such “coupled oscil-
lator” models of neural oscillations exist, aiming to explain how temporal signatures
are shared among populations of neurons (Moon et al., 2015; Naze et al., 2015). Recent
empirical work as well as work in computational modeling has converged to suggest
that such coupling involves not only action potentials but a range of mechanisms that
also involve subthreshold and synaptic events, along with potential changes at glia cells
(Buzsáki et al., 2012).
It is now well established that spike timing, that is, the timing of action potentials,
is related to synaptic oscillations measured by local field potentials, likely both driving
and being driven by these synaptic fields. This research has also demonstrated that
subthreshold oscillations at membranes and post- synaptic potentials constrain
the spiking rate of individual neurons and may also play a role in coordinating the
firing among different neurons (Mazzoni et al., 2010). This is important because it
demonstrates the interplay of all- or-
nothing neuronal communication and post-
synaptic events, opening avenues for research that identifies the convergent versus
complementary roles of oscillations at different levels, within and across neurons, in be-
havioral and cognitive processes.
Research in human participants has increasingly made use of intracranial data
obtained from patients under pre-surgical evaluation for neurological disorders, with
implanted sensor arrays (Parvizi & Kastner, 2018). In these studies, one striking dis-
crepancy arises between levels of observation regarding the frequency content of
intracranial recordings vis-à-vis extracranial recordings. Intracranial recordings, for
example, electrocorticogram (ECoG) data, tend to contain robust high-frequency
oscillations (Osipova et al., 2008), which are small and less reliably observed in extra-
cranial recordings such as EEG or MEG (Yuval-Greenberg et al., 2008). This salient
difference between intracranial and scalp recordings may reflect the orders of magni-
tude difference in the summation of electrical potentials that comprise each signal. For
example, the local field potential recorded from an intra-cranially sensor embedded
in cortical tissue is thought to measure the extracellular electrical fluctuations of a few
hundreds to thousands of neurons, while EEG data is thought to require synchrony
across a few millimeters of cortex before any fluctuations are observed on the scalp
(Nunez & Srinivasan, 2006). Based on simulations and in vitro studies, some authors
suggest that postsynaptic changes at dendritic trees of 40,000 to 100,000 pyramidal cells
are required to cause EEG changes in the 1–2 microvolt range—if the dendrites of these
neurons are oriented in parallel, generating an open electric field, required for extra-
cranial measurement. By contrast, ECoG and LFP data are thought to reflect activity
in differentially oriented cells, including contributions from interneurons, pyramidal
cells, and glial cells (Buzsáki et al., 2012). Thus, EEG oscillations reflect activity in a more
specific subset of neurons than ECoG and LFP but integrated over a wider distribution
of space.
48 ANDREAS KEIL and NINA THIGPEN
From its inception, the study of brain oscillations and their role in cognition has been
multi-disciplinary, with a strong emphasis not only on physiological measurements,
FROM NEURAL OSCILLATIONS TO COGNITIVE PROCESSES 49
3.7.4 Oscillatory Hierarchies
Many prominent models of the functional role of oscillatory brain activity have
considered interactions between oscillations at different temporal rates. A classical
model of working memory proposed that hippocampal neurons encode memory con-
tent using timed gamma-frequency bursts, which in turn receive their temporal struc-
ture from the phase of low-frequency oscillations in the theta band (Lisman & Jensen,
2013). This account has received substantial support as well as extension and refinement
based on computational and empirical studies.
In another widely recognized account, systematic relations across different
frequencies are used by the brain to organize the interplay between behavioral and cog-
nitive processes—the concept of “active sensing” emphasizes the active aspect of per-
ceptual sampling (Schroeder et al., 2010). Examples for such active sensing include
whisking and sniffing in rodents and saccadic sampling of visual scenes in primates. In
all these cases, perception is tightly linked to the motor activity that causes the stimu-
lation of sensory receptors as it initiates inflow of information through the sensory
pathways. Each afferent volley of information meets ongoing brain states hypothesized
to be partly determined by past experience and current goals, as well as the physiological
properties of the tissue. In these hypothetical hierarchies, neural oscillations at lower
frequencies (such as alpha or theta, 4–8 Hz as defined in adults) may provide temporal
structure for higher-frequency phenomena, for example, beta-(13–30 Hz) or gamma-
band (>30 Hz) oscillations. In a similar vein, studies in macaque monkey visual cortex
have suggested that high-amplitude gamma oscillations more likely occur during spe-
cific phases of the alpha cycle (Bonnefond & Jensen, 2015). These findings have been
taken to indicate that alpha phase represents the excitability of the neural tissue, pos-
sibly aligning excitability of sensory cortices with other events such as expectancy or
memory-driven signals (Kizuk & Mathewson, 2016; Mathewson et al., 2011).
columns, areas; Buzsáki & Draguhn, 2004). In earlier theories of oscillatory activity,
temporal synchrony was seen as the crucial feature for coding information that belongs
together but is distributed between different neurons or populations of neurons (von
der Malsburg & Buhmann, 1992). This notion is readily aligned with the Hebbian per-
spective discussed earlier, adding to the appeal it has held for many decades. At the
microscopic and mesoscopic level, synchronous firing in local circuits, especially in the
gamma frequency range, has been proposed as a mechanism for cognitive processes
as diverse as gestalt perception, predictive coding, motor preparation, and selective
attention (Bauer et al., 2014; Keil et al., 2001).
Extending the concept of synchrony to long-range communication and integration, re-
cent work focuses on the interaction between local and inter-area (long-range) synchrony,
increasingly on oscillatory interactions across different frequencies (Chapter 20). For in-
stance, oscillatory activity in the theta range (4–8 Hz) has been traditionally hypothesized
as a potential mechanism for transmitting information between distant brain areas
(Klimesch et al., 2005; Sarnthein et al., 1998). Empirical evidence in rodents, cats, and non-
human primates supports this notion, with theta oscillations characterizing long-range
signaling in large-scale networks, including the hippocampus, the amygdaloid, complex,
and extensive cortical areas, during tasks that involve learning and memory (Popescu
et al., 2009). Similar findings exist for alpha-band oscillations, discussed now as carriers
of prediction signals from higher-order to sensory cortices. Importantly, the concept of
synchrony in the context of these inter-area communications has been extended beyond
zero-lag co-activation at different sites, and has instead identified inter-area interactions
by the extent to which the phase lag between two recording sites is consistent over time,
that is, so-called inter-site phase locking (Brovelli et al., 2004).
Together, local-and inter-area oscillations hold great promise for informing and
constraining network-based models of human cognition. Because of their compati-
bility with findings obtained in the animal model, and the ability to relate neural activity
across different levels of observation, indices of oscillatory brain activity also represent
powerful dependent variables in cognitive neuroscience studies. The following example
illustrate these benefits by selectively reviewing examples for such studies and address
processes of perception, attention, learning, and memory.
of a visual cue, or the direction of attention toward a salient visual stimulus (Adrian &
Matthews, 1934; Klimesch, 1999; Pfurtscheller et al., 1996). Low levels of alpha power to-
gether with alpha phase also predict heightened performance in near-threshold detec-
tion tasks, where participants report weak sensory events (Mathewson et al., 2011; Weisz
et al., 2014). This evidence supports a long-held notion that scalp-recorded alpha activity
reflects a cortical state in which little sensory information is received (Pfurtscheller,
1992), mediated perhaps via a thalamic gating mechanism (Steriade et al., 1990).
Other findings are consistent with this notion.: Presenting a cue that directs attention
to one hemifield prompts reduction of alpha power in the contralateral hemisphere
(Foxe & Snyder, 2011). By contrast, alpha power over ipsilateral sensors (representing
the ignored hemifield) remains stable (Thut et al., 2006), or increases (Kelly et al., 2006),
taken to suggest suppression of irrelevant information. This interpretation is consistent
with research examining attention to target items embedded in a rapid serial visual pres-
entation stream of distractors. Accurate identification of rapidly presented targets amid
non-targets is associated with higher pre-target alpha power (Petro & Keil, 2015). Thus,
high alpha power may index behavioral states that benefit performance by reducing the
processing of irrelevant sensory information (Klimesch et al., 2006). Importantly, these
attention-related changes in alpha power are associated with faster and more accurate
responses for the task, suggesting that they possess a functional role in the active selec-
tion of target stimulus features (Foxe & Snyder, 2011).
A further role of alpha oscillations during attention tasks is under consideration in
the context of predictive coding. This has become a current topic in cognitive neuro-
science. Increased alpha power and heightened inter-site phase-locking during target
anticipation are proposed as a mechanism that optimizes the temporal organization of
sensory processing and thus facilitates the sensory analysis of expected, task-relevant,
visual stimuli (Samaha et al., 2015). Several of the different proposed functions of alpha-
band oscillations during perception and attention are not mutually exclusive and have
given rise to the idea that the alpha frequency is used by a wide range of neural processes,
in the service of different cognitive and behavioral goals (Chapter 10).
Work in the animal model abundantly demonstrates that associative learning induces
Hebbian plasticity in widely distributed brain networks, both sub-cortical and cor-
tical (Pape & Paré , 2010). These memory formation processes can be conceptualized as
changes in the neural communication between nearby and distant brain loci, which dy-
namically sculpt neural signaling pathways at multiple levels of analysis. Recent findings
suggest that neuronal oscillations are ideally suited to support the flexible formation
of such network-level changes in neural architecture (Popescu et al., 2009), including
those that support the acquisition and extinction of fear memories (Paré et al., 2002).
In addition, computational models using simulated agents in artificial evolutionary
environments show that network oscillations provide a fitness advantage, insofar as
54 ANDREAS KEIL and NINA THIGPEN
they help to promote rapid switches in perception and attention (Heerebout & Phaf,
2010). Oscillatory signals may therefore play an important role in the critical changes
that occur when a previously innocuous stimulus acquires relevance for controlling
behavior (Headley & Weinberger, 2011). At the level of neuronal populations, the as-
sociative principles that guide the formation of a newly acquired memory (e.g., the
association between light and electric shock during classical fear conditioning) must ef-
fectively coordinate activity between neural representations of the conditioned stimuli
(the light) and the systems that code biological value of the unconditioned event (the
shock). Although conditioning-induced changes in neuroarchitecture and function
occur on multiple spatio-temporal scales (Maren & Quirk, 2004), ranging from in-
dividual neurons to cortical sheets, and from minutes to days, the notion of a cell as-
sembly (Hebb, 1949) provides the necessary conceptual framework for bridging across
these multi-scale phenomena. Oscillatory synchronization between distributed neur-
onal assemblies in specific frequency bandwidths appears to represent a highly plaus-
ible substrate for synaptic plasticity transfer, that is, storing newly acquired memories
as changes in synaptic weights (Paré et al., 2002). In particular, increased large-scale
synchrony between subcortical and cortical networks may be crucial in producing the
experience-dependent changes in the representation of fear-conditioned cues.
Another example is provided by a series of studies conducted by Walter Freeman and
his colleagues (reviewed in Skarda & Freeman, 1987) in the olfactory system of the rabbit.
The basic paradigm involves the placement of an 8 × 8 electrode grid onto the olfactory
bulb in order to record electrocorticography (ECoG) signals associated with different
odorant stimuli. To begin with, each odorant elicits a pattern of wave activity that appears
as aperiodic noise with variability across trials. However, after the animal learns to asso-
ciate an odor with a motivationally relevant outcome, for example, the delivery of food or
aversive tactile stimulation, neural activity at the olfactory bulb undergoes a state tran-
sition whereby the odor comes to elicit a discriminant spatiotemporal pattern of ampli-
tude across the electrode grid array. The learning-related establishment of a new global
response pattern then acts to enslave the output of individual neurons (Freeman, 1994).
Freeman interprets such selective changes in spatiotemporal amplitude as embodying
the affective meaning of a stimulus for the animal, with meaning changing in accordance
with the momentary relevance of a stimulus (e.g., an animal responds differently to an
odorant associated with food once it is fed to satiety). A mechanism for stimulus selective
changes in the output of olfactory neurons is provided by alterations of synaptic effi-
ciency and neuropil structure, guided by Hebbian principles of association.
5
.4 cd/m2
–5
μV2
5 .7
4.9 cd/m2
–5 .2
5
70.7 cd/m2
–5
–200 0 200 400 600
Time (ms)
Figure 3.3 Effects of stimulus intensity on inter-trial phase locking of driven oscillations.
Participants viewed sinusoidal gratings flickering at 15 Hz. Gratings varied in luminance, with
40 low-luminance trials (.4 cd/m2), 40 medium-luminance trials (4.9 cd/m2), and 40 high-
luminance trials (70.7 cd/m2). Each colored line represents one single trial from sensor Oz. Inter-
trial phase locking increases with increasing luminance, whereas the overall magnitude of the
signal within the single trials does not change. The increased phase-locking is associated with
increases in 15-Hz power of the trial-averaged signal, as depicted in the topographies shown on
the right.
Data from Thigpen et al., 2018.
the driving stimulus across trials (Moratti et al., 2007). Figure 3.3 shows an example
of this phenomenon as it interacts with stimulus intensity, demonstrating increasing
phase alignment of single trial EEG traces with increasing stimulus contrast. The de-
bate of entrainment versus superposition perspectives is ongoing, and ssVEPs (be-
cause of their known frequency and pronounced signal) are well suited for examining
competing hypotheses regarding the interaction of ongoing oscillations and sensory
events.
This chapter aimed to illustrate how studies of oscillatory brain activity provide rich
opportunity for testing hypotheses relating brain function to cognitive processes.
Measuring oscillatory activity during cognitive task performance opens avenues into
developing and testing neuromechanistic accounts of some of the most central building
blocks of human behavior and experience. In conclusion, we discuss challenges for this
field moving forward, and provide key concepts towards an integrative framework of
oscillatory brain activity in the study of cognition.
FROM NEURAL OSCILLATIONS TO COGNITIVE PROCESSES 57
the formation of networks, local and distributed. The high signal-to-noise ratio of many
frequency-domain measures and the resulting potential for trial-by-trial analyses may
be leveraged to quantify neural changes that occur over the course of an experimental
session (McTeague et al., 2015). Such an approach complements trial averaging, which
is still the dominant approach in cognitive neuroscience, and sheds light on adaptation
dynamics and plasticity in response to the experimental situation, processes of high
ecological validity and with substantial clinical relevance.
to offer alternative models of inter-scale nonlinear interactions, which may provide crit-
ical predictions to be tested in empirical work (Neymotin et al., 2013).
References
Adesnik, H. (2018). Cell type-specific optogenetic dissection of brain rhythms. Trends in
Neurosciences, 41(3), 122–124. https://doi.org/10.1016/j.tins.2018.01.001
Adrian, E. D. & Matthews, B. H. (1934). The Berger rhythm: Potential changes from the oc-
cipital lobes in man. Brain, 57(4), 355–385.
Andersen, N., Krauth, N., & Nabavi, S. (2017). Hebbian plasticity in vivo: Relevance
and induction. Current Opinion in Neurobiology, 45, 188–192. https://doi.org/10.1016/
j.conb.2017.06.001
Awh, E., Belopolsky, A. V., & Theeuwes, J. (2012). Top-down versus bottom-up attentional con-
trol: A failed theoretical dichotomy. Trends in Cognitive Sciences, 16(8), 437–443. https://doi.
org/10.1016/J.Tics.2012.06.010
Baria, A. T., Maniscalco, B., & He, B. J. (2017). Initial-state-dependent, robust, transient
neural dynamics encode conscious visual perception. PLOS Computational Biology, 13(11),
e1005806. https://doi.org/10.1371/journal.pcbi.1005806
Başar, E. & Güntekin, B. (2012). A short review of alpha activity in cognitive processes and
in cognitive impairment. International Journal of Psychophysiology, 86(1), 25–38. https://doi.
org/10.1016/j.ijpsycho.2012.07.001
60 ANDREAS KEIL and NINA THIGPEN
Bauer, M., Stenner, M.-P., Friston, K. J., & Dolan, R. J. (2014). Attentional modulation of alpha/
beta and gamma oscillations reflect functionally distinct processes. Journal of Neuroscience,
34(48), 16117–16125. https://doi.org/10.1523/JNEUROSCI.3474-13.2014
Berger, H. (1929). Über das Elektrenkephalogramm des Menschen. Archiv Für Psychiatrie Und
Nervenkrankheiten, 87, 527–570.
Bernstein, J. (1868). Ueber den zeitlichen Verlauf der negativen Schwankung des Nervenstroms.
Archiv für die gesamte Physiologie des Menschen und der Tiere, 1(1), 173–207. https://doi.org/
10.1007/BF01640316
Bliss, T. V. P. & Lømo, T. (1973). Long-lasting potentiation of synaptic transmission in the den-
tate area of the anaesthetized rabbit following stimulation of the perforant path. The Journal
of Physiology, 232(2), 331–356. https://doi.org/10.1113/jphysiol.1973.sp010273
Bonnefond, M. & Jensen, O. (2015). Gamma activity coupled to alpha phase as a mechanism
for top-down controlled gating. PLOS ONE, 10(6), e0128667. https://doi.org/10.1371/journal.
pone.0128667
Boynton, G. M. (2011). Spikes, BOLD, attention, and awareness: A comparison of electro-
physiological and fMRI signals in V1. Journal of Vision, 11(5), 12–12. https://doi.org/10.1167/
11.5.12
Breakspear, M. (2017). Dynamic models of large-scale brain activity. Nature Neuroscience,
20(3), 340–352. https://doi.org/10.1038/nn.4497
Brovelli, A., Ding, M., Ledberg, A., Chen, Y., Nakamura, R., & Bressler, S. L. (2004). Beta
oscillations in a large-scale sensorimotor cortical network: directional influences revealed
by Granger causality. Proceedings of the National Academy of Sciences of the United States of
America, 101(26), 9849–9854. (15210971).
Buzsáki, G., Anastassiou, C. A., & Koch, C. (2012). The origin of extracellular fields and
currents—EEG, ECoG, LFP and spikes. Nature Reviews Neuroscience, 13(6), 407–420.
https://doi.org/10.1038/nrn3241
Buzsáki, G. & Draguhn, A. (2004). Neuronal oscillations in cortical networks. Science,
304(5679), 1926–1929. https://doi.org/10.1126/science.1099745
Buzsáki, G., Logothetis, N., & Singer, W. (2013). Scaling brain size, keeping timing: Evolutionary
preservation of brain rhythms. Neuron, 80(3), 751–764. https://doi.org/10.1016/j.neu
ron.2013.10.002
Capilla, A., Pazo-Alvarez, P., Darriba, A., Campo, P., & Gross, J. (2011). Steady-state visual
evoked potentials can be explained by temporal superposition of transient event-related
responses. PLoS One, 6(1), e14543. https://doi.org/10.1371/journal.pone.0014543
Cohen, M. X. (2014). Analyzing neural time series data: Theory and practice. MIT Press.
Donner, T. H. & Siegel, M. (2011). A framework for local cortical oscillation patterns. Trends in
Cognitive Sciences, 15(5), 191–199. https://doi.org/10.1016/j.tics.2011.03.007
Elbert, T., Ray, W. J., Kowalik, Z. J., Skinner, J. E., Graf, K. E., & Birbaumer, N. (1994). Chaos
and physiology: Deterministic chaos in excitable cell assemblies. Physiology Review,
74(1), 1–47.
Foxe, J. J. & Snyder, A. C. (2011). The role of alpha-band brain oscillations as a sensory
suppression mechanism during selective attention. Frontiers in Psychology, 2, 154. https://
doi.org/10.3389/fpsyg.2011.00154
Freeman, W. J. (1991). The physiology of perception. Scientific American, 264, 78–87.
Freeman, W. J. (1994). Role of chaotic dynamics in neural plasticity. Progress in Brain Research,
102, 319–333.
FROM NEURAL OSCILLATIONS TO COGNITIVE PROCESSES 61
Fuchs, A., Jirsa, V. K., & Kelso, J. A. (2000). Theory of the relation between human brain ac-
tivity (MEG) and hand movements. Neuroimage, 11(5 Pt 1), 359–369. https://doi.org/10.1006/
nimg.1999.0532
Galambos, R. (1992). A comparison of certain gamma-band (40 Hz) brain rhythms in cat
and man. In E. Basar & T. Bullock (Eds.), Induced rhythms in the brain (pp. 103–122).
Springer.
Giabbiconi, C.-M., Jurilj, V., Gruber, T., & Vocks, S. (2016). Steady-state visually evoked poten-
tial correlates of human body perception. Experimental Brain Research, 234(11),1–11. https://
doi.org/10.1007/s00221-016-4711-8
Gütig, R. (2014). To spike, or when to spike? Current Opinion in Neurobiology, 25, 134–139.
https://doi.org/10.1016/j.conb.2014.01.004
Hadjipapas, A., Lowet, E., Roberts, M. J., Peter, A., & De Weerd, P. (2015). Parametric variation
of gamma frequency and power with luminance contrast: A comparative study of human
MEG and monkey LFP and spike responses. NeuroImage, 112, 327–340. https://doi.org/
10.1016/j.neuroimage.2015.02.062
Haken, H., Kelso, J. A., & Bunz, H. (1985). A theoretical model of phase transitions in human
hand movements. Biological Cybernetics, 51(5), 347–356.
Harris, K. P. & Littleton, J. T. (2015). Transmission, development, and plasticity of synapses.
Genetics, 201(2), 345–375. https://doi.org/10.1534/genetics.115.176529
Headley, D. B. & Weinberger, N. M. (2011). Gamma-band activation predicts both associative
memory and cortical plasticity. Journal of Neuroscience, 31(36), 12748–12758. https://doi.org/
10.1523/JNEUROSCI.2528-11.2011
Hebb, D. (1949). The organization of behavior: A neuropsychological theory. Wiley.
Heerebout, B. T. & Phaf, R. H. (2010). Good vibrations switch attention: An affective function
for network oscillations in evolutionary simulations. Cognitive, Affective, & Behavioral
Neuroscience, 10(2), 217–229. https://doi.org/10.3758/CABN.10.2.217
Henley, J. M. & Wilkinson, K. A. (2016). Synaptic AMPA receptor composition in develop-
ment, plasticity and disease. Nature Reviews Neuroscience, 17(6), 337–350. https://doi.org/
10.1038/nrn.2016.37
Herrmann, C. S., Strüber, D., Helfrich, R. F., & Engel, A. K. (2016). EEG oscillations: From
correlation to causality. International Journal of Psychophysiology, 103, 12–21. https://doi.org/
10.1016/j.ijpsycho.2015.02.003
Keil, A., Gruber, T., & Muller, M. M. (2001). Functional correlates of macroscopic high-
frequency brain activity in the human visual system. Neuroscience and Biobehavioral
Reviews, 25(6), 527–534.
Keitel, C., Keitel, A., Benwell, C. S. Y., Daube, C., Thut, G., & Gross, J. (2019). Stimulus-driven
brain rhythms within the alpha band: The attentional-modulation conundrum. Journal of
Neuroscience, 39(16), 3119–3129. https://doi.org/10.1523/JNEUROSCI.1633-18.2019
Kelly, S. P., Lalor, E. C., Reilly, R. B., & Foxe, J. J. (2006). Increases in alpha oscillatory power
reflect an active retinotopic mechanism for distracter suppression during sustained visuo-
spatial attention. Journal of Neurophysiology, 95(6), 3844–3851. https://doi.org/10.1152/
jn.01234.2005
Kim, Y.- J., Grabowecky, M., Paller, K. A., & Suzuki, S. (2010). Differential roles of
frequency-following and frequency-doubling visual responses revealed by evoked neural
harmonics. Journal of Cognitive Neuroscience, 23(8), 1875–1886. https://doi.org/10.1162/
jocn.2010.21536
62 ANDREAS KEIL and NINA THIGPEN
Kizuk, S. A. D. & Mathewson, K. E. (2016). Power and phase of alpha oscillations reveal an
interaction between spatial and temporal visual attention. Journal of Cognitive Neuroscience,
29(3), 480–494. https://doi.org/10.1162/jocn_a_01058
Klimesch, W. (1999). EEG alpha and theta oscillations reflect cognitive and memory perform-
ance: A review and analysis. Brain Research Reviews, 29(2), 169–195. https://doi.org/10.1016/
S0165-0173(98)00056-3
Klimesch, W., Sauseng, P., & Hanslmayr, S. (2006). EEG alpha oscillations: The inhibition-
timing hypothesis. Brain Research Reviews, 53(1), 63–88.
Klimesch, W., Schack, B., & Sauseng, P. (2005). The functional significance of theta and upper
alpha oscillations. Experimental Psychology, 52(2), 99–108.
Lisman, J. E., & Jensen, O. (2013). The theta-gamma neural code. Neuron, 77(6), 1002–1016.
https://doi.org/10.1016/j.neuron.2013.03.007
Lopes da Silva, F. (1991). Neural mechanisms underlying brain waves: from neural membranes
to networks. Electroencephalography and Clinical Neurophysiology, 79(2), 81–93. https://doi.
org/10.1016/0013-4694(91)90044-5
Maren, S. & Quirk, G. J. (2004). Neuronal signalling of fear memory. Nature Reviews
Neuroscience, 5(11), 844–852. https://doi.org/10.1038/nrn1535
Marrazzi, A. S. & Lorente de No, R. (1944). Interaction of neighboring fibres in myelinated
nerve. Journal of Neurophysiology, 7(2), 83–101. https://doi.org/10.1152/jn.1944.7.2.83
Mathewson, K. E., Lleras, A., Beck, D. M., Fabiani, M., Ro, T., & Gratton, G. (2011). Pulsed out
of awareness: EEG alpha oscillations represent a pulsed-inhibition of ongoing cortical pro-
cessing. Frontiers in Psychology, 2, 99. https://doi.org/10.3389/fpsyg.2011.00099
Mazzoni, A., Whittingstall, K., Brunel, N., Logothetis, N. K., & Panzeri, S. (2010).
Understanding the relationships between spike rate and delta/gamma frequency bands of
LFPs and EEGs using a local cortical network model. NeuroImage, 52(3), 956–972. https://
doi.org/10.1016/j.neuroimage.2009.12.040
McTeague, L. M., Gruss, L. F., & Keil, A. (2015). Aversive learning shapes neuronal orientation
tuning in human visual cortex. Nature Communications, 6, 7823. https://doi.org/10.1038/nco
mms8823
Moon, J.-Y., Lee, U., Blain-Moraes, S., & Mashour, G. A. (2015). General relationship of
global topology, local dynamics, and directionality in large-scale brain networks. PLoS
Computational Biology, 11(4), e1004225. https://doi.org/10.1371/journal.pcbi.1004225
Moratti, S., Clementz, B. A., Gao, Y., Ortiz, T., & Keil, A. (2007). Neural mechanisms of evoked
oscillations: stability and interaction with transient events. Human Brain Mapping, 28(12),
1318–1333. https://doi.org/10.1002/hbm.20342
Morone, F., Roth, K., Min, B., Stanley, H. E., & Makse, H. A. (2017). Model of brain activa-
tion predicts the neural collective influence map of the brain. Proceedings of the National
Academy of Sciences of the United States of America, 114(15), 3849–3854. https://doi.org/
10.1073/pnas.1620808114
Muller, L., Chavane, F., Reynolds, J., & Sejnowski, T. J. (2018). Cortical travelling
waves: mechanisms and computational principles. Nature Reviews Neuroscience, 19(5), 255–
268. https://doi.org/10.1038/nrn.2018.20
Müller, M. M., Teder, W., & Hillyard, S. A. (1997). Magnetoencephalographic recording of
steady-state visual evoked cortical activity. Brain Topography, 9(3), 163–168.
Nadel, L. & Maurer, A. P. (2018). Recalling Lashley and reconsolidating Hebb. Hippocampus,
0( 0), 1–18. https://doi.org/10.1002/hipo.23027
FROM NEURAL OSCILLATIONS TO COGNITIVE PROCESSES 63
Nanou, E. & Catterall, W. A. (2018). Calcium channels, synaptic plasticity, and neuropsychi-
atric disease. Neuron, 98(3), 466–481. https://doi.org/10.1016/j.neuron.2018.03.017
Naze, S., Bernard, C., & Jirsa, V. (2015). Computational modeling of seizure dynamics using
coupled neuronal networks: Factors shaping epileptiform activity. PLoS Computational
Biology, 11(5), e1004209. https://doi.org/10.1371/journal.pcbi.1004209
Neymotin, S. A., Hilscher, M. M., Moulin, T. C., Skolnick, Y., Lazarewicz, M. T., & Lytton, W. W.
(2013). Ih tunes theta/gamma oscillations and cross-frequency coupling in an in silico CA3
model. PLoS One, 8(10), e76285. https://doi.org/10.1371/journal.pone.0076285
Norcia, A. M., Appelbaum, L. G., Ales, J. M., Cottereau, B. R., & Rossion, B. (2015). The steady-
state visual evoked potential in vision research: A review. Journal of Vision, 15(6), 4–4.
https://doi.org/10.1167/15.6.4
Nunez, P. L., & Srinivasan, R. (2006). Electric fields of the brain (2nd ed.). Oxford
University Press.
Olejniczak, P. (2006). Neurophysiologic basis of EEG. Journal of Clinical Neurophysiology,
23(3), 186–189. https://doi.org/10.1097/01.wnp.0000220079.61973.6c
Osipova, D., Hermes, D., & Jensen, O. (2008). Gamma power is phase-locked to posterior
alpha activity. PLoS One, 3(12), e3990. https://doi.org/10.1371/journal.pone.0003990
Pape, H.-C., & Paré, D. (2010). Plastic synaptic networks of the amygdala for the acquisition,
expression, and extinction of conditioned fear. Physiological Reviews, 90(2), 419–463. https://
doi.org/10.1152/physrev.00037.2009
Paré, D., Collins, D. R., & Pelletier, J. G. (2002). Amygdala oscillations and the consolidation
of emotional memories. Trends in Cognitive Sciences, 6(7), 306–314. https://doi.org/10.1016/
S1364-6613(02)01924-1
Parvizi, J. & Kastner, S. (2018). Promises and limitations of human intracranial electroenceph-
alography. Nature Neuroscience, 21(4), 474. https://doi.org/10.1038/s41593-018-0108-2
Petro, N. M. & Keil, A. (2015). Pre-target oscillatory brain activity and the attentional blink.
Experimental Brain Research, 233(12), 3583–3595. https://doi.org/10.1007/s00221-015-4418-2
Pfurtscheller, G. (1989). Spatiotemporal analysis of alpha frequency components with the ERD
technique. Brain Topography, 2(1–2), 3–8.
Pfurtscheller, G. (1992). Event-related synchronization (ERS): An electrophysiological correlate
of cortical areas at rest. Electroencephalography and Clinical Neurophysiology, 83, 62–69.
Pfurtscheller, G., Stancak, A., Jr., & Neuper, C. (1996). Event-related synchronization (ERS) in
the alpha band--an electrophysiological correlate of cortical idling: A review. International
Journal of Psychophysiology, 24(1–2), 39–46.
Poldrack, R. A. (2011). Inferring mental states from neuroimaging data: From reverse inference
to large-scale decoding. Neuron, 72(5), 692–697. https://doi.org/10.1016/j.neuron.2011.11.001
Popescu, A. T., Popa, D., & Paré, D. (2009). Coherent gamma oscillations couple the amygdala
and striatum during learning. Nature Neuroscience, 12(6), 801–807. https://doi.org/10.1038/
nn.2305
Pulvermuller, F., & Fadiga, L. (2010). Active perception: Sensorimotor circuits as a cortical
basis for language. Nature Reviews Neuroscience, 11(5), 351–360. https://doi.org/10.1038/
nrn2811
Regan, D. (1989). Human brain electrophysiology: Evoked potentials and evoked magnetic fields
in science and medicine. Elsevier.
Rossion, B., & Boremanse, A. (2011). Robust sensitivity to facial identity in the right human
occipito-temporal cortex as revealed by steady-state visual-evoked potentials. Journal of
Vision, 11(2), 16. https://doi.org/10.1167/11.2.16
64 ANDREAS KEIL and NINA THIGPEN
Samaha, J., Bauer, P., Cimaroli, S., & Postle, B. R. (2015). Top-down control of the phase of
alpha-band oscillations as a mechanism for temporal prediction. Proceedings of the National
Academy of Sciences, 112(27), 8439–8444. https://doi.org/10.1073/pnas.1503686112
Sarnthein, J., Petsche, H., Rappelsberger, P., Shaw, G. L., & von Stein, A. (1998). Synchronization
between prefrontal and posterior association cortex during human working memory.
Proceeding of the National Academy of Sciences of the United States of America, 95(12),
7092–7096.
Schroeder, C. E., & Lakatos, P. (2009). Low-frequency neuronal oscillations as instruments of
sensory selection. Trends in Neuroscience, 32(1), 9–18. https://doi.org/10.1016/j.tins.2008.09.012
Schroeder, C. E., Wilson, D. A., Radman, T., Scharfman, H., & Lakatos, P. (2010). Dynamics
of active sensing and perceptual selection. Current Opinion in Neurobiology, 20(2), 172–176.
https://doi.org/10.1016/j.conb.2010.02.010
Singer, W., Gray, C., Engel, A., König, P., Artola, A., & Brocher, S. (1990). Formation of cortical
cell assemblies. Cold Spring Harbor Symposium on Quantitative Biology, 55, 939–952.
Skarda, C. A., & Freeman, W. J. (1987). How brains make chaos in order to make sense of the
world. Behavioral and Brain Sciences, 10(2), 161–173. https://doi.org/10.1017/S0140525X0
0047336
Snyder, A. C., Morais, M. J., Willis, C. M., & Smith, M. A. (2015). Global network influences on
local functional connectivity. Nature Neuroscience, 18(5), 736–743. https://doi.org/10.1038/
nn.3979
Steriade, M., Gloor, P., Llinás, R. R., Lopes da Silva, F. H., & Mesulam, M.-M. (1990). Basic
mechanisms of cerebral rhythmic activities. Electroencephalography and Clinical
Neurophysiology, 76(6), 481–508. https://doi.org/10.1016/0013-4694(90)90001-Z
Tehovnik, E. J., Tolias, A. S., Sultan, F., Slocum, W. M., & Logothetis, N. K. (2006). Direct
and indirect activation of cortical neurons by electrical microstimulation. Journal of
Neurophysiology, 96(2), 512–521. https://doi.org/10.1152/jn.00126.2006
Thigpen, N. N., Bradley, M. M., & Keil, A. (2018). Assessing the relationship between pupil diam-
eter and visuocortical activity. Journal of Vision, 18(6), 7–7. https://doi.org/10.1167/18.6.7
Thut, G., Nietzel, A., Brandt, S. A., & Pascual- Leone, A. (2006). Alpha- band
electroencephalographic activity over occipital cortex indexes visuospatial attention bias
and predicts visual target detection. Journal of Neuroscience, 26(37), 9494–9502.
VanRullen, R. (2016). Perceptual cycles. Trends in Cognitive Sciences, 20(10), 723–735. https://
doi.org/10.1016/j.tics.2016.07.006
von der Malsburg, C. & Buhmann, J. (1992). Sensory segmentation with coupled neural
oscillators. Biological Cybernetics, 67(3), 233–242.
Weisz, N., Wühle, A., Monittola, G., Demarchi, G., Frey, J., Popov, T., & Braun, C. (2014).
Prestimulus oscillatory power and connectivity patterns predispose conscious somatosen-
sory perception. Proceedings of the National Academy of Sciences, 111(4), E417–E425. https://
doi.org/10.1073/pnas.1317267111
Wieser, M. J. & Keil, A. (2011). Temporal trade-off effects in sustained attention: Dynamics
in visual cortex predict the target detection performance during distraction. Journal of
Neuroscience, 31(21), 7784–7790. https://doi.org/10.1523/JNEUROSCI.5632-10.2011
Yuval-Greenberg, S., Tomer, O., Keren, A. S., Nelken, I., & Deouell, L. Y. (2008). Transient
induced gamma-band response in EEG as a manifestation of miniature saccades. Neuron,
58(3), 429–441.
CHAPTER 4
TIME-F REQU E NC Y
DE C OM P OSITION MET H OD S
F OR EVENT-R E L AT E D
P OTENTIAL A NA LYSI S
SELIN AVIYENTE
4.1 Introduction
EEG reflects the electrical activity of a collection of neural populations in the brain.
As such, it reflects the superposition of different simultaneously acting dynamical
systems. When a large number of parallel-oriented cortical neurons receive the same
repetitive synaptic input, their synchronous activity produces extracellular rhythmic
field potentials. These electrical fields are volume conducted throughout the brain and
recorded as EEG from the scalp (Nunez & Srinivasan, 2006). In recent years, it has been
suggested that this synchronization optimizes relations between spike-mediated “top-
down” and “bottom-up” communication, both within and between brain areas. This
optimization might have particular importance during motivated anticipation of, and
attention to, meaningful events and associations and in response to their anticipated
consequences (Von Stein et al., 2000; Fries et al., 2001; Salinas & Sejnowski, 2001; Makeig
et al., 2004). This new theory of cortical and scalp-recorded field dynamics requires new
data analysis approaches for spatially distributed event-related EEG dynamics.
The standard analysis method to study EEG in an event-related fashion is to focus
on event-related potentials (ERPs) by averaging. ERPs are positive or negative voltage
deflections seen in the averages of EEG epochs time-locked to a class of repeated stimulus
or response events. However, this approach assumes that ERPs are superimposed on
ongoing background EEG “noise” with amplitude and phase distributions that are
completely unrelated to the task. This view of ERPs has been challenged in the last two
decades. First, time-frequency analysis of single-trial EEG signals shows that EEG is
66 SELIN AVIYENTE
not solely random noise; rather, there are event-related changes in the magnitude and
phase of EEG oscillations at specific frequencies (Makeig et al., 2004). The early work in
this area focused on spectral analysis of EEG signals, which breaks down the oscillatory
EEG waveforms into sinusoids at different frequencies, reflecting the role of different
frequency bands in cognitive function. Second, ERPs themselves may represent tran-
sient phase resetting of ongoing EEG by experimental events, leading to transient time-
and phase locking of frequency specific oscillations (Makeig et al., 2004; Makeig et al.,
2002). Makeig and colleagues (2002) have introduced the event-related brain dynamics
framework, which emphasizes the spectral decomposition of single-trial event-related
EEG epochs in order to separately examine event-related changes in the magnitude
and phase of oscillations at different frequencies. However, spectral analysis assumes
stationarity of the signal, in spite of the fact that information processing in the brain is
mostly reflected by fast dynamic changes in EEG. Even though the Fourier transform
provides a complete representation of the time series, the resulting power spectrum is
only interpretable for stationary signals; non-stationarities are encoded in the phase
spectrum, which is typically impossible to interpret visually. Spectral analysis is only
capable of reflecting the average or global frequency content, rather than illustrating the
local activity. The non-stationarities in the EEG signal are the primary motivation for
methods that can capture simultaneously the variation of signal energy across time and
frequency.
Time- frequency analyses of EEG provide additional information about neural
synchrony not apparent in ongoing EEG activity or in ERPs. They can tell us which
frequencies have the most power at different time points and how their phase
synchronizes across time and space. Thus, using time-frequency analyses, we can assess
changes in power and synchronization of EEG within or between spatial locations
across trials with respect to the onset of tasks. However, it is important to note that time-
frequency analysis cannot by itself determine the cause of the changes in EEG power,
that is, whether the change is due to changes in the magnitude of the oscillations or to
changes in their degree of synchronization (Yeung, 2004; Roach & Mathalon, 2008).
The goal of this chapter is to give an overview of different time-frequency analysis
tools for both quantifying the changes in EEG power across time and frequency and the
changes in phase synchrony across time, frequency, and space. The different methods
discussed in this chapter can be divided into three categories: linear, nonlinear, and
data-driven or adaptive methods. The first category of methods focus on simple exten-
sion of Fourier transform to the non-stationary signal that is, short-time Fourier trans-
form or sliding window approach, and the continuous wavelet transform. The second
category of methods focus on Cohen’s class of time-frequency distributions (TFDs) that
compute the spectrum of the time-varying auto-correlation function of the signal, ra-
ther than the signal itself. In this manner, these distributions are highly nonlinear but
offer high-resolution visualizations of EEG dynamics. The last category of methods
stem from recent advances in signal processing that focus on sparse signal represen-
tation. The two methods reviewed under this category are matching pursuit (MP) and
empirical mode decomposition (EMD).
TIME-FREQUENCY DECOMPOSITION METHODS 67
1
All integrals are from −∞ to ∞ unless otherwise noted.
68 SELIN AVIYENTE
window function should be large, that is, the rate of decay the window function in the
frequency domain, G (ω ), should be fast. Some common window functions include
the rectangular, Hamming, Hanning, and Gaussian window functions. Rectangular
windows have the worst frequency resolution due to the sharp transition of the window
function. Gaussian window achieves the best joint time and frequency localization as it
meets the lower bound of the uncertainty principle. In STFT, the duration of the time
window is fixed in contrast to the wavelet transform, which in turn provides uniform
frequency resolution. STFT is a complex valued time-frequency representation, and an
energy distribution can be obtained as| S (t , ω ) |2 , known as the spectrogram.
1 τ −t
W (t , s ) = ∫ x ( τ ) ψ* dτ, (4.2)
s s
where ψ is the wavelet function and s is the scale parameter. Similar to STFT defined
in Eq. 4.1, the wavelet transform finds the decomposition of a given signal in terms of
time-frequency atoms. However, in this case, the time-frequency atoms are the time-
shifted and scaled versions of the mother wavelet, ψ . For this reason, wavelet transform
should be thought of as a time-scale decomposition rather than a TFD, as the signal is
not decomposed in terms of sinusoids across different frequencies. One of the major
differences between STFT and CWT is the time-frequency localization. In STFT, the
frequency localization is uniform due to the fixed window size. On the other hand, for
CWT the time resolution is variable, with shorter time windows for higher frequencies
and longer time windows for lower frequencies. This variable time resolution closely
matches the structural properties of ERP signals and has made CWT an attractive
choice for time-frequency analysis of ERPs (Roach & Mathalon, 2008; Demiralp et al.,
2001; Polikar et al., 2007). Using CWT, the ERP can be decomposed across several or-
thogonal functions, wavelets, with overlapping time courses at different scales.
While any number of wavelet functions can be used for ERP analysis, the wavelet
must provide a biologically plausible fit to the signal being modeled. Some commonly
used wavelets for EEG and ERP analysis are real wavelets, such as splines, and analytic
wavelets, such as the Morlet wavelet and the Generalized Morse Wavelet (GMW). In
particular, the analytic wavelets are useful to obtain both energy and phase information
of the underlying oscillations. The Morlet wavelet is one that has been commonly used
TIME-FREQUENCY DECOMPOSITION METHODS 69
for EEG/ERP analysis. The Morlet wavelet is defined as a complex sinusoid tapered by a
Gaussian given as:
t2
−
ψ (t ) = e j 2 π ft
e 2 σ2
, (4.3)
where σ is the width of the Gaussian and is reciprocally related to the frequency in order
to retain the wavelet’s scaling properties. By this scaling, one obtains the same number
of significant wavelet cycles at all frequencies. Thus, the Morlet wavelet transform has
a different time and frequency resolution at each scale. Therefore, at high frequencies
the temporal resolution of a wavelet is better than at low frequencies. However, the in-
verse is true for the frequency resolution of the wavelet transform. Convolution with
the complex Morlet wavelet results in a complex-valued signal from which instantan-
eous power and phase can be extracted at each time point. Wavelet convolution can be
conceptualized as a template matching or bandpass filtering. Convolutions with Morlet
wavelets can be computed for multiple frequencies in order to yield a time-frequency
representation. Since Morlet wavelet is a complex function, the corresponding wavelet
transform is also complex, where the real part corresponds to the bandpass filtered
signal and the imaginary part corresponds to Hilbert transform of the signal. There are
several advantages of Morlet wavelets for EEG/ERP analysis (Cohen, 2018). One is that
the Morlet wavelet is Gaussian shaped in the frequency domain and the absence of sharp
edges minimizes ripple effects. Second, the results of Morlet convolution retain the tem-
poral resolution of the original signal. Third, wavelet convolution is computationally
efficient and can be implemented using the fast Fourier transform. The wavelet power
spectrum, also known as the scalogram, can be obtained as | W (t , s ) |2 .
For a signal, x (t ), a bilinear TFD, C (t, ω ), from Cohen’s class can be expressed as
(Cohen, 1995):
τ τ j θu − θt − τω)
C (t , ω ) = ∫∫∫ φ (θ, τ ) x u + x * u − e ( du dθ dτ, (4.4)
2 2
where φ (θ, τ ) is the kernel function in the ambiguity domain (θ, τ ). Unlike linear
transforms, which first multiply the signal with a time-limited window function before
70 SELIN AVIYENTE
computing its Fourier transform, Cohen’s class of distributions compute the Fourier
transform of the local autocorrelation function. In this manner, the signal acts like a
window on itself. The kernel φ (θ, τ ) acts like a filter on this local autocorrelation
function and determines which parts to preserve. TFDs represent the energy distri-
bution of a signal over time and frequency, simultaneously. Some of the most desired
properties of TFDs are the energy preservation, the marginals, and the reduced interfer-
ence. Some common TFDs used for EEG and ERP analysis include the Wigner distribu-
tion and its filtered versions (Tağluk et al., 2005; Abdulla & Wong, 2011).
τ τ
W (t , ω ) = ∫ x t + x * t − e − jωτ dτ, (4.5)
2 2
where x (t ) is the signal and τ is the time lag variable. The Wigner distribution
computes the Fourier transform of the local autocorrelation of the signal as the kernel
φ (θ, τ ) = 1. As the kernel acts as an all pass filter, Wigner distribution keeps the whole
autocorrelation function before transforming it. This provides a high time-frequency
resolution. In fact, it has been shown that the Wigner distribution provides the highest
time-frequency resolution among all Cohen’s class of TFDs (Cohen, 1995). It overcomes
some of the shortcomings of the spectrogram as there is no dependency on the choice
of the window. Moreover, the Wigner distribution provides accurate instantaneous fre-
quency estimates and is an accurate energy distribution as it satisfies the marginals.
However, Wigner distribution is a bilinear function of signals. Therefore, Wigner dis-
tribution of multi-component signals or mono-component signals with curved time-
frequency supports will be cluttered by spurious terms called cross-terms. The existence
of cross-terms may decrease the interpretability of the TFD. ERPs, like other non-
stationary signals, require an analysis technique that is free from cross-terms.
of the ambiguity plane. This allows for interpretation of the kernel function as a low-
pass filter that attempts to reject the cross-terms and leave the auto-terms unchanged.
The shape of the filter can be adapted to the analysis of the desirable signal components
depending on the nature of the signal. For RIDs, φ (θ, τ ) = h (θτ ), where h is a monot-
onously decreasing function to ensure that the parameterization function is a low-pass
filter in the ambiguity plane. Depending on the canonical form of h, the resultant RIDs
have different cross-term rejection capabilities. Some commonly used RIDs include
Choi–Williams (CW) (Jeong & Williams, 1992), B-distribution (Zhao et al., 1990),
Born–Jordan, and Zhao–Atlas–Marks (Zhao et al., 1990).
Since these distributions will be implemented in discrete-time, discrete-time TFD of
size 2N + 1 × 2N + 1 can be defined as:
N N
n +n − jω n −n
TFD (n, ω; ψ ) = ∑ ∑ x (n + n ) x (n + n ) ψ − 2
n1 =− N n2 =− N
1
*
2
1 2
, n1 − n2 e ( 1 2 ) , (4.6)
where ψ is the discrete-time kernel in the time and time-lag domain. The most commonly
used discrete-time kernel for EEG/ERP analysis is the binomial kernel, which is given by
m
ψ (n, m) = 2 −m
m for n ≤ m .
n + 2
2
Unlike the first two types of time-frequency transforms, the methods reviewed in this
section do not directly result in time-frequency energy distributions, but rather result
in a description of the ERP dynamics as a sum of a few numbers of distinct components.
These components may be either data-driven, that is, derived directly from the signals,
or selected from a large dictionary of atoms. Usually, these components are localized
in time and frequency and thus can be thought of as a decomposition of the event-
related EEG signals into a few distinct time-frequency elements. Thus, the resulting
components can be visualized using any transform discussed in the previous sections
to obtain a time-frequency spectrum. For this reason, these are sometimes considered
as time-frequency analysis tools in the literature. By the nature of the algorithms used to
derive these components, these methods are nonlinear. The first one of these methods,
EMD, is data-driven and the resulting components usually represent the different fre-
quency bands of EEG. The second method, MP, can be thought of as an extension of
wavelet transform. However, unlike the wavelet functions, which usually form an
72 SELIN AVIYENTE
orthonormal basis for the signal space, the time-frequency atoms used in MP form an
overcomplete dictionary. This means that the set of functions used to express the signal
is more than necessary. This overcompleteness results in sparse representation of the
EEG/ERP signals. These atoms are pre-determined similar to the wavelet transform and
can be selected by the user based on the characteristics of the signals. Some example
atoms could be a combination of sinusoids and wavelets, Gabor functions, etc. This ef-
ficiency in representation of the signals with a few time-frequency atoms comes at the
expense of computational complexity. Unlike wavelet transform, which is computation-
ally efficient, MP usually relies on greedy algorithms (Mallat & Zhang, 1993; Tropp &
Gilbert, 2007).
are the IMFs and r (t ) is the residue. The algorithm can be described as follows:
1. Let x (t ) = x (t ) .
2. Identify all local maxima and minima of x (t ) .
3. Find two envelopes emin (t ) and emax (t ) that interpolate through the local
minima and maxima, respectively.
1
( )
4. Let d (t ) = x (t ) − emin (t ) + emax (t ) as the detail part of the signal.
2
5. Let x (t ) = d (t ) and go to step 2 and repeat until d (t ) becomes an IMF. The two
IMF criteria are: a) the number of extrema and the number of zero-crossings must
either be equal or differ at most by one; b) At any point, the mean value of the en-
velope defined by the local maxima and the envelope defined by the local minima
is zero.
6. Compute the residue r (t ) = x (t ) − d (t ) and go back to step 1 until the energy of
the residue is below a threshold.
TIME-FREQUENCY DECOMPOSITION METHODS 73
Although EMD has the ability to extract ERP components, it suffers from the mode
mixing problem (Huang et al., 2003). That is, the different time-frequency components
may not directly correspond to the different IMFs, which makes it difficult to deter-
mine the distinct ERP components. To overcome the problem of mode mixing, Wu
and Huang (2009) recommend ensemble EMD (EEMD), a noise-assisted data-analysis
method. The output of EEMD is a set of IMFs generated from ensemble means of trials
by repeating EMD on the same signal with different sets of Gaussian noise. A further
potential drawback of EMD has been put forth by Flandrin and colleagues (2004), who
showed that EMD behaves as a dyadic filter bank. This poses the concern that EMD
may naturally lead to such a decomposition in all data, which implies that important
oscillations may not be identified if they do not adhere to a dyadic frequency relation-
ship with one another.
Once the IMFs are obtained, Hilbert spectral analysis, also known as Hilbert–Huang
Transform (HHT), can be used to evaluate the frequency content of each IMF. Hilbert
spectral analysis provides the instantaneous frequency of each IMF. According to Huang
and colleagues (2011), the instantaneous frequency could represent the nonlinear and
non-stationary signals without resorting to the mathematical artifact of harmonics. Like
measuring ERPs, averaging IMFs across trials provides event-related modes (ERMs).
Based on the instantaneous frequency of ERMs, ERP components can be extracted by
summing ERMs (Cong et al., 2009; Williams et al., 2011; Wu et al., 2012) or using an
ERM within a frequency range.
M
x (t ) = ∑ ai g i (t ) . (4.7)
i =1
The problem of choosing M functions, which explain the largest proportion of en-
ergy of a given signal, is a computationally complex problem and is known to belong to
the class of non-deterministic polynomial-time hard (NP-hard) problems. MP offers a
tractable suboptimal solution, obtained by an iterative algorithm.
74 SELIN AVIYENTE
In the first step of the iterative procedure, we choose the element of the dictionary
that gives the largest inner product with the signal, that is, g 1 = argmax g i ∈D x , g i . This
first element of the dictionary is subtracted from the signal to obtain the residue. The
iterative procedure is repeated on the subsequent residual, R k x . This procedure can be
summarized as:
g k = argmax g i ∈D R k x , g i . (4.8)
R k +1 x = R k x − R k x , g k g k . (4.9)
x = ∑ R k x , g k g k + R M +1 x. (4.10)
k =1
preserves signal energy. From this representation, one can derive a TFD of a signal’s en-
ergy by adding Wigner distributions of selected atoms.
The overcomplete dictionary, D , can be designed to fit the class of signals at hand.
Two important requirements for a dictionary are its descriptive power, that is, its
ability to represent the signals of interest with relatively few atoms (sparsity), and its
interpretability, that is, that the parameters indexing the atoms convey information.
Although overcomplete dictionaries do not provide uniqueness of decomposition,
they have more descriptive power than more classical, orthogonal dictionaries, such
as wavelets. Regarding interpretability, the choice of atoms and their parameters is
motivated by the types of activities that will be encountered. The dictionary is fur-
thermore supposed to depend continuously on the parameter space (even though
for numerical reasons, this space will obviously be discretized). This constraint is
imposed only for ease of presentation, but the approach could be generalized, at the
cost of some added complexity. For analyzing EEG and ERP signals, previous work
has shown that Gabor logons represent the signals with few coefficients (Durka &
Blinowska, 2001; Brown et al., 1994; Aviyente, 2007) . Gabor logons also have the
advantage of being the most concentrated signals on the time-frequency plane,
TIME-FREQUENCY DECOMPOSITION METHODS 75
achieving the lower bound of the time-bandwidth product. A dyadic Gabor dic-
tionary also allows for computationally effective implementation and has been used
widely in EEG analysis.
r
argmax g i ∈D ∑ R k x l , g i . (4.11)
l =1
R k +1 x l = R k x l − R k x l , g k g k . (4.12)
Different implementations of MMP have been employed for ERP analysis, that is,
evoked activity MP (EMP) and induced activity MP (IMP) (Bénar et al., 2009). In EMP,
the atoms are determined to maximize the correlation with the average signal, and the
amplitude is adapted to the individual trials. In this manner, the method accounts for
amplitude variability across trials, but not for variability in the parameter space. IMP,
on the other hand, maximizes the average energy across trials. Durka and colleagues
76 SELIN AVIYENTE
This section illustrates the performance of some of the time-frequency analysis methods
discussed earlier for an example ERP signal. In particular, we focus on the error-related
negativity (ERN), a neurophysiological marker of performance monitoring. ERN is a
negative deflection in the ERP that peaks within 100 ms of an erroneous response at
frontocentral recording sites (Moser, 2017; Moran et al., 2015). The ERN is generally
considered to be an index of cognitive control-related performance monitoring that is
involved in coordinating optimal responding following mistakes. For the purpose of
this illustration, we consider ERP obtained by averaging across trials recorded at the
FCz electrode for one subject.
First, we compare the performance of linear and nonlinear time-frequency energy
distributions, namely the continuous wavelet transform with the Morlet wavelet, and
RID with the binomial kernel. From Figure 4.1, a typical ERN waveform with a negative
potential occurring in the first 100 ms can be seen, along with different TFDs. From both
4
μV
Time 0
–4
–600 –400 –200 0 200 400
Binomial 10
Hz
RID
Amplitude
0
High
Morlet 10
Hz
Wavelet
0 Low
Time (ms)
Figure 4.1 Time-frequency Analysis of ERN waveform. From top to bottom: ERN signal in the
time domain, reduced interference distribution (binomial kernel), continuous wavelet transform
with Morlet wavelet.
TIME-FREQUENCY DECOMPOSITION METHODS 77
distributions, it is clear that there is high energy corresponding to ERN and P3e, P300
following the error response. The biggest difference between the different distributions
is the time and frequency resolution of this component in the time-frequency plane. For
RID, there are two distinct time-frequency components; one during ERN localized in
the theta band (4–8Hz) and another distributed in time up till 400 ms in the delta (0–2
Hz) band. Morlet wavelet produces a distribution that captures the theta band activity
without highlighting the delta band energy. It can be seen from these results that RID
has the better time and frequency resolution, whereas the wavelet transform has poorer
time and frequency localization. From this comparison, it can be concluded that RID
provides better delineation of different ERP components. The results of the CWT may
be improved by changing the wavelet type and wavelet scales. Unlike RID, which does
not depend on the selection of any parameters, CWT is highly dependent on the user
provided parameter values.
Next, we compare the performance of data-driven component extraction methods,
EMD and MP. Figure 4.2 illustrates the first four IMFs, along with their Hilbert spectra.
This figure shows that the first two IMFs correspond to the ERN time interval, whereas
the third and fourth IMFs correspond to the P3e potential. The corresponding time-
frequency energy distribution has a high energy concentration in the 100–200 ms and
300–400 ms time intervals within the delta frequency band. These components corres-
pond to the P3e. However, the ERN component is not easily detected from this distribu-
tion as it has less energy than P3e.
In a similar manner, Figure 4.3 illustrates an MP spectrum obtained using a Gabor
dictionary. The dictionary is constructed using Gabor functions, that is, Gaussian
window functions shifted in time and modulated in frequency. The figure illustrates the
time-frequency localization of the selected atoms. The figure also shows a high energy
atom in the 100–200 ms time interval in the delta band. There are also Gabor functions
with negative weights in the 0–100 ms time window around 10–12 Hz frequency band.
This corresponds to ERN time window and alpha frequency band.
From these comparisons, it can be seen that RID is the best in terms of separating
different ERP oscillations from each other with high resolution. The component-
based methods work better in separating different oscillatory components, such as
the IMFs extracted from EMD. However, EMD is not a true time-frequency energy
distribution method, as it does not produce an actual spectrum like RID. The current
method of visualizing IMFs in the time-frequency plane relies on the Hilbert trans-
form of extracting the individual envelope and instantaneous frequency for each IMF.
As the Hilbert transform is not a true time-frequency localization method, the IMFs
can also be transformed to the time-frequency plane using different distributions to
obtain higher-resolution visualizations. MP, on the other hand, is highly dependent on
the selected dictionary atoms. The more suitable the dictionary is for the underlying
signal, the sparser the resulting representation. In this particular example, MP wrongly
detects alpha components for the ERN time window while missing the theta activity for
the same time window. As EMD is data-driven and independent of priors, it performs
better than MP for ERP analysis.
78 SELIN AVIYENTE
–20
–0.2 –0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
5
IMF 1
–5
–0.2 –0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
10
IMF 2
–10
–0.2 –0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
10
IMF 3
–10
–0.2 –0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
2
IMF 4
–2
–0.2 –0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
10
Residual
–10
–0.2 –0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Time (second)
(b) 100
40
35
80
30
Frequency (Hz)
25 60
20
15 40
10
20
5
0
–0.2 –0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Time (second)
Figure 4.2 Empirical mode decomposition for ERP signal: (A) The first four intrinsic
mode functions (IMFs) extracted from the ERP signal. (B) The time-frequency spectrum
corresponding to the extracted IMFs.
TIME-FREQUENCY DECOMPOSITION METHODS 79
400
12
300
10
200
8
Frequency (Hz)
100
6
0
4 –100
2 –200
–0.2 –0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Time (second)
Figure 4.3 Time-frequency analysis using matching pursuit with a Gabor dictionary.
Although time-frequency energy distribution is effective for studying the spectral con-
tent of EEG and ERP oscillations, most of the current approaches focus only on the
magnitude spectrum of the oscillations and ignore the phase information. Recent re-
search shows that ERPs result from event-related partial phase resetting of ongoing os-
cillatory activity along with transient increases in the magnitude of oscillations that are
time-locked to the experimental events (Roach & Mathalon, 2008; Makeig et al., 2004).
Therefore, it is important to characterize the change in phase information across time
and frequency.
Phase synchrony quantifies the relation between the temporal structures of the
signals, regardless of signal amplitude (Rosenblum et al., 2000; 2001). Two signals
are said to be synchronous if their rhythms coincide. The amount of synchrony be-
tween two signals is usually quantified by first estimating the instantaneous phase
of the individual signals around the frequency of interest. Traditionally, the instant-
aneous phase of a given signal is estimated first, and then the phase synchrony is
quantified using statistical measures. Conventionally, the instantaneous phase is
estimated using the Hilbert transform. However, the Hilbert transform does not
80 SELIN AVIYENTE
Wx (t , f ) = ∫ x (u ) ψ t*, f (u ) du (4.14)
−∞
where ψ t*, f (u ) represents the complex conjugate of the wavelet function. Using the
( u − t )2
−
Morlet wavelet ψ t , f (u ) =
j 2 π f (u − t )
fe e 2 σ2
, the phase spectrum of x (t ) can be
evaluated as follows:
W (t , f )
Φx (t , f ) = arg x . (4.15)
Wx (t , f )
for quantifying phase synchrony between signal pairs. A problem with current syn-
chrony detection methods such as the Hilbert transform is that they depend on a priori
selection of bandpass filters. In response to this problem, EMD-based phase synchrony
measures have been defined (Rutkowski et al., 2008; Looney et al., 2009; Sweeney-Reed
& Nasuto, 2007; Mutlu & Aviyente, 2011). In most of these approaches, the IMFs for
each time series were extracted individually and were compared individually against
the IMFs from the other time series for computing phase synchrony. This approach
has multiple shortcomings. First, the IMFs from the different time series do not neces-
sarily correspond to the same frequency, thus making it hard to compute exact within-
frequency phase synchronization across different EEG channels. Second, the different
time series may end up having different numbers of IMFs, which makes it difficult to
match the different IMFs for synchrony computation. Finally, Looney and colleages
(2009) showed that univariate EMD is not robust under noise and may suffer from
mode mixing, which refers to the phenomenon of different modes (or frequencies)
existing in a single IMF due to noise or intermittent signal activity.
Recently, extensions of EMD to the multivariate case have been developed including
Complex EMD (Tanaka & Mandic, 2007), Rotation Invariant EMD (RIEMD)
(Altaf et al., 2007), and Bivariate EMD (BEMD) (Rilling et al., 2007). These complex
extensions of EMD decompose data from different sources simultaneously. Looney and
colleages (2009) showed that the IMFs obtained in this fashion are matched, not only in
number, but also in frequency overcoming problems of uniqueness and mode mixing,
and first suggested the idea of using bivariate EMD to compute phase synchrony be-
tween two signals and the BEMD was shown to perform better than univariate EMD for
quantifying bivariate synchrony. However, this approach still has some shortcomings, as
the frequency bands corresponding to IMFs from different bivariate pairs are not neces-
sarily the same. As such, this method is mostly limited to bivariate synchrony analysis
and is not as easily generalizable to the whole brain synchrony analysis. Moreover, the
multivariate extensions of EMD have been shown to be computationally expensive for
computing functional connectivity across the whole brain (Mutlu & Aviyente, 2011).
where X * (ω ) is the complex conjugate of the Fourier transform of the signal and
C (t, ω ) measures the complex energy of a signal around time, t and frequency, ω . The
complex energy density function provides a fuller appreciation of the properties of
phase-modulated signals that is not available with other TFDs. Rihaczek distribution
is a bilinear, time-and frequency-shift covariant, complex-valued TFD belonging to
Cohen’s class. This distribution satisfies the marginals and preserves the energy of the
signal. Rihaczek distribution provides both a time-varying energy spectrum as well as a
phase spectrum, and thus is useful for estimating the phase synchrony between any two
signals. One of the disadvantages of Rihaczek distribution, similar to other quadratic
TFDs, is the existence of cross-terms for multicomponent signals. In order to get rid of
these cross-terms, in previous work (Aviyente et al., 2011), we proposed to apply a kernel
function such as the CW kernel to filter the cross-terms. The resulting distribution can
be written as:
(θτ)2 θτ
C (t , ω ) = ∫ ∫ exp − exp j A (θ, τ ) e (
− j θt + τω )
dτ dθ,(4.16)
σ 2
CWkernel Rihaczekkernel
θτ
j
where e is the kernel function for the Rihaczek distribution. This new distribution,
2
the corresponding distribution will both satisfy the marginals and preserve the en-
ergy, as well as be a complex energy distribution at the same time. The value of σ can
be adjusted to achieve a desired trade-off between resolution and the number of cross-
terms retained. The phase estimates from RID-Rihaczek distribution can be obtained as
C (t , ω )
φ (t , ω ) = arg . This phase estimate has been shown to be more robust to noise
C (t , ω )
and has uniformly high resolution in time and frequency compared to wavelet-based
phase estimates (Aviyente & Mutlu, 2011).
have been proposed to quantify the synchrony based on the relative phase difference,
that is, φ x , y (t ) is wrapped into the interval [0, 2π ) . The first index uses an information-
theoretic criterion to quantify synchronization. This measure studies the distribu-
tion of φ x , y (t ) by partitioning the interval [0, 2π ) into L bins and comparing it with
the distribution of the cyclic relative phase obtained from two series of independent
phases. This comparison is carried out by estimating the Shannon entropy of both
distributions, that is, that of the original phases, and that of the independent phases, and
computing the normalized difference (Quiroga et al., 2002). The second metric, phase
synchronization index, is also known as the mean phase coherence and computed as
1 N −1 jΦ (t )
( ) ( )
< cos Φxy (t ) > 2 + < sin Φxy (t ) > 2 = ∑ e xy k . It is a measure of how the rela-
N k =0
tive phase is distributed over the unit circle. If the two signals are phase synchronized,
the relative phase will occupy a small portion of the circle and mean phase coherence
is high. This measure is equal to 1 for the case of complete phase synchronization and
tends to zero for independent oscillators.
These different measures of synchrony can be used to quantify three different types
of synchronization from EEG/ERP signals, depending on the application. The major
difference between the different implementations of phase synchrony is whether the
consistency of phase differences is measured across time, trials, or channels. Phase
Locking Value (PLV) quantifies the consistency of the phase differences across trials
1 N
( )
between two channels as follows: PLV (t , ω ) = ∑ exp jφkx , y (t , ω ) , where k is the
N k =1
trial number and N is the number of trials. This measure is also known as the inter-
channel phase synchrony (ICPS). It is commonly used to estimate phase-locking in ex-
perimental situations, common in neurocognitive studies, where a subject is presented
with a sequence of similar stimuli. Single trial phase-locking value (S-PLV), on the
other hand, allows us to measure the significance of synchronies in single trials, and
does not depend on block repetition of events. The variability of phase-difference is
not measured across trials, but across successive time-steps, around a target latency.
Specifically, a smoothed or S-PLV is defined for each individual trial. Finally, inter-trial
phase synchrony (ITPS) quantifies the consistency of phase values for a given frequency
band at each point in time over trials, in one particular electrode. ITPC is defined as
1 N
( )
ITPC (t , ω ) = ∑ exp jφk (t , ω ) , where N is the number of trials and φk (t, ω ) is the
N k =1
phase of the k th trial for each time and each frequency point. ITPC thus reflects the
extent to which oscillation phase values are consistent over trials at that point in time-
frequency plane. Note that this measure of phase coherence does not differentiate be-
tween possible biophysical mechanisms underlying phase consistency, such as phase
reset or phase “smearing”. Rather, this measure simply indicates the statistical prob-
ability of increased phase consistency between trial and baseline epochs.
84 SELIN AVIYENTE
4.7 Summary
This chapter reviewed some of the basic time-frequency methods for analyzing ERP
dynamics. We first introduced the different transforms that have been used to analyze
the transient ERP activity. These transforms can be divided into three categories: linear
methods such as STFT and CWT, data-adaptive methods such as EMD, and nonlinear
methods such as Cohen’s class of distributions. Although the overall goal of all of
these methods is to determine the dynamics of transient activity, the mathematical
principles upon which they rely are quite different. These differences in mathematical for-
mulation lead to different computational complexities. Therefore, it is important to under-
stand how the different methods can be used for different applications and purposes. For
example, if the goal is to visualize the energy distribution of the transient ERP activity in
time and frequency simultaneously, then Cohen’s class of distributions (such as the RID)
performs the best. Even though this method is the most complex in terms of computa-
tional complexity, it offers very high time-frequency resolution, which in turn provides
a way to delineate different ERP components from each other. However, if the goal is
to obtain a compact representation of the ERP signals in terms of a few physiologically
meaningful components, then data-adaptive methods like EMD and MP are more suit-
able and computationally efficient. This chapter also shows how the same mathematical
transforms can be used to study ERP dynamics through both energy distributions and
phase synchrony in the time-frequency domain. While the energy distributions focus
on the univariate activity, that is, dynamics within a certain brain region or electrode,
the phase synchrony quantifies the dynamics across brain regions. In this manner, it is
possible to obtain a better understanding of ERP timing, synchronization across time,
frequency, and space.
REFERENCES
Abdulla, W. & Wong, L. (2011). Neonatal EEG signal characteristics using time frequency ana-
lysis. Physica A: Statistical Mechanics and its Applications, 390(6), 1096–1110.
Altaf, M. U. B., Gautama, T., Tanaka, T., & Mandic, D. P. (2007). Rotation invariant complex
empirical mode decomposition. In 2007 IEEE international conference on acoustics, speech
and signal processing-ICASSP’07, vol. 3 (pp. III–1009). IEEE.
Aviyente, S. (2007). Compressed sensing framework for EEG compression. In Proceedings
of IEEE/SP: 14th workshop on statistical signal processing (pp. 181–184). IEEE. Doi: 10.1109/
SSP.2007.4301243.
Aviyente, S., Bernat, E. M., Evans, W. S., & Sponheim, S. R. (2011). A phase synchrony measure
for quantifying dynamic functional integration in the brain. Human Brain Mapping,
32(1), 80–93.
Aviyente, S. & Mutlu, A. Y. (2011). A time-frequency-based approach to phase and phase syn-
chrony estimation. IEEE Transactions on Signal Processing, 59(7), 3086–3098.
TIME-FREQUENCY DECOMPOSITION METHODS 85
Aviyente, S., Tootell, A., & Bernat, E. M. (2017). Time-frequency phase-synchrony approaches
with ERPs. International Journal of Psychophysiology, 111, 88–97.
Bénar, C. G., Papadopoulo, T., Torrésani, B., & Clerc, M. (2009). Consensus matching pursuit
for multi-trial EEG signals. Journal of Neuroscience Methods, 180(1), 161–170.
Brown, M. L., Williams, W. J., & Hero, A. O. (1994). Non-orthogonal Gabor representa-
tion of biological signals. In Proceedings of ICASSP ’94: IEEE international conference
on acoustics, speech and signal processing, vol. 4 (pp. IV/305-IV/308). IEEE. doi: 10.1109/
ICASSP.1994.389742.
Cohen, L. (1995). Time-frequency analysis, vol. 778. Prentice Hall.
Cohen, M. X. (2018). A better way to define and describe Morlet wavelets for time-frequency
analysis. bioRxiv [online], 397182.
Cong, F., Sipola, T., Huttunen-Scott, T., Xu, X., Ristaniemi, T., & Lyytinen, H. (2009). Hilbert-
Huang versus Morlet wavelet transformation on mismatch negativity of children in uninter-
rupted sound paradigm. Nonlinear Biomedical Physics, 3(1), 1.
Demiralp, T., Ademoglu, A., Istefanopulos, Y., Başar-Eroglu, C., & Başar, E. (2001). Wavelet
analysis of oddball p300. International Journal of Psychophysiology, 39(2–3), 221–227.
Durka, P. J. & Blinowska, K. J. (2001). A unified time-frequency parametrization of EEGs. IEEE
Engineering in Medicine and Biology, 20(5), 47–53.
Durka, P. J., Matysiak, A., Montes, E. M., Valdés Sosa, P., & Blinowska, K. J. (2005).
Multichannel matching pursuit and EEG inverse solutions. Journal of Neuroscience
Methods, 148(1), 49–59.
Flandrin, P., Rilling, G., & Goncalves, P. (2004). Empirical mode decomposition as a filter
bank. IEEE Signal Processing Letters, 11(2), 112–114.
Fries, P., Reynolds, J. H., Rorie, A. E., & Desimone, R. (2001). Modulation of oscillatory neur-
onal synchronization by selective visual attention. Science, 291(5508), 1560–1563.
Gribonval, R., Rauhut, H., Schnass, K., & Vandergheynst, P. (2008). Atoms of all channels,
unite! Average case analysis of multi-channel sparse recovery using greedy algorithms.
Journal of Fourier Analysis and Applications, 14(5–6), 655–687.
Hassanpour, H., Mesbah, M., & Boashash, B. (2004). Time-frequency feature extraction of
newborn EEG seizure using SVD-based techniques. EURASIP Journal on Applied Signal
Processing, 2004, 2544–2554.
Huang, N. E., Shen, Z., Long, S. R., Wu, M. C., Shih, H. H., Zheng, Q., et al. (1998). The empir-
ical mode decomposition and the Hilbert spectrum for nonlinear and non-stationary time
series analysis. Proceedings of the Royal Society of London. Series A: Mathematical, Physical
and Engineering Sciences, 454(1971), 903–995.
Huang, N. E., Wu, M-L., Qu, W., Long, S. R., & Shen, S. S. P. (2003). Applications of Hilbert–
Huang transform to non-stationary financial time series analysis. Applied Stochastic Models
in Business and Industry, 19(3), 245–268.
Huang, Y. X., Schmitt, F. G., Hermand, J- P., Gagne, Y., Lu, Z. M., Liu, Y. L.,
et al. (2011). Arbitrary-order Hilbert spectral analysis for time series possessing scaling
statistics: Comparison study with detrended fluctuation analysis and wavelet leaders.
Physical Review E, 84(1), 016208.
Jeong, J. & Williams, W. J. (1992). Kernel design for reduced interference distributions. IEEE
Transactions on Signal Processing, 40(2), 402–412.
Lachaux, J-P., Rodriguez, E., Martinerie, J., & Varela, F. J. (1999). Measuring phase synchrony in
brain signals. Human Brain Mapping, 8(4), 194–208.
86 SELIN AVIYENTE
Lelic, D., Gratkowski, M., Hennings, K., & Drewes, A. M. (2011). Multichannel matching pur-
suit validation and clustering—a simulation and empirical study. Journal of Neuroscience
Methods, 196(1), 190–200.
Le Van Quyen, M., Foucher, J., Lachaux, J-P., Rodriguez, E., Lutz, A., Martinerie, J., & Varela, F.
J. (2001). Comparison of Hilbert transform and wavelet methods for the analysis of neuronal
synchrony. Journal of Neuroscience Methods, 111(2), 83–98.
Looney, D., Park, C., Kidmose, P., Ungstrup, M., & Mandic, D. P. (2009). Measuring phase
synchrony using complex extensions of EMD. In 2009 IEEE/SP 15th workshop on statistical
signal processing (pp. 49–52). IEEE.
Makeig, S., Debener, S., Onton, J., & Delorme, A. (2004). Mining event-related brain dynamics.
Trends in Cognitive Sciences, 8(5), 204–210,
Makeig, S., Westerfield, M., Jung, T-P., Enghoff, S., Townsend, J., Courchesne, E., & Sejnowski,
T. J. (2002). Dynamic brain sources of visual evoked responses. Science, 295(5555), 690–694.
Mallat, S. (1999). A wavelet tour of signal processing. Elsevier.
Mallat, S. G. & Zhang, Z. (1993). Matching pursuits with time-frequency dictionaries. IEEE
Transactions on Signal Processing, 41(12), 3397–3415.
Moran, T. P., Bernat, E. M., Aviyente, S., Schroder, H. S., & Moser, J. S. (2015). Sending mixed
signals: Worry is associated with enhanced initial error processing but reduced call for sub-
sequent cognitive control. Social Cognitive and Affective Neuroscience, 10(11), 1548–1556.
Moser, J. S. (2017). The nature of the relationship between anxiety and the error-related nega-
tivity across development. Current Behavioral Neuroscience Reports, 4(4), 309–321.
Mutlu, A. Y. & Aviyente, S. (2011). Multivariate empirical mode decomposition for quantifying
multivariate phase synchronization. EURASIP Journal on Advances in Signal Processing,
2011(1), 615717.
Nunez, P. L. & Srinivasan, R. (2006). Electric fields of the brain: The neurophysics of EEG. Oxford
University Press.
Polikar, R., Topalis, A., Green, D., Kounios, J., & Clark, C. M. (2007). Comparative multiresolution
wavelet analysis of ERP spectral bands using an ensemble of classifiers approach for early diag-
nosis of Alzheimer’s disease. Computers in Biology and Medicine, 37(4), 542–558.
Quiroga, R. Q., Kraskov, A., Kreuz, T., & Grassberger, P. (2002). Performance of different syn-
chronization measures in real data: a case study on electroencephalographic signals. Physical
Review E, 65(4), 041903.
Rihaczek, A. (1968). Signal energy distribution in time and frequency. IEEE Transactions on
Information Theory, 14(3), 369–374.
Rilling, G., Flandrin, P., Gonçalves, P., & Lilly, J. M. (2007). Bivariate empirical mode decom-
position. IEEE signal processing letters, 14(12), 936–939.
Roach, B. J. & Mathalon, D. H. (2008). Event-related EEG time-frequency analysis: An over-
view of measures and an analysis of early gamma band phase locking in schizophrenia.
Schizophrenia Bulletin, 34(5), 907–926.
Rosenblum, M., Pikovsky, A., Kurths, J., Schäfer, C., & Tass, P. A. (2001). Phase synchroniza-
tion: From theory to data analysis. In F. Moss & S. Gielen (Eds.), Handbook of biological
physics: Neuro-informatics and neural modelling, vol. 4 (pp. 279–321). Elsevier.
Rosenblum, M., Tass, P., Kurths, J., Volkmann, J., Schnitzler, A., & Freund, H-J. (2000).
Detection of phase locking from noisy data: Application to magnetoencephalography. In
K. Lehnertz, C. E. Elger, J. Arnhold, & P. Grassberger (Eds.), Proceedings of the workshop on
chaos in brain (pp. 34–51). World Scientific.
TIME-FREQUENCY DECOMPOSITION METHODS 87
Rutkowski, T. M., Mandic, D. P., Cichocki, A., & Przybyszewski, A. W. (2008). EMD approach
to multichannel EEG data-the amplitude and phase synchrony analysis technique. In D-S.
Huang, D. C. Wunsch II, D. S. Levine, & K-H. Jo (Eds.), Proceedings of the international
conference on intelligent computing: Advanced intelligent computing theories and applications.
With aspects of theoretical and methodological issues (pp. 122–129). Springer.
Salinas, E. & Sejnowski, T. J. (2001). Correlated neuronal activity and the flow of neural infor-
mation. Nature Reviews Neuroscience, 2(8), 539.
Sieluzycki, C., Konig, R., Matysiak, A., Kus, R., Ircha, D., & Durka, P. J. (2008). Single-trial
evoked brain responses modeled by multivariate matching pursuit. IEEE Transactions on
Biomedical Engineering, 56(1), 74–82.
Sponheim, S. R., McGuire, K. A., Kang, S. S., Davenport, N. D., Aviyente, S., Bernat, E. M., &
Lim, K. O. Evidence of disrupted functional connectivity in the brain after combat-related
blast injury. Neuroimage, 54, S21–S29, 2011.
Sweeney-Reed, C. M. & Nasuto, S. J. (2007). A novel approach to the detection of synchron-
isation in EEG based on empirical mode decomposition. Journal of Computational
Neuroscience, 23(1), 79–111.
Tağluk, M. E., Çakmak, E. D., & Karakaş, S. (2005). Analysis of the time-varying energy of
brain responses to an oddball paradigm using short-term smoothed Wigner–Ville distribu-
tion. Journal of Neuroscience Methods, 143(2), 197–208.
Tanaka, T. & Mandic, D. P. (2007). Complex empirical mode decomposition. IEEE Signal
Processing Letters, 14(2), 101–104.
Tropp, J. A. & Gilbert, A. C. (2007). Signal recovery from random measurements via orthog-
onal matching pursuit. IEEE Transactions on Information Theory, 53(12), 4655–4666.
Von Stein, A., Chiang, C., & König, P. (2000). Top-down processing mediated by interareal
synchronization. Proceedings of the National Academy of Sciences, 97(26), 14748–14753.
Williams, N., Nasuto, S. J., & Saddy, J. D. (2011). Evaluation of empirical mode decomposition
for event-related potential analysis. EURASIP Journal on Advances in Signal Processing,
2011(1), 965237.
Wu, C-H., Lee, P-L., Shu, C-H., Yang, C-Y., Lo, M-T., Chang, C-Y., & Hsieh, J-C. (2012).
Empirical mode decomposition-based approach for intertrial analysis of olfactory event-
related potential features. Chemosensory Perception, 5(3–4), 280–291.
Wu, Z. & Huang, N. E. (2009). Ensemble empirical mode decomposition: A noise-assisted data
analysis method. Advances in Adaptive Data Analysis, 1(1), 1–41.
Yeung, N., Bogacz, R., Holroyd, C. B., & Cohen, J. D. (2004). Detection of synchronized
oscillations in the electroencephalogram: An evaluation of methods. Psychophysiology,
41(6), 822–832.
Zhao, Y. L., Atlas, E., & Marks, R. J. (1990). The use of cone-shaped kernels for generalized
time-frequency representations of nonstationary signals. IEEE Transactions on Acoustics,
Speech, and Signal Processing, 38(7), 1084–1091.
CHAPTER 5
T IM E FREQU EN C Y A NA LYSE S
IN EVENT-R E L AT E D
P OTENTIAL M ETH OD OL O G I E S
5.1 Introduction
Event-related potentials (ERPs) are powerful tools for measuring the dynamics of
human brain activity, and they have been used for decades to measure sensory, cog-
nitive, motor, and emotion-related processes—as well as individual differences in
these processes—across the lifespan (Cohen, 2014; De Haan, 2013; Hajcak et al., 2012;
Kappenman & Luck, 2016; Luck, 2014). ERPs are defined by voltage fluctuations in the
ongoing electroencephalogram (EEG) that are time-locked to specific events. As we
discuss further in the studies we describe, these events can be the onset of an external
stimulus (e.g., a picture, tone, or feedback about performance), or the generation of
motor response (e.g., a button press).
Typically, tasks are designed so that these events are repeated across multiple trials,
which are then averaged within conditions, presumably canceling out substantial
amounts of trial-level noise and yielding the “prototypical” waveform that is common
across trials (Figure 5.1A shows an example waveform). Specific ERP components are
then identified within these averaged waveforms, and they are quantified numeric-
ally as deviations from a pre-event baseline period. A component is defined “a set of
voltage changes that are consistent with a single neural generator site and that sys-
tematically vary in amplitude across conditions, time, [and] individuals,” (Luck, 2014,
p. 68). Thus, an ERP component is a portion of the overall waveform—often, a peak
deflection in the waveform—that captures the brain response (or set of processes) that
are of interest. ERPs are typically described in terms of their amplitude (measured in
TIME FREQUENCY ANALYSES 89
Figure 5.1 (A) Example data depicting how multiple underlying ERP components (left) con-
tribute to the observed grand averaged ERP waveforms (middle), which are then compared
across conditions (right). (B) ERPs elicited by feedback delivery on the social incentive delay task.
conventions then frequently reflect both the polarity and latency of the component.
Thus, the P3 is the third major positive-going peak in the ERP waveform after a stimulus
is presented (and often peaks in the neighborhood of 300 ms, hence the common alter-
native name “P300”). At times, ERP components receive functionally descriptive names
(e.g., the error-related negativity or ERN).
Because they measure the electrical activity of the brain, the speed of which
approaches the speed of light, ERPs capture neural responses in the time-frame in which
cognition occurs. This millisecond temporal resolution makes it theoretically possible
to isolate dozens of individual neural processes that occur in even very close temporal
proximity. ERPs have thus been particularly useful in studies requiring high temporal
resolution to identify, for example, the transition from sensory-driven stimulus pro-
cessing to higher-order cognitive functions (De Cesarei & Codispoti, 2006; Wiens
et al., 2011), or specific cognitive or affective impairments in a given diagnostic group
(Duncan et al., 2009; Kuperberg et al., 2018).
In comparison with other neuroimaging modalities, EEG techniques are well-
suited as assessment tools. EEG/ERP data collection is relatively efficient, economical,
noninvasive, (Kappenman & Luck, 2016), and is well-tolerated by most participants
across the developmental spectrum (De Haan, 2013). While ERP studies have tradition-
ally been conducted in tightly controlled environments to minimize noise and electrical
interference, advances in technology have allowed for progress in ERP data collection
outside the laboratory and in remote field settings (Tarullo et al., 2017). Additionally,
few contraindications exist for EEG research: for example, people with braces, pregnant
women, and awake infants—all of whom are frequently excluded from MRI studies—
typically can participate in EEG studies, making it feasible to collect neuroimaging data
from large and relatively diverse groups of participants. Finally, multiple studies over the
years have explored the psychometric properties of ERPs and found them to be highly
reliable measures of neural activity, comparable to other common assessment methods
both in terms of internal consistency and test-retest reliability (Baldwin, Larson, &
Clayson, 2015; Ethridge & Weinberg, 2018; Foti et al., 2016; Kujawa et al., 2018; Weinberg
& Hajcak, 2011).
Naturally, these strengths are accompanied by a number of limitations. The
assumption underlying measurement of ERP components in the time domain is that
(a) each component represents distinct sensory and/or cognitive processes (or a small
set of related processes), and (b) components reflect the activity of a single brain region
(or a small network of closely related brain regions). In line with these assumptions,
ERP components are typically scored based on where on the scalp and when in time
specific peaks in the waveform occur, taking the average amplitude within a specified
time window or the peak deflection (Figure 5.1A). This technique reduces the multi-
dimensional EEG signal down to two dimensions, and has many advantages (Cohen,
2014; Luck, 2014). Yet the reality of neural activity is that multiple sensory, cogni-
tive, affective, and motor processes can and do occur simultaneously. Because the
observed trial-averaged waveform represents the sum of all activity measured at a
particular site on the head within a particular time-window, ERPs representing these
TIME FREQUENCY ANALYSES 91
unique processes are summed together in the waveform. Figure 5.1B shows how mul-
tiple observed components that overlap both spatially and temporally contribute to
the grand averaged waveform. Traditional component-scored methods make it dif-
ficult to isolate the contribution each underlying component makes to the observed
average ERP.
Additionally, the electrical signals captured by ERPs are conducted through the brain,
meninges, skull, and scalp, and this signal is subject to spread as it seeks paths of low
resistance; precise identification of primary neural contributors to any one component
is therefore often difficult, particularly for brain regions that are relatively far from the
scalp. Furthermore, neural activity recorded from an electrode exterior to the skull
can reflect the simultaneous and summed activation of many, many thousands—even
millions—of neurons (Luck, 2014). Combined with the difficulty of effectively isolating
different components, this can often make source localization of time-domain ERP
components a dicey proposition (Cohen, 2014).
Processing techniques that isolate unique sources of systematic variance within the
trial-averaged waveform may allow for both more accurate identification of distinct
neuroelectric signals and better description of their anatomical origins. These signals
can be differentiated based on their temporal and spatial variance, as is done with typical
time-window averages as well as more advanced techniques like principal component
analysis (PCA) and independent components analysis (ICA; Dien, Spencer, & Donchin,
2003; Foti, Weinberg, Dien, & Hajcak, 2011; Spencer, Dien, & Donchin, 2001). Critically,
distinct neural signals can also be differentiated based on their spectral properties. This
is because the electrical activity measured by the EEG contains rhythmic oscillations.
These oscillations reflect fluctuations in the activity of populations of neurons, and
the different properties of these oscillations can be helpful in differentiating cognitive
operations.
Oscillations are described by their frequency, power, and phase. Frequency describes
the speed of the oscillation, or how many times a sine wave repeats, or cycles, in a given
period of time, and is measured in hertz (Hz). A wave that repeats four times per second
has a frequency of 4 Hz, a wave that repeats 50 times per second has a frequency of 50
Hz, and a wave that repeats once every two seconds has a frequency of 0.5 Hz. Power is
a measure of how much energy is present in a frequency band and is represented as the
amplitude (or height) of the peaks, squared. Finally, phase is measured in degrees, or
radians, and is a measure of when in time any given part of the sine wave exists. EEG
data, as well as the ERPs derived from these EEG data, are composed of oscillations at
multiple frequencies, each present with different relative power at different time-points.
Time-frequency techniques then attempt to deconvolve, or separate, these signals, to
identify, for example, how much power is present at a given frequency at specific points
in time.
There are many different signal processing techniques available for decomposing
neuroelectric signals to describe the spectral characteristics of ERPs, including moving
window Fourier transforms, wavelet transforms, and Cohen’s class of time-frequency
distributions. These all involve representing a given ERP waveform in terms of a set
92 ANNA WEINBERG, PAIGE ETHRIDGE, BELEL AIT OUMEZIANE, and DAN FOTI
The temporal resolution of ERPs makes them well suited for studying the earliest
stages of sensory processing. One widely studied sensory ERP component is the
MMN, a neural index of change detection that is automatically elicited by a de-
viant stimulus presented within a repetitive sequence (Näätänen, 1995; Näätänen,
Paavilainen, Rinne, & Alho, 2007). While the MMN can be elicited within any
sensory modality, it is commonly studied as part of auditory processing. For ex-
ample, within a sequence of repeating standard tones, the presentation of a tone
that differs with regard to pitch, duration, or some other stimulus characteristics
TIME FREQUENCY ANALYSES 93
will automatically elicit the MMN. The auditory MMN typically occurs between
150 and 250 ms following the deviant stimulus and is maximal at frontocentral
electrodes, with a smaller positive-going potential often apparent at temporal/
mastoid electrodes. While MMN morphology is dependent upon the stimulus
characteristics, studies generally indicate that the MMN emanates from a combin-
ation of activity in supratemporal and frontal cortical regions (i.e., “temporal” and
“frontal” MMN subcomponents), likely related to pre-p erceptual stimulus pro-
cessing and involuntary attentional switch, respectively (Alho, 1995). In addition
to the large basic neuroscience literature applying the MMN to the study of sen-
sory functioning, individual differences in MMN amplitude have been examined
with regard to cognitive decline in aging, and cognitive impairment in psychiatric
disorders (Näätänen et al., 2011). For example, the MMN is reduced by approxi-
mately one standard deviation among individuals with schizophrenia (Umbricht
& Krljes, 2005). This impact of schizophrenia on MMN amplitude is equivalent
to approximately 30 years of cognitive aging (i.e., the MMN of an individual with
schizophrenia at age 20 is comparable to that of a non-psychotic individual at age
50; (Kiang, Braff, Sprock, & Light, 2009). Overall, traditional time-domain analyses
of MMN amplitude have been useful for examining the time course of early sensory
processing and deficits therein.
Complementing these time-domain studies, time-frequency studies have been
useful for isolating the temporal and frontal subcomponents of the MMN, as well as
contextualizing individual differences in MMN amplitude. The frontal portion of the
MMN has been linked primarily to an increase in theta power (i.e., the amplitude of
theta waves across trials), whereas the temporal portion has been linked to theta phase
coherence (i.e., the alignment of theta waves across trials; also see Chapter 9) but not
to theta power (Fuentemilla, Marco-Pallarés, Münte, & Grau, 2008; Ko et al., 2012).
Time-frequency approaches have also helped clarify the nature of impaired sen-
sory processing in schizophrenia. While a reduced frontal MMN in schizophrenia is
well-documented, there is some evidence from time-domain analyses that the tem-
poral (mastoid) subcomponent may be less affected (Baldeweg, Klugman, Gruzelier,
& Hirsch, 2002). Subsequent studies in schizophrenia have used time-frequency
approaches to clarify how abnormal MMN amplitude is explained by alterations in
theta power and phase coherence. For example, one study found that time-domain
MMN amplitude was strongly correlated with frontal theta power among healthy
controls but not among individuals with schizophrenia, suggesting a decoupling be-
tween these signals (Hong, Moran, Du, O’Donnell, & Summerfelt, 2012). Other work
shows that reduced MMN in schizophrenia is characterized by reductions in both
theta power and phase coherence, with differential deficits based on the type of deviant
stimulus (Lee et al., 2017). Thus, time-frequency decomposition of the MMN has been
useful for teasing apart distinct neurological signals involved in sensory processing
that would be difficult to capture within the time domain, with relevance to both basic
science and clinical applications.
94 ANNA WEINBERG, PAIGE ETHRIDGE, BELEL AIT OUMEZIANE, and DAN FOTI
5.3 The P3
The P3 was first reported by Sutton and colleagues (1965) and has since been
among the most well-researched components in the ERP literature (Polich, 2012). It is
among the canonical ERP components in that it is commonly observed across a wide
range of stimuli and laboratory tasks, generally manifesting as the third major positive-
going deflection in the waveform and maximal at parietal electrodes. A P3 is typically
elicited by stimuli that are relatively salient in the local context due to being infrequent,
unexpected, or because they are designated “targets” that are task-relevant. Thus, the P3
is often present within the waveform for most ERP studies, even if it is not the compo-
nent of primary interest. When the P3 is among the primary outcomes measures, one of
the most common laboratory paradigms used is the “oddball” task (Donchin, Ritter, &
McCallum, 1978). During an oddball task, participants are asked to respond to or other-
wise keep track of a designated target stimulus and otherwise disregard other non-target
stimuli (i.e., standard stimuli). In the case of a two-stimulus, auditory oddball task, for
example, two different tones may be presented within an ongoing sequence, with dif-
ferential likelihood (e.g., 80% for standards, 20% for targets). Participants are required
to distinguish between these tones by responding to the occurrence of the target (e.g.,
button press or mentally counting) and not responding to the standard (Polich & Kok,
1995). Discriminating this infrequent target stimulus from the frequently occurring
standard elicits a robust P3, which is increased (i.e., more positive) for the target vs.
standard stimuli (Polich, 2012). Many studies also distinguish between the P3a and P3b
subcomponents, which are clearly differentiated on three-stimuli oddball tasks which
include frequent standard, infrequent target, and infrequent novel stimuli (i.e., a third
stimulus type that is novel and task-irrelevant). In this case, the classic parietal P3 in
response to target stimuli is referred to as the P3b, in order to distinguish it from an
overlapping, frontocentral positivity to novel stimuli that is referred to as the P3a, (or
Novelty P3; Simons, Graham, Miles, & Chen, 2001; N. Squires, Squires, & Hillyard,
1975). Broadly, the P3a is thought to reflect frontal lobe functioning in response to nov-
elty, whereas the P3/P3b is thought to reflect temporal-parietal brain activity associated
with attention and memory processing (Polich, 2007). For the remainder of this chapter,
we generally use the term “P3” rather than “P3b.”
Some theoretical accounts of the P3 within the oddball task posit that it captures
the updating of an individual’s mental representation prompted by incoming stimuli
(i.e., Context Updating Theory; Donchin, 1981). Following an initial sensory input, it
is thought that an attention-driven comparison between the current stimulus and pre-
vious mental representation in working memory is made. If no change in stimulus
attributes have been distinguished, the prior mental representation or “schema” of
the stimulus is maintained, thereby evoking sensory potentials. However, when a
new stimulus attribute is perceived (e.g., a target) the mental representation of the
stimulus context in working memory is updated to elicit the P3 (Donchin et al., 1986).
TIME FREQUENCY ANALYSES 95
Rapidly identifying the mistakes that we make and altering our behavior in response
to these errors is critical for adaptive functioning. The error-related negativity (ERN;
also referred to as the error negativity or Ne) is an ERP component that has commonly
been used to study neural processes associated with identifying and adapting to errors
(Falkenstein, Hohnsbein, Hoormann, & Blanke, 1991; Gehring, Goss, Coles, Meyer,
& Donchin, 1993). The ERN is a response-locked ERP often elicited in speeded reac-
tion tasks that peaks approximately 100 ms following erroneous responses, is max-
imal at frontocentral electrode sites, and is thought to be generated in medial frontal
cortex regions including the anterior cingulate cortex (ACC; for reviews see Gehring
et al., 2011; Holroyd & Coles, 2002; Olvet & Hajcak, 2008). Time-frequency analyses
have contributed to our knowledge of the unique structure of the ERN (Bernat et al.,
2005; Munneke, Nap, Schippers, & Cohen, 2015; Riesel, Weinberg, Moran, & Hajcak,
2012; Yordanova, Falkenstein, Hohnsbein, & Kolev, 2004), have elucidated possible
mechanisms by which it is generated (Luu, Tucker, & Makeig, 2004; Trujillo & Allen,
2007), and have provided empirical support for some existing theories of its functional
significance (Cavanagh, Cohen, & Allen, 2009; Cavanagh, Zambrano‐Vazquez, &
Allen, 2012; Luu & Tucker, 2001; Luu, Tucker, Derryberry, Reed, & Poulsen, 2003).
With regard to the structure of the component, time-frequency analyses have
identified how the ERN is both linked to, and separable from, other related ERP
components (Cavanagh et al., 2012; Di Gregorio, Maier, & Steinhauser, 2018; Gehring
& Willoughby, 2004; Steele et al., 2016; Yordanova et al., 2004). For instance, it has been
debated whether neural responses to errors (ERN) and to correct responses (correct-
related negativity; CRN or Nc), which overlap in time and scalp topography, reflect
the same or distinct processes (Vidal, Hasbroucq, Grapperon, & Bonnet, 2000). While
evidence indicates that the ERN and CRN demonstrate similar characteristics in the
time-domain, frequency analyses have identified error-specific signals with unique
scalp topographies (Yordanova et al., 2004), suggesting that the ERN and CRN are not
identical processes and may have distinct neural generators. Specifically, Yordanova and
TIME FREQUENCY ANALYSES 97
colleagues (2004) identified a delta-frequency component (1.5–3.5 Hz) that was specific
to erroneous responses, as well as a theta-frequency component (4–8 Hz) that emerged
for both erroneous and correct responses but demonstrated different scalp topographies
depending on accuracy and response side (left or right hand). Similarly, time-frequency
analyses have been used to demonstrate distinctions between the ERN and the error-
related positivity (Pe) by demonstrating that the Pe can emerge in the absence of an ERN
(Di Gregorio et al., 2018; Steele et al., 2016), as well as between the ERN and the feedback
negativity (FN, see Section 5.5) by identifying distinct scalp topographies of the ERN
and FN (Gehring & Willoughby, 2004). Nevertheless, theta frequency activity that is
common to several ERP components (ERN, CRN, FN, and N2) has been interpreted
to mean that each of these ERPs reflects similar (though not identical) neural processes
related to performance monitoring and behavioral control (Cavanagh & Shackman,
2015; Cavanagh et al., 2012).
Time-frequency analyses are also uniquely placed to enhance our understanding of
the mechanisms by which the ERN is generated. The classic view of ERPs suggests that
they reflect phasic bursts of neural activation, while more recently it has been suggested
that ERPs might reflect phase-resetting of ongoing oscillatory activity (see Keil &
Thigpen, this volume). Although some argue that time-frequency techniques cannot
distinguish between these two hypotheses (Yeung, Bogacz, Holroyd, & Cohen, 2004;
Yeung, Bogacz, Holroyd, Nieuwenhuis, & Cohen, 2007), evidence suggests that the
ERN is best explained by the combination of an amplitude increase and phase-resetting
(Trujillo & Allen, 2007). This is an important contribution, partly because under-
standing how the ERN is generated provides insight into its functional significance. For
example, Cavanagh and colleagues (2009) suggested that theta oscillatory dynamics
reflected in the ERN may represent a mechanism by which neural regions responsible
for performance monitoring and for cognitive control communicate with each other
(see Chapter 3 for a discussion of neural oscillations; Keil & Thigpen).
As with all of the ERPs discussed in this chapter, understanding how the ERN is
associated with individual difference variables is an active area of research. Time-
window analyses of ERN responses have identified links between the ERN and multiple
forms of psychopathology, including generalized anxiety disorder (GAD) (Weinberg,
Olvet, & Hajcak, 2010), obsessive compulsive disorder (Endrass et al., 2010; Riesel,
Kathmann, & Endrass, 2014), substance use disorder (Euser, Evans, Greaves-Lord,
Huizink, & Franken, 2013), and depression (Weinberg, Liu, & Shankman, 2016). Time-
frequency analyses have begun to meaningfully advance our understanding of such
individual differences. For instance, Cavanagh and colleagues (2017) determined that
error-related theta power was a unique predictor of GAD status over and above the
time-window scored ERN, and when they included theta network dynamics in their
model, they were able to classify clinical participants with an impressive 66% accuracy.
These data suggest that time-frequency analyses may have a powerful role to play in
using neural data to classify clinical populations. While time-frequency analyses have
led to several critical advances in our understanding of the ERN, as noted here, this is
likely to remain a fruitful avenue of continued study.
98 ANNA WEINBERG, PAIGE ETHRIDGE, BELEL AIT OUMEZIANE, and DAN FOTI
might then summate to create the observed negative-going component following feed-
back indicating losses that had typically been observed in the time domain (i.e., the
classic FN/FRN/MFN).
Studies using time-frequency methods to explore neural activity in this time-window
have tended to bear this out, as there is evidence that activity in both the theta-and
delta-bands make unique contributions to the amplitude of the FN/RewP (Bernat
et al., 2015; Bernat et al., 2011). In particular, theta activity in this time range appears
to be primarily sensitive to loss/negative outcomes, with losses eliciting enhanced ac-
tivity compared to gains in this frequency range (Bernat, Nelson, Holroyd, Gehring, &
Patrick, 2008; Bernat et al., 2011; Gheza, De Raedt, Baeken, & Pourtois, 2018; Harper
et al., 2011; L. Nelson, Patrick, Collins, Lang, & Bernat, 2011; Olson, Harper, Golosheykin,
Bernat, & Anokhin, 2011; Webb et al., 2017). Delta has more often been linked to activity
in the time-range of the P3 that follows the FN/RewP at more parietal sites (Gehring
& Willoughby, 2002; Holroyd & Coles, 2002; Miltner et al., 1997). However, there is
increasing evidence that delta activity also underlies the observed reward positivity
occurring in the time-range and spatial location of the FN/RewP (Bernat et al., 2015;
Bernat et al., 2008; Foti, Weinberg, Bernat, & Proudfit, 2014; Harper et al., 2011). Indeed,
delta activity in this time-range is enhanced for gains compared to losses (Bernat et al.,
2015; Bernat et al., 2008; Bernat et al., 2011; Cavanagh, 2015; Cavanagh, Masters, Bath,
& Frank, 2014; Foti, Weinberg, et al., 2014; Gheza et al., 2018; Harper et al., 2011; Leicht
et al., 2013; L. Nelson et al., 2011; Olson et al., 2011; Pornpattananangkul & Nusslock,
2016; Webb et al., 2017), and appears to drive the reward positivity observed in the trial-
averaged waveforms (Bernat et al., 2015; Foti et al., 2011; Harper et al., 2011).
The activity of these two frequency bands also appears to be dissociable, insofar as the
differences between loss and gain activity in the theta and delta band tend to be at best
weakly correlated (e.g., Bernat et al., 2015; Bernat et al., 2011; Foti, Weinberg, et al., 2014).
Moreover, source analysis suggests unique generators, with loss-related theta localizing
to the ACC and gain-related delta to a possible source in the basal ganglia (Foti,
Weinberg, et al., 2014). Combined, these data suggest that changes in theta and delta are
not yoked expressions of the same underlying process, but instead may represent dis-
tinct cognitive processes and contributions to the FN/RewP (Bernat et al., 2015; Bernat
et al., 2008; Bernat et al., 2011; Gheza et al., 2018; Harper et al., 2011; Olson et al., 2011).
Consistent with this, theta response to losses appears to reflect a relatively low-level re-
sponse to negative outcomes that is frequently insensitive to stimulus parameters and
experimental context (Bernat et al., 2015; Bernat et al., 2011; Harper et al., 2011; Watts,
Bachman, & Bernat, 2017; Watts & Bernat, 2018). This increased theta (Section 5.4)
appears to then act as a signal to recruit attentional and executive resources to respond
to mistakes, novelty, and negative feedback (Aviyente, Tootell, & Bernat, 2017; Cavanagh
& Frank, 2014; Cavanagh et al., 2012; Van Noordt, Campopiano, & Segalowitz, 2016; van
Noordt, Desjardins, Gogo, Tekok-Kilic, & Segalowitz, 2017). In contrast, in monetary
reward paradigms, delta appears sensitive not only to loss vs. gain distinctions, but also
higher-level secondary stimulus attributes, such as magnitude (Bernat et al., 2015), con-
text (Bernat et al., 2015; Watts & Bernat, 2018), expectancy violations (Cavanagh, 2015;
100 ANNA WEINBERG, PAIGE ETHRIDGE, BELEL AIT OUMEZIANE, and DAN FOTI
Gheza et al., 2018; Watts et al., 2017), and reward uncertainty (Gheza et al., 2018) and
thus may reflect more elaborative processing of the feedback beyond the most salient
dimension of the stimuli (see, however: Leicht et al., 2013).
Identification of dissociable and functionally distinct processes in the time-window
of the FN/RewP has also proven to be useful in studies seeking to understand indi-
vidual differences in neural responses to feedback, including describing develop-
mental influences on more specific neural processes, as well as identifying specific
cognitive-affective deficits in clinical samples. For instance, there is evidence from
MRI research that brain regions supporting the neural response to rewards undergo
considerable developmental changes from childhood through adolescence and into
adulthood. In particular, fMRI studies tend to find an adolescent-specific peak in
striatal activation to reward feedback (Braams, van Duijvenvoorde, Peper, & Crone,
2015; J. R. Cohen et al., 2010; Ernst et al., 2005; Galvan et al., 2006; Somerville, Hare,
& Casey, 2011; Van Leijenhorst et al., 2009). Most studies examining developmental
changes in the FN/RewP, however, have failed to find evidence for an adolescent-
specific peak, or indeed even significant developmental changes (Kujawa et al.,
2018; Larson, South, Krauskopf, Clawson, & Crowley, 2011; Lukie, Montazer-Hojat,
& Holroyd, 2014; Santesso, Dzyundzyak, & Segalowitz, 2011); but see also (Arbel,
McCarty, Goldman, Donchin, & Brumback, 2018). In a recent study of 8–17-year-old
participants, however, Bowers and colleagues (2018) examined both time-domain
ERPs and time-frequency-derived theta and delta band activity. They found that,
consistent with previous work, the time-domain-scored RewP was not associated
with participants’ age. Yet, theta power—which was enhanced for losses relative to
gains—decreased with age, whereas delta power—which was greater for gains than
losses—increased. In a study investigating whether family history of psychopath-
ology might influence these normative developmental effects, we also collected
a sample of never-depressed daughters of mothers with or without a history of de-
pression (Ethridge et al., 2021). In this sample, we found that the association between
delta power and daughters’ developmental stage differed depending on maternal risk
status. For daughters of mothers who had never been depressed, increased develop-
ment was associated with increases in delta power following rewards. For daughters
of mothers with a history of depression, increased development was associated with
decreased delta power following rewards, suggesting high-risk daughters may become
increasingly vulnerable across the course of adolescence. This effect was not observed
for power in the theta frequency.
Work identifying functional differences between delta and theta activation in re-
sponse to external feedback has laid the foundation for identifying specific cognitive and
affective deficits in various psychopathological groups. For instance, there is extensive
data suggesting that individuals high on externalizing-proneness (including alcohol and
substance use/abuse, rule-breaking, and personality measures) show broad and non-
specific amplitude reductions in multiple ERPs (Hall, Bernat, & Patrick, 2007; Polich
et al., 1994). This work has been further clarified by time-frequency decompositions
suggesting these individuals do not in fact show deficits related to theta power elicited
TIME FREQUENCY ANALYSES 101
Schryer-Praga, & Foti, 2017). ERP data were collected from 27 adults during a social
incentive delay task, a modified version of the commonly used monetary incentive
delay task (Knutson, Westdorp, Kaiser, & Hommer, 2000; B. K. Novak, Novak, Lynam,
& Foti, 2016; K. D. Novak & Foti, 2015). The task is designed to tease apart anticipa-
tory and consummatory stages of reward processing. On each trial, participants are first
presented with a cue indicating whether it is an incentive or neutral trial. On incen-
tive trials, participants have the opportunity to earn a reward by rapidly responding to
a target stimulus (and are punished for slow reaction times), whereas on neutral trials
participants break even regardless of their reaction time. Following their behavioral re-
sponse to the target stimulus, participants are then shown feedback indicating the result
of that trial (i.e., win or loss on incentive trials, break-even on neutral trials). On the
traditional monetary version of the task the rewards are nominal amounts of money,
yet more recently a social reward version has been developed in which the “wins” are
instead positive social feedback putatively administered by the experimenter. In a re-
cent study, it was shown that social reward feedback elicits a RewP and feedback-P3
(described as such in order to differentiate it from the oddball P3, described earlier) that
is of similar morphology and amplitude to monetary reward feedback (Ait Oumeziane
et al., 2017). These original analyses focused exclusively on traditional time-domain
approaches to scoring the RewP and feedback-P3. Here, we revisit these data using time-
frequency analysis.
The time-domain ERP waveforms and scalp topographies are presented in Figure
5.1A. Consistent with previous research, social reward (“wins”) elicited a more
positive-going waveform than social punishment (“losses”). The RewP difference
score (contrasting the valence of uncertain feedback: win vs. loss) peaked at 325 ms,
whereas the feedback-P3 difference score (contrasting the salience of uncertain vs.
certain feedback: win vs. break-even, loss vs. break-even) peaked at approximately 400
ms. A challenge in isolating the RewP and feedback-P3 in this case, however, is that
the waveform to social rewards is modulated within a relatively broad time window
of approximately 200–800 ms at centroparietal electrodes, thereby encompassing
the P2-RewP-P3-Slow Wave complex. It is unclear whether the difference between
conditions represents multiple, overlapping effects or is instead better interpreted as
a single, sustained increase in the ERP waveform. These possibilities are difficult to
disentangle using traditional time-window averages but can be addressed using time-
frequency analysis. Building upon previous time-frequency decompositions of the
RewP (Bernat et al., 2011; Foti et al., 2015), we sought to isolate three distinct neural
responses:
For illustrative purposes, we focus here on the P3 to wins; a similar pattern of results was
also observed for the P3 to losses.
Standard signal processing procedures were applied to the data, including re-
referencing to the mastoid electrodes, filtering from 0.1–30 Hz, ocular correction,
and artifact rejection (for details, see Ait Oumeziane et al., 2017). The continuous
EEG was segmented relative to the onset of feedback stimuli using a relatively broad
time window of −1500 to 1500 ms, allowing for edge artifacts (i.e., distortion of the
signal near the edges of the window following time-frequency transform; note this
is a broader time-window than is typical for most ERP research). ERP averages were
created for each of the three conditions (win, loss, break-even), and then the time-
frequency transform was applied. Complex Morlet wavelets were calculated within
BrainVision Analyzer (Brain Products), using a frequency range of 0.5–20 Hz and
linear steps of 0.25 Hz. A relatively narrow frequency range was chosen here due to the
a priori focus on activity in the delta and theta bands, whereas a broader range could be
chosen for more exploratory analyses. The Morlet parameter was set at c =3.5 (i.e., 3.5
cycles in the wavelet), and wavelet functions were normalized using Gabor normaliza-
tion. Baseline correction was performed for each wavelet layer using a baseline window
of −500 to −300 ms; the average amplitude in the baseline window was subtracted
from each time point in the layer. These time-frequency transforms were applied to
the averaged ERP data for each subject and then averaged across subjects to create the
grand averaged spectrogram.
Of interest were two contrasts: wins vs. losses, which ought to elicit reward-related
delta (RewP) and loss-related theta (FN) from approximately 200–400 ms; and win vs.
break even, which ought to elicit delta-band activity in the time range of the feedback-
P3 (300–500 ms). The spectrogram for the win vs. loss contrast is presented in Figure
5.2. As expected, activity in the delta frequency band was increased for wins vs. losses
(the red portion of the spectrogram within the delta band), which peaked at approxi-
mately 2.5 Hz and 295 ms. Overlapping with this response was an increase in activity
within the theta frequency band for losses vs. wins (the blue portion of the spectrogram
within the theta band), which peaked at approximately 5.25 Hz and at 235 ms. Both of
these responses occurred within the time range of the FN/RewP but at distinct fre-
quency bands, presenting the opportunity to extract separate scores for each. We scored
the delta and theta responses by extracting wavelet layers centered at 2.5 Hz and 5.25 Hz,
respectively. Wavelet layers yield a waveform representing power at that frequency band
over time, which can then be scored by taking a time-window average of peak score
akin to time-domain ERPs (Figure 5.2, right). We scored each response as the average
power within a 50-ms window surrounding the peak of the win minus loss difference,
which was somewhat different for each frequency band: 270–320 ms for delta and 210–
260 ms for theta. Notably, these wavelets each exhibited more focal scalp topographies
than the time-domain FN/RewP, with maxima at frontocentral electrodes. To maintain
consistency with the time-domain FN/RewP, we scored the delta and theta responses
at electrode Cz. The effects of condition (win vs. loss) were significant for both delta
(t(26) =3.04, p < .01) and theta (t(26) =3.11, p < .01). Critically, the difference scores
104 ANNA WEINBERG, PAIGE ETHRIDGE, BELEL AIT OUMEZIANE, and DAN FOTI
Theta
Delta layer at 2.5 Hz
Delta
Delta
Figure 5.2 Time-frequency analysis of the reward positivity elicited by the social incen-
tive delay task. Top: The contrast of wins versus losses, isolating delta-and theta-band activity
corresponding to the time-domain FN/RewP. Bottom: The contrast of wins versus break-even
outcomes, isolating delta-band activity corresponding to the feedback-P3.
for delta and theta were also both significantly correlated with the time-domain RewP
difference score (Table 5.1). As is common in the literature, the delta and theta effects
were not significantly correlated with one another, further suggesting that they each
capture a distinct portion of the FN/RewP.
The analogous spectrogram for the win vs. break-even contrast is presented in Figure
5.2. In comparison with the FN/RewP analyses, here we expected to find prominent
delta-band activity corresponding to the feedback-P3. As seen in the spectrogram,
activity in the delta band was indeed increased for wins vs. break-even outcomes, an
effect which peaked at 1.75 Hz and at 365 ms. We extracted the wavelet layer at 1.75 Hz
and scored this delta response as the average power from 340–390 ms. This wavelet
exhibited a scalp topography highly similar to the time-domain feedback-P3, with a
peak at centroparietal electrodes. To be consistent with our time-domain analyses, we
scored this delta effect at electrode Pz. The increase in delta-band activity for win vs.
break-even outcomes was statistically significant (t(26) =7.90, p < .001), and the delta-
band difference score was significantly correlated with the time-domain feedback-P3
difference score (Table 5.1).
TIME FREQUENCY ANALYSES 105
Note: FN/RewP variables are the contrast of win vs. loss; feedback-P3 variables are the contrast of win
vs. break-even. Coefficients are Spearman’s rho. Values in bold are significant at p < .05.
It is notable that superior separation of the FN/RewP and feedback-P3 was achieved
in this case by using time-frequency analyses as compared to traditional time-domain
analyses. Specifically, the time-domain FN/RewP and feedback-P3 were significantly
correlated with one another despite being scored at non-overlapping time windows
(300–350 ms and 370–420 ms) and electrodes (Cz and Pz). This was not the case for the
time-frequency measures: the FN-theta and RewP-delta scores were not significantly
correlated with the time-domain feedback-P3 score, and P3-delta scores were not sig-
nificantly correlated with the time-domain FN/RewP scores. We further tested this
pattern of specificity using multiple regression, predicting each time-domain ERP from
the three time-frequency variables. When entered as simultaneous predictors, time-
domain FN/RewP amplitude was significantly predicted by a combination of RewP-
delta (β =.46, p < .05) and FN-theta (β =−.38, p < .05), but not P3-delta (β =.11, p =.53).
Conversely, time-domain feedback-P3 amplitude was significantly predicted only by
P3-delta (β =.66, p < .001) and not by RewP-delta (β =.15, p=.33) or FN-theta (β= −.15,
p =.28). Thus, time-frequency analysis facilitated the isolation of three distinct neural
signals involved in social reward processing that would have been difficult to achieve
solely within the time domain, where signal overlap is more problematic in this experi-
mental context.
The example described shows how time-frequency analyses can aid in the interpret-
ation of traditional ERP effects. In these data, positive social feedback modulated the
ERP waveform in a time-range spanning the FN/RewP and P3, which overlapped with
each other. The traditional time-domain approach of scoring these ERPs as the average
amplitude within separate time windows could not adequately address the problem of
component overlap. Time-frequency decomposition, on the other hand, was able to iso-
late three distinct signals: RewP-delta, FN-theta, and P3-delta. Similar to the time do-
main, these three signals were all sensitive to outcome type: RewP-delta and FN-theta
106 ANNA WEINBERG, PAIGE ETHRIDGE, BELEL AIT OUMEZIANE, and DAN FOTI
were modulated by outcome valence, and P3-delta was modulated by outcome certainty.
In contrast with the time domain, however, these three signals achieved superior sep-
aration: RewP-delta and FN-theta each captured unique portions of the time-domain
FN/RewP component, and they were unrelated to the time-domain P3 component.
Likewise, P3-delta was strongly related to the time-domain P3 component but was unre-
lated to the time-domain FN/RewP component. Overall, time-frequency analyses help
us rule out the interpretation that social reward broadly increases the ERP waveform in
a nonspecific fashion; instead, this broad modulation clearly represents a composite of
multiple neural signals. This also lays the groundwork for more precise characterization
of individual differences, whereby we would expect specific associations with the time-
frequency variables (which are relatively uncorrelated) as compared to the time-domain
variables (which have substantial overlap).
Comprehensive descriptions of different methodological approaches are covered
in other chapters (e.g., Chapter 4, this volume). However, one important step to con-
sider is whether to conduct these analyses on single-trial or averaged data, as was done
in our practical description earlier. As noted, ERPs are typically derived from averages
composed of data from multiple trials. This practice reflects the typically low signal-
to-noise ratio of ERPs: the amplitude of many ERPs is no more than a few µV, while
the amplitude of “background” neuroelectric signals (see chapter 6, this volume),
other electrophysiological signals, and electrical interference from non- biological
sources can be closer to tens of µV. When averaged together, this presumably randomly
distributed noise cancels itself out, whereas the systematic signal, or ERPs, in the data
will not, resulting in a legible ERP component. Single-trial analysis of ERP data typically
requires relatively large-amplitude components (e.g., the P3). However, time-frequency
techniques can in some cases make data more amenable to single-trial analysis, as
electrical “noise” often has distinct spectral properties from the signal of interest. For
instance, in studies interested in how error or feedback processing might influence trial-
level adjustments of behavior, time-frequency techniques could allow the researcher to
focus narrowly on theta and delta power on each trial, with higher frequency activity
(e.g., alpha, 60-Hz line noise) isolated from these signals. Nonetheless, it is common
practice to conduct time-frequency decompositions on averages of many trials. And be-
cause averages improve signal-to-noise ratio, they are also a helpful basis for identifying
spectral power at different frequency ranges, particularly in exploratory research where
the spectral characteristics of ERPs are less well-understood. As is often the case, the op-
timal method will depend on the research question.
5.7 Conclusions
ERPs are powerful and flexible tools for understanding sensory, cognitive, and af-
fective processes in the brain, but they are not without limitations. As discussed, time-
frequency techniques can be helpful in addressing some of these limitations. They can
TIME FREQUENCY ANALYSES 107
better leverage the multi-dimensional nature of EEG data—and, in principle, better rep-
resent the underlying neural signals—by accounting for not only voltage changes over
time and site on the scalp, but also frequency, power, and phase, and may reveal mul-
tiple dissociable processes folded within the time-window-scored ERP. However, time-
frequency techniques are not a magic bullet. As is discussed at length in other chapters,
time-frequency decompositions can reduce temporal precision, a chief advantage of the
ERP technique (though, as described in Chapter 4, there are techniques to minimize
this loss). When applied to trial-averaged ERP data, time-frequency decompositions
are best thought of conceptually as isolating potentially relevant ERP subcomponents.
That is, it should be possible to “recreate” the pattern of observed ERP findings within
the time-frequency domain, perhaps based on a combination of activity across multiple
frequency bands. The time-frequency measures may be more precise than their time-
domain counterparts to the extent that they isolate the relevant neural signal of interest,
but in cases where there is no apparent modulation of the ERP waveform across ex-
perimental conditions, it is unlikely that time-frequency approaches will uncover “new”
findings. Therefore, we encourage the reader to think of time-frequency decompositions
applied to ERP data as the conceptual equivalent of deriving subscales within a self-
report questionnaire, which can often enhance the precision of measurement and help
clarify the nature of associations with other measures. This is particularly useful where
there are already well-established associations between a time-domain ERP and an ex-
ternal measure (e.g., clinical diagnosis, behavior, or personality trait), whereby time-
frequency decompositions can help explain the nature of those associations. Thus,
time-frequency analyses can be an effective approach for revisiting published data or
extending the results of prior studies.
REFERENCES
Ait Oumeziane, B., Schryer- Praga, J., & Foti, D. (2017). “why don't they ‘like’me
more?”: Comparing the time courses of social and monetary reward processing.
Neuropsychologia, 107, 48–59.
Alho, K. (1995). Cerebral generators of mismatch negativity (MMN) and its magnetic counter-
part (MMNm) elicited by sound changes. Ear and hearing, 16(1), 38–51.
Arbel, Y., McCarty, K. N., Goldman, M., Donchin, E., & Brumback, T. (2018). Developmental
changes in the feedback related negativity from 8 to 14 years. International Journal of
Psychophysiology.
Aviyente, S., Tootell, A., & Bernat, E. M. (2017). Time-frequency phase-synchrony approaches
with ERPs. International Journal of Psychophysiology, 111, 88–97.
Baker, T., & Holroyd, C. (2011). Dissociated roles of the anterior cingulate cortex in reward and
conflict processing as revealed by the feedback error-related negativity and N200. Biological
Psychology, 87, 25–34.
Baldeweg, T., Klugman, A., Gruzelier, J. H., & Hirsch, S. R. (2002). Impairment in frontal but
not temporal components of mismatch negativity in schizophrenia. International Journal of
Psychophysiology, 43(2), 111–122.
108 ANNA WEINBERG, PAIGE ETHRIDGE, BELEL AIT OUMEZIANE, and DAN FOTI
Baldwin, S. A., Larson, M. J., & Clayson, P. E. (2015). The dependability of electrophysiological
measurements of performance monitoring in a clinical sample: A generalizability and deci-
sion analysis of the ERN and P e. Psychophysiology, 52(6), 790–800.
Başar-Eroglu, C., Başar, E., Demiralp, T., & Schürmann, M. (1992). P300-response: possible
psychophysiological correlates in delta and theta frequency channels. A review. International
Journal of Psychophysiology, 13(2), 161–179.
Bernat, E., Malone, S., Williams, W., Patrick, C., & Iacono, W. (2007). Decomposing delta, theta,
and alpha time-frequency ERP activity from a visual oddball task using PCA. International
Journal of Psychophysiology, 64(1), 62–74.
Bernat, E., Nelson, L., & Baskin-Sommers, A. (2015). Time-Frequency Theta and Delta
Measures Index Separable Components of Feedback Processing in a Gambling Task.
Psychophysiology, 52(5), 626–637.
Bernat, E., Nelson, L., Holroyd, C., Gehring, W., & Patrick, C. (2008). Separating cognitive
processes with principal components analysis of EEG time-frequency distributions. Paper
presented at the Proceedings of SPIE.
Bernat, E., Nelson, L., Steele, V., Gehring, W., & Patrick, C. (2011). Externalizing psychopath-
ology and gain–loss feedback in a simulated gambling task: Dissociable components of brain
response revealed by time-frequency analysis. Journal of abnormal psychology, 120(2), 352.
Bernat, E., Williams, W., & Gehring, W. (2005). Decomposing ERP time-frequency energy
using PCA. Clinical Neurophysiology, 116(6), 1314–1334.
Bogdan, R., Santesso, D., Fagerness, J., Perlis, R., & Pizzagalli, D. (2011). Corticotropin-
Releasing Hormone Receptor Type 1 (CRHR1) Genetic Variation and Stress Interact to
Influence Reward Learning. The Journal of Neuroscience, 31(37), 13246–13254.
Bress, J., Smith, E., Foti, D., Klein, D., & Hajcak, G. (2011). Neural response to reward and de-
pressive symptoms in late childhood to early adolescence. Biological Psychology, 89(1), 156–
162. doi:doi.org/10.1016/j.biopsycho.2011.10.004
Carlson, J., Foti, D., Harmon-Jones, E., Mujica-Parodi, L., & Hajcak, G. (2011). The neural pro-
cessing of rewards in the human striatum: Correlation of fMRI and event-related potentials.
NeuroImage, 57, 1608–1616. doi:10.1016/j.neuroimage.2011.05.037
Cavanagh, J. F. (2015). Cortical delta activity reflects reward prediction error and related behav-
ioral adjustments, but at different times. NeuroImage, 110, 205–216.
Cavanagh, J. F., Bismark, A. W., Frank, M. J., & Allen, J. J. (2018). Multiple Dissociations be-
tween Comorbid Depression and Anxiety on Reward and Punishment Processing: Evidence
from Computationally Informed EEG. Computational Psychiatry, 1–17.
Cavanagh, J. F., Cohen, M. X., & Allen, J. J. B. (2009). Prelude to and resolution of an error: EEG
phase synchrony reveals cognitive control dynamics during action monitoring. Journal of
Neuroscience, 29(1), 98–105. doi:10.1523/jneurosci.4137-08.2009
Cavanagh, J. F., & Frank, M. J. (2014). Frontal theta as a mechanism for cognitive control.
Trends in cognitive sciences, 18(8), 414–421.
Cavanagh, J. F., Masters, S. E., Bath, K., & Frank, M. J. (2014). Conflict acts as an implicit cost in
reinforcement learning. Nature communications, 5, 5394.
Cavanagh, J. F., & Shackman, A. J. (2015). Frontal midline theta reflects anxiety and cogni-
tive control: Meta-analytic evidence. Journal of Physiology –Paris, 109, 3–15. doi:10.1016/
j.jphysparis.2014.04.003
Cavanagh, J. F., Zambrano‐Vazquez, L., & Allen, J. J. (2012). Theta lingua franca: A common
mid‐frontal substrate for action monitoring processes. Psychophysiology, 49(2), 220–238.
Cohen, M. (2014). Analyzing neural time series data: theory and practice: MIT press.
TIME FREQUENCY ANALYSES 109
Cohen, M., Elger, C., & Ranganath, C. (2007). Reward expectation modulates feedback-related
negativity and EEG spectra. NeuroImage, 35(2), 968–978.
Di Gregorio, F., Maier, M. E., & Steinhauser, M. (2018). Errors can elicit an error positivity
in the absence of an error negativity: Evidence for independent systems of human error
monitoring. NeuroImage, 172, 427–436. doi:10.1016/j.neuroimage.2018.01.081
Dien, J., Spencer, K. M., & Donchin, E. (2003). Localization of the event-related potential nov-
elty response as defined by principal components analysis. Cognitive Brain Research, 17(3),
637–650.
Donchin, E. (1981). Surprise!. . . surprise? Psychophysiology, 18(5), 493–513.
Donchin, E., Ritter, W., & McCallum, W. (1978). Cognitive psychophysiology: The endogenous
components of the ERP. Event-related brain potentials in man, 349–411.
Duncan‐Johnson, C. C., & Donchin, E. (1977). On quantifying surprise: The variation of event‐
related potentials with subjective probability. Psychophysiology, 14(5), 456–467.
Endrass, T., Schuermann, B., Kaufmann, C., Spielberg, R., Kniesche, R., & Kathmann,
N. (2010). Performance monitoring and error significance in patients with obsessive-
compulsive disorder. Biological Psychology, 84, 257–263. doi:10.1016/j.biopsycho.2010.02.002
Ergen, M., Marbach, S., Brand, A., Başar-Eroğlu, C., & Demiralp, T. (2008). P3 and delta
band responses in visual oddball paradigm in schizophrenia. Neuroscience letters, 440(3),
304–308.
Ethridge, P., & Weinberg, A. (2018). Psychometric Properties of Neural Responses to Monetary
and Social Rewards across Development. International Journal of Psychophysiology, in press.
Euser, A. S., Evans, B. E., Greaves-Lord, K., Huizink, A. C., & Franken, I. H. A. (2013).
Diminished error-related brain activity as a promising endophenotype for substance use
disorders: Evidence from high-risk offspring. Addiction Biology, 18(970-984). doi:10.1111/
adb.12002
Falkenstein, M., Hohnsbein, J., Hoormann, J., & Blanke, L. (1991). Effects of crossmodal
divided attention on late ERP components: II. Error processing in choice reaction
tasks. Electroencephalography & Clinical Neurophysiology, 78(6), 447– 455. doi:10.1016/
0013-4694(91)90062-9
Foti, D., Carlson, J., Sauder, C., & Proudfit, G. (2014). Reward dysfunction in major de-
pression: Multimodal neuroimaging evidence for refining the melancholic phenotype.
NeuroImage, 101, 50–58.
Foti, D., & Hajcak, G. (2009). Depression and reduced sensitivity to non-rewards versus
rewards. Biological Psychology, 81(1), 1–8. doi:10.1016/j.biopsycho.2008.12.004,
Foti, D., Perlman, G., Hajcak, G., Mohanty, A., Jackson, F., & Kotov, R. (2016). Impaired error
processing in late-phase psychosis: Four-year stability and relationships with negative
symptoms. Schizophrenia research, 176(2-3), 520–526.
Foti, D., & Weinberg, A. (2018). Reward and feedback processing: State of the field, best
practices, and future directions. In.
Foti, D., Weinberg, A., Bernat, E., & Proudfit, G. (2014). Anterior cingulate activity to mon-
etary loss and basal ganglia activity to monetary gain uniquely contribute to the feedback
negativity. Clinical Neurophysiology.
Foti, D., Weinberg, A., Bernat, E., & Proudfit, G. (2015). Anterior cingulate activity to monetary
loss and basal ganglia activity to monetary gain uniquely contribute to the feedback nega-
tivity. Clin Neurophysiol, 126(7), 1338–1347. doi:10.1016/j.clinph.2014.08.025
Foti, D., Weinberg, A., Dien, J., & Hajcak, G. (2011). Event‐related potential activity in the basal
ganglia differentiates rewards from nonrewards: Temporospatial principal components
110 ANNA WEINBERG, PAIGE ETHRIDGE, BELEL AIT OUMEZIANE, and DAN FOTI
analysis and source localization of the feedback negativity. Human brain mapping, 32(12),
2207–2216.
Fuentemilla, L., Marco-Pallarés, J., Münte, T., & Grau, C. (2008). Theta EEG oscillatory activity
and auditory change detection. Brain research, 1220, 93–101.
Gehring, W., Goss, B., Coles, M. G. H., Meyer, D. E., & Donchin, E. (1993). A neural system for
error detection and compensation. Psychological Science, 4(6), 385–390. doi:10.1111/j.1467-
9280.1993.tb00586.x
Gehring, W., Liu, Y., Orr, J. M., & Carp, J. (Eds.). (2011). The error-related negativity (ERN/Ne).
New York, NY: Oxford University Press.
Gehring, W., & Willoughby, A. (2002). The medial frontal cortex and the rapid pro-
cessing of monetary gains and losses. Science, 295(5563), 2279–2282. doi:10.1126/
science.1066893
Gehring, W., & Willoughby, A. (2004). Are all medial frontal negativities created equal?
Toward a richer empirical basis for theories of action monitoring. Errors, conflicts, and the
brain. Current opinions on performance monitoring, 14, 20.
Gheza, D., De Raedt, R., Baeken, C., & Pourtois, G. (2018). Integration of reward with cost
anticipation during performance monitoring revealed by ERPs and EEG spectral
perturbations. NeuroImage, 173, 153–164.
Gilmore, C., Malone, S., Bernat, E., & Iacono, W. (2010). Relationship between the P3 event‐
related potential, its associated time‐frequency components, and externalizing psychopath-
ology. Psychophysiology, 47(1), 123–132.
Hajcak, G., Moser, J., Holroyd, C., & Simons, R. (2006). The feedback-related negativity reflects
the binary evaluation of good versus bad outcomes. Biological Psychology, 71(2), 148–154.
doi:10.1016/j.biopsycho.2005.04.001
Hall, J. R., Bernat, E. M., & Patrick, C. J. (2007). Externalizing psychopathology and the error-
related negativity. Psychological science, 18(4), 326–333.
Harper, J., Olson, L., Nelson, L., & Bernat, E. (2011). Delta-Band Reward Processing Occurring
During the Feedback-Related Negativity (FRN). Poster presented at annual meeting of the
Society for Psychophysiological Research. Boston, MA.
Herrmann, C. S., Rach, S., Vosskuhl, J., & Strüber, D. (2014). Time–frequency analysis of event-
related potentials: a brief tutorial. Brain topography, 27(4), 438–450.
Hewig, J., Kretschmer, N., Trippe, R., Hecht, H., Coles, M., Holroyd, C., & Miltner, W. (2010).
Hypersensitivity to reward in problem gamblers. Biological psychiatry, 67(8), 781–783.
Holroyd, C., & Coles, M. (2002). The neural basis of human error processing: Reinforcement
learning, dopamine, and the error-related negativity. Psychological review, 109(4), 679–709.
doi:10.1037//0033-295X.109.4.67
Holroyd, C., Krigolson, O., & Lee, S. (2011). Reward positivity elicited by predictive cues.
Neuroreport, 22(5), 249.
Holroyd, C., Nieuwenhuis, S., Yeung, N., & Cohen, J. (2003). Errors in reward prediction are
reflected in the event-related brain potential. Neuroreport, 14(18), 2481–2484. doi:10.10 97/
01.wnr.0 0 0 0 0 9 9 6 01.414 03.a
Holroyd, C., Pakzad-Vaezi, K., & Krigolson, O. (2008). The feedback correct-related posi-
tivity: sensitivity of the event-related brain potential to unexpected positive feedback.
Psychophysiology, 45(5), 688–697. doi:10.1111/j.1469-8986.2008.00668.x
Hong, L. E., Moran, L. V., Du, X., O’Donnell, P., & Summerfelt, A. (2012). Mismatch negativity
and low frequency oscillations in schizophrenia families. Clinical Neurophysiology, 123(10),
1980–1988.
TIME FREQUENCY ANALYSES 111
Isreal, J. B., Chesney, G. L., Wickens, C. D., & Donchin, E. (1980). P300 and tracking diffi-
culty: Evidence for multiple resources in dual‐task performance. Psychophysiology, 17(3),
259–273.
Jeon, Y. W., & Polich, J. (2003). Meta‐analysis of P300 and schizophrenia: Patients, paradigms,
and practical implications. Psychophysiology, 40(5), 684–701.
Kappenman, E. S., & Luck, S. J. (2016). Best practices for event-related potential research in
clinical populations. Biological Psychiatry: Cognitive Neuroscience and Neuroimaging, 1(2),
110–115.
Kiang, M., Braff, D. L., Sprock, J., & Light, G. A. (2009). The relationship between preattentive
sensory processing deficits and age in schizophrenia patients. Clinical Neurophysiology,
120(11), 1949–1957.
Knutson, B., Westdorp, A., Kaiser, E., & Hommer, D. (2000). FMRI visualization of brain ac-
tivity during a monetary incentive delay task. NeuroImage, 12(1), 20–27.
Ko, D., Kwon, S., Lee, G.-T., Im, C. H., Kim, K. H., & Jung, K.-Y. (2012). Theta oscillation related
to the auditory discrimination process in mismatch negativity: oddball versus control para-
digm. Journal of Clinical Neurology, 8(1), 35–42.
Kolev, V., Demiralp, T., Yordanova, J., Ademoglu, A., & Isoglu-Alkaç, Ü. (1997). Time–
frequency analysis reveals multiple functional components during oddball P300.
Neuroreport, 8(8), 2061–2065.
Kreussel, L., Hewig, J., Kretschmer, N., Hecht, H., Coles, M., & Miltner, W. (2011). The influence
of the magnitude, probability, and valence of potential wins and losses on the amplitude of
the feedback negativity. Psychophysiology, 49, 207–219. doi:10.1111/j.1469-8986.2011.01291.x
Kujawa, A., Carroll, A., Mumper, E., Mukherjee, D., Kessel, E., Olino, T., . . . Klein, D. (2018).
A longitudinal examination of event- related potentials sensitive to monetary reward
and loss feedback from late childhood to middle adolescence. International Journal of
Psychophysiology, 132, 323–330.
Kujawa, A., Proudfit, G., & Klein, D. (2014). Neural reactivity to rewards and losses in offspring
of mothers and fathers with histories of depressive and anxiety disorders. Journal of ab-
normal psychology, 123(2), 287.
Larson, M. J., South, M., Krauskopf, E., Clawson, A., & Crowley, M. J. (2011). Feedback and re-
ward processing in high-functioning autism. Psychiatry research, 187(1-2), 198–203.
Lee, M., Sehatpour, P., Hoptman, M. J., Lakatos, P., Dias, E. C., Kantrowitz, J. T., . . . Javitt, D. C.
(2017). Neural mechanisms of mismatch negativity dysfunction in schizophrenia. Molecular
psychiatry, 22(11), 1585.
Leicht, G., Troschütz, S., Andreou, C., Karamatskos, E., Ertl, M., Naber, D., & Mulert, C. (2013).
Relationship between oscillatory neuronal activity during reward processing and trait im-
pulsivity and sensation seeking. PloS one, 8(12), e83414.
Luck, S. J. (2014). An introduction to the event-related potential technique: MIT press.
Lukie, C. N., Montazer-Hojat, S., & Holroyd, C. B. (2014). Developmental changes in the re-
ward positivity: An electrophysiological trajectory of reward processing. Developmental
cognitive neuroscience, 9, 191–199.
Luu, P., & Tucker, D. M. (2001). Regulating action: Alternating activation of midline frontal
and motor cortical networks. Clinical Neurophysiology, 112(7), 1295– 1306. doi:10.1016/
S1388-2457(01)00559-4
Luu, P., Tucker, D. M., Derryberry, D., Reed, M., & Poulsen, C. (2003). Electrophysiological
responses to errors and feedback in the process of action regulation. Psychological Science,
14(1), 47–53. doi:10.1111/1467-9280.01417
112 ANNA WEINBERG, PAIGE ETHRIDGE, BELEL AIT OUMEZIANE, and DAN FOTI
Luu, P., Tucker, D. M., & Makeig, S. (2004). Frontal midline theta and the error-related nega-
tivity: Neurophysiological mechanisms of action regulation. Clinical Neurophysiology, 115,
1821–1835. doi:10.1016/j.clinph.2004.03.031
Miltner, W., Braun, C., & Coles, M. (1997). Event-related brain potentials following incorrect
feedback in a time-estimation task: Evidence for a “generic” neural system for error detec-
tion. Journal of Cognitive Neuroscience, 9(6), 788–798. doi:10.1162/jocn.1997.9.6.788
Mueller, E. M., Panitz, C., Pizzagalli, D. A., Hermann, C., & Wacker, J. (2015). Midline theta
dissociates agentic extraversion and anhedonic depression. Personality and Individual
Differences, 79, 172–177.
Munneke, G.-J., Nap, T. S., Schippers, E. E., & Cohen, M. X. (2015). A statistical comparison of
EEG time-and time-frequency domain representations of error processing. Brain Research,
1618(222-230). doi:10.1016/j.brainres.2015.05.030
Näätänen, R. (1995). The mismatch negativity: a powerful tool for cognitive neuroscience. Ear
and hearing, 16(1), 6–18.
Näätänen, R., Kujala, T., Kreegipuu, K., Carlson, S., Escera, C., Baldeweg, T., & Ponton, C.
(2011). The mismatch negativity: an index of cognitive decline in neuropsychiatric and
neurological diseases and in ageing. Brain, 134(12), 3435–3453.
Näätänen, R., Paavilainen, P., Rinne, T., & Alho, K. (2007). The mismatch negativity (MMN)
in basic research of central auditory processing: a review. Clinical Neurophysiology, 118(12),
2544–2590.
Nelson, B., Infantolino, Z., Klein, D., Perlman, G., Kotov, R., & Hajcak, G. (2018). Time-
frequency reward-related delta prospectively predicts the development of adolescent-onset
depression. Biological Psychiatry: Cognitive Neuroscience and Neuroimaging, 3(1), 41–49.
Nelson, L., Patrick, C., Collins, P., Lang, A., & Bernat, E. (2011). Alcohol impairs brain reactivity
to explicit loss feedback. Psychopharmacology, 1–10.
Novak, B. K., Novak, K. D., Lynam, D. R., & Foti, D. (2016). Individual differences in the time
course of reward processing: stage-specific links with depression and impulsivity. Biological
Psychology, 119, 79–90.
Novak, K. D., & Foti, D. (2015). Teasing apart the anticipatory and consummatory pro-
cessing of monetary incentives: An event‐related potential study of reward dynamics.
Psychophysiology, 52(11), 1470–1482.
Olson, L., Harper, J., Golosheykin, S., Bernat, E., & Anokhin, A. (2011). Development of the
Feedback Negativity in early adolescence: A longitudinal study of time-frequency theta and
delta components. Poster presented at the 2011 meeting of the Society for Psychophysiological
Research.
Olvet, D. M., & Hajcak, G. (2008). The error-related negativity (ERN) and psychopath-
ology: Toward an endophenotype. Clinical Psychology Review, 28(8), 1343–1354. doi:10.1016/
j.cpr.2008.07.003
Polich, J. (2007). Updating P300: an integrative theory of P3a and P3b. Clinical Neurophysiology,
118(10), 2128–2148.
Polich, J. (2012). Neuropsychology of P300. Oxford handbook of event- related potential
components, 159–188.
Polich, J., & Kok, A. (1995). Cognitive and biological determinants of P300: an integrative re-
view. Biological Psychology, 41(2), 103–146.
Polich, J., Pollock, V., & Bloom, F. (1994). Meta-analysis of P300 amplitude from males at risk
for alcoholism. Psychological bulletin, 115(1), 55.
TIME FREQUENCY ANALYSES 113
Watts, A. T., & Bernat, E. M. (2018). Effects of reward context on feedback processing as
indexed by time‐frequency analysis. Psychophysiology, e13195.
Webb, C., Auerbach, R., Bondy, E., Stanton, C., Foti, D., & Pizzagalli, D. (2017). Abnormal
neural responses to feedback in depressed adolescents. Journal of abnormal psychology,
126(1), 19.
Weinberg, A., & Hajcak, G. (2011). Longer term test–retest reliability of error‐related brain ac-
tivity. Psychophysiology, 48(10), 1420–1425.
Weinberg, A., Liu, H., Hajcak, G., & Shankman, S. (2015). Blunted neural response to rewards
as a vulnerability factor for depression: Results from a family study. Journal of abnormal
psychology, 124(4), 878.
Weinberg, A., Liu, H., & Shankman, S. A. (2016). Blunted neural response to errors as a
trait marker of melancholic depression. Biological Psychology, 113, 100–107. doi:10.1016/
j.biopsycho.2015.11.012
Weinberg, A., Olvet, D. M., & Hajcak, G. (2010). Increased error- related brain ac-
tivity in generalized anxiety disorder. Biological Psychology, 85, 472– 480. doi:10.1016/
j.biopsycho.2010.09.011
Weinberg, A., & Shankman, S. (2016). Blunted reward processing in remitted melancholic de-
pression. Clinical Psychological Science, in press.
Wickens, C., Kramer, A., Vanasse, L., & Donchin, E. (1983). Performance of concurrent
tasks: a psychophysiological analysis of the reciprocity of information-processing resources.
Science, 221(4615), 1080–1082.
Yeung, N., Bogacz, R., Holroyd, C. B., & Cohen, J. D. (2004). Detection and synchronized
oscillations in the electroencephalogram: An evaluation of methods. Psychophysiology, 41,
822–832. doi:10.1111/j.1469-8986.2004.00239.x
Yeung, N., Bogacz, R., Holroyd, C. B., Nieuwenhuis, S., & Cohen, J. D. (2007). Theta phase
resetting and the error- related negativity. Psychophysiology, 44, 39– 49. doi:10.1111/
j.1469-8986.2006.00482.x
Yordanova, J., Falkenstein, M., Hohnsbein, J., & Kolev, V. (2004). Parallel systems of error pro-
cessing in the brain. NeuroImage, 22, 590–602. doi:10.1016/j.neuroimage.2004.01.040
CHAPTER 6
ALI MAZAHERI
6.1 INTRODUCTION
For more than a century now researchers have been examining the electrical potentials
and magnetic fields measured at the scalp to understand what is happening inside our
brains when we perform various cognitive tasks. Researchers’ primary approach is to
characterize how the electro/magnetoencephalogram (E/MEG) signal changes in re-
sponse to a particular “event”, whether it be a button press or the onset of an auditory
tone. These changes are historically labelled as either evoked or induced, with each label
making assumptions about the origins of the change. The rationale behind evoked ac-
tivity is that the brain produces a new response as a consequence of processing the event.
This response is both time-locked and phase-locked to the experimental event. Induced
activity, on the other hand, assumes that the brain has ongoing brain activity (i.e., ac-
tivity that is always there) independent of any additive activity, and the event modulates
this ongoing activity, in a time-locked, but not necessarily phase-locked manner. My
hope is that the reader, through what I discuss in this chapter, will understand that these
“labels” (while at times useful) can paint an incomplete picture of what is going on in the
brain during cognitive processing. According to Andy Warhol, “The moment you label
something, you take a step—I mean, you can never go back again to seeing it unlabeled”.
I further argue that for the field to move forward in gaining a richer understanding of
116 ALI MAZAHERI
the link between brain and cognition, we need to rethink how we label the different
types of EEG responses.
(c)
30
10
–0.2 0 0.5
Time (S)
Figure 6.1 (A) The onset of an event (e.g., auditory stimulus) can evoke activity that is both
phase and time-locked to the onset of the event as well as induce activity that is time-locked but
not phase-locked. (B) Time-domain averaging of multiple data epochs would result in the at-
tenuation of the non-phase locked activity due to destructive interference. The activity remaining
after the averaging reflects the brain’s transient phase-locked response to an event.
118 ALI MAZAHERI
The rather traditional view of evoked and ongoing activity (Figure 6.2) is that they re-
flect rather separate distinct neural phenomena. According to this view, the evoked ac-
tivity that always has a consistent phase-locked to the onset of an experimental event
rides on top of the ongoing activity. This is also sometimes referred to as the “additive
view” of how ERPs are generated. Taken to the extreme, it is possible to view the evoked
activity as completely independent of the ongoing activity (figure 6.3A). An alternate
theory, referred to as a phase-resetting theory (figure 6.3B), postulates that there is no
additive evoked activity elicited by the onset of an event, but that rather, the ongoing ac-
tivity adjusts its phase to the onset of the experimental event (Makeig et al., 2002). Here,
by averaging trials locked to the event, the ongoing activity before the onset of the event
which has random phases is averaged out, while the event-related phase perturbed ac-
tivity emerges as the evoked response.
Given that the predominant ongoing activity in the EEG signal is the alpha rhythm,
it is believed that its phase-reset (or adjustment) to the onset of the experimental
event plays a particular role in the formation of evoked responses (Makeig et al., 2002;
Klimesch et al., 2007; Gruber et al., 2005; Hanslmayr et al., 2006). However, the phase-
reset of the ongoing rhythms is not exclusive to the alpha activity, with the theta rhythm
RELATIONSHIP BETWEEN EVOKED AND INDUCED EEG/MEG CHANGES 119
Cortex Cortex
RE RE
RE RE
TCR TCR
TCR TCR
Figure 6.2 Schema for the generation of induced (ERD/ERS) and evoked (ERP) activity
whereby the former is highly frequency-specific.
Adapted from Pfurtscheller & Lopes da Silva, 1999.
also proposed to be involved in the formation of specific evoked responses such as the
error-related negativity (Luu et al., 2004).
There has been a fair amount of controversy over whether phase-resetting can
account for the formation of ERPs (Mazaheri & Jensen, 2006). The primary evidence
for the occurrence of a phase-reset is that the phase of the ongoing activity at the time
of the evoked response would be consistent across trials. However, the addition of
a signal with a consistent phase across trials (i.e., a traditional additive evoked re-
sponse) would also make the phase of the ongoing activity appear consistent across
Figure 6.3 The additive versus phase-resetting theory of evoked response generation. (A) The
additive theory assumes that evoked and ongoing activities are distinct neuronal phenomena.
The experimental event “evokes” an additive, phase-locked response in each trial. (B) According
to the phase-resetting view, the ongoing and evoked activity are the same neuronal phenomena,
with no “new” additive response. Here the phases of the ongoing background oscillations be-
come aligned (phase-reset or partial phase-reset) to an experimental event. The phase-locked
(i.e., adjusted) oscillatory activity emerges as an evoked component when averaging the event-
locked trials.
120 ALI MAZAHERI
p=0.79 p<0.001
Average Evoked Potential Power Changes
+
10 μV 10
– p = 0.12
Number
0
– 0 +
–200 800ms Power Relative to Baseline
Figure 6.4 Model of generation. The upper left of the figure shows 20 superimposed trials of
a model evoked potential consisting of a single cycle of activity added to ongoing activity of the
same frequency with variable phase and amplitude. Below is given the average evoked potential
over 100 trials. At the upper right are the polar plots showing the phase distributions of the fre-
quency of the evoked potential (and background activity) during the baseline and at the middle
of the evoked potential. There is significant phase synchronization at the time of the evoked po-
tential. At the bottom is a histogram of the power measurements in the middle of the evoked po-
tential across the 100 trials (because of the Mortlet filtering effect this gives the maximum power).
There is no significant change in power.
Reprinted by permission from Mazaheri & Picton, 2005.
trials (Figure 6.4) (Mazaheri & Picton, 2005; van Diepen & Mazaheri, 2018; Yeung
et al., 2004).
Moreover, a rather convincing argument has been raised that the additive and phase-
resetting model cannot be mathematically distinguished at the scalp level without inva-
sive electrophysiological recordings (Telenczuk et al., 2010).
While the additive and phase-resetting theories offer a contradictory account of
evoked and ongoing activity, they do share two common elements. Both theories
assume across- trials averaging results in the attenuation of ongoing activity.
However, this assumption has been challenged. There is now compelling evidence
that the alpha rhythm, the dominant ongoing signal detected at the scalp, is non-
sinusoidal, and across-trials averaging never really makes it go away. This obser-
vation greatly blurs the line between evoked and ongoing activity. Furthermore,
it is important to note that the additive and phase-resetting debate has exclusively
focused on the early-stimulus evoked responses (P1, N1, or the ERN) and fails to
provide a complete account for the brain responses occurring 200 ms after an ex-
perimental event. These sustained responses (Figure 6.5), often lasting 100–200 ms,
are believed to reflect neural processing related to high-level cognitive constructs
ranging from working memory representation (Ikkai et al., 2010) to language com-
prehension (Kutas & Federmeier, 2011).
RELATIONSHIP BETWEEN EVOKED AND INDUCED EEG/MEG CHANGES 121
Figure 6.5 Grand averaged ERP difference waves (contralateral activity minus ipsilateral ac-
tivity) timelocked to the memory array averaged across the lateral occipital and posterior parietal
electrode sites and divided across the high and low memory capacity groups.
From Vogel et al., 2005.
(a) (b)
100 ms
(c) (d)
ER ER
Figure 6.6 The amplitude modulation of neuronal oscillatory activity is conventionally viewed
as being symmetric at approximately zero. (B) We propose that the amplitude modulations of the
oscillatory activity are asymmetric such that the peaks are more strongly modulated than the
troughs. For the 10 Hz alpha activity, this could be explained by bouts of activity every ~100 ms.
(C) The conventional view ignoring asymmetric modulations of oscillatory activity would mean
that averaging across trials (the arrow representing the start of the evoked response) would not
result in the generation of slow fields. (D) As a direct consequence of amplitude asymmetry, a
depression (or increase) in alpha activity in response to a stimulus will result in the generation of
slow fields when multiple trials are averaged. Adapted from Mazaheri and Jensen (2008).
400
0
Amplitude [fT/cm]
–400
–800
40
20
0
–20
1 2 3 4
Time (seconds)
Figure 6.7 Baseline shifts in ongoing oscillations. Upper trace: spatially filtered (with inde-
pendent component analysis) broadband signal from a channel above the right sensorimotor
area during rest. Lower trace: the mean values in three time intervals. Clearly, there are baseline
shifts in the ongoing activity associated with oscillations changing from large to small and back
to large amplitude. If many epochs with similar amplitude dynamics are averaged, oscillatory
patterns would disappear whereas the baseline shifts would remain leading to the appearance of
an evoked response.
Nikulin et al., 2007.
low amplitude. We found that despite the stimulus being the same, the sorting of the
trials based on alpha amplitude resulted in the formation of slow-evoked responses
(Figure 6.9). Across participants the amplitude, and polarity of these slow responses was
highly correlated with the direction of the amplitude asymmetry of the ongoing alpha
activity. Thus we were able to demonstrate (albeit with simple grating stimuli) that it was
(in principle) possible to form slow-evoked responses in the trial averaged EEG epochs
if there were systematic changes in the amplitude of the ongoing alpha activity.
Mazaheri and Jensen (2010) proposed four prerequisites for linking modulations of
oscillatory activity to evoked component generation.
(a)
S1(t) = A(t)(1+sin(2πft) + noise AFAindex = 0.96
Amplitude
8–12 Hz
Amplitude
8–12 Hz
(c)
S3(t) = A(t)+sin(2πft) + noise AFAindex = –0.07
Amplitude
8–12 Hz
Peaks
Throughs
Figure 6.8 Various simulations in which surrogate signals were used to test the AFAindex.
(A) The signal, s1(t), was designed to have an amplitude asymmetry. The amplitude modulation
was determined by a slower signal A(t). Clearly the peaks (red dots) are more modulated than the
troughs (blue dots) yielding a strong AFAindex. (B) The signal, s2(t), was designed such that the
slow modulations, A(t), affected the alpha rhythm in a multiplicative manner. Thus peaks and
troughs are modulated symmetrically over time yielding an AFAindex close to 0. (C) In signal
s3(t) slow modulations were added to the alpha oscillations (DC-like offset of the signal). This
affected peaks and troughs in the same direction producing an AFAindex close to 0.
Adapted from Mazaheri & Jensen, 2008.
Iemi and colleagues’ (2019) comprehensive study systematically examined the re-
lationship of pre-stimulus power of oscillatory, amplitude asymmetry, and the for-
mation of evoked responses. In particular, Iemi and colleagues’ (2019) focused on
differentiating the impact of pre-stimulus functional inhibition (a sensory state being
in a less-responsive state) from amplitude asymmetry on both the early and late sensory
evoked responses.
Here, with functional inhibition, the authors were referring to the currently widely
held view that an increase in alpha activity in a sensory system reflects its functional
inhibition and consequently results in attenuated evoked responses (evidence recently
reviewed in Van Diepen et al., 2019). They found that the early evoked (<0.200 s: e.g.,
the C1/N1 components) were indeed modulated by the amplitude of the pre-stimulus
RELATIONSHIP BETWEEN EVOKED AND INDUCED EEG/MEG CHANGES 125
Frequency (Hz)
10 40
(fT2/Hz)
Power
30% high α trials
20
–40
10
–0.5 0 0.5
Time (s)
Figure 6.9 Time-frequency representations of the trials with the 30% lowest and 30% highest
modulations of alpha power (TFRs baseline corrected; −0.6 < t < −0.1 s) in a representative sub-
ject. The respective ERFs (right) reveal a clear difference in the sustained modulation with respect
to low-(thin line) and high-alpha-power changes (thick line).
Adapted from Mazaheri & Jensen, 2008.
Stimulus
Onset
ERD ERS
ERP
What are the consequences of re-labelling the CDA as a change in ongoing ac-
tivity rather than a purely additive response? For one thing, this could have profound
implications on how we believe the brain carries out working memory processes. As
mentioned, one popular view of the role of alpha modulation in cognition is the
suppression of task-irrelevant regions (Van Diepen et al., 2019). Thus, the CDA, rather
than being an additive neural process involved in memory maintenance, could instead
be reflecting the inhibition of task-irrelevant brain areas. Additionally, unifying on-
going and event-related activity has the potential to mechanistically account for some
rather intriguing ERP findings, for which the origins of the responses remain a mystery.
For example, a now-classic study (Otten et al., 2006) found that the amplitude of slow
event-related potentials locked to the onset of a cue, but peaking before the onset of a
word to be remembered, could predict if the word was later remembered. By linking
the slow ERPs to the modulation of ongoing alpha activity, one simple interpretation of
the observed difference between the remembered versus forgotten words could be that
alpha activity is higher (i.e., the brain is in a more inhibited state) prior to the onset of
forgotten words. This is indeed in line with several experiments observing pre-stimulus
alpha oscillations to modulate perception (van Dijk et al., 2008) as well as reflect slips of
sustained attention (Bengson et al., 2012).
I would certainly not advocate any researchers to dismiss the event-related averaging
approach in exchange for looking at changes in the brain’s ongoing activity. However,
strictly viewing ongoing activity and evoked responses as separate unique entities is
implicitly believing the brain was doing nothing before the onset of the experimental
event. Such a view is particularly limited when it comes to trying to understand how the
brain tries to make sense of the outside world.
As an example of how the brain endeavors to make sense of the world, one rather in-
fluential theory proposes that the brain is constantly making predictions about what
is going to happen next (reviewed in detail in Friston, 2010). Specifically, this theory,
referred to as “predictive coding”, postulates that brain sets expectations and predictions
about upcoming sensory input and then subsequently updates these expectations after
the onset of the sensory input. Here, the discrepancy between the expectation and actual
sensory input is referred to as prediction error. While evoked responses can reveal in-
formation about the degree of prediction error and the perceived mismatch between ex-
pectation and reality, they are not directly informative about the neurophysiology of the
predictive processes themselves, since, by definition, the evoked response emerges after
the sensory input. By removing the separate labels (going back to the Warhol quote) of
evoked responses and ongoing activity, it is possible to get a richer, but at the same time
more parsimonious, picture of the neural processes underlying cognition.
Finally, I paraphrase Warhol for one last time: I hope some of the mystery behind
event-related responses and ongoing activity is gone, but the amazement is just starting.
REFERENCES
Bae, G.-Y. & Luck, S. J. (2018). Dissociable decoding of spatial attention and working memory
from EEG oscillations and sustained potentials. The Journal of Neuroscience, 38(2), 409–422.
Bengson, J. J., Mangun, G. R., & Mazaheri, A. (2012). The neural markers of an imminent
failure of response inhibition. NeuroImage, 59(2), 1534–1539.
Cohen, M. X. (2017). Where does EEG come from and what does it mean? Trends in
Neuroscience, 40(4), 208–218.
Friston, K. (2010). The free- energy principle: A unified brain theory? Nature Reviews
Neuroscience, 11(2), 127–138.
RELATIONSHIP BETWEEN EVOKED AND INDUCED EEG/MEG CHANGES 129
Fukuda, K., Mance, I., & Vogel, E. K. (2015). α-power modulation and event-related slow wave
provide dissociable correlates of visual working memory. The Journal of Neuroscience, 35(41),
14009–14016.
Gruber, W. R., Klimesch, W., Sauseng, P., & Doppelmayr, M. (2005). Alpha phase synchroniza-
tion predicts P1 and N1 latency and amplitude size. Cerebral Cortex, 15(4), 371–377.
Hämäläinen, M., Hari, R., Ilmoniemi, R. J., Knuutila, J., & Lounasmaa, O. V. (1993).
Magnetoencephalography: Theory, instrumentation, and applications to noninvasive
studies of the working human brain. Reviews of Modern Physics, 65, 413.
Hanslmayr, S., Klimesch, W., Sauseng, P., Gruber, W., Doppelmayr, M., Freunberger, R.,
Pecherstorfer, T., & Birbaumer, N. (2006). Alpha phase reset contributes to the generation of
ERPs. Cerebral Cortex, 17(1), 1–8.
Hanslmayr, S., Staresina, B. P., & Bowman, H. (2016). Oscillations and episodic
memory: Addressing the synchronization/ desynchronization conundrum. Trends in
Neuroscience, 39(1), 16–25.
Iemi, L., Busch, N. A., Laudini, A., Haegens, S., Samaha, J., Villringer, A., & Nikulin, V. V.
(2019). Multiple mechanisms link prestimulus neural oscillations to sensory responses.
Elife, 8, e43620.
Ikkai, A., McCollough, A. W., & Vogel, E. K. (2010). Contralateral delay activity provides a
neural measure of the number of representations in visual working memory. Journal of
Neurophysiology, 103(4), 1963–1968.
Kappenman, E. S., & Luck, S. J. (Eds.). (2011). Oxford handbook of event-related potential
components. Oxford University Press.
Klimesch, W., Sauseng, P., Hanslmayr, S., Gruber, W., & Freunberger, R. (2007). Event-related
phase reorganization may explain evoked neural dynamics. Neuroscience & Biobehavioral
Reviews, 31(7), 1003–1016.
Kutas, M. & Federmeier, K. D. (2011). Thirty years and counting: Finding meaning in the N400
component of the event-related brain potential (ERP). Annual Review of Psychology, 62,
621–647.
Luu, P., Tucker, D. M., & Makeig, S. (2004). Frontal midline theta and the error-related nega-
tivity: neurophysiological mechanisms of action regulation. Clinical Neurophysiology, 115(8),
1821–1835.
Makeig, S., Westerfield, M., Jung, T. P., Enghoff, S., Townsend, J., Courchesne, E., & Sejnowski,
T. J. (2002). Dynamic brain sources of visual evoked responses. Science, 295(5555), 690–694.
Mazaheri, A. & Jensen, O. (2006). Posterior α activity is not phase-reset by visual stimuli.
Proceedings of the National Academy of Sciences of the United States of America, 103(8),
2948–2952.
Mazaheri, A. & Jensen, O. (2008). Asymmetric amplitude modulations of brain oscillations
generate slow evoked responses. The Journal of Neuroscience, 28(31), 7781–7787.
Mazaheri, A. & Jensen, O. (2010). Rhythmic pulsing: Linking ongoing brain activity with
evoked responses. Frontiers in Human Neuroscience [online], 4(177). doi: 10.3389/
fnhum.2010.00177
Mazaheri, A. & Picton, T. W. (2005). EEG spectral dynamics during discrimination of auditory
and visual targets. Brain Research Cognitive Brain Research, 24(1), 81–96.
Nikulin, V. V., Linkenkaer-Hansen, K., Nolte, G., Lemm, S., Müller, K. R., Ilmoniemi, R. J., &
Curio, G. (2007). A novel mechanism for evoked responses in the human brain. European
Journal of Neuroscience, 25(10), 3146–3154.
Otten, L. J., Quayle, A. H., Akram, S., Ditewig, T. A., & Rugg, M. D. (2006). Brain activity be-
fore an event predicts later recollection. Nature Neuroscience, 9(4), 489–491.
130 ALI MAZAHERI
STEVEN L. BRESSLER
7.1 Introduction
The neocortex of the macaque monkey is very similar to that of the human in its archi-
tectonics. The six neocortical laminae show similar variation with region in the monkey
and human. In fact, it appears that all mammalian species share a common microstruc-
ture, which makes distinguishing neuroanatomical slices from different mammalian
species under the microscope nearly impossible. What appears to be more different be-
tween the macaque monkey and human neocortices is at the macroscopic level. Overall,
the number of neocortical areas is larger in the human brain, and the between-area
connectivity is more complex. However, the macroscopic structure of the two species is
highly similar in certain systems, for example, the visual system.
The similarity between the human and macaque monkey neocortex was first
recognized for the visual system, where neocortical oscillations are a prominent product
of the visual architecture of both humans and monkeys. Visual neocortical local field
potentials (LFPs) show oscillations in both low-frequency (delta, theta, and alpha) and
high-frequency (beta and gamma) bands. Many studies report the importance of the
neocortical LFP oscillations in the boundary frequency region between the low beta
(13–20 Hz) and alpha (8–12 Hz) frequency bands for top-down neocortical processing
in vision (Liang et al., 2002; Bressler et al., 1993; Bressler et al., 2007; Engel & Fries, 2010;
Bressler & Richter, 2015; Bastos et al., 2015; Richter et al.,2018), and in the theta (4–7 Hz)
and gamma (above >30 Hz) bands for bottom-up visual processing (Markov et al., 2013;
Bastos et al., 2015; Michalareas et al., 2016).
A great deal of research has also involved oscillatory activity in the prefrontal cortex
of the macaque monkey, which has important analogies to the human prefrontal cortex,
132 STEVEN L. BRESSLER
which is the most developed of all the mammalian species and is therefore unique.
The macaque monkey prefrontal cortex is clearly different from that of the human,
but it shares many of the same features. As working memory is an essential compo-
nent of different cognitive functions requiring the prefrontal cortex in both macaque
monkeys and humans (Miller et al., 2018), studies of working memory in the macaque
monkey prefrontal cortex are essential for understanding human working memory.
Monkey studies first demonstrated a role for prefrontal neocortical oscillatory activity
in working memory (Siegel et al.,2009), and these provided the theory that neocortical
oscillations are important for working memory processes in the beta-alpha and gamma
bands (Pesaran et al., 2002; Salazar et al., 2012; Antzoulatos & Miller, 2016). This specu-
lation has now been verified (discussed later; see also Lundqvist et al.,2016, 2018; Miller
et al., 2018).
The study of working memory oscillations in the monkey prefrontal cortex gives
impetus to the investigation of prefrontal neocortical oscillations in human working
memory (D’Esposito et al., 1995), and recent human studies demonstrate the import-
ance of prefrontal oscillatory activity in human working memory (Jensen et al., 2007;
Alekseichuk et al., 2016). Research by Miller and colleagues (2018) involving macaque
monkeys verifies that working memory oscillations exist in the theta, alpha-beta, and
gamma frequency bands. These oscillations are best studied in monkeys, which dem-
onstrate excellent working memory capability, have associated oscillations that can be
studied invasively, and which have a prefrontal neuroanatomy that is similar to humans.
Questions about working memory and high-frequency (beta and gamma) oscillations
tend to be addressed using monkeys due to the relatively low signal-to-noise ratio at
higher frequencies in the human electroencephalogram (EEG) (Crone et al., 2006). The
purpose of this report is to present evidence on neocortical oscillations from the ma-
caque monkey engaged in cognitive functions that rely on working memory.
1975; Buzsaki et al., 2012). The MEG is thought to arise from those same currents
passing through the dendritic shafts of the same neurons (Murakami & Okada, 2006).
Oscillatory activity is a prominent feature of both electric and magnetic signals, higher
frequency oscillations likely reflecting interactions of excitatory and inhibitory neurons
(Kopell, 2000).
There is growing recognition that the dendrites of neocortical neurons are thus es-
sential for the genesis of neocortical oscillations, and there is speculation that the dy-
namics of cognitive processing, including working memory (Voytek & Knight, 2015),
depends on oscillations. However, the recording and characterization of dendritic ac-
tivity is problematic. It is usually not possible to record from single dendrites of neurons
in the brain due to their thinness, or from the dendritic trees of single neurons due to the
lack of extracellular potentials from single-neuron dendrites. Also, neither the single-
neuron dendritic branch response nor the single-neuron dendritic tree response can be
extracted from the compound dendritic response. For these reasons, much of neuro-
physiology has focused on the action potential (spike) as the essential neuronal signal.
This focus is not because of the neuron doctrine, which, holding only that the neuron is
the central signaling cell in the nervous system, is neutral with respect to the parts of the
neuron that carry out particular aspects of that signaling.
Dendritic activity, and hence oscillatory activity, is typically recorded at the popula-
tion level—from groups of neurons rather than from single neurons. In most neurons,
the resultant sum of synaptic actions from an entire dendritic tree, contributed to by
thousands of synaptic potentials, is delivered to the initial segment of the axon, where
it causes the resultant axonal membrane potential to be graded in intensity. The mem-
brane potential of the initial segment has a low threshold for generating a spike because
voltage-sensitive Na +membrane channels are concentrated there. The recorded axonal
membrane potential may be supra-threshold if the membrane potential is above the
threshold, in which case spike trains are generated and travel down the axon, or sub-
threshold, in which case the axon may be affected but no spike trains are generated.
The spike trains may contribute to the neuronal synchronization underlying the LFP
(Murthy & Fetz, 1996).
The EEG, electrocorticogram (ECoG), intracranial EEG (iEEG),
magnetoencephalogram (MEG), and LFP signals all are generated by the dendritic ac-
tivity of neuronal populations, and all display oscillations. Except for the MEG, which
is magnetic, all these signals are electric and are recorded with respect to a reference
potential. The EEG and MEG are usually recorded from outside the cranium, whereas
the ECoG, iEEG, and LFP are recorded from inside the cranium. The ECoG is recorded
from outside the cortical tissue, whereas the iEEG and LFP are recorded from inside
the cortex. The iEEG is recorded from macroscopic electrodes and the LFP is usually
recorded with indwelling microelectrodes. The LFP recording may be monopolar, in
which case a single microelectrode records the LFP, or it may be bipolar, in which case
the LFP is the difference in potential between two nearby microelectrodes. In fact, the
LFP is often recorded from the same microelectrodes used to record neuronal spikes,
134 STEVEN L. BRESSLER
and at the same time (Perelman & Ginosar, 2007; Salazar et al., 2012). However, the LFP
typically shows oscillations whereas they are usually not obvious in spike activity.
Since these signal types all reveal oscillatory activity, which is considered by many to
be essential for cognition in the brain (Voytek & Knight, 2015), and are typically present
when humans perform cognitive tasks, the dendritic activity of neurons in the brain
should not be overlooked in the search for neural correlates of cognition, even though it
cannot currently be studied in the single neuron. Oscillations in the summed dendritic
activity of neuronal populations are strongly related to cognitive function in the neo-
cortex (Donner & Siegel, 2011).
Cs Cs
IPs IPs
6 6
2 1 Scales 2 1
4 4
Granger
3 Coherence 3
Causality
0.5 0.25
As As
0.4 0.2
5 0.3 0.15 5
STs 0.2 0.1 STs
Ls Ls
0.1 0.05
4 4
Cs Cs
As As
IPs IPs
1 2 3 Ls 1 2 3 Ls
LFPs, simultaneously recorded from the striate and extrastriate (V4—visual area 4 and
TEO—temporal-occipital area) visual cortices prior to appearance of the visual stimulus
in the same visual pattern discrimination task (Bressler & Richter, 2015), also show os-
cillatory activity in the beta frequency range. The same phase-coupling and neural caus-
ality metrics have been computed from the visual LFPs as from somatosensory-motor
LFPs. The main finding is that extrastriate–striate site pairs are beta-frequency phase
coupled and carried by strong top-down (extrastriate-to-striate) beta influences, in an-
ticipation of visual processing (Figures 7.2 and 7.3). Furthermore, behavioral context is
conveyed to primary visual cortex prior to appearance of the visual stimulus (Richter
et al., 2018). Thus, information about the visual stimulus appears in the visual cortex
(a)
Residual
30
V1/V2
1
20
0
10 20 30 40 50 60 70 80 90
(b)
40 2
power (dB μV)
Extrastriate
Residual
30 1
20
0
10 20 30 40 50 60 70 80 90
(c)
0.05
Coherence
0
10 20 30 40 50 60 70 80 90
(d)
0.04
0.02
GC
0
10 20 30 40 50 60 70 80 90
Frequency (Hz)
Figure 7.2 Prestimulus beta-frequency power, coherence, and (conditional spectral Wiener-
Granger causality spectra of V1/V2 and V4/TEO LFPs. (A) Average striate power spectrum over
sites (black line ± s.e.m.), and the residual power spectrum after 1/f removal (red line ± s.e.m.) for
V1 sites. (B) Average extrastriate power spectrum over sites and monkeys (black line ± s.e.m.),
and the residual power spectrum after 1/f removal (red line ± s.e.m.) for the V4/TEO sites.
(C) Average coherence spectrum over V1/V2-extrastriate site pairs ± s.e.m. for V1/V2-extrastriate
pairs. (D) Average top-down (red line ± s.e.m.), and bottom-up (blue line ± s.e.m.) GC spectra for
V1-extrastriate pairs. Shaded grey rectangular region denotes the frequencies (8–23 Hz) where
top-down and bottom-up sGC were significantly different (p<0.001).
FREQUENCY ANALYSIS OF THE MONKEY 137
M1 GC M2
0.20
0.15
0.10
0.05
M1 Coh M2
0.20
0.15
0.10
0.05
Figure 7.3 Prestimulus beta-frequency coherence and top-down (conditional spectral Wiener-
)Granger causality maps. Top: Maps of the recording sites for M1 and M2. V1/V2 electrode
locations are marked by yellow circles, and extrastriate (V4 and TEO) locations by gray circles.
Middle: enlarged maps of visual cortex showing top-down sGC at 16 Hz as arrows for V1/V2-
extrastriate pairs. Bottom: corresponding maps of coherence for the same site pairs. Thickness of
the top-down sGC arrows and coherence bars is proportional to the magnitude of sGC or coher-
ence at 16 Hz.
before the actual onset of the visual stimulus. This result validates previous proposals
that top-down visual processing depends on interareal synchronization in visual neo-
cortex (von Stein et al., 2000), and suggests that it acts to “prime” visual cortex to pre-
pare it for receiving the visual stimulus.
The same basic methodology, namely conditional spectral WG causality analysis, has
subsequently been applied to visual neocortical LFPs in macaque monkeys performing
a visuospatial attention task (Bastos et al., 2015). In the visual system, neocortical areas
are seen to interact in both bottom-up and top-down directions, with bottom-up
gamma-band influences conveying sensory signals, and top-down beta-band influences
modulating those bottom-up influences according to behavioral context. In the ma-
caque monkey visual neocortex, bottom-up influences are carried by theta-band (~4
Hz) and gamma-band (~60–80 Hz) synchronization, and top-down influences by
beta-band (~14–18 Hz) synchronization (Bastos et al. 2015). Furthermore, neocor-
tical hierarchies (Felleman & Van Essen, 1991; Hilgetag et al., 1996) created based on
138 STEVEN L. BRESSLER
7.5 Oscillations
in Prefrontal Neocortex
Monkey LFP oscillations have been reported during working memory in the theta,
alpha, beta, and gamma bands. The gamma band is associated with sensory information,
FREQUENCY ANALYSIS OF THE MONKEY 139
and gamma-band power correlates with the number of objects held in working
memory, whereas the beta band is associated with top-down information, and beta-
band synchrony correlates with task rules (Liang et al., 2002; Miller et al., 2018; Richter
et al., 2018). Gamma bursting is anti-correlated with alpha-beta bursting (Lundqvist
et al., 2016).
Information in prefrontal neuronal spiking has been linked to brief bursts in the
gamma band in monkeys performing a working memory task (Lundqvist et al., 2016).
Prefrontal laminar LFP data has been obtained from monkeys using prefrontal laminar
probes (Bastos et al., 2018). Gamma-band activity is strongest in the superficial layers,
and beta and alpha LFP power is strongest in the deep layers. In keeping with pre-
vious studies (Lundqvist et al., 2016), gamma-band bursting is most informative about
working memory in the superficial layers.
Gamma-band bursts, varying in time and frequency, were reported to accompany
both the encoding and re-activation of sensory information (Bastos et al., 2015). The
conclusion is that gamma bursts could gate access to, and prevent sensory interfer-
ence with, working memory, since only the neuronal activity that was associated with
working memory encoding and decoding was correlated with gamma-band burst rate
changes. Bursts in the beta band were also brief and variable, but they reflected a “default
state” that was interrupted by encoding and decoding.
The prefrontal cortex is not the only region of neocortex involved in working
memory. To test the involvement of posterior parietal cortex in working memory,
LFPs were recorded from distributed sites in the prefrontal cortex and in the pos-
terior parietal cortex of macaque monkeys performing a working memory delayed-
match-to-sample visual identity task (Salazar et al., 2012). Phase coupling between
prefrontal and posterior parietal cortices, examined by computing spectral coher-
ence between prefrontal-posterior parietal LFP pairs, was found in the beta frequency
band. Analysis of prefrontal-posterior parietal interactions in working memory by
conditional spectral WG causality applied to the delay period of the delayed match-
to-sample visual identity task showed that the beta frequency band was dominant.
Causal influences in the beta band were roughly balanced between the two directions,
that is, beta causal influences in the prefrontal-to-posterior parietal and posterior
parietal-to-prefrontal directions were roughly the same (slightly greater from pos-
terior parietal cortex to prefrontal cortex). This finding indicates that prefrontal
and posterior parietal cortices are roughly in balance during working memory task
performance.
140 STEVEN L. BRESSLER
7.7 Conclusions
References
Alekseichuk, I., Turi, Z., Amador de Lara, G., Antal, A., & Paulus, W. (2016). Spatial working
memory in humans depends on theta and high gamma synchronization in the prefrontal
cortex. Current Biology, 26, 1513–1521.
Antzoulatos, E. & Miller, E. (2016). Synchronous beta rhythms of frontoparietal networks
support only behaviorally relevant representations. eLife, 5, e17822.doi: 10.7554/eLife.17822
Bastos, A., Litvak, V., Moran, R., Bosman, C., Fries, P., & Friston, K. (2015). A DCM study of
spectral asymmetries in feedforward and feedback connections between visual areas V1 and
V4 in the monkey. Neuroimage, 108, 460–475.
Bastos, A., Loonis, R., Kornblith, S., Lundqvist, M., & Miller, E. (2018). Laminar recordings
in frontal cortex suggest distinct layers for maintenance and control of working memory.
Proceedings of the National Academy of Sciences of the United States of America, 115, 1117–1122.
Bastos, A., Usrey, W., Adams, R., Mangun, G., Fries, P., & Friston, K. (2012). Canonical
microcircuits for predictive coding. Neuron, 76, 695–7 11.
FREQUENCY ANALYSIS OF THE MONKEY 141
Bressler, S., Coppola, R., & Nakamura, R. (1993). Episodic multiregional cortical coherence at
multiple frequencies during visual task performance. Nature, 366, 153–156.
Bressler, S., Richter, C., Chen, Y., & Ding, M. (2007). Cortical functional network organization
from autoregressive modeling of local field potential oscillations. Statistics in Medicine, 26,
3875–3885.
Bressler, S. & Richter, C. (2015). Interareal oscillatory synchronization in top-down neocortical
processing. Current Opinion in Neurobiology, 31C, 62–66.
Brovelli, A., Ding, M., Ledberg, A., Chen, Y., Nakamura, R., & Bressler, S. (2004). Beta
oscillations in a large-scale cortical network: Directional influences revealed by Granger
causality. Proceedings of the National Academy of Sciences of the United States of America, 101,
9849–9854.
Buffalo, E., Fries, P., Landman, R., Buschman, T., & Desimone, R. (2011). Laminar differences
in gamma and alpha coherence in the ventral stream. Proceedings of the National Academy of
Sciences of the United States of America, 108, 11262–11267.
Buzsaki, G., Anastassiou, C., & Koch, C. (2012). The origin of extracellular fields and currents—
EEG, ECoG, LFP and spikes. Nature Reviews Neuroscience, 13, 407–420.
Crone, N., Sinai, A., & Korzeniewska, A. (2006). High-frequency gamma oscillations
and human brain mapping with electrocorticography. Progress in Brain Research, 159,
275–295.
D’Esposito M, Detre J, Alsop D, Shin R, Atlas S, Grossman M (1995) The neural basis of the cen-
tral executive system in working memory. Nature, 378, 279–281.
Ding, M., Chen, Y., & Bressler, S. (2006). Granger causality: Basic theory and application to
neuroscience. In B. Schelter, M. Winterhalder, & J. Timmer (Eds.), Handbook of time series
analysis: Recent theoretical developments and applications, (pp. 437–460). Wiley.
Donner, T. & Siegal, M. (2011). A framework for local cortical oscillation patterns. Trends in
Cognitive Science, 15, 191–199.
Engel, A. & Fries, P. (2010). Beta-band oscillations –signaling the status quo? Current Opinion
in Neurobiology, 20, 156–165.
Felleman, D. & Van Essen, D. (1991). Distributed hierarchical processing in the primate cere-
bral cortex. Cerebral Cortex, 1, 1–47.
Freeman, W. (1975). Mass action in the nervous system. Academic.
Hilgetag, C., O’Neill, M., & Young, M. (1996). Indeterminate organization of the visual system.
Science, 271, 776–777.
Howard, M., Rizzuto, D., Caplan, J., Madsen, J., Lisman, J., Aschenbrenner-Scheibe, R.,
Schulze-Bonhage, A., & Kahana, M. (2003). Gamma oscillations correlate with working
memory load in humans. Cerebral Cortex, 13, 1369–1374.
Jensen, O., Kaiser, J., & Lachaux, J. (2007). Human gamma-frequency oscillations associated
with attention and memory. Trends in Neuroscience, 30, 317–324.
Jokisch, D. & Jensen, O. (2007). Modulation of gamma and alpha activity during a working
memory task engaging the dorsal or ventral stream. Journal of Neuroscience, 27, 3244–3251.
Kopell, N. (2000). We got rhythm: Dynamical systems of the nervous system. Notices of the
American Mathematical Society, 47, 6–16.
Liang, H., Bressler, S., Ding, M., Truccolo, W., & Nakamura, R. (2002). Synchronized activity in
prefrontal cortex during anticipation of visuomotor processing. Neuroreports, 13, 2011–2015.
Lundqvist, M., Rose, J., Herman, P., Brincat, S., Buschman, T., & Miller, E. (2016). Gamma and
beta bursts underlie working memory. Neuron, 90, 152–164.
Lundqvist, M., Herman, P., Warden, M., Brincat, S., & Miller, E. (2018). Gamma and beta
bursts during working memory readout suggest roles in its volitional control. Nature
Communications, 9, 394.
142 STEVEN L. BRESSLER
Markov, N., Ercsey-Ravasz, M., Van Essen, D., Knoblauch, K., Toroczkai, Z., & Kennedy, H.
(2013). Cortical high-density counterstream architectures. Science, 342, 1238406.
Markov, N., Vezoli, J., Chameau, P., Falchier, A., Quilodran, R., Huissod, C., Lamy, C., Misery,
P., . . . Kennedy, H. (2014). Anatomy of hierarchy: Feedforward and feedback pathways in
macaque visual cortex. Journal of Comparative Neurology, 522, 225–229.
Michalareas, G., Vezoli, J., Van Pelt, S., Schoffelen, J-M., Kennedy, H., & Fries, P. (2016). Alpha-
beta and gamma rhythms subserve feedback and feedforward influences among human
visual cortical areas. Neuron, 46, 60528.
Miller, E., Lundqvist, M., & Bastos, A. (2018). Working memory 2.0. Neuron, 100, 463–475.
Murakami, S. & Okada, Y. (2006). Contributions of principal neocortical neurons to
magnetoencephalography and electroencephalography signals. Journal of Physiology, 575,
925–936.
Murthy, V. & Fetz, E. (1996). Synchronization of neurons during local field potential oscillations
in sensorimotor cortex of awake monkeys. Journal of Neurophysiology, 76, 3968–3982.
Perelman, Y. & Ginosar, R. (2007). An integrated system for multichannel recording with spike/
LFP separation, integrated A/D conversion, and threshold detection. IEEE Transactions on
Biomedical Engineering, 54, 130–137.
Pesaran, B., Pezaris, J., Sahani, M., Mitra, P., & Anderson, R. (2002). Temporal structure in neuronal
activity during working memory in macaque parietal cortex. Nature Neuroscience, 5, 805–81.
Raghavachari, S., Kahana, M., Rizzuto, D., Caplan, J., Kirschen, M., Bourgeois, B., Madsen, J.,
& Lisman, J. (2001). Gating of human theta oscillations by a working memory task. Journal
of Neuroscience, 21, 3175–3183.
Richter, C., Copolla, R., & Bressler, S. (2018). Top-down beta oscillatory signaling conveys be-
havioral context in early visual cortex. Scientific Reports, 8, 6991.
Roberts, M., Lowet, E., Brunet, N., Ter Wal, M., Tiesinga, P., Fries, P., & De Weerd, P. (2013).
Robust gamma coherence between macaque V1 and V2 by dynamic frequency matching.
Neuron, 78, 523–536
Salazar, R., Dotson, N., Bressler, S., & Gray, C. (2012). Content-specific fronto-parietal syn-
chronization during visual working memory. Science, 338, 1097–1100
Siegel, M., Warden, M., & Miller, E. (2009). Phase-dependent neuronal coding of objects in
short-term memory. Proceedings of the National Academy of Sciences of the United States of
America, 106, 21341–21346
Van Kerkoerle, T., Self, M., Dagnino, B., Gariel-Mathis, M-A., Poort, J., van der Togt, C., &
Roelfsema, P. (2014). Alpha and gamma oscillations characterize feedback and feedforward
processing in monkey visual cortex. Proceedings of the National Academy of Sciences of the
United States of America, 111, 14332–14341.
von Stein, A., Chiang, C., & Konig, P. (2000). Top-down processing mediated by interareal syn-
chronization. Proceedings of the National Academy of Sciences of the United States of America,
97, 14748–14753.
Voytek, B. & Knight, R. (2015). Dynamic network communication as a unifying neural basis for
cognition, development, aging, and disease. Biol Psych 77, 1089–1097.
Wang, X-J. (2010). Neurophysiological and computational principles of cortical rhythms in
cognition. Physiological Reviews, 90, 1195–1268.
Xing, B., Guo, J., Meng, X., Wei, S-G., & Li, S-B. (2012). The dopamine D1 but not D3 receptor
plays a fundamental role in spatial working memory and BDNF expression in prefrontal
cortex of mice. Behavioral Brain Research, 235, 36–41.
Pa rt I I
CHAPTER 8
GAMM A ACT I V I T Y I N
SE NSORY AND C O G NI T I V E
PRO CES SI NG
When the German psychiatrist Hans Berger reported his discovery of the human EEG
in 1929, he described two types of waves with the larger ones oscillating at 10–11 Hz and
the smaller ones at 20–30 Hz (Berger, 1929)—the well-known alpha-and beta-waves,
respectively. However, in view of the results from the first Fourier analysis carried out
on the human EEG by his physicist co-worker Dietsch (1932), Berger had to acknow-
ledge that his beta-waves contained a lot more components than originally thought with
frequencies up to 125 Hz (Berger, 1934;1936). By comparing spontaneous EEG recordings
with recordings during mental calculation, Berger (1937) observed effects predomin-
antly in the 40–90 Hz range which led him to conclude that beta-waves of this specific
(high-)frequency band “are the physical effects that accompany mental processes”. One
year later, the term “gamma waves” was proposed for higher frequencies at 35–45 Hz,
that is, beyond the traditional beta-band (Jasper & Andrews, 1938). However, this first-
time labeling went mostly unnoticed at that time and “gamma” was later reintroduced
by other authors (e.g., Başar & Özesmi, 1972; Bressler & Freeman, 1980). Instead of
“gamma”, the term “40-Hz oscillation” is also widely used.
The first experimental evidence for a possible role of gamma oscillations in early
sensory processing in the mammalian brain was provided by Adrian (1942; 1950), who
recorded oscillatory responses of the olfactory bulb of hedgehogs, cats, and rabbits
to odorous substances. He obtained oscillatory responses in the 30–60 Hz frequency
range which he termed “induced waves” to differentiate these events from “intrinsic
146 DANIEL STRÜBER and CHRISTOPH S. HERRMANN
Gamma oscillations are part of the total EEG energy to a varying degree at any moment
in time. Such spontaneous or ongoing activity occurs independently of any external
148 DANIEL STRÜBER and CHRISTOPH S. HERRMANN
stimulation and may signal alterations of internal states like arousal or alertness (Strüber
et al., 2001). If gamma oscillations are related to the processing of external stimuli, a
distinction is made between “evoked” and “induced” activity (Başar-Eroǧlu, Strüber,
Schürmann, et al., 1996). Evoked gamma responses are strictly time-and phase-locked
to stimulus onset, that is, each single trial occurs at the same latency with zero phase
lag across trials, therefore summing up after averaging. Thus, evoked activity can be
analyzed by transforming the averaged single-trials (i.e., the averaged evoked potential,
ERP) into the frequency domain (see Figure 8.1, left panel). This type of gamma activity
is typically observed within an early time window of ~50–150 ms following stimula-
tion onset. In contrast, induced activity occurs later, following stimulation by at least
200–300 ms, and it cancels out almost completely if averaged due to its inter-trial vari-
ation of phase and latency. Therefore, to detect induced gamma oscillations, each single
trial needs to be subjected to a wavelet transform, and then the resulting absolute power
2
transform
wavelet
...
averaging averaging
Total Activity
frequency
100
[Hz]
ERP
50
0 100 300
time [ms]
Evoked Activity
frequency
100 0 0.5 1
[Hz]
50
0 100 300 amplitude [a.u.]
time [ms]
Figure 8.1 Analysis of evoked and induced gamma activity. Left: Averaging all single trials
containing evoked and induced gamma oscillations yields the ERP. Transforming the ERP into
the time-frequency domain (wavelet transform) leaves only the evoked gamma activity, because
the induced gamma activity cancels out in the ERP due to phase jitter. Right: Averaging the ab-
solute values of each single trial’s wavelet transform yields both evoked and induced activity (i.e.,
the total activity). Thus, evoked activity is obtained with both types of analyses, whereas induced
gamma activity appears only in the average of the single trial wavelet transforms.
Adapted from Herrmann et al., 2014.
NB: The latency jitter of the induced activity induces temporal smearing in the total activity. a.u., arbitrary units.
GAMMA ACTIVITY IN SENSORY AND COGNITIVE PROCESSING 149
values are averaged across trials without cancellation (see Figure 8.1, right panel). Note,
however, that this average contains not only the induced but also the evoked activity and
has, therefore, been termed “total activity” (Herrmann et al., 2014). Within this plot of
total activity, all activity that is not also present in the evoked plot, can be referred to as
induced activity. While Figure 8.1 depicts the differences between the different types of
gamma activity in principle, Figure 8.2 represents real data from human experiments.
In traditional EEG research, the range of gamma frequencies is given as 30–80 Hz
with evoked activity often oscillating nearby 40 Hz, whereas induced responses can
also reach higher frequencies (Herrmann, Munk, et al., 2004). However, in the con-
text of subdural electrocorticography (ECoG) in epilepsy patients, a broad range of
very high gamma frequencies (around 80–200 Hz) was discovered unexpectedly and
called “high gamma” to distinguish this broadband response from the traditional “low
gamma” oscillations in narrower bands (for a review, see Crone et al., 2011). Although
60
Reversal
Pattern
40
(b)
80
frequency [Hz]
60
Motion
40
[a.u.]
0 1
Figure 8.2 Time-frequency plots and topographies of evoked and induced gamma activity.
White box: Evoked gamma activity for pattern reversal (A) and motion (B). Yellow box: Induced
gamma activity for pattern reversal (A) and motion (B) within the plot of total activity. Red
box: Steady-state visual evoked potential (SSVEP) in the motion condition as a response to the
single images representing the motion of the gratings. The green line represent stimulus onset.
C: Topographic maps of evoked (left) and induced (right) gamma activity during an object recog-
nition task. a.u.: arbitrary units.
(A, B) Adapted from Naue et al., 2011; (C) Adapted from Busch, Herrmann, et al., 2006.
150 DANIEL STRÜBER and CHRISTOPH S. HERRMANN
this broadband activity often overlaps with the band-limited (low) gamma oscillations,
it is possible that such broadband responses reflect mere increases of (non-oscillatory)
gamma power and/or spiking activity rather than a true oscillation (Buzsáki & Wang,
2012; Ray & Maunsell, 2011), since the latter would be indicated as a clear peak in the
power spectrum. There is a recent debate about how to assess and interpret high gamma
activity in the absence of a spectral peak (Brunet et al., 2014; Hermes et al., 2015). Further
research is needed to elucidate the distinction between “low and high” or “narrow-and
broadband” gamma activity in terms of their neurophysiological underpinnings and
functional roles. In this context, note that the observation of high gamma responses is
not bound by its superior signal-to-noise ratio to invasive depth recordings and ECoG
in patients, as gamma responses in the 90-250 Hz range have also been recorded from
conventional scalp EEGs of healthy volunteers (Darvas et al., 2010; Lenz, Jeschke,
et al., 2008).
In light of the latency difference between evoked (~50–150 ms) and induced gamma ac-
tivity (~200–300 ms), it has been suggested that evoked activity reflects an early pro-
cessing stage at the level of primary visual cortices (Zaehle et al., 2009), whereas induced
activities occur at later stages (Herrmann et al., 2010). Findings of evoked gamma ac-
tivity modulated by low-level physical features of visual stimuli support this view (for
the sake of brevity, we focus on the visual domain in the remainder of this chapter).
However, there is a remarkable overlap between evoked and induced gamma activity in
reflecting sensory vs. cognitive effects, as demonstrated in the sequel.
Regarding stimulus-driven effects, early evoked gamma activity has been observed
to increase with stimulus size and with central as compared to peripheral stimulation
(Busch et al., 2004). Also high visual contrast (Schadow et al., 2007) and low spatial
frequencies (Fründ et al., 2007) of simple grating stimuli increased the power of evoked
gamma activity. With regard to motion, evoked gamma power did not differ between
stationary and moving gratings (Naue et al., 2011; Swettenham et al., 2009). However,
inverting the black and white stripes of a stationary grating (pattern reversal) led to a
threefold increase of the evoked gamma amplitude compared to stationary and moving
gratings, probably due to related contrast effects (Naue et al., 2011, see Figure 8.2).
These exemplary findings not only indicate an early processing stage of evoked
gamma responses but also implicate that finding evoked gamma activity requires a
corresponding design of physical stimulus parameters. Relatedly, before a cognitive
effect on evoked gamma activity can be inferred from two experimental conditions, care
must be taken that physically identical stimuli were used. Otherwise, differences of the
evoked gamma response cannot unequivocally be attributed to the cognitive process
under study.
GAMMA ACTIVITY IN SENSORY AND COGNITIVE PROCESSING 151
Cognitive effects on the evoked gamma activity have been observed in visual discrim-
ination tasks where participants had to attend to a target and to ignore distractor stimuli
that were physically more or less similar to the target (Herrmann & Mecklinger, 2001;
Herrmann et al., 1999). The results not only demonstrated a target-effect (i.e., higher
evoked gamma activity to attended stimuli), but also a graded effect of similarity be-
tween target and distractor stimuli (i.e., the more features a distractor shared with the
target the stronger was the evoked gamma activity). This led to the suggestion that each
stimulus is compared to a short-term memory template of the target and that the de-
gree of matching between stimulus and template determines the strength of the evoked
gamma activity. Similarly for long-term memory (LTM), a match between line drawings
of real- world objects and their well- consolidated memory representation evoked
stronger gamma responses than the perception of unfamiliar (non-)objects without an
LTM representation (Herrmann et al., 2004).
Modulations of evoked gamma activity have also been reported in the context of
visual semantic memory (Oppermann et al., 2012). Simultaneously presented pairs
of conceptually coherent scenes (e.g., mouse–cheese) evoked stronger gamma-band
responses than semantically unrelated object pairs (e.g., camel–magnet). This effect
occurred within a widespread network of bilateral occipital as well as right temporal and
frontal regions in a time window between 70–130 ms after stimulus onset, indicating a
role of early gamma activity for a rapid memory-based extraction of the gist of a scene.
Together, studies on early evoked gamma oscillations demonstrate not only their sen-
sitivity to low-level stimulus features but also an influence of cognitive effects occurring
as early as 50–150 ms after stimulus onset, thereby revealing an early interaction between
bottom-up and top-down processes in the gamma band (Busch et al., 2006).
Especially the induced type of gamma activity has been consistently related to higher
cognitive processes (Kaiser & Lutzenberger, 2005), motivated by the initial findings
in animals suggesting a possible role of induced synchronous neural discharges in
bottom-up feature binding (see Section 8.1; Eckhorn et al., 1988; Gray et al., 1989). By
using protocols very similar to these animal studies, their main finding of an increase in
the strength of induced gamma synchrony during passive viewing of coherent moving
bars could be replicated in the human scalp EEG (Lutzenberger et al., 1995; Müller et al.,
1996). Increases of induced gamma activity have also been reported in response to co-
herent versus incoherent static stimuli during a visual discrimination task. In a classical
study, presenting an illusory Kanizsa triangle, a real triangle, and a no-triangle stimulus
with the black inducer disks rotated outwards, resulted in specific enhancements of the
induced gamma activity for the coherent triangles (illusory and real) as compared to the
no-triangle stimuli (Tallon-Baudry et al., 1996).
152 DANIEL STRÜBER and CHRISTOPH S. HERRMANN
How do gamma oscillations exert their functions in basic sensory and higher cogni-
tive functions and what are the neurophysiological mechanisms underlying oscillatory
gamma activity?
The central concept which is thought to underlie the functional relevance of gamma
activity is referred to as “oscillatory synchrony”, that is, the periodic co-occurrence
of electrical impulses from a group of neurons on a fine temporal scale. In the con-
text of EEG scalp recordings, the obtained electrophysiological data inevitably reflect
synchronized activity from large neuronal populations, because otherwise the weak
synaptic currents would not be measurable at the scalp level. Local oscillatory syn-
chrony resulting from within-area interactions would then appear as increased power
at individual electrodes, whereas oscillatory synchrony related to large-scale integration
is best characterized by phase coherence between two distant sources (Siegel et al., 2012;
Varela et al., 2001).
Oscillatory synchrony is thought to temporally structure the occurrence of spike
trains without changing the mean firing rates of neurons and, thereby, to have a func-
tional role in the processing of incoming inputs and the emergence of functional
networks by gating the information flow (Salinas & Sejnowski, 2001). Converging
evidence from experimental work in animals (in vivo and in vitro) and modeling
approaches suggests that cortical gamma oscillations and their local synchronization
result from interactions of reciprocally connected excitatory pyramidal cells and in-
hibitory gamma-aminobutyric acid (GABA)-ergic interneurons (Bartos et al., 2007;
Buzsáki & Wang, 2012; Wang, 2010; Whittington et al., 2011; Whittington et al., 2000).
In such a network, the pyramidal cells activate the interneurons, which self-
generate synchronized gamma oscillations. This gamma-synchronized activity of the
interneurons is then imposed onto the pyramidal cells, resulting in rhythmic inhib-
ition of the pyramidal cells, which in turn leads to a rhythmic synchronization of their
discharges. Thus, during each cycle of these excitatory-inhibitory feedback loops, the
time window during which the pyramidal cells are able to discharge is restricted by the
decay time of their inhibitory input from the fast-spiking interneurons, introducing a
phase delay of a few milliseconds between pyramidal and interneuron discharges (Fries
et al, 2007). From this basic mechanism of gamma synchronization several mechanistic
consequences arise that are thought to be instrumental for the formation of neuronal
cell assemblies, cortical signal transmission, and, thereby, the implementation of cogni-
tive functions (Bosman et al., 2014; Cannon et al., 2014; Vinck & Bosman, 2016).
One consequence of gamma synchronization that relates to assembly formation is
its involvement in the regulation of synaptic plasticity. The precise timing of pre-and
post-synaptic processes emerging from gamma synchronization acts on a time scale
of a few tens of milliseconds, which is relevant for spike-timing-dependent plasticity
and, thus, for the induction of long-term potentiation or depression (Fell & Axmacher,
154 DANIEL STRÜBER and CHRISTOPH S. HERRMANN
2011; Paulsen & Sejnowski, 2000; Sejnowski & Paulsen, 2006) and associative learning
(Miltner et al., 1999). It has been proposed that the period of gamma oscillations (~25
msec) is “designed” to match the time course of calcium fluctuations in dendrites and,
therefore, to facilitate learning (Bibbig et al., 2001).
With regard to local within-area interactions, synchronization in the gamma-band
with its high temporal precision in the millisecond range is hypothesized to increase the
impact of presynaptic neurons on their target cells because the gamma-synchronized
spike packages arrive close together, thus summing up more effectively to initiate
postsynaptic discharges. This effect of precise spike timing is referred to as feedforward
coincidence detection (Fries, 2009; Salinas & Sejnowski, 2001), which has been linked
to the integration (binding) of different stimulus features during object recognition
(Bosman et al., 2014).
This concept has also been transferred to inter-areal communication between mul-
tiple groups of neurons, each oscillating in the gamma frequency range. In this context,
the “communication through coherence” (CTC) hypothesis (Fries, 2005; 2015) claims
that the communication between two groups of neurons can be facilitated by gamma
synchronization in the two groups, if the spikes from the presynaptic group (sender)
arrive at the postsynaptic group (receiver) at the appropriate phase, that is, during a min-
imal amount of GABAergic inhibition. This phase-coupling of oscillations constitutes a
window of “opportunity” in which neural networks jointly involved in signal processing
can communicate. This mechanism allows for gain modulation of pre-synaptic inputs
that compete for activating higher-level post-synaptic targets as is the case with se-
lective attention. In this scenario, visual attention might selectively increase the effective
strength of those synaptic inputs from lower level neurons that process attended stimuli
at the expense of inputs from the non-attended stimuli (Bosman et al., 2014; Fries, 2015).
In this way, gamma synchronization might dynamically route the information flow be-
tween higher and lower level areas within the visual hierarchy as has been observed lo-
cally between multiple visual areas (see, for review, Bosman et al., 2014).
However, it is less clear whether gamma synchronization also serves as a mech-
anism for large-scale integration across distant brain regions that also include, for ex-
ample, frontal or parietal areas. It has been suggested that the spatial distance between
interacting brain areas and, hence, the conduction delay may define the communication
frequency, with gamma oscillations acting more locally and lower frequencies more glo-
bally (Kopell et al., 2000; von Stein & Sarnthein, 2000). A reason for this may be that, in
contrast to the millisecond precision of gamma oscillations, lower-frequency bands are
more robust to spike timing delays (Buschman & Miller, 2007). Indeed, several findings
in monkeys have shown that long-distance top-down processes are carried by inter-
areal synchrony in the alpha and low-beta frequency range, whereas gamma oscillations
index a local encoding of information and the bottom-up transfer of low-level sensory
information to higher-level areas (for reviews, see, Bressler & Richter, 2015; Gregoriou
et al., 2015; Siegel et al., 2012; Wang, 2010).
On the other hand, there is also increasing evidence for a role of gamma synchrony
in large-scale interaction during various perceptual and attentional processes (see
GAMMA ACTIVITY IN SENSORY AND COGNITIVE PROCESSING 155
for a review, Gregoriou et al., 2015), indicating that physical distance might not be the
only factor that determines the frequency band used for inter-areal synchronization.
Moreover, gamma-synchronized long-range signal transmission seems to work both
upstream (i.e., bottom-up) and downstream (i.e., top-down) within the cortical hier-
archy. For example, long-range gamma coupling between prefrontal and visual areas
during directed attention was found to be initiated frontally, thereby signaling top-
down attentional influences on the visual cortex (Gregoriou et al., 2009). Another study
reported long-range bottom-up directed gamma synchrony between posterior parietal
and prefrontal regions during automatic attention driven by salient stimuli (Buschman
& Miller, 2007). Intriguingly, the same areas synchronized in top-down direction if
attention was focused volitionally, but in this case in the beta-range (Buschman &
Miller, 2007), indicating that different frequencies are used for feedforward and feed-
back signaling.
Such frequency-specific differences in the direction of inter-areal interaction (i.e.,
bottom-up vs. top-down) have been related to the different directions of information
flow in the cortical layers with feedforward and feedback connections originating pri-
marily in superficial and deep layers, respectively (Felleman & Van Essen, 1991; Markov
et al., 2014). Accumulating evidence suggests that alpha/beta oscillations originate
in deeper layers of the visual cortex and support feedback signaling, whereas gamma
synchrony emerges from superficial layers and signals feedforward information flow
(Buffalo et al., 2011; Michalareas et al., 2016; van Kerkoerle et al., 2014; von Stein et al.,
2000). Together, these findings clearly suggest a role for gamma synchrony in local
bottom-up interaction between visual cortices (Siegel et al., 2012). However, it remains
an open question how this function can be reconciled with the documented long-range
gamma synchrony during top-down attention (Gregoriou et al., 2015).
Overall, the recurrent excitatory-inhibitory network interactions underlying local
gamma synchronization are thought to establish low-level circuit functions (e.g.,
synaptic plasticity, coincidence detection, gain modulation, phase coding, dynamic
routing) that may act as elementary building blocks of cognitive functions (e.g., visual
feature integration, selective attention, learning, and memory) in a task-specific com-
bination (Bosman et al., 2014; Siegel et al., 2012). Nevertheless, there are multiple meth-
odological concerns that need to be considered to reliably assess EEG gamma activity as
outlined in the next section.
single images are presented in order to show a movie (cf. Figure 8.2 right panel, SSVEP).
Second, physiological artefacts like eye movements and electromyogenic (EMG) ac-
tivity generated by muscles from the scalp, face, and neck (see, for review, Nottage &
Horder, 2016).
Power line noise occurs at a frequency of 50 Hz in Europe and 60 Hz in the US; both
frequencies are in the gamma range. Therefore, any power supply like electric cables,
wall sockets, or electrically operated equipment inside the recording cabin (lamps)
results in a 50-or 60-Hz peak in the EEG power spectrum. To effectively avoid frequency
interference, experiments should be conducted in an electrically shielded room with all
devices inside the cabin operated on batteries. The stimulation monitor can be placed
outside the cabin behind an electrically shielded window and a fiberoptic cable can be
used to transfer the EEG data to a computer outside the recording cabin. If it is not pos-
sible to prevent power line noise from being recorded, a low-pass or “notch” filter at
50 or 60 Hz might be applied to the data. However, the use of such filters is problem-
atic because they might induce “ringing” (i.e., the induction of spurious oscillations by
transients in the EEG) and distort the phase of neural oscillations. Alternative strategies
have been developed on the basis of noise cancellation (see Nottage, 2010 for details).
In addition to power line noise, every screen refresh generates an electromagnetic
signal which might be seen as a narrow band in the EEG gamma range, typically at
anterior electrodes. The exact frequency depends on the primary refresh rate and its
harmonics. Nottage & Horder (2016) explore possible ways to remove this type of
artefact.
The difficulty of properly removing EMG artifacts is particularly important for studies
on gamma activity, given the fact that the broad frequency spectrum of EMG activity
substantially overlaps with the gamma frequency range. An amplitude maximum
of EMG contamination in the gamma range (40–80 Hz) was found at temporal sites
during phasic contraction of facial muscles (Goncharova et al., 2003). Because of their
high amplitude, occasionally occurring phasic muscle contractions (e.g., chewing
or jaw clenching) can be detected relatively easily by visual inspection or mathemat-
ical algorithms and then be omitted from further analysis (e.g., Fitzgibbon et al., 2015).
However, in addition to large phasic contractions, the head and neck muscles are con-
stantly active to maintain posture, which might result in tonic EMG activity throughout
an EEG session.
GAMMA ACTIVITY IN SENSORY AND COGNITIVE PROCESSING 157
Critically, low-amplitude tonic muscle activity occurs even in “relaxed” states and
is, thus, difficult to detect in scalp recordings but still contributes significantly to EMG
contamination in the gamma range, as has been shown in pharmacologically paralyzed
human volunteers (Whitham et al., 2007). Paralysis compared to pre-paralyzed states
of the same participants was found to result in a marked reduction of spectral power in
the gamma range, with 84% of the power derived from EMG, mostly over peripheral
scalp regions. Moreover, the execution of different cognitive tasks induced a broadband
increase in the gamma frequency range in unparalyzed but not paralyzed participants,
indicating that EMG activity was responsible for the task-related spectral power
increases (Whitham et al., 2008). However, when employing a standard oddball task
that allows for stimulus-locked data analysis, significant gamma activity was identified
in both the pre-paralyzed and the paralyzed condition, with stronger activity for rare
target than for more frequent standard stimuli (Pope et al., 2009).
While these studies clearly demonstrate a large contribution of EMG activity to the
overall EEG spectral power, they also show that (at least under experimental conditions
where time-and stimulus-locked analysis is possible, and in contrast to, for example,
analyzing states of varying cognitive load during mental arithmetic) neuronal gamma
activity can still be measured at the scalp EEG. Thus, it seems possible to extract neural
gamma activity from EMG noise if it is analyzed in response to discrete stimuli (e.g., co-
herent stimuli vs. incoherent visual objects). For this, however, methods for the effective
removal of EMG artefacts are required.
There is a multitude of methods for dealing with scalp EMG (see, for review, Nottage
& Horder, 2016). One novel approach that specifically addresses the tonic nature of scalp
and neck muscle artefacts uses mathematical modeling to fit individual muscle spikes
and subtract these from the signal, resulting in an effective correction of the gamma ac-
tivity associated with a self-paced motor task (Nottage et al., 2013). Recently, Janani and
colleagues (2018) used datasets that are free of muscle activity due to paralysis (taken
from Whitham et al., 2007, 2008) to identify limitations of traditional approaches and,
then, to evaluate the improvements of a newly developed algorithm for tonic muscle
artefact removal. With this method, high-frequency EMG artefacts were reduced con-
siderably, although a residual artefact still remains (compared to paralysis). Given the
variety of techniques that are based on different mathematical concepts, the authors
suggested using a combination of complementary algorithms for a further improve-
ment of EMG artefact removal (Janani et al., 2018).
Ocular activity also generates muscle-related artefacts. The main sources of eye
movement artefacts are blinks and saccades. It is standard practice to measure the
electro-oculogram (EOG) from two channels of the left and right eye. For rejecting eye
blinks and large saccades, an amplitude threshold is usually set (e.g., 50 µV), which is
then used by an algorithm to exclude contaminated EEG trials in any channel from fur-
ther analysis. This automatic amplitude threshold procedure is then complemented by
visual inspection of all epochs. For algorithms based on eye tracking data that detect
and correct eye blinks and other ocular artifacts in a fully automated fashion, see Plöchl
et al., 2012.
158 DANIEL STRÜBER and CHRISTOPH S. HERRMANN
Blink artefacts elicit a large potential with a predominantly frontal topography and are
easy to identify in the vertical electro-oculogram (VEOG). Saccades are generated by a
pair of extra-ocular muscles which contract to move the eyeball, thereby inducing a sac-
cadic spike potential in the EEG at the onset of each saccade (Thickbroom & Mastaglia,
1985). Saccadic spike potentials are too small to be detected by standard amplitude
thresholds for automatic EOG-artefact removal, especially in case of microsaccades
which are produced during attempted fixation (Martinez-Conde et al., 2013). In their
seminal paper, Yuval-Greenberg and colleagues (2008) suspected that high-frequency
EEG activity that had been regarded as induced gamma activity in fact reflects
microsaccade-induced muscle artefacts. Naturally, this report triggered an intense de-
bate within the field of induced gamma activity research (Schwartzman & Kranczioch,
2011). Section 8.9 summarizes the main points.
Microsaccades are small (up to 1 degree), jerk-like saccades with a duration of about
25 ms and an average rate of 1–2 per second (Martinez-Conde et al., 2013). There are at
least three important similarities between microsaccades and induced gamma activity
that need to be considered. First, in response to a sudden change in stimulus input, the
microsaccade rate shows a characteristic biphasic time course, with an early inhibition
phase peaking at 100–150 ms after stimulus onset, followed by a rebound phase peaking
between 200 and 400 ms after stimulus onset (Engbert, 2006). Thus, the rebound phase
of microsaccades corresponds to the time window where induced gamma activity typ-
ically appears.
Second, the rate of microsaccades is modulated by perceptual and cognitive factors
that have also been linked to induced gamma activity, including attention and memory-
related processes (Engbert & Kliegl, 2003; Valsecchi et al., 2007; Valsecchi et al., 2009),
physical stimulus features like color and luminance contrast (Rolfs et al., 2008), and the
coherence of objects (Yuval-Greenberg et al., 2008). Thus, microsaccade rate represents
a true confound in experiments on gamma activity since modulations of microsaccade
rate and gamma activity might generate an identical pattern of results across experi-
mental conditions. For example, both microsaccade rate and induced gamma power
show increased responses to a coherent object in comparison to an incoherent object.
Third, depending on the location of the reference electrode, the frontally occurring
microsaccade-induced saccadic spike potential translates into a broadband (~20–90
Hz) gamma power increase with a maximum at centro-parietal and occipital electrodes,
that is, regions where it strongly coincides with induced gamma activity (Reva &
Aftanas, 2004; Yuval-Greenberg et al., 2008). Thus, saccadic spike potentials mimic not
only the frequency content of induced gamma activity but also its typical topography.
GAMMA ACTIVITY IN SENSORY AND COGNITIVE PROCESSING 159
Notably, this can be avoided by using MEG, since it does not require a reference (Gruber
et al., 2008).
These similarities between microsaccades and induced gamma activity regarding
their sensitivity to cognitive effects as well as their temporal, topographical and spec-
tral properties have been convincingly demonstrated and have led to the suggestion
that—at least in some cases—induced gamma activity reflect saccadic spike potentials
rather than a neural response (Yuval-Greenberg et al., 2008). This also concerns earlier
EEG findings that might have been erroneously interpreted as neuronal gamma activity,
when in fact they reflected microsaccade-induced spike potential artefacts (Keren et al.,
2010). How strong previous EEG findings might be affected by microsaccades depends
on several aspects, for example, the exact time window of the effects (in relation to the
saccadic rebound), the broadness of the gamma response (in relation to the 20–90 Hz
spike potential), and the choice of the reference electrode, given its influence on the top-
ography, to name but a few ( for further details, see Schwartzman & Kranczioch, 2011).
Notably, evoked gamma activity is not affected by microsaccade-induced spike po-
tential artefacts. One reason is that evoked gamma activity occurs much earlier than
the microsaccade rebound after sensory stimulation. Another reason is that spike
potentials, in contrast to the evoked gamma activity, are not tightly time-locked to
the stimulus and, thus, cancel out during averaging of the evoked potential (Yuval-
Greenberg & Deouell, 2009).
Although Yuval-Greenberg and colleagues (2008) did not deny the existence of
induced gamma activity and its role in perception and cognition in general, their findings
posed a challenge to develop methods for separating the effects of microsaccades and
induced gamma responses in EEG scalp recordings. One immediate consequence of
this study was the necessity to use eye tracking with sufficient spatial and temporal reso-
lution for detecting microsaccades as small as 0.15° visual angle, given that even precise
fixation of a continuously present fixation point does not preclude microsaccade-related
brain activity (Dimigen et al., 2009). The rationale behind a combined recording of EEG
and eye tracking data is to demonstrate that differences of induced gamma activity and
microsaccade rate do not covary across conditions and, thus, effects of induced gamma
activity cannot easily be explained by microsaccadic muscle potentials. Indeed, such
differential modulations of gamma activity and microsaccades time courses have been
reported, for example, during object motion (Naue et al., 2011) and memory-based ob-
ject recognition (Hassler et al., 2013).
However, high- resolution eye trackers are expensive and not always available.
Alternatively, there are methods available allowing for offline artefact correction
that identify saccadic spike potentials via EOG sensors, that is, without the need of
an eye tracker (Hassler et al., 2011; Keren et al., 2010; Nottage, 2010). However, Keren
and colleagues (2010) used eye tracking in addition to EOG-based detection of
microsaccades to evaluate the hit and false alarm rates of several spatiotemporal filters
that were applied to the EOG data to identify the sharp amplitude increase at the onset
of saccadic spike potentials. With this method, detection rates of 80% on average could
be achieved for microsaccades of at least 0.2°, which were then efficiently attenuated by
means of mathematical algorithms. For artifact correction methods based on combined
160 DANIEL STRÜBER and CHRISTOPH S. HERRMANN
EEG and eye tracking recordings, see also Plöchl et al. (2012). Nottage (2010) and
Hassler and colleagues (2011) used different mathematical approaches but both have
been demonstrated to effectively remove microsaccadic muscle artefacts, resulting in a
more sustained and narrow-band gamma signal of the residual activity compared to the
frequency plot of the saccadic spike potential (see, for review, Nottage & Horder, 2016).
Although the available methods are relatively easy to apply and effective in reducing
saccadic spike potentials, they do not remove the artefact completely and they do not
attenuate saccade-related visual brain activity (Dimigen et al., 2009; Plöchl et al., 2012).
Also, the fundamental problem of the confounding co-modulation of induced gamma
activity and microsaccade rate cannot be solved by artefact correction techniques.
Therefore, it has been suggested recently to reduce the occurrence of microsaccades and
related confounds in the EEG at the level of experimental design instead of applying off-
line correction methods (Tal & Yuval-Greenberg, 2018). Considerably reducing the in-
cidence of microsaccades during experimentation would allow rejecting microsaccades
rather than just correcting artefactual trials and still leaving enough (then artefact-free)
trials for analysis. As a first step, Tal and Yuval-Greenberg (2018) were able to reduce the
average number of saccades (including microsaccades) by 10–25% through manipula-
tion of different task characteristics (e.g., by adding a foveal task, whereas the stimulus
of interest was parafoveal). However, it remains to be seen how this approach can be
applied to the diverse experimental setups covering the full range of induced gamma ac-
tivity related cognitive processes.
In summary, together with other muscle artefacts, microsaccades impose a serious
difficulty on the analysis and interpretation of induced gamma activity in human EEG.
To separate activity related to saccadic spike potentials from induced gamma activity,
using high-resolution eye tracking is recommended as the most reliable way to iden-
tify microsaccades. However, there are also techniques available to use EOG data for
saccadic spike potential detection. In a second step, off-line artefact removal methods
should be used to correct artefactual trials. Rejecting all artefactual trials is currently
not possible due to the high prevalence of microsaccades in typical visual experiments,
but there are suggestions how to reduce the number of microsaccades through experi-
mental design. Current evidence from artefact-corrected data confirms the influence of
microsaccades on induced gamma activity, but also indicates the existence of induced
gamma activity that survived artifact suppression, thereby replicating earlier findings
on the functional roles of induced gamma activity in, for instance, object representation
(Hassler et al., 2013; 2011). Finally, depending on the research question at hand, a further
option to prevent microsaccadic contamination of gamma activity might be to restrict
the analyses to the evoked gamma activity, which is not affected (e.g., Lally et al., 2014).
Given its role in multiple cognitive processes and neural information integration,
disturbed gamma activity might reflect an important pathophysiological mechanism
GAMMA ACTIVITY IN SENSORY AND COGNITIVE PROCESSING 161
Notably, however, there is recent evidence from a mouse model of AD that impaired
gamma oscillations are specifically related to a hallmark of AD, that is, the abnormal
aggregation of plaque-forming proteins (amyloid-beta protein) in the brain (Iaccarino
et al., 2016). The authors found that inducing gamma oscillations by a 40-Hz flickering
light led to a major reduction of plaques in the visual cortex of the exposed mice and
to an activation of microglia (i.e., immune cells in the brain) that degrade the plaques
(Iaccarino et al., 2016). Crucially, later studies using different forms of sensory gamma
stimulation in multiple AD mouse models could not only demonstrate a clear link be-
tween gamma activity and cellular metabolism but also improvements of cognitive
functions like learning and memory, suggesting a therapeutic role of gamma entrain-
ment in AD (see, for review, Adaikkan & Tsai, 2020). There is emerging evidence that
similar cognitive effects can be achieved in humans by employing non-invasive brain
stimulation techniques that could, in contrast to rhythmic sensory stimulation, target
higher-order cognitive brain areas linked more directly to core symptoms of the dis-
order than sensory cortices (Benussi et al., 2021; see, for reviews, Bréchet et al., 2021, and
Strüber & Herrmann, 2020).
There are a number of challenges in relation to the ubiquitous role of gamma oscillations
for cognitive functions. For instance, it has been criticized that the power of gamma
oscillations is low and inconsistent and that its frequency and power depend on low-
level stimulus features like size or contrast, which seems to be incompatible with the
proposed role of gamma synchronization in cortical processing (Ray & Maunsell, 2015).
However, there is also recent reconciling evidence in support of a functional role for
gamma synchronization (Singer, 2018).
One related issue regarding the functions of gamma activity in sensory and cogni-
tive processing as well as in clinical contexts is that most of the evidence is correlative
in nature. Therefore, it remains unclear whether gamma oscillations and their syn-
chronization are truly relevant for information processing in the brain or whether they
merely reflect a by-product of brain organization. It has been argued that even gamma
oscillations as such may be a functional epiphenomenon arising from network activities
supporting the excitatory-inhibitory balance necessary for normal brain functioning
(Merker, 2013). Although this argumentation might reflect a “category mistake” since
gamma oscillations cannot be functionally separated from the circuit mechanisms that
generate them (Bosman et al., 2014), the necessity remains to go beyond correlations
and to demonstrate that gamma activity is causally involved in cortical information pro-
cessing and cognitive functions.
Causal evidence for a role of gamma oscillation in cognitive processes can be
provided by demonstrating that modulating gamma power or coherence improves cog-
nitive outcomes. There are several established methods to modulate brain oscillations,
GAMMA ACTIVITY IN SENSORY AND COGNITIVE PROCESSING 165
REFERENCES
Adaikkan, C. & Tsai, L.-H. (2020). Gamma entrainment: Impact on neurocircuits, glia, and
therapeutic opportunities. Trends in Neurosciences, 43(1), 24–41. https://doi.org/10.1016/
j.tins.2019.11.001
Adrian, E. D. (1942). Olfactory reactions in the brain of the hedgehog. Journal of Physiology,
100(4), 459–473.
Adrian, E. D. (1950). The electrical activity of the mammalian olfactory bulb.
Electroencephalography and Clinical Neurophysiology, 2, 377–388.
Barry, R. J., Clarke, A. R., Hajos, M., McCarthy, R., Selikowitz, M., & Dupuy, F. E. (2010).
Resting-state EEG gamma activity in children with attention-deficit/hyperactivity disorder.
Clinical Neurophysiology, 121(11), 1871–1877. http://doi.org/10.1016/j.clinph.2010.04.022
166 DANIEL STRÜBER and CHRISTOPH S. HERRMANN
Bartos, M., Vida, I., & Jonas, P. (2007). Synaptic mechanisms of synchronized gamma
oscillations in inhibitory interneuron networks. Nature Reviews Neuroscience, 8(1), 45–56.
http://doi.org/10.1038/nrn2044
Başar, E. (2013). A review of gamma oscillations in healthy subjects and in cognitive impair-
ment. International Journal of Psychophysiology, 90(2), 99–117. http://doi.org/10.1016/j.ijpsy
cho.2013.07.005
Başar, E., Başar-Eroǧlu, C., Güntekin, B., & Yener, G. G. (2013). Brain’s alpha, beta, gamma,
delta, and theta oscillations in neuropsychiatric diseases: Proposal for biomarker
strategies. Supplements to Clinical Neurophysiology, 62, 19–54. http://doi.org/10.1016/
B978-0-7020-5307-8.00002-8
Başar, E., Emek-Savaş, D. D., Güntekin, B., & Yener, G. G. (2016). Delay of cognitive gamma
responses in Alzheimer’s disease. NeuroImage: Clinical, 11, 106–115. http://doi.org/10.1016/
j.nicl.2016.01.015
Başar, E., Femir, B., Emek-Savaş, D. D., Güntekin, B., & Yener, G. G. (2017). Increased long dis-
tance event-related gamma band connectivity in Alzheimer’s disease. NeuroImage: Clinical,
14, 580–590. http://doi.org/10.1016/j.nicl.2017.02.021
Başar, E., Gönder, A., Özesmi, C., & Ungan, P. (1975). Dynamics of brain rhythmic and evoked
potentials. II. Studies in the auditory pathway, reticular formation and hippocampus during
the waking stage. Biological Cybernetics, 20(3-4), 145–160.
Başar, E., Gönder, A., & Ungan, P. (1976). Important relation between EEG and brain evoked
potentials. I. Resonance phenomena in subdural structures of the cat. Biological Cybernetics,
25(1), 27–40.
Başar, E. & Özesmi, C. (1972). The hippocampal EEG-activity and a systems analytical inter-
pretation of averaged evoked potentials of the brain. Kybernetik, 12, 45–54.
Başar, E., Schmiedt-Fehr, C., Mathes, B., Femir, B., Emek-Savaş, D. D., Tülay, E., Tan, D.,
Düzgün, A., . . . Başar-Eroǧlu, C. (2016). What does the broken brain say to the neuroscien-
tist? Oscillations and connectivity in schizophrenia, Alzheimer’s disease, and bipolar dis-
order. International Journal of Psychophysiology, 103, 135–148. http://doi.org/10.1016/j.ijpsy
cho.2015.02.004
Başar-Eroǧlu, C., Brand, A., Hildebrandt, H., Karolina Kedzior, K., Mathes, B., & Schmiedt,
C. (2007). Working memory related gamma oscillations in schizophrenia patients.
International Journal of Psychophysiology, 64(1), 39–45. http://doi.org/10.1016/j.ijpsy
cho.2006.07.007
Başar-Eroǧlu, C., Mathes, B., Brand, A., & Schmiedt- Fehr, C. (2011). Occipital γ re-
sponse to auditory stimulation in patients with schizophrenia. International Journal of
Psychophysiology, 79, 3–8.
Başar-Eroǧlu, C., Strüber, D., Kruse, P., Başar, E., & Stadler, M. (1996). Frontal gamma-band en-
hancement during multistable visual perception. International Journal of Psychophysiology,
24(1-2), 113–125. http://doi.org/10.1016/S0167-8760(96)00055-4
Başar-Eroǧlu, C., Strüber, D., Schürmann, M., Stadler, M., & Başar, E. (1996). Gamma-band
responses in the brain: A short review of psychophysiological correlates and functional sig-
nificance. International Journal of Psychophysiology, 24(1-2), 101–112. http://doi.org/10.1016/
S0167-8760(96)00051-7
Benussi, A., Cantoni, V, Cotelli, M. S., Cotelli, M., Brattini, C., Datta, A., Thomas, C.,
Santarnecchi, E., Pascual-Leone, A., & Borroni, B. (2021). Exposure to gamma tACS in
Alzheimer's disease: A randomized, double-blind, sham-controlled, crossover, pilot study.
Brain Stimulation, 14, 531–540. https://doi.org/10.1016/j.brs.2021.03.007
GAMMA ACTIVITY IN SENSORY AND COGNITIVE PROCESSING 167
Berger, H. (1929). Über das Elektrenkephalogramm des Menschen, I. Mitteilung. Archiv Für
Psychiatrie Und Nervenkrankheiten, 87(4), 527–570.
Berger, H. (1934). Über das Elektrenkephalogramm des Menschen, IX. Mitteilung. Archiv Für
Psychiatrie Und Nervenkrankheiten, 102, 538–557. http://doi.org/10.1007/BF01797193
Berger, H. (1936). Über das Elektrenkephalogramm des Menschen, XI. Mitteilung. Archiv Für
Psychiatrie Und Nervenkrankheiten, 104, 678–689. http://doi.org/10.1007/BF01797193
Berger, H. (1937). Über das Elektrenkephalogramm des Menschen, XII. Mitteilung. Archiv Für
Psychiatrie Und Nervenkrankheiten, 106, 165–187. http://doi.org/10.1007/BF01797193
Bibbig, A., Faulkner, H. J., Whittington, M. A., & Traub, R. D. (2001). Self-organized synaptic
plasticity contributes to the shaping of gamma and beta oscillations in vitro. Journal of
Neuroscience, 21(22), 9053–9067. http://doi.org/21/22/9053 [pii]
Bosman, C. A., Lansink, C. S., & Pennartz, C. M. A. (2014). Functions of gamma-band syn-
chronization in cognition: From single circuits to functional diversity across cortical and
subcortical systems. European Journal of Neuroscience, 39(11), 1982–1999. http://doi.org/
10.1111/ejn.12606
Bouyer, J. J., Montaron, M. F., & Rougeul, A. (1981). Fast fronto-parietal rhythms during
combined focused attentive behaviour and immobility in cat: Cortical and thalamic
localizations. Electroencephalography and Clinical Neurophysiology, 51(3), 244–252. http://
doi.org/10.1016/0013-4694(81)90138-3
Bréchet, L., Michel, C. M., Schacter, D. L., & Pascual-Leone, A. (2021). Improving auto-
biographical memory in Alzheimer's disease by transcranial alternating current stimu-
lation. Current Opinion in Behavioral Sciences, 40, 64–71. https://doi.org/10.1016/j.cob
eha.2021.01.003
Bressler, S. L. & Freeman, W. J. (1980). Frequency analysis of olfactory system EEG in cat,
rabbit, and rat. Electroencephalography and Clinical Neurophysiology, 50(1-2), 19–24. http://
doi.org/10.1016/0013-4694(80)90319-3
Bressler, S. L. & Richter, C. G. (2015). Interareal oscillatory synchronization in top-down neo-
cortical processing. Current Opinion in Neurobiology, 31, 62–66. http://doi.org/10.1016/
j.conb.2014.08.010
Brunet, N., Vinck, M., Bosman, C. A., Singer, W., & Fries, P. (2014). Gamma or no gamma,
that is the question. Trends in Cognitive Sciences, 18(10), 507–509. http://doi.org/10.1016/
j.tics.2014.08.006
Buffalo, E. A., Fries, P., Landman, R., Buschman, T. J., & Desimone, R. (2011). Laminar
differences in gamma and alpha coherence in the ventral stream. Proceedings of the National
Academy of Sciences of the United States of America, 108(27), 11262–11267.
Busch, N. A., Debener, S., Kranczioch, C., Engel, A. K., & Herrmann, C. S. (2004). Size
matters: Effects of stimulus size, duration and eccentricity on the visual gamma-
band response. Clinical Neurophysiology, 115(8), 1810–1820. http://doi.org/10.1016/j.cli
nph.2004.03.015
Busch, N. A., Herrmann, C. S., Müller, M. M., Lenz, D., & Gruber, T. (2006). A cross-
laboratory study of event-related gamma activity in a standard object recognition paradigm.
NeuroImage, 33(4), 1169–1177. http://doi.org/10.1016/j.neuroimage.2006.07.034
Busch, N. A., Schadow, J., Fründ, I., & Herrmann, C. S. (2006). Time-frequency analysis of
target detection reveals an early interface between bottom-up and top-down processes
in the gamma- band. NeuroImage, 29(4), 1106–1116. http://doi.org/10.1016/j.neuroim
age.2005.09.009
168 DANIEL STRÜBER and CHRISTOPH S. HERRMANN
Gregoriou, G. G., Paneri, S., & Sapountzis, P. (2015). Oscillatory synchrony as a mechanism
of attentional processing. Brain Research, 1626, 165–182. http://doi.org/10.1016/j.brain
res.2015.02.004
Gruber, T., Maess, B., Trujillo-Barreto, N. J., & Müller, M. M. (2008). Sources of synchronized
induced gamma-band responses during a simple object recognition task: A replica-
tion study in human MEG. Brain Research, 1196, 74–84. http://doi.org/10.1016/j.brain
res.2007.12.037
Gruber, T., Müller, M. M., Keil, A., & Elbert, T. (1999). Selective visual-spatial attention induced
gamma band responses in the human EEG. Clinical Neurophysiology, 110(1999), 2074–2085.
Gruber, T., Trujillo-Barreto, N. J., Giabbiconi, C. M., Valdés-Sosa, P. A., & Müller, M. M.
(2006). Brain electrical tomography (BET) analysis of induced gamma band responses
during a simple object recognition task. NeuroImage, 29(3), 888–900. http://doi.org/10.1016/
j.neuroimage.2005.09.004
Gruber, T., Tsivilis, D., Montaldi, D., & Müller, M. M. (2004). Induced gamma band
responses: An early marker of memory encoding and retrieval. NeuroReport, 15(11), 1837–
1841. http://doi.org/10.1097/01.wnr.0000137077.26010.12
Haenschel, C., Bittner, R. A., Waltz, J., Haertling, F., Wibral, M., Singer, W., . . . Rodriguez, E.
(2009). Cortical oscillatory activity is critical for working memory as revealed by deficits
in early-onset schizophrenia. Journal of Neuroscience, 29(30), 9481–9489. http://doi.org/
10.1523/JNEUROSCI.1428-09.2009
Hassler, U., Friese, U., Martens, U., Trujillo-Barreto, N., & Gruber, T. (2013). Repetition priming
effects dissociate between miniature eye movements and induced gamma-band responses
in the human electroencephalogram. European Journal of Neuroscience, 38(3), 2425–2433.
http://doi.org/10.1111/ejn.12244
Hassler, U., Trujillo Barreto, N., & Gruber, T. (2011). Induced gamma band responses in human
EEG after the control of miniature saccadic artifacts. NeuroImage, 57(4), 1411–1421. http://
doi.org/10.1016/j.neuroimage.2011.05.062
Hermes, D., Miller, K. J., Wandell, B. A., & Winawer, J. (2015). Gamma oscillations in visual
cortex: The stimulus matters. Trends in Cognitive Sciences, 19(2), 57–58. http://doi.org/
10.1016/j.tics.2014.12.009
Herrmann, C. S. (2001). Human EEG responses to 1-100 Hz flicker: Resonance phenomena in
visual cortex and their potential correlation to cognitive phenomena. Experimental Brain
Research, 137(3-4), 346–353. http://doi.org/10.1007/s002210100682
Herrmann, C. S. & Demiralp, T. (2005). Human EEG gamma oscillations in neuropsychi-
atric disorders. Clinical Neurophysiology, 116(12), 2719–2733. http://doi.org/10.1016/j.cli
nph.2005.07.007
Herrmann, C. S., Fründ, I., & Lenz, D. (2010). Human gamma-band activity: A review on
cognitive and behavioral correlates and network models. Neuroscience and Biobehavioral
Reviews, 34(7), 981–992. http://doi.org/10.1016/j.neubiorev.2009.09.001
Herrmann, C. S., Lenz, D., Junge, S., Busch, N. A., & Maess, B. (2004). Memory-matches evoke
human gamma-responses. BMC Neuroscience, 5, 13.
Herrmann, C. S. & Mecklinger, A. (2001). Gamma activity in human EEG is related to
highspeed memory comparisons during object selective attention. Visual Cognition, 8(3-5),
593–608. http://doi.org/10.1080/13506280143000142
Herrmann, C. S., Mecklinger, A., & Pfeifer, E. (1999). Gamma responses and ERPs in a visual
classification task. Clinical Neurophysiology, 110(4), 636–642. http://doi.org/10.1016/
S1388-2457(99)00002-4
GAMMA ACTIVITY IN SENSORY AND COGNITIVE PROCESSING 171
Herrmann, C. S., Munk, M. H. J., & Engel, A. K. (2004). Cognitive functions of gamma-band
activity: Memory match and utilization. Trends in Cognitive Sciences, 8(8), 347–355. http://
doi.org/10.1016/j.tics.2004.06.006
Herrmann, C. S., Rach, S., Vosskuhl, J., & Strüber, D. (2014). Time-frequency analysis of event-
related potentials: A brief tutorial. Brain Topography, 27(4), 438–450. http://doi.org/10.1007/
s10548-013-0327-5
Herrmann, C. S. & Strüber, D. (2017). What can transcranial alternating current stimulation
tell us about brain oscillations? Current Behavioral Neuroscience Reports, 4(2), 128–137.
http://doi.org/10.1007/s40473-017-0114-9
Herrmann, C. S., Strüber, D., Helfrich, R. F., & Engel, A. K. (2016). EEG oscillations: From
correlation to causality. International Journal of Psychophysiology, 103, 12–21. http://doi.org/
10.1016/j.ijpsycho.2015.02.003
Iaccarino, H. F., Singer, A. C., Martorell, A. J., Rudenko, A., Gao, F., Gillingham, T. Z., . . . Tsai, L.
H. (2016). Gamma frequency entrainment attenuates amyloid load and modifies microglia.
Nature, 540(7632), 230–235. http://doi.org/10.1038/nature20587
Jadi, M. P., Behrens, M. M., & Sejnowski, T. J. (2016). Abnormal gamma oscillations in N-
Methyl-D-Aspartate receptor hypofunction models of schizophrenia. Biological Psychiatry,
79(9), 716–726. http://doi.org/10.1016/j.biopsych.2015.07.005
Janani, A. S., Grummett, T. S., Lewis, T. W., Fitzgibbon, S. P., Whitham, E. M., DelosAngeles,
D., . . . Pope, K. J. (2018). Improved artefact removal from EEG using canonical correl-
ation analysis and spectral slope. Journal of Neuroscience Methods, 298, 1–15. http://doi.org/
10.1016/j.jneumeth.2018.01.004
Jasper, H. H. & Andrews, H. L. (1938). Electro-encephalography III. Normal differentiation of
occipital and precentral regions in man. Archives of Neurology & Psychiatry, 39(1), 96–115.
Jensen, O., Kaiser, J., & Lachaux, J. P. (2007). Human gamma-frequency oscillations associated
with attention and memory. Trends in Neurosciences, 30(7), 317–324. http://doi.org/10.1016/
j.tins.2007.05.001
Kaiser, J. & Lutzenberger, W. (2005). Human gamma-band activity: A window to cognitive pro-
cessing. NeuroReport, 16(3), 207–211. http://doi.org/10.1097/00001756-200502280-00001
Keizer, A. W., Verment, R. S., & Hommel, B. (2010). Enhancing cognitive control through
neurofeedback: A role of gamma-band activity in managing episodic retrieval. NeuroImage,
49(4), 3404–3413. http://doi.org/10.1016/j.neuroimage.2009.11.023
Keizer, A. W., Verschoor, M., Verment, R. S., & Hommel, B. (2010). The effect of gamma
enhancing neurofeedback on the control of feature bindings and intelligence measures.
International Journal of Psychophysiology, 75(1), 25–32. http://doi.org/10.1016/j.ijpsy
cho.2009.10.011
Keren, A. S., Yuval-Greenberg, S., & Deouell, L. Y. (2010). Saccadic spike potentials in gamma-
band EEG: Characterization, detection and suppression. NeuroImage, 49(3), 2248–2263.
http://doi.org/10.1016/j.neuroimage.2009.10.057
Kopell, N., Ermentrout, G. B., Whittington, M. A., & Traub, R. D. (2000). Gamma rhythms
and beta rhythms have different synchronization properties. Proceedings of the National
Academy of Sciences of the United States of America, 97(4), 1867–1872. http://doi.org/10.1073/
pnas.97.4.1867
Kreiter, A. K. & Singer, W. (1996). Stimulus-dependent synchronization of neuronal responses
in the visual cortex of the awake macaque monkey. Journal of Neuroscience, 16(7), 2381–2396.
Lally, N., Mullins, P. G., Roberts, M. V., Price, D., Gruber, T., & Haenschel, C. (2014).
Glutamatergic correlates of gamma- band oscillatory activity during cognition: A
172 DANIEL STRÜBER and CHRISTOPH S. HERRMANN
Minzenberg, M. J., Firl, A. J., Yoon, J. H., Gomes, G. C., Reinking, C., & Carter, C. S. (2010).
Gamma oscillatory power is impaired during cognitive control independent of medication
status in first-episode schizophrenia. Neuropsychopharmacology, 35(13), 2590–2599. http://
doi.org/10.1038/npp.2010.150
Müller, M. M., Bosch, J., Elbert, T., Kreiter, A., Sosa, M., Sosa, P., & Rockstroh, B. (1996).
Visually induced gamma-band responses in human electroencephalographic activity: A
link to animal studies. Experimental Brain Research, 112(1), 96–102. http://doi.org/10.1007/
BF00227182
Naue, N., Strüber, D., Fründ, I., Schadow, J., Lenz, D., Rach, S., Körner, U., & Herrmann, C.
S. (2011). Gamma in motion: Pattern reversal elicits stronger gamma-band responses than
motion. NeuroImage, 55(2), 808–817. http://doi.org/10.1016/j.neuroimage.2010.11.053
Nimmrich, V., Draguhn, A., & Axmacher, N. (2015). Neuronal network oscillations in
neurodegenerative diseases. Neuromolecular Medicine, 17(3), 270–284. http://doi.org/
10.1007/s12017-015-8355-9
Nottage, J. F. (2010). Uncovering gamma in visual tasks. Brain Topography, 23(1), 58–7 1. http://
doi.org/10.1007/s10548-009-0129-y
Nottage, J. F. & Horder, J. (2016). State-of-the-art analysis of high-frequency (gamma range)
electroencephalography in humans. Neuropsychobiology, 72(3-4), 219–228. http://doi.org/
10.1159/000382023
Nottage, J. F., Morrison, P. D., Williams, S. C. R., & Ffytche, D. H. (2013). A novel method for
reducing the effect of tonic muscle activity on the gamma band of the scalp EEG. Brain
Topography, 26(1), 50–61. http://doi.org/10.1007/s10548-012-0255-9
Oppermann, F., Hassler, U., Jescheniak, J. D., & Gruber, T. (2012). The rapid extraction of gist—
early neural correlates of high-level visual processing. Journal of Cognitive Neuroscience,
24(2), 521–529.
Palop, J. J. & Mucke, L. (2016). Network abnormalities and interneuron dysfunction in
Alzheimer disease. Nature Reviews Neuroscience, 17(12), 777–792. http://doi.org/10.1038/
nrn.2016.141
Paulsen, O. & Sejnowski, T. J. (2000). Natural patterns of activity and long-term syn-
aptic plasticity. Current Opinion in Neurobiology, 10(2), 172–179. http://doi.org/10.1016/
S0959-4388(0 0)00076-3
Plöchl, M., Ossandón, J. P., & König, P. (2012). Combining EEG and eye
tracking: Identification, characterization, and correction of eye movement artifacts in
electroencephalographic data. Frontiers in Human Neuroscience, 6, 278. http://doi.org/
10.3389/fnhum.2012.00278
Pope, K. J., Fitzgibbon, S. P., Lewis, T. W., Whitham, E. M., & Willoughby, J. O. (2009). Relation
of gamma oscillations in scalp recordings to muscular activity. Brain Topography, 22(1), 13–
17. http://doi.org/10.1007/s10548-009-0081-x
Prehn-Kristensen, A., Wiesner, C. D., & Baving, L. (2015). Early gamma-band activity during
interference predicts working memory distractibility in ADHD. Journal of Attention
Disorders, 19(11), 971–976. http://doi.org/10.1177/1087054712459887
Pulvermüller, F., Birbaumer, N., Lutzenberger, W., & Mohr, B. (1997). High-frequency brain
activity: Its possible role in attention, perception and language processing. Progress in
Neurobiology, 52(5), 427–445.
Ray, S. & Maunsell, J. H. R. (2011). Different origins of gamma rhythm and high-gamma ac-
tivity in macaque visual cortex. PLoS Biology, 9(4), e1000610. http://doi.org/10.1371/journal.
pbio.1000610
174 DANIEL STRÜBER and CHRISTOPH S. HERRMANN
Ray, S. & Maunsell, J. H. R. (2015). Do gamma oscillations play a role in cerebral cortex? Trends
in Cognitive Sciences, 19(2), 78–85. http://doi.org/10.1016/j.tics.2014.12.002
Regan, D. (1968). A high frequency mechanism which underlies visual evoked potentials.
Electroencephalography and Clinical Neurophysiology, 25, 231–237.
Regan, D. & Spekreijse, H. (1986). Evoked potentials in vision research 1961–86. Vision
Research, 26(9), 1461–1480. http://doi.org/10.1016/B978-0-12-385157-4.00529-7
Reilly, T. J., Nottage, J. F., Studerus, E., Rutigliano, G., De Micheli, A. I., Fusar-Poli, P., &
McGuire, P. (2018). Gamma band oscillations in the early phase of psychosis: A systematic
review. Neuroscience & Biobehavioral Reviews, 90, 381–399. http://doi.org/10.1016/j.neubio
rev.2018.04.006
Reva, N. V. & Aftanas, L. I. (2004). The coincidence between late non- phase- locked
gamma synchronization response and saccadic eye movements. International Journal of
Psychophysiology, 51(3), 215–222. http://doi.org/10.1016/j.ijpsycho.2003.09.005
Rolfs, M., Kliegl, R., & Engbert, R. (2008). Toward a model of microsaccade generation: The
case of microsaccadic inhibition. Journal of Vision, 8(5), 1–23.
Rossi, S., Hallett, M., Rossini, P. M., Pascual-Leone, A., Avanzini, G., Bestmann, S., . . . Ziemann,
U. (2009). Safety, ethical considerations, and application guidelines for the use of transcranial
magnetic stimulation in clinical practice and research. Clinical Neurophysiology, 120(12),
2008–2039. http://doi.org/10.1016/j.clinph.2009.08.016
Ruden, J. B., Dugan, L. L., Konradi, C. (2021). Parvalbumin interneuron vulnerability and
brain disorders. Neuropsychopharmacology, 46, 279– 287. https://doi.org/10.1038/s41
386-020-0778-9
Salari, N., Büchel, C., & Rose, M. (2014). Neurofeedback training of gamma band oscillations
improves perceptual processing. Experimental Brain Research, 232(10), 3353–3361. http://doi.
org/10.1007/s00221-014-4023-9
Salinas, E. & Sejnowski, T. J. (2001). Correlated neuronal activity and the flow of neural infor-
mation. Nature Reviews Neuroscience, 2, 539–550. http://doi.org/10.1038/35086012
Schadow, J., Lenz, D., Thaerig, S., Busch, N. A., Fründ, I., Rieger, J. W., & Herrmann, C. S.
(2007). Stimulus intensity affects early sensory processing: Visual contrast modulates
evoked gamma-band activity in human EEG. International Journal of Psychophysiology,
66(1), 28–36. http://doi.org/10.1016/j.ijpsycho.2007.05.010
Schwartzman, D. J. & Kranczioch, C. (2011). In the blink of an eye: The contribution of
microsaccadic activity to the induced gamma band response. International Journal of
Psychophysiology, 79(1), 73–82. http://doi.org/10.1016/j.ijpsycho.2010.10.006
Sejnowski, T. J. & Paulsen, O. (2006). Network oscillations: Emerging computational
principles. Journal of Neuroscience, 26(6), 1673–1676. http://doi.org/10.1523/JNEURO
SCI.3737-05d.2006
Senkowski, D. & Gallinat, J. (2015). Dysfunctional prefrontal gamma-band oscillations reflect
working memory and other cognitive deficits in schizophrenia. Biological Psychiatry, 77(12),
1010–1019.
Sheer, D. E. (1984). Focused arousal, 40-Hz EEG, and dysfunction. In T. Elbert, B. Rockstroh,
W. Lutzenberger, & N. Birbaumer (Eds.), Self-regulation of the brain and behavior (pp. 64–
84). Springer.
Siegel, M., Donner, T. H., & Engel, A. K. (2012). Spectral fingerprints of large-scale neuronal
interactions. Nature Reviews Neuroscience, 13(2), 121–134. http://doi.org/10.1038/nrn3137
Singer, W. (2018). Neuronal oscillations: Unavoidable and useful? European Journal of
Neuroscience, 48(7), 2389–2398. http://doi.org/10.1111/ejn.13796
GAMMA ACTIVITY IN SENSORY AND COGNITIVE PROCESSING 175
Singer, W. & Gray, C. M. (1995). Visual feature integration and the temporal correlation hy-
pothesis. Annual Review of Neuroscience, 18, 555–586. http://doi.org/10.1146/annurev.
ne.18.030195.003011
Spencer, K. M. (2012). Baseline gamma power during auditory steady- state stimula-
tion in schizophrenia. Frontiers in Human Neuroscience, 5, 190. http://doi.org/10.3389/
fnhum.2011.00190
Spencer, K. M., Niznikiewicz, M. A., Shenton, M. E., & McCarley, R. W. (2008). Sensory-
evoked gamma oscillations in chronic schizophrenia. Biological Psychiatry, 63(8), 744–747.
Spydell, J. D., Ford, J. D., & Sheer, D. E. (1979). Task dependent cerebral lateralization of the 40
Hertz EEG rhythm. Psychophysiology, 16(4), 347–350.
Stroganova, T. A., Butorina, A. V., Sysoeva, O. V., Prokofyev, A. O., Nikolaeva, A. Y., Tsetlin, M.
M., & Orekhova, E. V. (2015). Altered modulation of gamma oscillation frequency by speed
of visual motion in children with autism spectrum disorders. Journal of Neurodevelopmental
Disorders, 7(1), 1–17. http://doi.org/10.1186/s11689-015-9121-x
Strüber, D., Başar-Eroǧlu, C., Hoff, E., & Stadler, M. A. (2000). Reversal-rate dependent
differences in the EEG gamma-band during multistable visual perception. International
Journal of Psychophysiology, 38(3), 243–252.
Strüber, D., Başar-Eroǧlu, C., Miener, M., & Stadler, M. A. (2001). EEG gamma-band response
during the perception of Necker cube reversals. Visual Cognition, 8(3), 609–621. http://doi.
org/10.1080/13506280143000151
Strüber, D. & Herrmann, C. S. (2020). Modulation of gamma oscillations as a possible thera-
peutic tool for neuropsychiatric diseases: A review and perspective. International Journal of
Psychophysiology, 152, 15–25. https://doi.org/10.1016/j.ijpsycho.2020.03.003
Summerfield, C. & Mangels, J. A. (2006). Dissociable neural mechanisms for encoding pre-
dictable and unpredictable events. Journal of Cognitive Neuroscience, 18(7), 1120–1132.
Swettenham, J. B., Muthukumaraswamy, S. D., & Singh, K. D. (2009). Spectral properties of
induced and evoked gamma oscillations in human early visual cortex to moving and sta-
tionary stimuli. Journal of Neurophysiology, 102(2), 1241–1253. http://doi.org/10.1152/
jn.91044.2008
Tal, N. & Yuval-Greenberg, S. (2018). Reducing saccadic artifacts and confounds in brain im-
aging studies through experimental design. Psychophysiology, 55(11), 1–20. http://doi.org/
10.1111/psyp.13215
Tallon-Baudry, C. (2009). The roles of gamma-band oscillatory synchrony in human visual
cognition. Frontiers in Bioscience, 14, 321–332.
Tallon-Baudry, C. & Bertrand, O. (1999). Oscillatory gamma activity in humans and its role in
object representation. Trends in Cognitive Sciences, 3(4), 151–162.
Tallon-Baudry, C., Bertrand, O., Delpuech, C., & Pernier, J. (1997). Oscillatory γ-band (30–70
Hz) activity induced by a visual search task in humans. The Journal of Neuroscience, 17(2),
722–734. https://doi.org/10.1523/JNEUROSCI.17-02-00722.1997
Tallon-Baudry, C., Bertrand, O., Delpuech, C., & Pernier, J. (1996). Stimulus specificity of
phase-locked and non-phase-locked 40 Hz visual responses in human. The Journal of
Neuroscience, 16(13), 4240–4249. http://doi.org/10.1016/j.neuropsychologia.2011.02.038
Tallon-Baudry, C., Bertrand, O., Peronnet, F., & Pernier, J. (1998). Induced gamma-band ac-
tivity during the delay of a visual short-term memory task in humans. The Journal of
Neuroscience, 18(11), 4244–4254. http://doi.org/20026318
Thickbroom, G. W. & Mastaglia, L. (1985). Presaccadic “spike” potential: Investigation of top-
ography and source. Brain Research, 339, 271–280.
176 DANIEL STRÜBER and CHRISTOPH S. HERRMANN
Tombor, L., Kakuszi, B., Papp, S., Réthelyi, J., Bitter, I., & Czobor, P. (2019). Decreased resting
gamma activity in adult attention deficit/hyperactivity disorder. The World Journal of
Biological Psychiatry, 20(9), 691-702. http://doi.org/10.1080/15622975.2018.1441547
Uhlhaas, P. J. & Singer, W. (2006). Neural synchrony in brain disorders: Relevance for cogni-
tive dysfunctions and pathophysiology. Neuron, 52(1), 155–168. http://doi.org/10.1016/j.neu
ron.2006.09.020
Uhlhaas, P. J. & Singer, W. (2010). Abnormal neural oscillations and synchrony in schizo-
phrenia. Nature Reviews Neuroscience, 11(2), 100–113. http://doi.org/10.1038/nrn2774
Uhlhaas, P. J. & Singer, W. (2012). Neuronal dynamics and neuropsychiatric disorders: Toward
a translational paradigm for dysfunctional large-scale networks. Neuron, 75(6), 963–980.
http://doi.org/10.1016/j.neuron.2012.09.004
Valsecchi, M., Betta, E., & Turatto, M. (2007). Visual oddballs induce prolonged microsaccadic
inhibition. Experimental Brain Research, 177(2), 196–208. http://doi.org/10.1007/s00
221-006-0665-6
Valsecchi, M., Dimigen, O., Kliegl, R., Sommer, W., & Turatto, M. (2009). Microsaccadic in-
hibition and P300 enhancement in a visual oddball task. Psychophysiology, 46(3), 635–644.
http://doi.org/10.1111/j.1469-8986.2009.00791.x
Van Diessen, E., Senders, J., Jansen, F. E., Boersma, M., & Bruining, H. (2015). Increased power
of resting-state gamma oscillations in autism spectrum disorder detected by routine elec-
troencephalography. European Archives of Psychiatry and Clinical Neuroscience, 265(6), 537–
540. http://doi.org/10.1007/s00406-014-0527-3
Van Kerkoerle, T., Self, M. W., Dagnino, B., Gariel-Mathis, M.-A., Poort, J., van der Togt, C.,
& Roelfsema, P. R. (2014). Alpha and gamma oscillations characterize feedback and feed-
forward processing in monkey visual cortex. Proceedings of the National Academy of Sciences
of the United States of America, 111(40), 14332–14341.
Varela, F., Lachaux, J. P., Rodriguez, E., & Martinerie, J. (2001). The brainweb: Phase synchron-
ization and large-scale integration. Nature Reviews Neuroscience, 2(4), 229–239. http://doi.
org/10.1038/35067550
Vinck, M. & Bosman, C. A. (2016). More gamma more predictions: Gamma-synchronization
as a key mechanism for efficient integration of classical receptive field inputs with surround
predictions. Frontiers in Systems Neuroscience, 10, 35. http://doi.org/10.3389/fnsys.2016.00035
Von Stein, A., Chiang, C., & König, P. (2000). Top-down processing mediated by interareal
synchronization. Proceedings of the National Academy of Sciences of the United States of
America, 97(26), 14748–14753. http://doi.org/10.1073/pnas.97.26.14748
Von Stein, A. & Sarnthein, J. (2000). Different frequencies for different scales of cortical inte-
gration: From local gamma to long range alpha/theta synchronization. International Journal
of Psychophysiology, 38(3), 301–313. http://doi.org/10.1016/S0167-8760(0 0)00172-0
Wang, J., Barstein, J., Ethridge, L. E., Mosconi, M. W., Takarae, Y., & Sweeney, J. A. (2013). Resting
state EEG abnormalities in autism spectrum disorders. Journal of Neurodevelopmental
Disorders, 5(1), 24. http://doi.org/10.1186/1866-1955-5-24
Wang, X.-J. (2010). Neurophysiological and computational principles of cortical rhythms in
cognition. Physiological Reviews, 90(3), 1195–1268. http://doi.org/10.1152/physrev.00035.2008
White, R. S. & Siegel, S. J. (2016). Cellular and circuit models of increased resting-state network
gamma activity in schizophrenia. Neuroscience, 321, 66–76.
Whitham, E. M., Lewis, T., Pope, K. J., Fitzgibbon, S. P., Clark, C. R., Loveless, S., . . . Willoughby,
J. O. (2008). Thinking activates EMG in scalp electrical recordings. Clinical Neurophysiology,
119(5), 1166–1175. http://doi.org/10.1016/j.clinph.2008.01.024
GAMMA ACTIVITY IN SENSORY AND COGNITIVE PROCESSING 177
Whitham, E. M., Pope, K. J., Fitzgibbon, S. P., Lewis, T., Clark, C. R., Loveless, S., . . . Willoughby,
J. O. (2007). Scalp electrical recording during paralysis: Quantitative evidence that EEG
frequencies above 20 Hz are contaminated by EMG. Clinical Neurophysiology, 118(8), 1877–
1888. http://doi.org/10.1016/j.clinph.2007.04.027
Whittington, M. A., Cunningham, M. O., LeBeau, F. E. N., Racca, C., & Traub, R. D. (2011).
Multiple origins of the cortical gamma rhythm. Developmental Neurobiology, 71(1), 92–106.
http://doi.org/10.1002/dneu.20814
Whittington, M. A., Traub, R. D., Kopell, N., Ermentrout, B., & Buhl, E. H. (2000). Inhibition-
based rhythms: Experimental and mathematical observations on network dynamics.
International Journal of Psychophysiology, 38(3), 315–336.
Yordanova, J., Banaschewski, T., Kolev, V., Woerner, W., & Rothenberger, A. (2001). Abnormal
early stages of task stimulus processing in children with attention-deficit hyperactivity dis-
order: Evidence from event-related gamma oscillations. Clinical Neurophysiology, 112(6),
1096–1108. http://doi.org/10.1016/S1388-2457(01)00524-7
Yuval-Greenberg, S. & Deouell, L. Y. (2009). The broadband-transient induced gamma-band
response in scalp EEG reflects the execution of saccades. Brain Topography, 22(1), 3–6. http://
doi.org/10.1007/s10548-009-0077-6
Yuval-Greenberg, S., Tomer, O., Keren, A. S., Nelken, I., & Deouell, L. Y. (2008). Transient
induced gamma-band response in EEG as a manifestation of miniature saccades. Neuron,
58(3), 429–441. http://doi.org/10.1016/j.neuron.2008.03.027
Zaehle, T., Fründ, I., Schadow, J., Thärig, S., Schoenfeld, M. A., & Herrmann, C. S. (2009).
Inter-and intra-individual covariations of hemodynamic and oscillatory gamma responses
in the human cortex. Frontiers in Human Neuroscience, 3, 8. http://doi.org/10.3389/
neuro.09.008.2009
CHAPTER 9
9.1 History
episodic memory encoding and retrieval (Jacobs et al., 2006; Kahana et al., 1999;
Klimesch, 1999; Klimesch et al., 2000; Nyhus & Curran, 2010; Rizzuto et al., 2006;
Sauseng et al., 2004). These findings indicated that theta is generated across human neo-
cortical areas (Caplan et al., 2003; Jacobs et al., 2006; Raghavachari et al., 2006) and
that it reflects a multitude of active cortical processes. Here we reiterate our caveat for
determining the relationship between a brain rhythm and a specific cognitive pro-
cess: rhythmic processes reflect common neural operations that have distinct repre-
sentational content depending on the generative neural system. Theta-band dynamics
reflect a non-specific marker of active cortical operation. The frontal midline variant of
theta is particularly prevalent in the human EEG, making it a good example for under-
standing broader cortical theta activities.
Experimental findings have accumulated an increasingly well- defined set of
processes associated with FMT. Talairach and colleagues (1973) described how electrical
stimulation of the human anterior cingulate cortex (ACC) elicited motor actions that
were integrated with environmental context, sometimes accompanied by FMT oscilla-
tory activities. Throughout the 1990s, the qualitative appearance of scalp-recorded FMT
during cognitive effort could be reliably evoked (Asada et al., 1999; Inouye et al., 1994;
Inouye, Shinosaki, Iyama, Matsumoto, Toi, et al., 1994) and a corresponding relation-
ship with anxiety was often observed (Mizuki et al., 1997; Mizuki et al., 1992; Mizuki
et al., 1996). These two effective and affective facets of vigilance have been increas-
ingly associated with FMT throughout the past few decades. Paralleling the evolving
capabilities of experimental neurophysiology, a growing area of research has moved be-
yond qualitative assessment towards a detailed quantification of the generators, elicitors,
and moderators involved in the genesis of the FMT rhythm.
9.2 Characteristics
Mental effort is a reliable elicitor of FMT (Smit et al., 2004; Smit et al., 2005). FMT
increases during perseverance and decreases during fatigue (Wascher et al., 2014).
Working memory load scales with FMT power (Gevins & Smith, 2000; Ishii et al., 1999;
Itthipuripat et al., 2013; Onton et al., 2005; Sauseng et al., 2010), although it should be
examined if this relationship is simply due to increased effort or if it represents specific
information content (see Hsieh et al., 2011; Roberts et al., 2013). Memory encoding and
retrieval are associated with broad cortical theta, including FMT (Hsieh & Ranganath,
2014; Jacobs et al., 2006; Kahana et al., 1999; Klimesch, 1999; Klimesch et al., 2000; Nyhus
& Curran, 2010; Rizzuto et al., 2003, 2006; Sauseng et al., 2004). It remains unknown if
the role of FMT is specific to control processes in memory rather than encoding per se
(see Hanslmayr et al., 2010; Staudigl et al., 2010), particularly since other cortical theta is
specifically associated with encoding (see Rizzuto et al., 2006; Wang et al., 2018).
FMT has been localized to broad medial frontal cortical areas, including the ACC and
the midcingulate cortex (MCC) using MEG (Beaton et al., 2018; Ishii et al., 1999; Jensen
180 JAMES F. CAVANAGH and MICHAEL X COHEN
& Tesche, 2002) and EEG (Cohen & Ridderinkhof, 2013; Gevins & Smith, 2000; Gevins
et al., 1999; Gevins et al., 1997; Onton et al., 2005). The MCC generates oscillations in the
theta band in human intracranial recording (Cohen et al., 2008; Wang et al., 2005) as well
as in non-human primates (Tsujimoto et al., 2010; Tsujimoto et al., 2006; Womelsdorf,
Johnston, et al., 2010; Womelsdorf, Vinck, et al., 2010). Recent reviews have summarized
distinct functional aspects of FMT, including its modulators (Mitchell, McNaughton,
Flanagan, & Kirk, 2008), its role in memory (Hsieh & Ranganath, 2014), and its broader
role in cognitive control (Cavanagh & Frank, 2014). We turn now to this defining area
of cognitive control and the history of linking FMT processes to frontal midline ERP
components intricately related to the need for control.
Stimulus-Locked Response-Locked
Novelty Conflict Punishment Conflict Errors
(a)
Event-Related
Potentital
μV
26 26 26 26 26
15 15 15 15 15
Hz 8
Power
8 8 8 8
5 5 5 5 5
3 3 3 3 3 +/–4
1 +/–2 1 +/–1 1 +/–2 1 +/–1 1
–250 0 250 500 750 –250 0 250 500 750 –250 0 250 500 750 –500 –250 0 250 750 –500 –250 0 250 750
ms peri-stimulus ms peri-stimulus ms peri-stimulus ms peri-stimulus ms peri-stimulus
(c)
Theta Increase
Topography of
Figure 9.1 The need for cognitive control is associated with a similar frontal midline theta
signature across a variety of eliciting events. (A) Phase-locked EEG activities (ERPs). While
these ERP components (i.e., peaks and troughs in the signal locked to particular external events
and averaged across trials) are related to learning and adaptive control, they represent a small
fraction of ongoing neural dynamics. (B) Time-frequency plots show richer spectral dynamics of
event-related neuro-electrical activity by averaging activities regardless of phase-locking. Here,
significant increases in power to novelty, conflict, punishment, and error are outlined in black,
revealing a common frontal midline theta band feature during events that signal a need for con-
trol. (C) Scalp topography of event-related frontal midline theta activity. The distribution of theta
power bursts is consistently maximal over the frontal midline.
N2, a component elicited by novelty or stimulus/response conflict; Feedback related negativity
(FRN), A similar N2-like component elicited by external feedback signaling that one’s actions were
incorrect or yielded a loss; Correct-related negativity (CRN), a small, obligatory component evoked
by motor responses even when these are correct according to the task and enhanced by response
conflict; Error related negativity (ERN), A component evoked by motor commission errors.
have a spectral representation in the theta band (Başar-Eroğlu et al., 1992; Başar, 1998a;
Yordanova et al., 2002). The ERN and the FRN/N2 are estimated to have MCC sources
via EEG source estimation (Gehring et al., 2012; Gruendler et al., 2011; van Noordt &
Segalowitz, 2012; Walsh & Anderson, 2012) and EEG-informed fMRI (Becker et al.,
2014; Debener et al., 2005; Edwards et al., 2012; Hauser et al., 2014; Huster et al., 2011),
suggesting at least some common processes linking the two.
A parsimonious summary could propose that both the stimulus-and response-
related fronto-central negativities reflect common features of the processing demands
of the MCC. These features are varied across systems related to cognitive and motor
control, attention, and reinforcement learning, but are especially sensitive to mismatch
signals of conflict, punishment, and error in the service of behavioral adaptation. The
commonality of these theta-band processes led to an integrative theory of a common
182 JAMES F. CAVANAGH and MICHAEL X COHEN
language, a theta lingua franca, for the realization of the need for control (Cavanagh
et al., 2009; 2012). Stimulus-and response-locked obligatory theta-band phase dy-
namics were proposed to represent a biophysical mechanism for the common tem-
poral organization of neural processes during stimulus or response processes. Variation
on this theme, such as power enhancement, reflects the realization of these reactive
responses (Figure 9.1). These computations appear to be used to merge attentive, af-
fective, and cognitive functions with motor selection in order to utilize environmental
context during action monitoring. FMT therefore appears to reflect general operations
of the MCC during action monitoring, particularly as an initial orienting response to a
novel or surprising event (Wessel, 2018).
did not affect the conflict-related FMT. This is consistent with other reports that have
reported a qualitative absence of phase-locking associated with response-conflict FMT
(Nigbur et al., 2012; Pastötter et al., 2010). These findings can be contrasted with similar
analyses of the FMT response to erroneous button presses, which contains theta-band
phase-locking (Trujillo & Allen, 2007) that is significantly affected by removing the ERP
(Munneke et al., 2015).
In some fields of neuroscience research, the term “theta” implicitly implies rodent
hippocampal activity (~4–12 Hz) that has been associated with learning, memory, and
spatial navigation (Buzsáki, 2006). Rodent hippocampal theta is not a unitary construct
(Colgin, 2013; Pignatelli et al., 2012), with separate theta rhythms occurring due to septal
drive as well as an intrinsic generative process that seems to be common to many types
of cortical and sub-cortical excitatory-inhibitory networks (Womelsdorf et al., 2014).
Mediofrontal spikes are phase- locked to both mediofrontal theta as well as
hippocampal theta (Benchenane et al., 2010; Hyman et al., 2011; Jones & Wilson, 2005b;
2005a; Paz et al., 2008; Pignatelli et al., 2012; Siapas et al., 2005). However, this evidence
of hippocampal interaction with frontal theta processes could reflect a generic phenom-
enon whereby cortical theta synchronizes disparate neural areas in a global workspace,
possibly via travelling waves (Lubenov & Siapas, 2009; Zhang et al., 2018). Many cortical
areas have shown phase-synchronous relationships with frontal theta, including visual
cortex (Lee et al., 2005; Liebe et al., 2012; Phillips et al., 2013), amygdala (Taub et al.,
2018), and ventral tegmental areas (Fujisawa & Buzsáki, 2011). The ubiquity of theta-
band findings across species has led to the suggestion that FMT reflects a non-specific
mechanism for organizing neural processes around “decision points”, such as action se-
lection (Womelsdorf, Vinck, et al., 2010).
The translational potential for using theta to infer similar cognitive processes between
species is promising but needs additional clarification. Some studies show that non-
human primates have similar error, conflict, and feedback ERPs at the skull (Phillips &
Everling, 2014), the scalp (Godlove et al., 2011), the dura (Vezoli & Procyk, 2009), and
within the cingulate cortex (Emeric et al., 2010), although the spectral representation
of these signals has not been defined. Rats have a FMT-dominant control network that
is transiently instantiated following an imperative tone, affording a chance to causally
manipulate this network and draw parallel conclusions to humans (Narayanan et al.,
2013) although this is non-specifically spectrally localized to the theta band compared
to typical human EEG findings. This common FMT electrophysiological response is
diminished in Parkinson’s patients as well as in a dopamine depletion rodent model
(Parker et al., 2015), suggesting a novel model of cognitive dysfunction in Parkinsonism.
Ample evidence suggests that FMT is sensitive to dopamine in humans, but it appears
to also be sensitive to other monoamines like norepinephrine and acetylcholine (see
184 JAMES F. CAVANAGH and MICHAEL X COHEN
(a)
Motor
IPFC
MCC
Sensory
BG
WM
Conflict
Error
(b) e e
Cu ons back
IPFC sp ed
Re Fe
age schizophrenia
tDCS age
in-phase tACS out-phase tACS
anxiety mild TBI
(c)
IPFC: L>R IPFC: R>L
Figure 9.2 Theta band phase consistency between mid-frontal and distal sites is transiently
increased following events that indicate a need for control. (A) Twenty five separate studies (A
through Y) have replicated the finding of theta-band phase synchrony between the frontal mid-
line (presumably MCC) and varied cortical areas, including lateral prefrontal cortex (lPFC),
motor cortex, sensory cortices, and basal ganglia (BG). WM =working memory. (B) A var-
iety of moderators affect medio-lateral phase synchrony, with increases due to age, anxiety and
transcranial direct or alternating current stimulation (tD/ACS), and decreases due to schizo-
phrenia, age, anti-phase tACS, and mild traumatic brain injury (TBI). (C) Studies often-
times report a nominal pattern of hemisphericity in medio-lateral phase synchrony. Errors
and punishments consistently evoked a right>left pattern, whereas conflict tended towards a
left>right pattern. However, the studies contributing to the left>right pattern were more com-
plex and involved proactive and reactive processes compared to error realization, which is ra-
ther straightforward and reactive. Future studies should formally investigate the moderators of
hemisphericity in medio-lateral phase synchrony.
Citations for fi
gure 9.2:
A: Hanslmayr et al., 2008; B: Cavanagh et al., 2009; C: Cohen et al., 2009; D: Cavanagh et al., 2010; E: Cohen & Cavanagh,
2011; F: Cohen & van Gaal, 2013; G: Nigbur et al., 2012; H: van de Vijver et al., 2011; I: van Driel et al., 2012; J: Narayanan et
al., 2013; K: Anguera et al., 2013; L: Cohen & van Gaal, 2014; M: Tóth et al., 2014; N: Van de Vijver et al., 2014; O: Moran et
al., 2014; P: Reinhart et al., 2015; Q: Zavala et al., 2013; R: Cavanagh et al., 2017; S: Reinhart, 2017; T: Vissers et al., 2018: U:
Ryman et al., 2018; V: Swart et al., 2018; W: Buzzell et al., 2018; X: Oehrn et al., 2014; Y: Cavanagh et al., 2020.
FRONTAL MIDLINE THETA AS A MODEL SPECIMEN OF CORTICAL THETA 187
by out-of-phase alternating current (Reinhart, 2017). These networks are also altered in
psychiatric distress, being increased in anxiety (Cavanagh et al., 2017) and diminished
in schizophrenia (Ryman et al., 2018), see Figure 9.2b. Many questions remain to be
addressed about the function of this theta-band phase synchrony, but there is correlative
(Anguera et al., 2013; Cavanagh et al., 2009; Cavanagh et al., 2017; Swart et al., 2018) and
causal evidence (Narayanan et al., 2013; Reinhart, 2017; Reinhart et al., 2015) that this
network directly affects behaviors related to the ability to learn from and adapt to the
need for control.
Many studies have observed a right-sided dominance of medio-lateral phase syn-
chrony, but this issue of hemisphericity requires further rigorous testing. Figure 9.2c
sorts studies that reported even a nominal pattern of hemisphericity, and it can be
seen that errors and punishments consistently evoked a right>left bias, whereas con-
flict tended towards a left>right bias. Although this distinction appears straightfor-
ward, it is likely overly simplistic to suggest a simple conflict vs. error dissociation on
hemisphericity. The studies contributing to the left>right pattern were more complex
and involved both proactive and reactive processes whereas error realization is rather
straightforward and reactive. Future studies should be formally test for hemispheric
bias in medio-lateral phase synchrony to better address questions about the functional
role of this signal.
FMT and related ERP features have compelling characteristics for clinical advance-
ment. The majority of the units of analysis in the National Institute of Mental Health
(NIMH) Research Domain Criteria (Insel et al., 2010) are EEG-based, and many of
these are FMT-family responses. Lower FMT appears to be a reliable endophenotype
for substance abuse and externalizing disorders (Gilmore et al., 2010; Kamarajan et al.,
2015; Kang et al., 2012; Rangaswamy et al., 2007; Zlojutro et al., 2011). Higher FMT is
reliably associated with anxious temperament (Cavanagh & Shackman, 2014; Moser
et al., 2013; Riesel et al., 2017), and ERN amplitude can even predict treatment response
in anxiety disorder patients (Gorka et al., 2018) . Future studies should derive the sen-
sitivity and specificity of FMT to determine the biomarker potential in select clinical
applications.
FMT can be elicited in simple tasks that are viable within a clinical environment.
Aberrant auditory orienting responses have already been advanced as candidate
biomarkers, like diminished mismatch negativity (MMN) in schizophrenia (Javitt
et al., 2018; Light et al., 2015) or diminished novelty habituation in Parkinson’s disease
(Cavanagh, Kumar, et al., 2018). The MMN is a theta-dominant response with sep-
arable frontal and temporal processes (Fuentemilla et al., 2008; Ko et al., 2012) that
may interact via theta-band phase synchrony (Choi et al., 2013). The neural systems
underlying auditory novelty detection are well detailed in rodent models (Escera &
188 JAMES F. CAVANAGH and MICHAEL X COHEN
Malmierca, 2014; Featherstone et al., 2018; Lee et al., 2018), facilitating cross-species
translation. Auditory-evoked responses are already routinely used in brainstem audi-
tory testing for hearing acuity in newborns, demonstrating that clinical infrastructure
and expertise exists for applying relevant tasks to patient groups when using EEG as a
diagnostic tool.
References
Allain, S., Carbonnell, L., Falkenstein, M., Burle, B., & Vidal, F. (2004). The modulation of the
Ne-like wave on correct responses foreshadows errors. Neuroscience Letters, 372(1–2), 161–
166. https://doi.org/S0304-3940(04)01171-1 [pii]10.1016/j.neulet.2004.09.036
Amarante, L. M., Caetano, M. S., & Laubach, M. (2017). Medial frontal theta is entrained to
rewarded actions. The Journal of Neuroscience, 37(44), 1965–1917. https://doi.org/10.1523/
JNEUROSCI.1965-17.2017
Anguera, J. A., Boccanfuso, J., Rintoul, J. L., Al-Hashimi, O., Faraji, F., Janowich, J., . . . Gazzaley,
A. (2013). Video game training enhances cognitive control in older adults. Nature, 501(7465),
97–101. https://doi.org/10.1038/nature12486
Arellano, A. P. & Schwab, R. S. (1950). Scalp and basal recording during mental activity.
Proceedings of the 1st International Congress of Psychiatry, September 18–27, Paris.
Asada, H., Fukuda, Y., Tsunoda, S., Yamaguchi, M., & Tonoike, M. (1999). Frontal midline
theta rhythms reflect alternative activation of prefrontal cortex and anterior cingulate cortex
in humans. Neuroscience Letters, 274(1), 29–32. https://doi.org/S0304-3940(99)00679-5
Bartoli, E., Conner, C. R., Kadipasaoglu, C. M., Yellapantula, S., Rollo, M. J., Carter, C. S., &
Tandon, N. (2017). Temporal dynamics of human frontal and cingulate neural activity
during conflict and cognitive control. Cerebral Cortex, 28(11), 3842–3856. https://doi.org/
10.1093/cercor/bhx245
Başar-Eroğlu, C., Başar, E., Demiralp, T., & Schurmann, M. (1992). P300-response: Possible
psychophysiological correlates in delta and theta frequency channels. A review. International
Journal of Psychophysiology, 13, 161–179.
Başar, E. (1998a). Brain function and oscillations: Brain oscillations–principles and approaches.
Springer-Verlag. https://doi.org/10.1007/978-3-642-72192-2
FRONTAL MIDLINE THETA AS A MODEL SPECIMEN OF CORTICAL THETA 189
Cavanagh, J. F. & Frank, M. J. (2014). Frontal theta as a mechanism for cognitive control. Trends
in Cognitive Sciences, 18(8), 1–8. https://doi.org/10.1016/j.tics.2014.04.012
Cavanagh, J. F., Frank, M. J., Klein, T. J., & Allen, J. J. B. (2010). Frontal theta links prediction
errors to behavioral adaptation in reinforcement learning. NeuroImage, 49(4), 3198–3209.
https://doi.org/10.1016/j.neuroimage.2009.11.080
Cavanagh, J. F., Kumar, P., Mueller, A. A., Pirio Richardson, S., Mueen, A., & Richardson, S. P.
(2018). Diminished EEG habituation to novel events effectively classifies Parkinson’s patients.
Clinical Neurophysiology, 129(2), 409–418. https://doi.org/10.1016/j.clinph.2017.11.023
Cavanagh, J. F., Meyer, A., & Hajcak, G. (2017). Error-specific cognitive control alterations
in generalized anxiety disorder. Biological Psychiatry: Cognitive Neuroscience and
Neuroimaging, 2(5), 413–420. https://doi.org/10.1016/j.bpsc.2017.01.004
Cavanagh, J. F., Rieger, R. E., Wilson, J. K., Gill, D., Fullerton, L., Brandt, E., & Mayer, A. R.
(2020). Joint analysis of frontal theta synchrony and white matter following mild traumatic
brain injury. Brain Imaging and Behavior, 14(6), 2210–2223.
Cavanagh, J. F., & Shackman, A. J. (2014). Frontal midline theta reflects anxiety and cogni-
tive control: Meta-analytic evidence. Journal of Physiology, 109(1–3), 3–15. https://doi.org/
10.1016/j.jphysparis.2014.04.003
Cavanagh, J. F., Zambrano-Vazquez, L., & Allen, J. J. B. (2012). Theta lingua franca: A common
mid-frontal substrate for action monitoring processes. Psychophysiology, 49(2), 220–238.
https://doi.org/10.1111/j.1469-8986.2011.01293.x
Choi, J. W., Lee, J. K., Ko, D., Lee, G. T., Jung, K. Y., & Kim, K. H. (2013). Fronto-temporal
interactions in the theta-band during auditory deviant processing. Neuroscience Letters, 548,
120–125. https://doi.org/10.1016/j.neulet.2013.05.079
Cohen, M. X. (2014). A neural microcircuit for cognitive conflict detection and signaling.
Trends in Neurosciences, 37(9), 480–490. https://doi.org/10.1016/j.tins.2014.06.004
Cohen, M. X. (2016). Midfrontal theta tracks action monitoring over multiple interactive time
scales. NeuroImage, 141, 262–272. https://doi.org/10.1016/J.NEUROIMAGE.2016.07.054
Cohen, M. X. & Cavanagh, J. F. (2011). Single-trial regression elucidates the role of prefrontal
theta oscillations in response conflict. Frontiers in Psychology [online], 2, 30. https://doi.org/
10.3389/fpsyg.2011.00030
Cohen, M. X. & Donner, T. H. (2013). Midfrontal conflict-related theta-band power reflects
neural oscillations that predict behavior. Journal of Neurophysiology, 110(12), 2752–2763.
https://doi.org/10.1152/jn.00479.2013
Cohen, M. X. & Ridderinkhof, K. R. (2013). EEG source reconstruction reveals frontal-parietal
dynamics of spatial conflict processing. PLoS One, 8(2), e57293. https://doi.org/10.1371/jour
nal.pone.0057293
Cohen, M. X., Ridderinkhof, K. R., Haupt, S., Elger, C. E., & Fell, J. (2008). Medial frontal cortex
and response conflict: Evidence from human intracranial EEG and medial frontal cortex
lesion. Brain Research, 1238, 127–142. https://doi.org/S0006-8993(08)01872-6 [pii]10.1016/
j.brainres.2008.07.114
Cohen, M. X, & van Gaal, S. (2013). Dynamic interactions between large-scale brain networks
predict behavioral adaptation after perceptual errors. Cerebral Cortex (New York, N.Y.: 1991),
23(5), 1061–1072. https://doi.org/10.1093/cercor/bhs069
Cohen, M. X, & van Gaal, S. (2014). Subthreshold muscle twitches dissociate oscillatory neural
signatures of conflicts from errors. NeuroImage, 86, 503–513. https://doi.org/10.1016/J.NEU
ROIMAGE.2013.10.033
FRONTAL MIDLINE THETA AS A MODEL SPECIMEN OF CORTICAL THETA 191
Cohen, M. X, van Gaal, S., Ridderinkhof, K. R., & Lamme, V. A. F. (2009). Unconscious errors
enhance prefrontal-occipital oscillatory synchrony. Frontiers in Human Neuroscience, 3, 54.
https://doi.org/10.3389/neuro.09.054.2009
Coles, M. G., Scheffers, M. K., & Holroyd, C. B. (2001). Why is there an ERN/Ne on correct trials?
Response representations, stimulus-related components, and the theory of error-processing.
Biological Psychology, 56(3), 173–189. https://doi.org/10.1016/S0301-0511(01)00076-X
Colgin, L. L. (2013). Mechanisms and functions of theta rhythms. Annual Review of
Neuroscience, 36(1), 295–312. https://doi.org/10.1146/annurev-neuro-062012-170330
Debener, S., Ullsperger, M., Siegel, M., Fiehler, K., von Cramon, D. Y., & Engel, A. K. (2005).
Trial-by-trial coupling of concurrent electroencephalogram and functional magnetic reson-
ance imaging identifies the dynamics of performance monitoring. Journal of Neuroscience,
25(50), 11730–11737. https://doi.org/25/50/11730 [pii]10.1523/JNEUROSCI.3286-05.2005
Dehaene, S., Posner, M. I., & Tucker, D. M. (1994). Localization of a neural system for error de-
tection and compensation. Psychological Science, 5(5),303–305.
Edwards, B. G., Calhoun, V. D., & Kiehl, K. A. (2012). Joint ICA of ERP and fMRI during
error-monitoring. NeuroImage, 59(2), 1896–1903. https://doi.org/10.1016/j.neuroim
age.2011.08.088
Emeric, E. E., Leslie, M. W., Pouget, P., & Schall, J. D. (2010). Performance monitoring local
field potentials in the medial frontal cortex of primates: Supplementary eye field. Journal of
Neurophysiology, 104, 1523–1537. https://doi.org/jn.01001.2009
Escera, C. & Malmierca, M. S. (2014). The auditory novelty system: An attempt to integrate
human and animal research. Psychophysiology, 51(2), 111–123. https://doi.org/10.1111/
psyp.12156
Falkenstein, M., Hohnsbein, J., Hoormann, J., & Blanke, L. (1991). Effects of crossmodal
divided attention on late ERP components. II. Error processing in choice reaction tasks.
Electroencephalography and Clinical Neurophysiology, 78(6), 447–455. https://pubmed.ncbi.
nlm.nih.gov/1712280/
Featherstone, R. E., Melnychenko, O., & Siegel, S. J. (2018). Mismatch negativity in preclinical
models of schizophrenia. Schizophrenia Research, 191, 35–42. https://doi.org/10.1016/j.sch
res.2017.07.039
Folstein, J. R. & Van Petten, C. (2008). Influence of cognitive control and mismatch on the N2
component of the ERP: a review. Psychophysiology, 45(1), 152–170. https://doi.org/PSYP602
[pii]10.1111/j.1469-8986.2007.00602.x
Fries, P. (2005). A mechanism for cognitive dynamics: neuronal communication through
neuronal coherence. Trends in Cognitive Science, 9(10), 474–480. https://doi.org/S1364-
6613(05)00242-1 [pii]10.1016/j.tics.2005.08.011
Fries, P. (2015). Rhythms for cognition: Communication through coherence. Neuron, 88(1),
220–235. https://doi.org/10.1016/j.neuron.2015.09.034
Fuentemilla, L., Marco-Pallarés, J., Münte, T. F., & Grau, C. (2008). Theta EEG oscillatory ac-
tivity and auditory change detection. Brain Research, 1220, 93–101. https://doi.org/10.1016/
j.brainres.2007.07.079
Fujisawa, S. & Buzsáki, G. (2011). A 4 Hz oscillation adaptively synchronizes prefrontal,
VTA, and hippocampal activities. Neuron, 72(1), 153–165. https://doi.org/10.1016/j.neu
ron.2011.08.018
Fukai, T. (1999). Sequence generation in arbitrary temporal patterns from theta-nested gamma
oscillations: A model of the basal ganglia-thalamo-cortical loops. Neural Networks, 12(7–8),
975–987. https://pubmed.ncbi.nlm.nih.gov/12662640/
192 JAMES F. CAVANAGH and MICHAEL X COHEN
Gehring, W. J., Goss, B., Coles, M. G. H., Meyer, D. E., & Donchin, E. (1993). A neural system
for error-detection and compensation. Psychological Science, 4(6), 385–390.
Gehring, W. J., Himle, J., & Nisenson, L. G. (2000). Action-monitoring dysfunction in
obsessive-compulsive disorder. Psychological Science, 11(1), 1–6. https://pubmed.ncbi.nlm.
nih.gov/11228836/
Gehring, W. J., Liu, Y., Orr, J. M., & Carp, J. (2012). The error-related negativity (ERN/Ne). In S.
J. Luck & E. Kappenman (Eds.), Oxford handbook of event-related potential components (pp.
231–291). Oxford University Press.
Gehring, W. J. & Willoughby, A. R. (2002). The medial frontal cortex and the rapid processing
of monetary gains and losses. Science, 295(5563), 2279–2282. https://doi.org/10.1126/scie
nce.1066893295/5563/2279
Gemba, H., Sasaki, K., & Brooks, V. B. (1986). “Error” potentials in limbic cortex (anterior
cingulate area 24) of monkeys during motor learning. Neuroscience Letters, 70(2), 223–227.
https://doi.org/0304-3940(86)90467-2
Gevins, A. & Smith, M. E. (2000). Neurophysiological measures of working memory and in-
dividual differences in cognitive ability and cognitive style. Cerebral Cortex, 10(9), 829–839.
https://pubmed.ncbi.nlm.nih.gov/10982744/
Gevins, A., Smith, M. E., McEvoy, L. K., Leong, H., & Le, J. (1999). Electroencephalographic
imaging of higher brain function. Philosophical Transactions of the Royal Society of London
Series B: Biological Sciences, 354(1387), 1125–1133. https://doi.org/10.1098/rstb.1999.0468
Gevins, A., Smith, M. E., McEvoy, L., & Yu, D. (1997). High-resolution EEG mapping of cor-
tical activation related to working memory: Effects of task difficulty, type of processing, and
practice. Cerebral Cortex, 7(4), 374–385. https://pubmed.ncbi.nlm.nih.gov/9177767/
Gilmore, C. S., Malone, S. M., & Iacono, W. G. (2010). Brain electrophysiological
endophenotypes for externalizing psychopathology: A multivariate approach. Behavior
Genetics, 40(2), 186–200. https://doi.org/10.1007/s10519-010-9343-3
Giraud, A.-L. L., Kleinschmidt, A., Poeppel, D., Lund, T. E., Frackowiak, R. S. J. J., & Laufs,
H. (2007). Endogenous cortical rhythms determine cerebral specialization for speech
perception and production. Neuron, 56(6), 1127–1134. https://doi.org/10.1016/j.neu
ron.2007.09.038
Giraud, A.-L. L., & Poeppel, D. (2012). Cortical oscillations and speech processing: emerging
computational principles and operations. Nature Neuroscience, 15(4), 511–517. https://doi.
org/10.1038/nn.3063
Godlove, D. C., Emeric, E. E., Segovis, C. M., Young, M. S., Schall, J. D., & Woodman, G. F.
(2011). Event-related potentials elicited by errors during the stop-signal task. I. Macaque
monkeys. The Journal of Neuroscience, 31(44), 15640–15649. https://doi.org/10.1523/JNEURO
SCI.3349-11.2011
Gorka, S. M., Burkhouse, K. L., Klumpp, H., Kennedy, A. E., Afshar, K., Francis, J., . . . Phan,
K. L. (2018). Error-related brain activity as a treatment moderator and index of symptom
change during cognitive-behavioral therapy or selective serotonin reuptake inhibitors.
Neuropsychopharmacology, 43(6), 1355–1363. https://doi.org/10.1038/npp.2017.289
Gratton, G. (2018). Brain reflections: A circuit-based framework for understanding informa-
tion processing and cognitive control. Psychophysiology, 55(3), 1–26. https://doi.org/10.1111/
psyp.13038
Gruber, W. R., Klimesch, W., Sauseng, P., & Doppelmayr, M. (2005). Alpha phase synchroniza-
tion predicts P1 and N1 latency and amplitude size. Cerebral Cortex, 15(4), 371–377. https://
doi.org/10.1093/cercor/bhh139
FRONTAL MIDLINE THETA AS A MODEL SPECIMEN OF CORTICAL THETA 193
Gruendler, T. O. J., Ullsperger, M., & Huster, R. J. (2011). Event-related potential correlates
of performance-monitoring in a lateralized time-estimation task. PloS One, 6(10), e25591.
https://doi.org/10.1371/journal.pone.0025591
Hajcak, G., Moser, J. S., Yeung, N., & Simons, R. F. (2005). On the ERN and the significance of
errors. Psychophysiology, 42(2), 151–160.
Hanslmayr, S., Pastotter, B., Bauml, K. H., Gruber, S., Wimber, M., & Klimesch, W. (2008). The
electrophysiological dynamics of interference during the Stroop task. Journal of Cognitive
Neuroscience, 20(2), 215–225. https://doi.org/10.1162/jocn.2008.20020
Hanslmayr, S., Staudigl, T., Aslan, A., & Bauml, K. H. (2010). Theta oscillations predict the
detrimental effects of memory retrieval. Cognitive, Affective, and Behavioral Neuroscience,
10(3), 329–338. doi: 10.3758/cabn.10.3.329
Hauser, T. U., Iannaccone, R., Stämpfli, P., Drechsler, R., Brandeis, D., Walitza, S., & Brem, S.
(2014). The feedback-related negativity (FRN) revisited: New insights into the localization,
meaning and network organization. NeuroImage, 84, 159–168. https://doi.org/10.1016/j.neu
roimage.2013.08.028
Helfrich, R. F., Fiebelkorn, I. C., Szczepanski, S. M., Lin, J. J., Parvizi, J., Knight, R. T., & Kastner,
S. (2018). Neural mechanisms of sustained attention are rhythmic. Neuron, 99(4), 854–865.
e5. https://doi.org/10.1016/j.neuron.2018.07.032
Holroyd, C. B. (2002). A note on the oddball N200 and the feedback ERN. In M. Ullsperger
& M. Falkenstein (Eds.), Errors, conflicts and the brain. Current opinions on performance
monitoring (pp. 211–218). Max Planck Institute for Human Cognitive and Brain Sciences.
Holroyd, C. B. & Coles, M. G. (2002). The neural basis of human error processing: reinforce-
ment learning, dopamine, and the error-related negativity. Psychological Review, 109(4),
679–709. https://pubmed.ncbi.nlm.nih.gov/12374324/
Holroyd, C. B., Coles, M. G. H., Nieuwenhuis, S., Gehring, W. J., & Willoughby, A. R. (2002).
Medial prefrontal cortex and error potentials. Science, 296(5573), 1610–1611.
Hsieh, L.-T., Ekstrom, A. D., & Ranganath, C. (2011). Neural oscillations associated with item
and temporal order maintenance in working memory. The Journal of Neuroscience, 31(30),
10803–10810. https://doi.org/10.1523/JNEUROSCI.0828-11.2011
Hsieh, L. T. & Ranganath, C. (2014). Frontal midline theta oscillations during working memory
maintenance and episodic encoding and retrieval. NeuroImage, 85, 721–729. https://doi.org/
10.1016/j.neuroimage.2013.08.003
Huang, Y., Chen, L., & Luo, H. (2015). Behavioral oscillation in priming: Competing percep-
tual predictions conveyed in alternating theta-band rhythms. Journal of Neuroscience, 35(6),
2830–2837. https://doi.org/10.1523/JNEUROSCI.4294-14.2015
Huster, R. J., Eichele, T., Enriquez-Geppert, S., Wollbrink, A, Kugel, H., Konrad, C., & Pantev, C.
(2011). Multimodal imaging of functional networks and event-related potentials in performance
monitoring. NeuroImage, 56(3), 1588–1597. https://doi.org/10.1016/j.neuroimage.2011.03.039
Hyman, J. M., Hasselmo, M. E., & Seamans, J. K. (2011). What is the functional relevance of pre-
frontal cortex entrainment to hippocampal theta rhythms? Frontiers in Neuroscience, 5, 1–13.
https://doi.org/10.3389/fnins.2011.00024
Inouye, T., Shinosaki, K., Iyama, A., Matsumoto, Y., & Toi, S. (1994). Moving potential field of
frontal midline theta activity during a mental task. Cognitive Brain Research, 2, 87–92.
Inouye, T., Shinosaki, K., Iyama, A., Matsumoto, Y., Toi, S., & Ishihar, T. (1994). Potential flow
of frontal midline theta activity during a mental task in the human electroencephalogram.
Neuroscience Letters, 169, 145–148.
194 JAMES F. CAVANAGH and MICHAEL X COHEN
Insel, T., Cuthbert, B., Garvey, M., Heinssen, R., Pine, D. S., Quinn, K., . . . Wang, P. (2010).
Research domain criteria (RDoC): Toward a new classification framework for research on
mental disorders. American Journal of Psychiatry, 167(7), 748–751. https://doi.org/167/7/748
[pii]10.1176/appi.ajp.2010.09091379
Ishii, R., Shinosaki, K., Ukai, S., Inouye, T., Ishihara, T., Yoshimine, T., . . . Takeda, M. (1999).
Medial prefrontal cortex generates frontal midline theta rhythm. Neuroreport, 10(4), 675–
679. https://pubmed.ncbi.nlm.nih.gov/10208529/
Itthipuripat, S., Wessel, J. R., & Aron, A. R. (2013). Frontal theta is a signature of successful
working memory manipulation. Experimental Brain Research, 224(2), 255–262. https://doi.
org/10.1007/s00221-012-3305-3
Jacobs, J., Hwang, G., Curran, T., & Kahana, M. J. (2006). EEG oscillations and recognition
memory: Theta correlates of memory retrieval and decision making. NeuroImage, 32(2),
978–987. https://pubmed.ncbi.nlm.nih.gov/16843012/
Javitt, D. C., Lee, M., Kantrowitz, J. T., & Martinez, A. (2018). Mismatch negativity as a bio-
marker of theta band oscillatory dysfunction in schizophrenia. Schizophrenia Research, 191,
51–60. https://doi.org/10.1016/j.schres.2017.06.023
Jensen, O. & Tesche, C. D. (2002). Frontal theta activity in humans increases with memory load
in a working memory task. European Journal of Neuroscience, 15(8), 1395–1399. https://pub
med.ncbi.nlm.nih.gov/11994134/
Jensen, O. & Colgin, L. L. (2007). Cross-frequency coupling between neuronal oscillations.
Trends in Cognitive Sciences, 11(7), 267–269. https://doi.org/10.1016/j.tics.2007.05.003
Jensen, O. & Lisman, J. E. (2005). Hippocampal sequence-encoding driven by a cortical multi-
item working memory buffer. Trends in Neurosciences, 28(2), 67–72. https://doi.org/10.1016/
j.tins.2004.12.001
Jocham, G. & Ullsperger, M. (2009). Neuropharmacology of performance monitoring. Neuroscience
and Biobehavioral Reviews, 33(1), 48–60. https://pubmed.ncbi.nlm.nih.gov/18789964/
Jones, M. W. & Wilson, M. A. (2005a). Phase precession of medial prefrontal cortical activity
relative to the hippocampal theta rhythm. Hippocampus, 15(7), 867–873. https://doi.org/
10.1002/hipo.20119
Jones, M. W. & Wilson, M. A. (2005b). Theta rhythms coordinate hippocampal-prefrontal
interactions in a spatial memory task. PLoS Biology, 3(12), e402. https://doi.org/10.1371/jour
nal.pbio.0030402
Jutras, M. J., Fries, P., & Buffalo, E. A. (2013). Oscillatory activity in the monkey hippocampus
during visual exploration and memory formation. Proceedings of the National Academy of
Sciences of the United States of America, 110(32), 13144–13149.
Kahana, M. J., Sekuler, R., Caplan, J. B., Kirschen, M., & Madsen, J. R. (1999). Human theta
oscillations exhibit task dependence during virtual maze navigation. Nature, 399(6738), 781–
784. https://doi.org/10.1038/21645
Kamarajan, C., Pandey, A. K., Chorlian, D. B., Manz, N., Stimus, A. T., Anokhin, A.
P., . . . Porjesz, B. (2015). Deficient event-related theta oscillations in individuals at risk for
alcoholism: A study of reward processing and impulsivity features. PLoS One, 10(11), 1–32.
https://doi.org/10.1371/journal.pone.0142659
Kang, S. J., Rangaswamy, M., Manz, N., Wang, J. C., Wetherill, L., Hinrichs, T., . . . Porjesz, B.
(2012). Family-based genome-wide association study of frontal theta oscillations identifies
potassium channel gene KCNJ6. Genes, Brain and Behavior, 11(6), 712–7 19. https://doi.org/
10.1111/j.1601-183X.2012.00803.x
FRONTAL MIDLINE THETA AS A MODEL SPECIMEN OF CORTICAL THETA 195
Killian, N. J., Jutras, M. J., & Buffalo, E. A. (2012). A map of visual space in the primate
entorhinal cortex. Nature, 491(7426), 761–764. https://doi.org/10.1038/nature11587
Klimesch, W. (1999). EEG alpha and theta oscillations reflect cognitive and memory per-
formance: a review and analysis. Brain Research: Brain Research Reviews, 29(2–3), 169–195.
https://pubmed.ncbi.nlm.nih.gov/10209231/
Klimesch, W., Doppelmayr, M., Schwaiger, J., Winkler, T., & Gruber, W. (2000). Theta
oscillations and the ERP old/new effect: Independent phenomena? Clinical Neurophysiology,
111(5), 781–793.
Ko, D., Kwon, S., Lee, G. T., Im, C. H., Kim, K. H., & Jung, K. Y. (2012). Theta oscillation related
to the auditory discrimination process in mismatch negativity: Oddball versus control para-
digm. Journal of Clinical Neurology (Korea), 8(1), 35–42. https://doi.org/10.3988/jcn.2012.8.1.35
Lee, H., Simpson, G. V., Logothetis, N. K., & Rainer, G. (2005). Phase locking of single neuron
activity to theta oscillations during working memory in monkey extrastriate visual cortex.
Neuron, 45(1), 147–156. https://doi.org/10.1016/j.neuron.2004.12.025
Lee, M., Balla, A., Sershen, H., Sehatpour, P., Lakatos, P., & Javitt, D. C. (2018). Rodent
mismatch negativity/theta neuro-oscillatory response as a translational neurophysio-
logical biomarker for N-methyl-D-aspartate receptor-based new treatment development
in schizophrenia. Neuropsychopharmacology, 43(3), 571–582. https://doi.org/10.1038/
npp.2017.176
Liebe, S., Hoerzer, G. M., Logothetis, N. K., & Rainer, G. (2012). Theta coupling between V4 and
prefrontal cortex predicts visual short-term memory performance. Nature Neuroscience,
15(3), 456–462, S1-2. https://doi.org/10.1038/nn.3038
Light, G. A., Swerdlow, N. R., Thomas, M. L., Calkins, M. E., Green, M. F., Greenwood, T.
A., . . . Turetsky, B. I. (2015). Validation of mismatch negativity and P3a for use in multi-
site studies of schizophrenia: Characterization of demographic, clinical, cognitive, and
functional correlates in COGS-2. Schizophrenia Research, 163(1–3), 63–72. https://doi.org/
10.1016/j.schres.2014.09.042
Lubenov, E. V. & Siapas, A. G. (2009). Hippocampal theta oscillations are travelling waves.
Nature, 459(7246), 534–539. https://doi.org/10.1038/nature08010
Luschei, E. S. (2006). Patterns of laryngeal electromyography and the activity of the respiratory
system during spontaneous laughter. Journal of Neurophysiology, 96(1), 442–450. https://doi.
org/10.1152/jn.00102.2006
Luu, P. & Tucker, D. M. (2001). Regulating action: Alternating activation of midline frontal and
motor cortical networks. Clinical Neurophysiology, 112(7), 1295–1306. https://doi.org/S1388-
2457(01)00559-4 [pii]
Luu, P., Tucker, D. M., & Makeig, S. (2004). Frontal midline theta and the error-related nega-
tivity: Neurophysiological mechanisms of action regulation. Clinical Neurophysiology,
115(8), 1821–1835. https://doi.org/10.1016/j.clinph.2004.03.031S1388245704001294 [pii]
MacNeilage, P. F. & Davis, B. L. (2001). Motor mechanisms in speech ontogeny: Phylogenetic,
neurobiological and linguistic implications. Current Opinion in Neurobiology, 11(6), 696–
700. https://doi.org/10.1016/S0959-4388(01)00271-9
Macrides, F., Eichenbaum, H., & Forbes, W. B. (1982). Temporal relationship between sniffing
and the limbic theta rhythm during odor discrimination reversal learning. Journal of
Neuroscience, 2(12), 1705–1717.
Miltner, W. H. R., Braun, C. H., & Coles, M. G. H. (1997). Event-related brain potentials
following incorrect feedback in a time-estimation task: Evidence for a “generic” neural
system for error detection. Journal of Cognitive Neuroscience, 9(6), 788–798.
196 JAMES F. CAVANAGH and MICHAEL X COHEN
Mitchell, D. J., McNaughton, N., Flanagan, D., & Kirk, I. J. (2008). Frontal-midline theta from
the perspective of hippocampal “theta”. Progress in Neurobiology, 86(3), 156–185. https://doi.
org/10.1016/j.pneurobio.2008.09.005
Mizuki, Y., Kajimura, N., Kai, S., Suetsugi, M., Ushijima, I., & Yamada, M. (1992). Differential
responses to mental stress in high and low anxious normal humans assessed by frontal mid-
line theta activity. International Journal of Psychophysiology, 12, 169–178.
Mizuki, Y., Suetsugi, M., Ushijima, I., & Yamada, M. (1996). Differential effects of noradrenergic
drugs on anxiety and arousal in healthy volunteers with high and low anxiety. Progress in
Neuro-Psychopharmacology & Biological Psychiatry, 20(8), 1353–1367.
Mizuki, Y, Suetsugi, M., Ushijima, I., & Yamada, M. (1997). Differential effects of dopamin-
ergic drugs on anxiety and arousal in healthy volunteers with high and low anxiety. Progress
in Neuro-Psychopharmacology & Biological Psychiatry, 21(4), 573–590. http://www.ncbi.nlm.
nih.gov/pubmed/9194141
Moran, T. P., Bernat, E. M., Aviyente, S., Schroder, H. S., & Moser, J. S. (2014). Sending mixed
signals: Worry is associated with enhanced initial error processing but reduced call for sub-
sequent cognitive control. Social Cognitive and Affective Neuroscience, 10(11), 1548–1556.
https://doi.org/10.1093/scan/nsv046
Moser, J. S., Moran, T. P., Schroder, H. S., Donnellan, M. B., & Yeung, N. (2013). On the relation-
ship between anxiety and error monitoring: A meta-analysis and conceptual framework.
Frontiers in Human Neuroscience, 7, 466. https://doi.org/10.3389/fnhum.2013.00466
Munneke, G.-J., Nap, T. S., Schippers, E. E., & Cohen, M. X. (2015). A statistical comparison of
EEG time-and time–frequency domain representations of error processing. Brain Research,
1618, 222–230. https://doi.org/10.1016/j.brainres.2015.05.030
Narayanan, N. S., Cavanagh, J. F., Frank, M. J., & Laubach, M. (2013). Common medial frontal
mechanisms of adaptive control in humans and rodents. Nature Neuroscience, 16, 1–10.
https://doi.org/10.1038/nn.3549
Nigbur, R., Cohen, M. X., Ridderinkhof, K. R., & Stürmer, B. (2012). Theta dynamics re-
veal domain-specific control over stimulus and response conflict. Journal of Cognitive
Neuroscience, 24(5), 1264–1274.
Nyhus, E., & Curran, T. (2010). Functional role of gamma and theta oscillations in episodic
memory. Neuroscience and Biobehavioral Reviews, 34(7), 1023–1035. https://doi.org/10.1016/
j.neubiorev.2009.12.014
Oehrn, C. R., Hanslmayr, S., Fell, J., Deuker, L., Kremers, N. A., Do Lam, A. T., . . . Axmacher,
N. (2014). Neural communication patterns underlying conflict detection, resolution, and
adaptation. Journal of Neuroscience, 34(31), 10438–10452. https://doi.org/10.1523/JNEURO
SCI.3099-13.2014
Onton, J., Delorme, A., & Makeig, S. (2005). Frontal midline EEG dynamics during working
memory. NeuroImage, 27(2), 341–356. https://pubmed.ncbi.nlm.nih.gov/15927487/
Pailing, P. E. & Segalowitz, S. J. (2004). The effects of uncertainty in error monitoring on
associated ERPs. Brain and Cognition, 56(2), 215–233.
Parker, K. L., Chen, K.-H., Kingyon, J. R., Cavanagh, J. F., & Narayanan, N. S. (2015). Medial
frontal ~4-Hz activity in humans and rodents is attenuated in PD patients and in rodents
with cortical dopamine depletion. Journal of Neurophysiology, 114(2), 1310–1320. https://doi.
org/10.1152/jn.00412.2015
Pastötter, B., Hanslmayr, S., & Bauml, K.-H. T. (2010). Conflict processing in the anterior cin-
gulate cortex constrains response priming. NeuroImage, 50(4), 1599–1605. https://doi.org/
10.1016/j.neuroimage.2010.01.095
FRONTAL MIDLINE THETA AS A MODEL SPECIMEN OF CORTICAL THETA 197
Paz, R., Bauer, E. P., & Pare, D. (2008). Theta synchronizes the activity of medial prefrontal
neurons during learning. Learning & Memory, 15(7), 524–531. https://doi.org/10.1101/
lm.932408
Pellegrino, F., Coupé, C., & Marsico, E. (2011). Across-language perspective on speech infor-
mation rate. Language, 87(3), 539–558. https://doi.org/10.1353/lan.2011.0057
Petajan, J. H. & Williams, D. D. (1972). Behavior of single motor units during pre-shivering
tone and shivering tremor. American Journal of Physical Medicine, 51(1), 16–22.
Phillips, J. M. & Everling, S. (2014). Event-related potentials associated with performance
monitoring in non-human primates. NeuroImage, 97, 308–320. https://doi.org/10.1016/j.neu
roimage.2014.04.028
Phillips, J. M., Vinck, M., Everling, S., & Womelsdorf, T. (2013). A long-range fronto-parietal
5-to 10-Hz network predicts “top-down” controlled guidance in a task-switch paradigm.
Cerebral Cortex, 24(8), 1996–2008. https://doi.org/10.1093/cercor/bht050
Pignatelli, M., Beyeler, A., & Leinekugel, X. (2012). Neural circuits underlying the generation
of theta oscillations. Journal of Physiology, 106(3–4), 81–92. https://doi.org/10.1016/j.jphyspa
ris.2011.09.007
Popov, T., Westner, B. U., Silton, R. L., Sass, S. M., Spielberg, J. M., Rockstroh, B., . . . Miller, G.
A. (2018). Time course of brain network reconfiguration supporting inhibitory control. The
Journal of Neuroscience, 38(18), 2639–17. https://doi.org/10.1523/JNEUROSCI.2639-17.2018
Raghavachari, S., Lisman, J. E., Tully, M., Madsen, J. R., Bromfield, E. B., & Kahana, M. J.
(2006). Theta oscillations in human cortex during a working-memory task: Evidence
for local generators. Journal of Neurophysiology, 95(3), 1630–1638. doi: 10.1152/
jn.00409.2005
Rajan, A., Siegel, S. N., Liu, Y., Bengson, J., Mangun, G. R., & Ding, M. (2018). Theta oscillations
index frontal decision-making and mediate reciprocal frontal–parietal interactions in
willed attention. Cerebral Cortex, 29(7), 2832–2843 . https://doi.org/10.1093/cercor/bhy149
Rangaswamy, M., Jones, K. A., Porjesz, B., Chorlian, D. B., Padmanabhapillai, A., Kamarajan,
C., . . . Begleiter, H. (2007). Delta and theta oscillations as risk markers in adolescent off-
spring of alcoholics. International Journal of Psychophysiology, 63(1), 3–15. https://doi.org/
10.1016/j.ijpsycho.2006.10.003
Reinhart, R. M. G. & Woodman, G. F. (2014). Causal control of medial-frontal cortex governs
electrophysiological and behavioral indices of performance monitoring and learning.
Journal of Neuroscience, 34(12), 4214–4227. https://doi.org/10.1523/JNEUROSCI.5421-13.2014
Reinhart, R. M. G. (2017). Disruption and rescue of interareal theta phase coupling and
adaptive behavior. Proceedings of the National Academy of Sciences of the United States of
America, 114(43), 11542–11547. https://doi.org/10.1073/pnas.1710257114
Reinhart, R. M. G., Zhu, J., Park, S., & Woodman, G. F. (2015). Synchronizing theta oscillations
with direct- current stimulation strengthens adaptive control in the human brain.
Proceedings of the National Academy of Sciences of the United States of America, 112(30),
9448–9453. https://doi.org/10.1073/pnas.1504196112
Riesel, A., Goldhahn, S., & Kathmann, N. (2017). Hyperactive performance monitoring
as a transdiagnostic marker: Results from health anxiety in comparison to obsessive–
compulsive disorder. Neuropsychologia, 96, 1–8. https://doi.org/10.1016/j.neuropsycholo
gia.2016.12.029
Rizzuto, D. S., Madsen, J. R., Bromfield, E. B., Schulze-Bonhage, A., & Kahana, M. J. (2006).
Human neocortical oscillations exhibit theta phase differences between encoding and re-
trieval. NeuroImage, 31(3), 1352–1358. doi: 10.1016/j.neuroimage.2006.01.009
198 JAMES F. CAVANAGH and MICHAEL X COHEN
Rizzuto, D. S., Madsen, J. R., Bromfield, E. B., Schulze-Bonhage, A., Seelig, D., Aschenbrenner-
Scheibe, R., & Kahana, M. J. (2003). Reset of human neocortical oscillations during a
working memory task. Proceedings of the National Academy of Sciences of the United States of
America, 100(13), 7931–7936. https://doi.org/10.1073/pnas.07320611000732061100 [pii]
Roberts, B. M., Hsieh, L. T., & Ranganath, C. (2013). Oscillatory activity during maintenance
of spatial and temporal information in working memory. Neuropsychologia, 51(2), 349–357.
https://doi.org/10.1016/j.neuropsychologia.2012.10.009
Rothé, M., Quilodran, R., Sallet, J., & Procyk, E. (2011). Coordination of high gamma activity
in anterior cingulate and lateral prefrontal cortical areas during adaptation. The Journal of
Neuroscience, 31(31), 11110–11117. https://doi.org/10.1523/JNEUROSCI.1016-11.2011
Ryman, S. G., Cavanagh, J. F., Wertz, C. J., Shaff, N. A., Dodd, A. B., Stevens, B., . . . Mayer, A.
R. (2018). Impaired midline theta power and connectivity during proactive cognitive con-
trol in schizophrenia. Biological Psychiatry, 84(9), 675–683. https://doi.org/10.1016/j.biops
ych.2018.04.021
Sato, K. (1952). On the theta rhythms in healthy human EEG. Folia Psychiatrica et Neurologica,
5(4), 283–297.
Sauseng, P., Griesmayr, B., Freunberger, R., & Klimesch, W. (2010). Control mechanisms
in working memory: A possible function of EEG theta oscillations. Neuroscience and
Biobehavioral Reviews, 34(7), 1015–1022. doi: 10.1016/j.neubiorev.2009.12.006
Sauseng, P., Klimesch, W., Doppelmayr, M., Hanslmayr, S., Schabus, M., & Gruber, W. R.
(2004). Theta coupling in the human electroencephalogram during a working memory task.
Neuroscience Letters, 354(2), 123–126. doi: 10.1016/j.neulet.2003.10.002
Schmidt, B., Kanis, H., Holroyd, C. B., Miltner, W. H. R., & Hewig, J. (2018). Anxious
gambling: Anxiety is associated with higher frontal midline theta predicting less risky
decisions. Psychophysiology, 55(10), 1–9. https://doi.org/10.1111/psyp.13210
Shackman, A. J., Salomons, T. V, Slagter, H. A, Fox, A. S., Winter, J. J., & Davidson, R. J. (2011).
The integration of negative affect, pain and cognitive control in the cingulate cortex. Nature
Reviews. Neuroscience, 12(3), 154–167. https://doi.org/10.1038/nrn2994
Siapas, A. G., Lubenov, E. V, & Wilson, M. A. (2005). Prefrontal phase locking to hippocampal
theta oscillations. Neuron, 46(1), 141–151. https://doi.org/10.1016/j.neuron.2005.02.028
Smit, A. S., Eling, P. A T. M., & Coenen, A. M. L. (2004). Mental effort affects vigilance en-
duringly: after-effects in EEG and behavior. International Journal of Psychophysiology, 53(3),
239–243. https://doi.org/10.1016/j.ijpsycho.2004.04.005
Smit, A. S., Eling, P. A T. M., Hopman, M. T., & Coenen, A. M. L. (2005). Mental and physical
effort affect vigilance differently. International Journal of Psychophysiology, 57(3), 211–217.
https://doi.org/10.1016/j.ijpsycho.2005.02.001
Smith, E. H., Banks, G. P., Mikell, C. B., Cash, S. S., Patel, S. R., Eskandar, E. N., & Sheth, S. A.
(2015). Frequency-dependent representation of reinforcement-related information in the
human medial and lateral prefrontal cortex. Journal of Neuroscience, 35(48), 15827–15836.
https://doi.org/10.1523/JNEUROSCI.1864-15.2015
Solomon, E. A., Kragel, J. E., Sperling, M. R., Sharan, A., Worrell, G., Kucewicz, M., . . . Kahana,
M. J. (2017). Widespread theta synchrony and high-frequency desynchronization underlies
enhanced cognition. Nature Communications, 8, 1704. https://doi.org/10.1038/s41
467-017-01763-2
Staudigl, T., Hanslmayr, S., & Bauml, K. H. (2010). Theta oscillations reflect the dynamics of
interference in episodic memory retrieval. Journal of Neuroscience, 30(34), 11356–11362.
doi:10.1523/JNEUROSCI.0637-10.2010
FRONTAL MIDLINE THETA AS A MODEL SPECIMEN OF CORTICAL THETA 199
Swart, J. C., Frank, M. J., Määttä, J. I., Jensen, O., Cools, R., & den Ouden, H. E. M. (2018).
Frontal network dynamics reflect neurocomputational mechanisms for reducing maladap-
tive biases in motivated action. PLoS Biology [online], 1–25. https://doi.org/10.1371/journal.
pbio.2005979
Talairach, J., Bancaud, J., Geier, S., Bordas-Ferrer, M., Bonis, A., Szikla, G., & Rusu, M.
(1973). The cingulate gyrus and human behaviour. Electroencephalography and Clinical
Neurophysiology, 34(1), 45–52. doi: 10.1016/0013-4694(73)90149-1
Tang, H., Yu, H. Y., Chou, C. C., Crone, N. E., Madsen, J. R., Anderson, W. S., & Kreiman, G.
(2016). Cascade of neural processing orchestrates cognitive control in human frontal cortex.
ELife, 5, e12352. https://doi.org/10.7554/eLife.12352
Taub, A. H., Perets, R., Kahana, E., & Paz, R. (2018). Oscillations synchronize amygdala-to-pre-
frontal primate circuits during aversive learning. Neuron, 97(2), 291–298.e3. https://doi.org/
10.1016/j.neuron.2017.11.042
Töllner, T., Wang, Y., Makeig, S., Müller, H. J., Jung, T.-P., & Gramann, K. (2017). Two inde-
pendent frontal midline theta oscillations during conflict detection and adaptation in a
Simon-type manual reaching task. The Journal of Neuroscience, 37(9), 2504–2515. https://doi.
org/10.1523/JNEUROSCI.1752-16.2017
Tóth, B., Kardos, Z., File, B., Boha, R., Stam, C. J., & Molnár, M. (2014). Frontal midline
theta connectivity is related to efficiency of WM maintenance and is affected by aging.
Neurobiology of Learning and Memory, 114, 58–69. https://doi.org/10.1016/j.nlm.2014.04.009
Trujillo, L. T. & Allen, J. J. (2007). Theta EEG dynamics of the error-related negativity. Clinical
Neurophysiology, 118(3), 645–668. doi: 10.1016/j.clinph.2006.11.009
Tsujimoto, S., Genovesio, A., & Wise, S. P. (2010). Evaluating self-generated decisions in frontal
pole cortex of monkeys. Nature Neuroscience, 13(1), 120–126. https://doi.org/10.1038/nn.2453
Tsujimoto, T., Shimazu, H., & Isomura, Y. (2006). Direct recording of theta oscillations in pri-
mate prefrontal and anterior cingulate cortices. Journal of Neurophysiology, 95(5), 2987–
3000. doi:10.1152/jn.00730.2005
van de Vijver, I., Cohen, M. X., & Ridderinkhof, K. R. (2014). Aging affects medial but not
anterior frontal learning-related theta oscillations. Neurobiology of Aging, 35(3), 692–704.
https://doi.org/10.1016/j.neurobiolaging.2013.09.006
van de Vijver, I., Ridderinkhof, K. R., & Cohen, M. X. (2011). Frontal oscillatory dynamics
predict feedback learning and action adjustment. Journal of Cognitive Neuroscience, 23(12),
4106–4121. https://doi.org/10.1162/jocn_a_00110
van Driel, J., Ridderinkhof, K. R., & Cohen, M. X. (2012). Not all errors are alike: Theta and
alpha EEG dynamics relate to differences in error-processing dynamics. The Journal of
Neuroscience, 32(47), 16795–16806. https://doi.org/10.1523/JNEUROSCI.0802-12.2012
van Noordt, S. J. R. & Segalowitz, S. J. (2012). Performance monitoring and the medial
prefrontal cortex: A review of individual differences and context effects as a window
on self-regulation. Frontiers in Human Neuroscience, 6, 197. https://doi.org/10.3389/
fnhum.2012.00197
Van Veen, V. & Carter, C. S. (2002). The timing of action-monitoring processes in the anterior
cingulate cortex. Journal of Cognitive Neuroscience, 14(4), 593–602. https://doi.org/10.1162/
08989290260045837
VanRullen, R. (2016). Perceptual cycles. Trends in Cognitive Sciences, 20(10), 723–735. https://
doi.org/10.1016/j.tics.2016.07.006
Verguts, T. (2017). Binding by random bursts: A computational model of cognitive control.
Journal of Cognitive Neuroscience, 29(6), 1103–1118. https://doi.org/10.1162/jocn_a_01117
200 JAMES F. CAVANAGH and MICHAEL X COHEN
Vezoli, J. & Procyk, E. (2009). Frontal feedback- related potentials in nonhuman pri-
mates: modulation during learning and under haloperidol. Journal of Neuroscience, 29(50),
15675–15683. doi:10.1523/JNEUROSCI.4943-09.2009
Vidal, F., Burle, B., Bonnet, M., Grapperon, J., & Hasbroucq, T. (2003). Error negativity on
correct trials: a reexamination of available data. Biological Psychology, 64(3), 265–282.
https://doi.org/S0301051103000978
Vidal, F., Hasbroucq, T., Grapperon, J., & Bonnet, M. (2000). Is the “error negativity”
specific to errors? Biological Psychology, 51(2–3), 109–128. https://doi.org/10.1016/
S0301-0511(99)00032-0
Vissers, M. E., Ridderinkhof, K. R., Cohen, M. X., & Slagter, H. A. (2018). Oscillatory
mechanisms of response conflict elicited by color and motion direction: An individual
differences approach. Journal of Cognitive Neuroscience, 30(4), 468–481. https://doi.org/
10.1162/jocn_a_01222
Walsh, M. M. & Anderson, J. R. (2012). Learning from experience: Event-related potential
correlates of reward processing, neural adaptation, and behavioral choice. Neuroscience and
Biobehavioral Reviews, 36(8), 1870–1884. https://doi.org/10.1016/j.neubiorev.2012.05.008
Wang, C., Ulbert, I., Schomer, D. L., Marinkovic, K., & Halgren, E. (2005). Responses of human
anterior cingulate cortex microdomains to error detection, conflict monitoring, stimulus-
response mapping, familiarity, and orienting. Journal of Neuroscience, 25(3), 604–613.
doi:10.1523/JNEUROSCI.4151-04.2005
Wang, D., Clouter, A., Chen, Q., Shapiro, K. L., & Hanslmayr, S. (2018). Single-trial phase entrain-
ment of theta oscillations in sensory regions predicts human associative memory performance.
The Journal of Neuroscience, 38(28), 0349–18. https://doi.org/10.1523/JNEUROSCI.0349-18.2018
Wascher, E., Rasch, B., Sänger, J., Hoffmann, S., Schneider, D., Rinkenauer, G., . . . Gutberlet, I.
(2014). Frontal theta activity reflects distinct aspects of mental fatigue. Biological Psychology,
96, 57–65. https://doi.org/10.1016/j.biopsycho.2013.11.010
Wessel, J. R. (2018). An adaptive orienting theory of error processing. Psychophysiology, 55(3),
1–21. https://doi.org/10.1111/psyp.13041
White, J. A., Banks, M. I., Pearce, R. A., & Kopell, N. J. (2000). Networks of interneurons with
fast and slow gamma-aminobutyric acid type A (GABAA) kinetics provide substrate for
mixed gamma–theta rhythm. Proceedings of the National Academy of Sciences of the United
States of America, 97(14), 8128–8133.
Wiecki, T. V. & Frank, M. J. (2013). A computational model of inhibitory control in frontal
cortex and basal ganglia. Psychological Review, 120(2), 329–355.
Womelsdorf, T., Johnston, K., Vinck, M., & Everling, S. (2010). Theta-activity in anterior cin-
gulate cortex predicts task rules and their adjustments following errors. Proceedings of the
National Academy of Sciences of the United States of America, 107(11), 5248–5253. doi: 10.1073/
pnas.0906194107
Womelsdorf, T., Schoffelen, J. M., Oostenveld, R., Singer, W., Desimone, R., Engel, A. K., &
Fries, P. (2007). Modulation of neuronal interactions through neuronal synchronization.
Science, 316(5831), 1609–1612. doi:10.1126/science.1139597
Womelsdorf, T., Valiante, T. A., Sahin, N. T., Miller, K. J., & Tiesinga, P. (2014). Dynamic circuit
motifs underlying rhythmic gain control, gating and integration. Nature Neuroscience, 17(8),
1031–1039. https://doi.org/10.1038/nn.3764
Womelsdorf, T., Vinck, M., Leung, L. S., & Everling, S. (2010). Selective theta-synchronization
of choice- relevant information subserves goal- directed behavior. Frontiers in Human
Neuroscience, 4, 210. https://doi.org/10.3389/fnhum.2010.00210
FRONTAL MIDLINE THETA AS A MODEL SPECIMEN OF CORTICAL THETA 201
Yamaguchi, M., Crump, M. J. C., & Logan, G. D. (2013). Speed-accuracy trade-off in skilled
typewriting: Decomposing the contributions of hierarchical control loops. Journal of
Experimental Psychology: Human Perception and Performance, 39(3), 678–699.
Yordanova, J, Falkenstein, M., Hohnsbein, J., & Kolev, V. (2004). Parallel systems of error
processing in the brain. NeuroImage, 22(2), 590–602. https://doi.org/10.1016/j.neuroim
age.2004.01.040S1053811904000850 [pii]
Yordanova, Juliana, Kolev, V., Rosso, O. A., Schürmann, M., Sakowitz, O. W., Özgören, M., &
Basar, E. (2002). Wavelet entropy analysis of event-related potentials indicates modality-
independent theta dominance. Journal of Neuroscience Methods, 117(1), 99–109. https://doi.
org/10.1016/S0165-0270(02)00095-X
Zavala, B. A., Tan, H., Little, S., Ashkan, K., Hariz, M., Foltynie, T., . . . Brown, P. (2014). Midline
frontal cortex low-frequency activity drives subthalamic nucleus oscillations during conflict.
Journal of Neuroscience, 34(21), 7322–7333. https://doi.org/10.1523/JNEUROSCI.1169-14.2014
Zavala, B., Brittain, J.-S., Jenkinson, N., Ashkan, K., Foltynie, T., Limousin, P., . . . Brown, P.
(2013). Subthalamic nucleus local field potential activity during the Eriksen Flanker task
reveals a novel role for theta phase during conflict monitoring. The Journal of Neuroscience,
33(37), 14758–14766. https://doi.org/10.1523/JNEUROSCI.1036-13.2013
Zavala, B., Tan, H., Little, S., Ashkan, K., Green, A. L., Aziz, T., . . . Brown, P. (2016). Decisions
made with less evidence involve higher levels of corticosubthalamic nucleus theta band
synchrony. Journal of Cognitive Neuroscience, 28(6), 811–825. https://doi.org/10.1162/jocn_
a_00934
Zhang, H., Watrous, A. J., Patel, A., & Jacobs, J. (2018). Theta and alpha oscillations are traveling
waves in the human neocortex. Neuron, 98(6), 1269–1281.e4. https://doi.org/10.1016/j.neu
ron.2018.05.019
Zlojutro, M., Manz, N., Rangaswamy, M., Xuei, X., Flury-Wetherill, L., Koller, D., . . . Almasy,
L. (2011). Genome-wide association study of theta band event-related oscillations identifies
serotonin receptor gene HTR7 influencing risk of alcohol dependence. American Journal of
Medical Genetics, Part B: Neuropsychiatric Genetics, 156(1), 44–58. https://doi.org/10.1002/
ajmg.b.31136
CHAPTER 10
In 1929 a German physician named Hans Berger recorded the first human EEG. The
first thing he noticed was a regular oscillation with a frequency of 10 Hz (Berger, 1929),
which he termed alpha oscillations. Many decades after Berger discovered these alpha
oscillations researchers use them as a first quality check of their EEG signal. After
attaching electrodes to the subject’s head, the researcher typically asks the subject to
close their eyes and relax; the researcher then sees beautiful alpha waves with a fre-
quency of around 10 Hz being maximal over posterior channels. When the subject then
opened their eyes, alpha oscillations largely reduce. This phenomenon is extremely re-
liable, such that not seeing the reduction in alpha amplitude when subjects open their
eyes would typically indicate that something went wrong. Reductions in alpha amp-
litude occur in all ranges of cognitive tasks such as visual processing (Adrian, 1944),
auditory processing (Krause et al., 1994; Obleser & Weisz, 2012), somatosensory pro-
cessing (Crone et al., 1998), memory encoding (Hanslmayr et al., 2009; Klimesch et al.,
1996), memory retrieval (Burgess & Gruzelier, 2000; Waldhauser et al., 2016), working
memory maintenance (Sauseng et al., 2009), decision making (Pornpattananangkul
THE ROLE OF ALPHA AND BETA OSCILLATIONS IN THE HUMAN EEG 203
et al., 2019), and motor preparation and execution (Pfurtscheller et al., 1997). The exact
frequency at which power reductions are maximal varies between tasks and often
involves the faster beta oscillation around 15 Hz, which can be considered the “fast” sib-
ling of alpha. In this chapter we therefore do not distinguish between alpha and beta
oscillations and refer to these oscillations as alpha/beta. Suppression of alpha/beta
oscillations is not only observed in humans, but also in a wide range of animals, for ex-
ample, non-human primates (Haegens et al., 2011), dogs (Lopes da Silva et al., 1980), cats
(von Stein et al., 2000), rodents (Wiest & Nicolelis, 2003), and even insects (Popov &
Szyszka, 2019). This ubiquity of alpha/beta power suppression across cognitive domains
and across the animal kingdom indicates that alpha/beta power reductions are a signa-
ture of an extremely general mechanism, which is called upon in almost any cognitive
task and has been retained over millions of years of evolution. What could this mech-
anism be? This chapter attempts to answer this question.
Before delving into the different physiological interpretations of alpha oscillations
it is worth pointing out a theoretical caveat that has become evident in cognitive
neuroscience and the way we can avoid these problems. Historically the job descrip-
tion for a cognitive neuroscientist was to pick a cognitive phenomenon (i.e., attention,
memory) and then find the “neural correlate” of that phenomenon. To demonstrate
this approach let us consider a short thought experiment involving two hypothetical
cognitive neuroscientists: AC and BD. Researcher AC is interested in attention. She
runs several meticulously controlled experiments that all manipulate certain aspects
of attention while she records EEG. Across this series of experiments, she finds that
alpha oscillations are very consistently modulated by attention. She goes to a confer-
ence, presents her results, and concludes that alpha oscillations are the neural correlate
of attention. At that same conference, researcher BD, who is interested in memory, also
presents his results. In a series of carefully controlled experiments he shows that alpha
oscillations are reliably modulated by various memory processes. He concludes his talk
with saying that alpha oscillations are the neural correlate of memory.
So, who is right? What cognitive function do alpha oscillations represent: memory
or attention? The answer is both are wrong. The error that both scientists make is to
assume that there is a one-to-one mapping between neural phenomena and cognitive
functions. This error has become known under the term “reverse inference error”, that
is, observing neural signature X has no predictive value for cognitive process Y to occur
(Poldrack, 2011). With respect to alpha oscillations, given their ubiquity in terms of cog-
nitive tasks (from attention to decision making) and species (from human to honeybees)
it is evident that the task of trying to attach one particular label from cognitive psych-
ology (attention vs. memory) to them is bound to fail. An alternative approach is
needed if we want to truly understand the function of alpha oscillations. This alternative
approach needs to avoid limiting itself by definitions of cognitive processes. Instead, this
approach needs to embrace the fact that a given neural phenomenon can be of service
to many different cognitive processes in many different species. One such alternative is
to assume that different oscillations implement canonical computations (Siegel et al.,
2012; Womelsdorf et al., 2014), which are basic neural operations that are called upon by
204 SEBASTIAN MICHELMANN et al.
different cognitive processes. Regulating the balance between excitation and inhibition,
for instance, would be one such basic operation. Another basic operation is to enable
the neural representation of information rich content. These operations arguably are
required by almost any cognitive task and species. The difference from this approach to
the traditional cognitive neuroscience approach is not in using different labels, but to try
to understand the computational utility of a neural operation for a given cognitive pro-
cess (see Buzsaki, 2019 for a similar line of argument). Instead of asking what the neural
correlate of attention is, we ask how a decrease in alpha oscillations can be of service for
attention, memory, decision making, etc. Section 10.2 provides a brief overview of the
behaviour of alpha oscillations in terms of frequency, power, and phase. It is critical to
understand these terms first before we can consider the physiological interpretations of
alpha oscillations.
10.2.1 Frequency
Analyzing alpha oscillations typically involves transforming an EEG signal into time-
frequency space (see Cohen, 2014 for an excellent overview of the different methods
and hands-on tutorials). Figure 10.1A shows a typical example of a raw EEG trace in a
healthy human subject. It is easy to spot the decrease in alpha amplitudes at time 0 (when
a stimulus was presented). Brain oscillations are defined by three physical properties: (i)
frequency (Figure 10.1B), (ii) amplitude (Figure 10.1C), and (iii) phase (Figure 10.1D).
Different brain networks are hypothesized to oscillate at different frequencies (Keitel
& Gross, 2016), with small networks oscillating at fast frequencies (>40 Hz) and large
networks oscillating at slower frequencies (<20 Hz) (von Stein & Sarnthein, 2000). Small
and large here refers to the number of neurons involved in generating the signal. This ana-
tomical property is reflected in the 1/F power ratio of EEG signals, which refers to the drop
in signal power with increasing frequency. Alpha oscillations are remarkable because
they stand out from the 1/F pattern (Figure 10.1B), which shows that they are a particu-
larly strong oscillation recruiting large pools of neurons. Frequencies are typically used
to distinguish between different types of oscillations, that is, theta ~ 4 Hz, beta ~15 Hz,
gamma ~ 40 Hz, etc. The frequency within an oscillation has been shown to be linked to
interindividual variability in memory processes (Cohen, 2011) or intelligence (Anokhin
& Vogel, 1996). Recent work demonstrates that the frequency of alpha can also change
within a subject, from trial-to-trial (Haegens et al., 2014) and may reflect cortical excit-
ability ( Cohen, 2014). Therefore, it is important to consider that the frequency of alpha
may vary from individual to individual and even from trial to trial within a participant.
Sometimes it can even be challenging to distinguish a “fast theta” from a “slow alpha”.
THE ROLE OF ALPHA AND BETA OSCILLATIONS IN THE HUMAN EEG 205
Frequency (Hz)
Amplitude (μV)
5 T.F. -
0 10 1
Analysis
–10
–15 0
2
–20 –1 –0.5 0 0.5 1 1.5
–1 –0.5 0 0.5 1 1.5 Time (sec)
Time (sec.)
10 –π/2
Alpha 4.4 Hz
(~10 Hz) 0
Power
35 –10
–20
13 –30
7
3 –0.2 0 –40
0.2 0.4 0.6 0.8 1 1.2 1.4
Time [sec]
Frequency
Figure 10.1 Alpha oscillations and their parameters. (A) An example of a raw signal as
recorded with a parietal EEG electrode is shown on the left. A stimulus was presented at time
0. The plot on the right shows the results of a time-frequency analysis in which power is depicted
for each time-point (x-axis) and frequency band (y-axis; a.u. =arbitrary units). (B) A sche-
matic of a typical EEG power spectrum is shown, with frequency on the x-axis and power on
the y-axis. The inverse relationship between the size of neural assemblies and power is depicted.
Note the peak at the alpha frequency which violates the 1/F relation between power and fre-
quency. (C) A typical time-frequency plot showing event related power increases (hot colors)
and decreases (cold colors) during processing of verbal information. Note the power increases
in theta (3–7 Hz) and gamma (35–100 Hz) and the power decreases in alpha (8–12 Hz) and beta
(13–35 Hz). (D) The relationship between EEG phase (top) and firing rates (bottom) is shown.
The differently colored lines show phase modulations in trials with high (red), medium (purple)
or low (blue) power.
(A) Reprinted with permission from Hanslmayr et al., 2011. (C) is modified and reprinted with permission from
Hanslmayr et al., 2012. (D) modified and reprinted with permission from Jacobs et al., 2007.
10.2.2 Power
The power of an oscillation refers to its signal strength. It is usually calculated by taking
the square of the signal. In human EEG, a signal is generated by the summation of several
millions of postsynaptic potentials (inhibitory and excitatory) over an area of some cm2
(Pfurtscheller et al., 1996). Importantly, the current changes induced by postsynaptic
potentials are tiny and we record these potentials with an electrode that is attached to
the scalp, separated by various layers of bone, cerebrospinal fluid, skin, etc., from the
brain. Therefore, in order to detect a signal that is large enough to be visible in the EEG,
tens of thousands of post-synaptic potentials must come together in time. To give an
analogy, recording EEG is a bit like recording the noise in a football stadium with one
206 SEBASTIAN MICHELMANN et al.
big microphone hanging from the ceiling. Using this coarse signal we cannot stand the
chance to tune in onto an individual conversation between two fans, but we can use the
signal to tell whether a goal was scored (because thousands of fans start shouting at the
same time). Accordingly, the strength of an oscillation is assumed to reflect the degree
of synchrony between inhibitory or excitatory postsynaptic potentials to an underlying
neural assembly. Power increases indicate increased local synchrony whereas power
decreases indicate de-synchronized local activity. This idea is reflected in the classic work
of Pfurtscheller & Aranibar (1977), who coined the terms event-related synchroniza-
tion and de-synchronization (ERS/ERD), which denotes power increases and decreases
in response to an event or stimulus, respectively. In EEG experiments absolute power
is usually transformed into power changes in response to a baseline (e.g., prestimulus
interval). Figure 10.1C shows a typical example of such data with stimulus driven power
increases in the lower (1–8 Hz: delta/theta) and higher (40–100 Hz; gamma) frequency
ranges, and power decreases in the middle frequency ranges (8–35 Hz; alpha/beta).
10.2.3 Phase
The phase of an oscillation specifies the current position in a given cycle (Figure 10.1D),
that is whether the oscillation exhibits a peak, trough, or zero crossing. Because the EEG
reflects the sum of postsynaptic input to a given neuron, we can assume that it impinges
on the neuron’s probability to fire an action potential. Figure 10.1D shows a single neuron
recorded in the human brain, and that the neuron is more likely to fire in the trough of
the oscillation, which indeed resembles the time point of maximally coinciding excita-
tory input (note that the LFP in this case is measured in the extracellular space; there-
fore, negativity indexes excitation). This same figure also shows that the modulation of
firing rate is stronger for trials of high power (red) and lower for trials of low power
(blue). We can therefore assume that the phase of alpha oscillations (or any other oscil-
lation for that matter) represent discrete time windows for neural firing, and that this
synchronizing effect scales with power. This is a very useful computational property as it
gives alpha oscillations the power to control the timing of neural firing in large groups of
neurons. Like the conductor of an orchestra, who tells individual instruments when to
play a particular note (and when to not play a note), alpha oscillations can synchronize
large groups of neurons to temporally structure neural processing. This aspect becomes
particularly important in Section 10.5, which covers the role of alpha phase for sampling
information and replaying this information from memory.
as one would usually expect to see an increase in brain signal strength when a subject
performs a challenging task, not a decrease. How can we functionally interpret this
negative relationship between alpha power and cognitive processing? Up until the early
2000s the prevailing view was that alpha oscillations reflect a state of “idling” or rest. In
their article, Pfurtscheller and colleagues (1996) give the example of the motor cortex,
in particular the visual cortex and the hand area during a reading task or a motor task
requiring finger movements, respectively. During reading, the visual cortex displays
profound alpha-power decreases, whereas the hand motor area shows an increase in
alpha power. This picture switches when the subject engages in a motor task requiring
finger movement. Within the idling hypothesis, one would interpret the increased alpha
power over areas that are not required by the task as “nil work”, that is, a passive res-
onance phenomenon of a part of cortex that has “nothing to do” (Adrian & Matthews,
1934). Since almost any cognitive task always involves specific activation of some regions
and de-activation of other regions (Fox et al., 2005) the EEG would always reflect some
areas that show alpha power decreases (or desynchronization) and some areas that show
alpha-power increases (or synchronization). The important emphasis of the idling hy-
pothesis is on the passive aspect of alpha synchronization which has no functional role
per se.
Klimesch and colleagues (2007) and Jensen and Mazaheri (2010) presented a
contrasting view to the idling hypothesis and suggested that alpha oscillations play
a critical role in cognitive processing. The seminal findings that led to this interpret-
ation were studies showing that alpha power increased with increasing cognitive load
in a working memory task (Jensen et al., 2002; Klimesch et al., 1999). This alpha-power
increase with cognitive load is difficult to reconcile with the idling hypothesis, which
led to the active inhibition hypothesis (Jensen & Mazaheri, 2010; Klimesch et al., 2007).
Within the active inhibition hypothesis an increase in alpha power reflects an active
inhibition process that serves to silence a particular region that is task irrelevant. This
silencing of task-irrelevant areas ensures that information is processed selectively in
task-relevant areas and protects the processing of this information from interference
or noise. A critical prediction that the inhibition account made was that an increase in
alpha oscillations narrows the time windows for neurons to fire (see red vs. blue lines
in Figure 10.1D) and therefore reduces neural firing. Thus, periods of high alpha power
should coincide with low neural firing, whereas periods of low alpha power should co-
incide with high neural firing. This prediction was confirmed in a non-human primate
study (Haegens et al., 2011), which suggests that an increase in alpha power in a given
area acts as a “silencing mechanism” which muffles neural assemblies that otherwise
might interfere.
Coming back to our example, alpha synchronization of the hand area while reading
this chapter ensures that you can focus on the text instead of moving around (or
thinking about movements). The more challenging a task, the more we need to tune out
task-irrelevant activity. The inhibition account has gained considerable support over the
last decades because it can accommodate the findings in the literature better than the
idling hypothesis. To give two examples, if subjects maintain visual content in working
208 SEBASTIAN MICHELMANN et al.
memory that was presented in the left hemifield, therefore being processed in the right
occipital cortex, alpha power increases over the left occipital cortex (Sauseng et al.,
2009). Externally enhancing alpha power with repetitive transcranial magnetic stimu-
lation over the irrelevant hemisphere then increases working memory performance.
Similar evidence comes from Bonnefond and Jensen (2012), who demonstrated that
subjects actively increase their alpha power in anticipation of a task-irrelevant distractor
presented during working memory maintenance. The more the subjects upregulated
alpha power the better the performance on the working memory task. These results rule
out a passive perspective of alpha oscillations and instead suggest that alpha oscillations
are very much an active process that regulates neural activity to ensure selective infor-
mation processing. Within this perspective, alpha oscillations serve the function of a
filter that tunes out task-irrelevant information to render the task-relevant signal more
salient.
The active inhibition account has been extremely useful in interpreting the role of alpha
power increases, or alpha synchronization during cognitive processing. This is because
any cognitive process requires selective information processing; alpha oscillations, by
inhibiting task irrelevant neural assemblies, ensure such selective information pro-
cessing. This functional interpretation of alpha oscillations is broad enough to ac-
commodate the fact that modulations of alpha oscillations are observed in a variety
of cognitive tasks and species. Returning to the thought experiment from earlier, the
attention scientist AC and memory scientist BD would interpret their findings to show
that both memory and attention crucially rely on an active filtering mechanism. Indeed,
alpha power increases over areas that hold the representation of task-irrelevant infor-
mation, regardless of whether this information is currently perceived (Thut et al., 2006;
Worden et al., 2000), held in working memory (Sauseng et al., 2009), or stored in long-
term memory (Waldhauser et al., 2012). The emphasis of the active inhibition hypoth-
esis is on alpha synchronization and its computational utility in terms of providing a
filter mechanism for selective information processing. What is less clear from this per-
spective, however, is what the computational utility of alpha power decreases are for
information processing (other than allowing for increased neural spike rates through
less inhibition). We therefore proposed an additional theory, which is complemen-
tary to the inhibition account. Within this framework we emphasized the role of alpha
power decreases in allowing for high-fidelity information to be represented in neural
assemblies. A key assumption of this account rests on the fact the alpha/beta power
decreases represent periods of de-correlated neural firing (see Murthy & Fetz, 1996 for
such a demonstration for beta oscillations in the motor cortex).
THE ROLE OF ALPHA AND BETA OSCILLATIONS IN THE HUMAN EEG 209
One way of interpreting the functional utility of de-correlated firing in alpha for cogni-
tive processing applies tenets of information theory (Shannon & Weaver, 1949) to neural
oscillations. This framework is known as the “information-via-desynchronization hy-
pothesis” (Hanslmayr et al., 2012), which proposes that synchronized alpha/beta states
are inherently bad for information representation as neuronal activity is highly redun-
dant. Take the simplified instance of two presynaptic neurons that act upon the same
postsynaptic neuron: if one presynaptic neuron fires in perfect synchrony with another,
what can this neuron add to the neuronal code that its synchronous partner does not
already (Schneidman et al., 2011)? If we expand this principle to networks of neurons,
we can postulate that highly synchronous networks are detrimental to information pro-
cessing, because they only represent redundancies. To overcome this limitation there-
fore, the networks must desynchronize. Through desynchronization, the underlying
neural code can be more complex and hence convey a more detailed representation of
information. The event-related desynchronization so commonly seen in alpha oscilla-
tion may be a prime example of this phenomenon.
Numerous studies support the idea that alpha/beta power decreases reflect the rep-
resentation of information within the cortex. One of the most direct lines of support
comes from a recent simultaneous EEG-fMRI experiment by Griffiths and colleagues
(2019), who asked participants to associate video clips with words, and to later recall
the clips using the words as a cue (Figure 10.2). For each trial, the researchers quantified
the amount of visual information present in the cortex by conducting representa-
tional similarity analysis (RSA) on the fMRI data (Kriegeskorte et al., 2008). RSA is
based on correlations of neural patterns and reasons that representations of the same
content should elicit neural patterns that are more alike than the patterns elicited by
representations of different content (Kriegeskorte et al., 2008). The researchers then
asked whether the power of the alpha and beta frequencies (8–30Hz) correlated with
the quantity of information (calculated via RSA) represented on a given trial. Indeed,
they found evidence to suggest a parametric link between alpha/beta power and infor-
mation: as power decreased, information increased (Figure 10.2b). Similarly, Hanslmayr
and colleagues (2009) presented participants with words and asked them to engage in
semantic processing (i.e., does this word represent a living entity or a non-living en-
tity?) or shallow processing (i.e., does the first letter of the word precede the last in the
alphabet?). As semantic processing involves much greater levels of information pro-
cessing (you must not only process the letters, but also what those letters mean), the
researchers hypothesized that alpha/beta power decreases would be greater during
this type of processing. Their results revealed just that—suggesting that alpha/beta
power decreases scale with the depth of information processing (see Fellner et al., 2018
for similar results contrasting familiar and unfamiliar stimuli). In conjunction with a
number of other studies, these results strongly implicate alpha/beta power reductions in
information processing (Hanslmayr et al., 2012).
While the alpha idling theory assumes that synchronous alpha oscillations mark a
default state in which the cortex does nothing (Pfurtscheller et al., 1996), the alpha in-
hibition theory highlights the active role of alpha synchronization in the suppression of
210 SEBASTIAN MICHELMANN et al.
Noise correlations
0
Info-via-desync
–0.2
–0.4
Perception Retrieval
Low Sync/High Info. High Sync/Low Info. Low Sync/High Info.
Figure 10.2 Alpha power and information processing. (A) Infographic depicting theories be-
hind alpha and information processing. Perception and memory retrieval involve the processing
of large quantities of information. The noise correlation account proposes that when task irrele-
vant neurons (in blue) synchronize (e.g., during the interval), they mask the signal generated
by task-relevant neurons (in red). When the task-irrelevant neurons desynchronize however
(i.e., during perception/retrieval), the signal can be detected above the background noise. The
information-via-desynchronization account proposes that oscillatory desynchronization allows
a more complex neural code to be generated. Such a complex code is necessary to process the
highly complex information encountered in daily life. In both instances, oscillatory desyn-
chronization benefits information processing (B) Reproduction of the results by Griffiths and
colleagues. Power decreases during both perception and retrieval negatively correlate with the
amount of stimulus specific information present on that trial.
task irrelevant information (Jensen & Mazaheri, 2010; Klimesch et al., 2007). It therefore
attributes an operative function to alpha synchrony. The information via desynchron-
ization hypothesis goes beyond the inhibition theory, in that it highlights the active role
of power decreases for the processing of information. Therein, power decreases are not
a mere absence of inhibition, but rather functionally involved in neural computations.
The crucial insight is that in order to process complex information or represent informa-
tion rich content, synchronous neural activity does not provide enough coding space.
Desynchronous neural activity on the other hand, which is marked by power decreases
in the alpha/beta band, provides the required coding space through locally decoupled
neural assemblies. The information via desynchronization theory therefore stresses
that, in order to perform cognitive operations that work on complex and information-
rich content, alpha/beta power must decrease.
Another way to interpret the inverse relation between alpha power and information
representation is offered by studies investigating “noise-correlations”, which refers to
correlated firing of task-irrelevant neurons—a process that can be detrimental to in-
formation representation (Mitchell et al., 2009). If two task-irrelevant neurons fire
together, their noise is amplified. Expand this principle to hundreds or thousands of
neurons and their noise becomes deafening. In such quantities, these noise correlations
THE ROLE OF ALPHA AND BETA OSCILLATIONS IN THE HUMAN EEG 211
mask the signal generated by neurons critical to the task at hand (Averbeck et al., 2006;
see also Figure 10.2a), leading to an impaired ability in processing and representing
information. To rectify this situation, the magnitude of noise correlations needs to be
attenuated. How, though, do noise correlations relate to alpha oscillations? Given that
the summed electric potential of the correlated neurons creates a spike in the amplitude
of local field potential (Averbeck et al., 2006), repeated and rhythmic patterns of noise
correlations would create repeated and rhythmic increases in the amplitude of local field
potential (LFP). As alpha oscillations dominate the neocortex, one may speculate that
the rhythmic noise correlations may resonate within this frequency band. Under this
assumption, periods of high-amplitude alpha oscillations would reflect periods of nu-
merous noise correlations where information processing is inhibited. This interpret-
ation conforms with the active inhibition account because it would allow alpha power
increases to suppress task-irrelevant information. Periods of alpha desynchrony, in con-
trast, would reflect periods of limited noise correlations, where there is a greatest poten-
tial for information representation.
In summary, there are two theoretical arguments that implicate alpha power
decreases in the representation of stimulus-specific information. The information-
via-desynchronization account suggests that alpha power reductions facilitate the
evolution of more complex neural codes, which then allows for the representation of
high-fidelity information in the cortex. The noise correlation account suggests that
alpha power decreases reduce the background noise in the cortex, allowing for key
signals to be more clearly communicated. Currently, empirical evidence supports both
ideas by demonstrating that alpha power decreases scale with the quantity of informa-
tion present in the cortex.
Recent evidence suggests that the continuous input that our brain receives from the
outside world is not sampled continuously, but in discrete rhythmic steps around the
alpha frequency (VanRullen, 2016). These studies show that the probability of detecting
a briefly presented visual stimulus fluctuates rhythmically (VanRullen et al., 2007).
This attentional sampling process has been suggested to operate at a frequency of
roughly 8 Hz (Landau & Fries, 2012), even when sustained attention to the same ob-
ject is maintained (Fiebelkorn & Kastner, 2019; Fiebelkorn et al., 2013). Overall, there
appears to be a close correlation between the phase of an oscillation in the lower alpha
band (around 8 Hz) and the perception of an incoming stimulus. In the perception of
continuous and dynamic stimuli, the stimulus identity can be reliably decoded from
the ongoing phase of neural oscillations. Therein, an individual continuous stimulus
is associated with a unique time course of neural activity. Specifically, the stimulus
212 SEBASTIAN MICHELMANN et al.
entrains the cortical rhythm in a content-specific way that organizes neural firing (Ng
et al., 2013). Interestingly, the phase of the ongoing oscillation contains more informa-
tion about stimulus identity than its power (Schyns et al., 2011).
Considering this information sampling role of alpha phase in perception, an immi-
nent question is whether or not the reinstatement of information from memory relies
on the replay of that information sampled in the alpha frequency. To this end, alpha
phase could index similar mechanisms during attention and memory, just like alpha
power decreases. But how can we measure the information sampling and the replay
thereof in the phase of a recorded EEG?
Similar to assessing the similarity of neural patterns in space, as is usually done in
fMRI, one can assess the similarity of neural patterns in phase (over time) via the use
of RSA methods (discussed earlier). Established measures of connectivity can help
quantify the similarity of neural patterns in the phase of the alpha-frequency band
(Greenblatt et al., 2012). Connectivity measures usually assess the similarity between
different channels in their time course of activity. This typically measures shared infor-
mation between regions. Yet, measures of connectivity will lend themselves perfectly to
quantify representational similarity between encoding and retrieval (Figure 10.3a).
In a first study, we applied such similarity measures of phase in a human EEG experi-
ment where subjects were instructed to encode and replay dynamic visual (short movie
clips) or auditory stimuli (short melodies played by different instruments). Indeed,
content-specific patterns of oscillatory phase in the lower alpha band during percep-
tion represent the identity of the visual or auditory stimuli in the visual and auditory
cortex, respectively. Strikingly, these phase patterns reappear when representations of
short video clips and short sound clips are replayed from memory (Michelmann et al.,
2016). Importantly, this replay takes place in the absence of the dynamic stimulus itself
and can be localized in sensory specific areas (Figure 10.3c). A recent study replicated
this effect in the visual domain and demonstrated that the reinstatement of temporal
patterns is only observed when content is successfully recalled, that is, temporal pattern
reinstatement is implicated in successful memory (Michelmann et al., 2019). Another
study replicated this effect during sleep, indicating that replay of stimulus-specific phase
patterns supports memory consolidation (Schreiner et al., 2018). Further evidence
documents that such content-specific phase patterns are also replayed when an asso-
ciation with a previously shown dynamic stimulus is formed (Michelmann et al., 2018).
Interestingly, the reinstatement of temporal patterns is not always beneficial for
memory but can in some cases interfere with memory as demonstrated by studies
which manipulated the contextual overlap between encoding and retrieval. For in-
stance, (Staudigl et al., 2015) manipulated the context in which a word is learned and
remembered by playing the same video clip in the background behind the word at
encoding and at retrieval (context match); or playing a different video clip in the back-
ground behind the word at encoding and at retrieval (context mismatch). The import-
ance of reinstatement of contextual information was observed via temporal and spatial
pattern similarity in the beta frequency band. Specifically, higher pattern similarity was
associated with better memory performance when contexts where matching. On the
THE ROLE OF ALPHA AND BETA OSCILLATIONS IN THE HUMAN EEG 213
(a)
μV
μV
stimulus onset cue onset cue onset response
time (t) time (t)
PI PI
phase
phase
angle
angle
0 0
–PI –PI
stimulus onset cue onset cue onset response
time (t) time (t)
sliding window
Phase similarity
(b) (c)
40
2
30
frequency [Hz]
T-values
20 0
0 –1 –2 –3 –4 –5
10 –2
t-values
0 1 2 3 4
time [seconds]
Figure 10.3 Alpha power decreases during memory retrieval code stimulus-specific infor-
mation. (A) Subjects first encoded a video (left). They later retrieved a vivid representation of
that video from memory (right). The phase time course was extracted from the EEG during
encoding and retrieval in order to calculate a similarity measure between encoding and retrieval.
(B) During retrieval, strong and sustained alpha power decreases were observed. (C) Reactivation
of stimulus-specific information, as measured with phase similarity, could be detected in the
alpha frequency band with a maximum in parietal regions. This reactivation was localized in par-
ietal cortex (C, left).
Reproduced from Michelmann et al., 2016.
of information during perception (VanRullen et al., 2007) and the replay of that infor-
mation during memory reinstatement.
Michelmann and colleagues (2016) showed that the frequency band that contains
content-specific temporal patterns is also the one that displays the most prominent
power decreases during retrieval (Figure 10.3b), and furthermore, they observed an
interaction such that sensory areas that were involved in the reinstatement of auditory
and visual temporal patterns also expressed stronger power decreases in the respective
condition. This suggests that power decreases and the representation of information in
oscillatory phase are not two separate processes but rather are intertwined. This raises
the question as to how the two signal properties of alpha—power and phase—interact in
the service of information representation.
We suggest that a decrease of power in an ongoing oscillation renders a signal less
stationary (and therefore also less predictable) and thereby allows for a flexible adjust-
ment of phase (Hanslmayr et al., 2016). These phase adjustments, or deviations from
stationarity, make it possible that time courses in phase can represent stimulus spe-
cific patterns, and to replay these patterns from memory (Figure 10.4C). From an in-
formation theoretic view, a stationary oscillation without phase adjustments wouldn’t
(a) Predictable Signal (High Power) (b) Phase per Period (c)
(every 100 ms)
Stimulus A
Stimulus B
Stimulus D
Figure 10.4 Information coding properties for stationary and non- stationary signals is
illustrated. (A) A stationary (high power) signal (orange) is shown together with its phase (blue)
in the upper row. The lower row shows a less stationary signal. (B) Phases for each signal are
shown every 100 ms (indicated by ticks in A). (C) Non-stationary signals allow for stimulus spe-
cific coding by assigning a different time course to each stimulus.
THE ROLE OF ALPHA AND BETA OSCILLATIONS IN THE HUMAN EEG 215
be able to represent much information in its phase because the time course of a sta-
tionary signal is perfectly predictable once the phase at one time point is known. For
example, the signal in the top row of Figure 10.4A visits the same phase every 100 ms
(top row in Figure 10.4B). Since information theory quantifies information as the in-
verse of the predictability (i.e., negative logarithm in the case of Shannon’s Entropy, see
Shannon & Weaver, 1949) we can infer that this signal has little potential to carry infor-
mation. Indeed, if we were to code the identity of a stimulus in such a perfectly predict-
able signal, we would not be able to distinguish between different stimuli. In contrast,
a non-stationary signal (lower row of Figure 10.4A), which has phase modulations, can
carry much more information. In this case, the phase cannot be predicted from pre-
vious time points (lower panel in Figure 10.4B). This allows us to code different stimuli
by assigning a phase time course to each stimulus (Figure 10.4C). This simple relation-
ship between the power of a signal and predictability of phase time courses elegantly
unifies the findings described in this chapter. To this end, a signal with high alpha power
leads to a more stationary time course and thus inhibits information coding. In con-
trast, a signal with lower alpha power allows for a less stationary time course and there-
fore the coding of information. Importantly, this idea is in line with the general notion
of an inhibitory role of alpha power increases but goes beyond the previous work in
ascribing specific computational roles to power and phase in the service of representing
information.
This chapter showed that alpha oscillations are ubiquitous as they are modulated by al-
most any task and can be observed in almost any animal. It therefore follows that alpha
oscillations must perform a basic neural operation, which is of service for many cognitive
operations. After an overview of the idling vs. inhibition hypothesis we then discussed
the computational utility of alpha power reductions for information representation. We
reviewed studies that demonstrated that alpha power reductions are intimately linked to
information coding—specifically, these studies showed that stimulus-specific informa-
tion is coded in the phase of alpha. Importantly, several studies demonstrated that this
phase information is replayed when a reminder to that stimulus is presented. Finally,
we illustrated how power decreases allow for less stationary phase time courses and
consequently for information representation. Together, we conclude that alpha power
decreases do play an important role for information representation—a neural operation
needed in almost any task and in almost any animal.
An important crucial future question refers to the nature of the relationship between
alpha power decreases and its role for coding information. If indeed, as we here suggest,
alpha power decreases are mechanisms for representing information we should be
able to manipulate perception, maintenance, and retrieval of information by directly
manipulating alpha oscillations via brain stimulation techniques (Hanslmayr et al.,
216 SEBASTIAN MICHELMANN et al.
2019). Such a demonstration of a causal relationship between alpha power decreases and
the representation of information is crucial in order to show that alpha oscillations are
indeed a mechanism for information representation, instead of a mere epiphenomenon.
References
Adrian, E. D. (1944). Brain rhythms. Nature, 153, 360–362.
Adrian, E. D. & Matthews, B. H. (1934). The interpretation of potential waves in the cortex.
Journal of Physiology, 81(4), 440–471.
Anokhin, A. & Vogel, F. (1996). EEG alpha rhythm frequency and intelligence in normal adults.
Intelligence, 23(1), 1–14. doi:10.1016/S0160-2896(96)80002-X
Averbeck, B. B., Latham, P. E., & Pouget, A. (2006). Neural correlations, population coding and
computation. Nature Reviews Neuroscience, 7(5), 358–366. doi:10.1038/nrn1888
Berger, H. (1929). Über das Elektrenkephalogramm des Menschen. Archiv für Psychiatrie und
Nervenkrankheiten, 87, 527–570.
Bonnefond, M. & Jensen, O. (2012). Alpha oscillations serve to protect working memory main-
tenance against anticipated distracters. Current Biology, 22(20), 1969–1974. doi:10.1016/
j.cub.2012.08.029
Burgess, A. P. & Gruzelier, J. H. (2000). Short duration power changes in the EEG during rec-
ognition memory for words and faces. Psychophysiology, 37(5), 596–606.
Buzsaki, G. (2019). The Brain from inside out. Oxford University Press.
Cohen, M. X. (2011). Hippocampal-prefrontal connectivity predicts midfrontal oscillations
and long-term memory performance. Current Biology, 21(22), 1900–1905. doi:10.1016/
j.cub.2011.09.036
Cohen, M. X. (2014). Analyzing neural time series data: Theory and practice. MIT Press.
Cohen, M. X. (2014). Fluctuations in oscillation frequency control spike timing and co-
ordinate neural networks. Journal of Neuroscience, 34(27), 8988– 8998. doi:10.1523/
JNEUROSCI.0261-14.2014
Crone, N. E., Miglioretti, D. L., Gordon, B., Sieracki, J. M., Wilson, M. T., Uematsu, S., & Lesser, R.
P. (1998). Functional mapping of human sensorimotor cortex with electrocorticographic spec-
tral analysis. I. Alpha and beta event-related desynchronization. Brain, 121 (Pt 12), 2271–2299.
Fellner, M. C., Gollwitzer, S., Rampp, S., Kreiselmeyr, G., Bush, D., Diehl, B., . . . Hanslmayr,
S. (2018). Spectral fingerprints or spectral tilt? Evidence for distinct oscillatory signatures
of memory formation. https://journals.plos.org/plosbiology/article?id=10.1371/journal.
pbio.3000403
Fiebelkorn, I. C. & Kastner, S. (2019). A rhythmic theory of attention. Trends in Cognitive
Sciences, 23(2), 87–101. doi:10.1016/j.tics.2018.11.009
Fiebelkorn, I. C., Saalmann, Y. B., & Kastner, S. (2013). Rhythmic sampling within and between
objects despite sustained attention at a cued location. Current Biology, 23(24), 2553–2558.
doi:10.1016/j.cub.2013.10.063
Fox, M. D., Snyder, A. Z., Vincent, J. L., Corbetta, M., Van Essen, D. C., & Raichle, M. E.
(2005). The human brain is intrinsically organized into dynamic, anticorrelated functional
networks. Proceedings of the National Academy of Sciences of the United States of America,
102(27), 9673–9678. doi:10.1073/pnas.0504136102
Greenblatt, R. E., Pflieger, M. E., & Ossadtchi, A. E. (2012). Connectivity measures applied
to human brain electrophysiological data. Journal of Neuroscience Methods, 207(1), 1–16.
doi:10.1016/j.jneumeth.2012.02.025
THE ROLE OF ALPHA AND BETA OSCILLATIONS IN THE HUMAN EEG 217
Griffiths, B. J., Mayhew, S. D., Mullinger, K. J., Jorge, J., Charest, I., Wimber, M., & Hanslmayr,
S. (2019). Alpha/beta power decreases track the fidelity of stimulus-specific information.
https://elifesciences.org/articles/49562
Haegens, S., Cousijn, H., Wallis, G., Harrison, P. J., & Nobre, A. C. (2014). Inter-and intra-
individual variability in alpha peak frequency. NeuroImage, 92, 46– 55. doi:10.1016/
j.neuroimage.2014.01.049
Haegens, S., Nacher, V., Luna, R., Romo, R., & Jensen, O. (2011). alpha-Oscillations in the
monkey sensorimotor network influence discrimination performance by rhythmical inhib-
ition of neuronal spiking. Proceedings of the National Academy of Sciences of the United States
of America, 108(48), 19377–19382. doi:10.1073/pnas.1117190108
Hanslmayr, S., Axmacher, N., & Inman, C. S. (2019). Modulating human memory via entrain-
ment of brain oscillations. Trends in Neuroscience, 42(7), 485–499.
Hanslmayr, S., Gross, J., Klimesch, W., & Shapiro, K. L. (2011). The role of α oscillations in tem-
poral attention. Brain Research Reviews, 67(1–2), 331–343.
Hanslmayr, S., Spitzer, B., & Bauml, K. H. (2009). Brain oscillations dissociate between se-
mantic and nonsemantic encoding of episodic memories. Cerebral Cortex, 19(7), 1631–1640.
doi:10.1093/cercor/bhn197
Hanslmayr, S., Staresina, B. P., & Bowman, H. (2016). Oscillations and episodic
memory: Addressing the synchronization/ desynchronization conundrum. Trends in
Neuroscience, 39(1), 16–25. doi:10.1016/j.tins.2015.11.004
Hanslmayr, S., Staudigl, T., & Fellner, M. C. (2012). Oscillatory power decreases and long-
term memory: The information via desynchronization hypothesis. Frontiers in Human
Neuroscience, 6, 74. doi:10.3389/fnhum.2012.00074
Jacobs, J., Kahana, M. J., Ekstrom, A. D., & Fried, I. (2007). Brain oscillations control timing of
single-neuron activity in humans. Journal of Neuroscience, 27(14), 3839–3844.
Jensen, O., Gelfand, J., Kounios, J., & Lisman, J. E. (2002). Oscillations in the alpha band (9-
12 Hz) increase with memory load during retention in a short-term memory task. Cerebral
Cortex, 12(8), 877–882.
Jensen, O. & Mazaheri, A. (2010). Shaping functional architecture by oscillatory alpha
activity: gating by inhibition. Frontiers in Human Neuroscience, 4, 186. doi:10.3389/
fnhum.2010.00186
Keitel, A. & Gross, J. (2016). Individual human brain areas can be identified from their charac-
teristic spectral activation fingerprints. PLoS Biology, 14(6), e1002498. doi:10.1371/journal.
pbio.1002498
Kerren, C., Linde-Domingo, J., Hanslmayr, S., & Wimber, M. (2018). An optimal oscillatory
phase for pattern reactivation during memory retrieval. Current Biology, 28(21), 3383–3392
e3386. doi:10.1016/j.cub.2018.08.065
Klimesch, W., Doppelmayr, M., Schwaiger, J., Auinger, P., & Winkler, T. (1999). ‘Paradoxical’
alpha synchronization in a memory task. Brain Research: Cognitive Brain Research, 7(4),
493–501.
Klimesch, W., Sauseng, P., & Hanslmayr, S. (2007). EEG alpha oscillations: the inhibition-timing
hypothesis. Brain Research Reviews, 53(1), 63–88. doi:10.1016/j.brainresrev.2006.06.003
Klimesch, W., Schimke, H., Doppelmayr, M., Ripper, B., Schwaiger, J., & Pfurtscheller, G.
(1996). Event-related desynchronization (ERD) and the Dm effect: Does alpha desyn-
chronization during encoding predict later recall performance? International Journal of
Psychophysiology, 24(1-2), 47–60.
Krause, C. M., Lang, H. A., Laine, M., Helle, S. I., Kuusisto, M. J., & Porn, B. (1994). Event-
related desynchronization evoked by auditory stimuli. Brain Topography, 7(2), 107–112.
218 SEBASTIAN MICHELMANN et al.
Kriegeskorte, N., Mur, M., & Bandettini, P. (2008). Representational similarity analysis -
connecting the branches of systems neuroscience. Frontiers in Systems Neuroscience, 2, 4.
doi:10.3389/neuro.06.004.2008
Landau, A. N. & Fries, P. (2012). Attention samples stimuli rhythmically. Current Biology,
22(11), 1000–1004. doi:10.1016/j.cub.2012.03.054
Lopes da Silva, F. H., Vos, J. E., Mooibroek, J., & Van Rotterdam, A. (1980). Relative
contributions of intracortical and thalamo-cortical processes in the generation of alpha
rhythms, revealed by partial coherence analysis. Electroencephalography and Clinical
Neurophysiology, 50(5-6), 449–456.
Michelmann, S., Bowman, H., & Hanslmayr, S. (2016). The temporal signature of
memories: Identification of a general mechanism for dynamic memory replay in humans.
PLoS Biology, 14(8), e1002528. doi:10.1371/journal.pbio.1002528
Michelmann, S., Bowman, H., & Hanslmayr, S. (2018). Replay of stimulus-specific temporal
patterns during associative memory formation. Journal of Cognitive Neuroscience, 30(11),
1577–1589. doi:10.1162/jocn_a_01304
Michelmann, S., Staresina, B. P., Bowman, H., & Hanslmayr, S. (2019). Speed of time-
compressed forward replay flexibly changes in human episodic memory. Nature Human
Behaviour, 3(2), 143–154. doi:10.1038/s41562-018-0491-4
Mitchell, J. F., Sundberg, K. A., & Reynolds, J. H. (2009). Spatial attention decorrelates intrinsic ac-
tivity fluctuations in macaque area V4. Neuron, 63(6), 879–888. doi:10.1016/j.neuron.2009.09.013
Murthy, V. N. & Fetz, E. E. (1996). Synchronization of neurons during local field potential
oscillations in sensorimotor cortex of awake monkeys. Journal of Neurophysiology, 76(6),
3968–3982. doi:10.1152/jn.1996.76.6.3968
Ng, B. S., Logothetis, N. K., & Kayser, C. (2013). EEG phase patterns reflect the selectivity of
neural firing. Cerebral Cortex, 23(2), 389–398. doi:10.1093/cercor/bhs031
Obleser, J. & Weisz, N. (2012). Suppressed alpha oscillations predict intelligibility of speech and
its acoustic details. Cerebral Cortex, 22(11), 2466–2477. doi:10.1093/cercor/bhr325
Pfurtscheller, G. & Aranibar, A. (1977). Event-related cortical desynchronization detected by
power measurements of scalp EEG. Electroencephalography and Clinical Neurophysiology,
42(6), 817–826.
Pfurtscheller, G., Neuper, C., Andrew, C., & Edlinger, G. (1997). Foot and hand area mu
rhythms. International Journal of Psychophysiology, 26(1–3), 121–135.
Pfurtscheller, G., Stancak, A., Jr., & Neuper, C. (1996). Event-related synchronization (ERS) in
the alpha band--an electrophysiological correlate of cortical idling: a review. International
Journal of Psychophysiology, 24(1-2), 39–46.
Poldrack, R. A. (2011). Inferring mental states from neuroimaging data: from reverse inference
to large-scale decoding. Neuron, 72(5), 692–697. doi:10.1016/j.neuron.2011.11.001
Popov, T., & Szyszka, P. (2019). Alpha oscillations govern interhemispheric spike timing
coordination in the honeybee brain. https://royalsocietypublishing.org/doi/10.1098/
rspb.2020.0115#:~:text=What%20is%20the%20function%20of,information%20
flow%20during%20spontaneous%20activity.
Pornpattananangkul, N., Grogans, S., Yu, R., & Nusslock, R. (2019). Single-trial EEG dissociates
motivation and conflict processes during decision-making under risk. NeuroImage, 188,
483–501. doi:10.1016/j.neuroimage.2018.12.029
Sauseng, P., Klimesch, W., Heise, K. F., Gruber, W. R., Holz, E., Karim, A. A., . . . Hummel, F. C.
(2009). Brain oscillatory substrates of visual short-term memory capacity. Current Biology,
19(21), 1846–1852. doi:10.1016/j.cub.2009.08.062
THE ROLE OF ALPHA AND BETA OSCILLATIONS IN THE HUMAN EEG 219
Schneidman, E., Puchalla, J. L., Segev, R., Harris, R. A., Bialek, W., & Berry, M. J., 2nd. (2011).
Synergy from silence in a combinatorial neural code. Journal of Neuroscience, 31(44), 15732–
15741. doi:10.1523/JNEUROSCI.0301-09.2011
Schreiner, T., Doeller, C. F., Jensen, O., Rasch, B., & Staudigl, T. (2018). Theta phase-coordinated
memory reactivation reoccurs in a slow-oscillatory rhythm during NREM sleep. Cell
Reports, 25(2), 296–301. doi:10.1016/j.celrep.2018.09.037
Schyns, P. G., Thut, G., & Gross, J. (2011). Cracking the code of oscillatory activity. PLoS Biology,
9(5), e1001064. doi:10.1371/journal.pbio.1001064
Shannon, C. E. & Weaver, W. (1949). A mathematical theory of communication. University of
Illinois Press.
Siegel, M., Donner, T. H., & Engel, A. K. (2012). Spectral fingerprints of large-scale neuronal
interactions. Nature Reviews Neuroscience, 13(2), 121–134. doi:10.1038/nrn3137
Staudigl, T. & Hanslmayr, S. (2019). Reactivation of neural patterns during memory reinstate-
ment supports encoding specificity. Cognitive Neuroscience, 10(4), 175–185.
Staudigl, T., Vollmar, C., Noachtar, S., & Hanslmayr, S. (2015). Temporal-pattern similarity
analysis reveals the beneficial and detrimental effects of context reinstatement on human
memory. Journal of Neuroscience, 35(13), 5373–5384. doi:10.1523/JNEUROSCI.4198-14.2015
Thut, G., Nietzel, A., Brandt, S. A., & Pascual- Leone, A. (2006). Alpha- band
electroencephalographic activity over occipital cortex indexes visuospatial attention bias
and predicts visual target detection. Journal of Neuroscience, 26(37), 9494–9502. doi:10.1523/
JNEUROSCI.0875-06.2006
VanRullen, R. (2016). Perceptual cycles. Trends in Cognitive Sciences, 20(10), 723–735.
doi:10.1016/j.tics.2016.07.006
VanRullen, R., Carlson, T., & Cavanagh, P. (2007). The blinking spotlight of attention.
Proceedings of the National Academy of Sciences of the United States of America, 104(49),
19204–19209. doi:10.1073/pnas.0707316104
von Stein, A., Chiang, C., & Konig, P. (2000). Top-down processing mediated by interareal syn-
chronization. Proceedings of the National Academy of Sciences of the United States of America,
97(26), 14748–14753. doi:10.1073/pnas.97.26.14748
von Stein, A. & Sarnthein, J. (2000). Different frequencies for different scales of cortical inte-
gration: from local gamma to long range alpha/theta synchronization. International Journal
of Psychophysiology, 38(3), 301–313.
Waldhauser, G. T., Braun, V., & Hanslmayr, S. (2016). Episodic memory retrieval functionally
relies on very rapid reactivation of sensory information. Journal of Neuroscience, 36(1), 251–
260. doi:10.1523/JNEUROSCI.2101-15.2016
Waldhauser, G. T., Johansson, M., & Hanslmayr, S. (2012). alpha/beta oscillations indi-
cate inhibition of interfering visual memories. Journal of Neuroscience, 32(6), 1953–1961.
doi:10.1523/JNEUROSCI.4201-11.2012
Wiest, M. C. & Nicolelis, M. A. (2003). Behavioral detection of tactile stimuli during 7-12 Hz
cortical oscillations in awake rats. Nature Neuroscience, 6(9), 913–914. doi:10.1038/nn1107
Womelsdorf, T., Valiante, T. A., Sahin, N. T., Miller, K. J., & Tiesinga, P. (2014). Dynamic circuit
motifs underlying rhythmic gain control, gating and integration. Nature Neuroscience, 17(8),
1031–1039. doi:10.1038/nn.3764
Worden, M. S., Foxe, J. J., Wang, N., & Simpson, G. V. (2000). Anticipatory biasing of visuo-
spatial attention indexed by retinotopically specific alpha-band electroencephalography
increases over occipital cortex. Journal of Neuroscience, 20(6), RC63.
CHAPTER 11
T heory and Re se a rc h
on Asymm etri c Fronta l
C ortical Ac t i v i t y
as Assesse d by E E G
F requ ency A na lyse s
Beginning in the 1930s, research using a variety of methods suggested that the left and
right frontal cortices are involved in different emotional (affective) responses. For ex-
ample, lesion studies as well as experiments that injected a barbiturate derivative
(amytal) into one of the internal carotid arteries (to suppress the activity of one hemi-
sphere) showed that the loss of activity in the left frontal cortex was associated with
depressed affect, whereas loss of activity in the right frontal cortex was associated with
manic affect and euphoria (Alema et al., 1961; Black, 1975; Gainotti, 1972; Gasparrini
et al., 1978; Goldstein, 1939; Perria et al., 1961; Robinson & Price, 1982; Rossi & Rosadini,
1967; Sackeim et al., 1982; Terzian & Cecotto, 1959). These outcomes are likely due to
the release of one hemisphere from contralateral inhibitory forces (Schutter & Harmon-
Jones, 2013). For example, when the right hemisphere was deactivated by damage or
amytal, the left hemisphere became more uninhibited and more active, which caused
manic affect.
222 EDDIE HARMON-JONES, TAYLOR POPP, AND PHILIP A. GABLE
These results can be interpreted in one of two ways. According to the affective va-
lence model of frontal asymmetry, the left frontal cortical region is involved in the ex-
perience and expression of positive affect, whereas the right frontal cortical region is
involved in the experience and expression of negative affect. According to the motiv-
ational direction model of frontal asymmetry, the left frontal cortical region is involved
in the experience and expression of approach motivation, whereas the right frontal cor-
tical region is involved in the experience and expression of withdrawal motivation. Until
the 1990s, the two conceptual models were considered to yield the same predictions be-
cause approach motivation was conceived of as being associated with positive affect and
withdrawal motivation was conceived of as being associated with negative affect. In the
late 1990s, researchers began using anger to tease predictions from these two conceptual
models apart, because anger is an approach-motivated but negative affect (Harmon-
Jones, 2003a). The following sections briefly review the research on the two models
along with research on anger and frontal asymmetry.
These early results inspired EEG researchers to examine EEG alpha power over the
frontal cortex and test whether it was associated with affective variables. Most of the first
EEG studies on asymmetric frontal asymmetry tested individuals in a resting, baseline
state (i.e., sitting quietly in the lab for four to eight minutes). Researchers assumed that
this resting, baseline EEG would tap into a personality trait, and they then related EEG
frontal asymmetry with other personality or individual difference measures.
Several studies found that depression was correlated with less relative left frontal
cortical activity (e.g., Allen et al., 1993; Henriques & Davidson, 1990; Jacobs & Snyder,
1996; Schaffer et al., 1983; see meta-analysis by Thibodeau et al., 2006). Other studies
suggested that trait negative affect was correlated with greater relative right frontal ac-
tivity, whereas trait positive affect was correlated with greater relative left frontal activity
(Tomarken et al., 1992).
The studies assessing resting baseline asymmetric frontal cortical activity were correl-
ational; therefore, it is impossible to infer that the differences in asymmetric frontal
cortical activity were causally involved in the observed affective processes. In an effort
ASYMMETRIC FRONTAL CORTICAL ACTIVITY 223
to provide evidence that would allow causal inferences, some researchers manipulated
asymmetric frontal cortical activity and tested whether manipulation influenced other
affective variables.
11.4.1 Neurofeedback
Studies have used neurofeedback training to manipulate asymmetric frontal cortical ac-
tivity (e.g., Allen et al., 2001; Harmon-Jones et al., 2008). In this research, operant con-
ditioning is used to produce certain patterns of brainwaves. Reward feedback (e.g., a
simple tone) corresponding to “desired” patterns of brainwave activity is presented to
participants. Over numerous presentations of this reward feedback, the brain learns
to produce the desired brainwave activity, such as greater relative left frontal cortical
activity. These neurofeedback-induced changes in brainwaves occur via nonconscious
learning processes (Kamiya, 1979; Siniatchkin et al., 2000).
In one example experiment, relative right versus relative left frontal activity was
manipulated using neurofeedback (Allen et al., 2001). In the experiment, participants
were instructed to attempt to make a particular tone play as much as possible.
Asymmetric frontal cortical activity (i.e., alpha power at F4 minus alpha power at F3)
was calculated during the first second of each two-second time period. This calculated
value was then compared against a criterion value that had been set for that session
(based on the individual’s previous asymmetry index). If the calculated value was larger
than the criterion value in the desired direction, a “reward” tone was played; if it was
not, a “non-reward” tone was played. After several days of neurofeedback training,
participants then viewed emotionally evocative film clips as zygomatic and corrugator
muscle region activity was recorded. Results revealed that the neurofeedback training
influenced asymmetric frontal cortical activity in the predicted direction (for a repli-
cation, see Quaedflieg et al., 2016). Moreover, the manipulated increase in relative right
frontal cortical activity, as compared to relative left frontal cortical activity, caused less
zygomatic and more corrugator muscle region activity in response to all film clips. These
results suggest that asymmetric frontal cortical activity is causally involved in emotional
responses.
contractions of the right hand cause increased feelings of positive affect and more posi-
tive perceptions and judgments (Schiff & Lamon, 1989, 1994).
Based on this research, experiments have tested whether unilateral hand contractions
would influence EEG and other emotional responses. In one experiment, participants
were instructed to squeeze a ball in their left or right hand for several minutes (Harmon-
Jones, 2006). Then, their self-reported affective responses were measured in response
to a radio editorial designed to evoke a moderate amount of positive affect. Results
revealed that the contraction of one hand increased cortical activity over the central and
frontal regions of the contralateral hemisphere (as measured by EEG alpha suppression;
see Gable et al., 2013 for replication). Moreover, this manipulation influenced positive
affect, such that right-hand contractions caused more positive affect than the left-hand
contractions.
Several studies suggest that the left frontal cortical region is involved in the experience
and expression of positive affect, whereas the right frontal cortical region is involved
in the experience and expression of negative affect. This research suggested that asym-
metric frontal cortical activity reflected positive or negative affect, or affective valence.
Research testing this affective valence model happened concurrently with other re-
search testing a motivational direction model, which posits that relative left frontal
cortical activity is associated with approach motivation and that relative right frontal
cortical activity is associated with withdrawal motivation. For several years, the affective
ASYMMETRIC FRONTAL CORTICAL ACTIVITY 225
valence and motivational direction models were viewed as compatible models that
made identical predictions. This compatibility resulted from viewing positive affect as
being always associated with approach motivation and negative affect as being always
associated with withdrawal motivation.
in relative left frontal cortical activity. This lack of a main effect of picture type on asym-
metric frontal activity is consistent with some other research (Elgavish et al., 2003;
Hagemann et al., 1998; see also reviews by Murphy et al., 2003 and Pizzagalli et al.,
2003). The lack of a main effect may have occurred because pictures evoke different
amounts of approach (withdrawal) motivation across participants. Some individuals
may respond with no motivation, and others may respond with much motivation
(Harmon-Jones, 2007).
Other research has suggested that some pictorial stimuli may evoke intense approach
(or withdrawal) motivation in all participants. For example, Schöne and colleagues
(2016) found that erotic stimuli caused greater relative left frontal activity than com-
parison stimuli of extreme sports, dressed women, and daily activities. The observed
difference in relative left frontal activity between erotic and extreme sports stimuli is
particularly interesting because both sets of stimuli evoked high and equal levels of self-
reported positive affect and arousal.
assessed behavioral persistence. As predicted by the motivational model (but not the
affective valence model), individuals who made the determination facial expression
(high approach-motivated positive affect) had greater relative left frontal activity than
individuals in the other two conditions. In addition, within the determination condi-
tion, greater relative left frontal cortical activity was associated with greater behavioral
persistence.
11.5.2.5 Final Thoughts
The reviewed evidence suggests that relative left frontal cortical activity is associated
with approach motivation. This association occurs even when approach motivation
is not confounded with positive affect. That is, in approach motivation, some positive
affects are low and some are high, but the research has revealed that it is primarily the
high approach-motivated positive affects that are associated with greater relative left
frontal activity.
Carver & Harmon-Jones, 2009; Harmon-Jones et al., 2009; Harmon-Jones et al., 2013).
To give just a few examples, research has revealed that individual differences in BAS
(Carver & White, 1994) are positively correlated with individual differences in anger
(Harmon-Jones, 2003b; Smits & Kuppens, 2005), greater anger responses to situational
anger manipulations (Carver, 2004), greater aggressive inclinations especially when
approach motivation is activated (Harmon-Jones & Peterson, 2008), and more atten-
tional engagement to angry faces, as in approach-based dominance confrontations
(Putman et al., 2004).
If asymmetric frontal cortical activity reflects affective valence, then anger should
be associated with relative right frontal activity because anger is a negative affect. In
contrast, if asymmetric frontal cortical activity reflects motivational direction, then
anger should be associated with relative left frontal activity because anger is approach
motivated. Research has tested these competing predictions.
to the other participants on trials they won. Replicating previous research, the con-
traction of the right hand led to greater relative left central and frontal cortical activity
than the contraction of left hand. More importantly, as compared to the left hand
contractions, the right hand contractions led to more aggression behavior (Peterson
et al., 2008).
Additional research revealed that compared to left- hand contractions, right-
hand contractions led to greater self-reported anger to being socially ostracized in
a Cyberball game (Peterson et al., 2011). These results suggest that manipulated
increases in relative left frontal cortical activity lead to increased anger-related
responses.
association between approach motivation and relative left frontal cortical activity
could be provided by manipulating approach motivation independent of anger. That is,
even though anger is often approach motivated, anger is not perfectly associated with
approach motivation and thus it should be possible to manipulate the two constructs
separately.
For example, one variable that should influence (approach) motivational intensity is
coping potential or the difficulty of engaging in behavior. That is, motivational intensity
increases with perceived task difficulty up to the point where the task is perceived as im-
possible, and then motivational intensity drops. So, when a task is perceived as impos-
sible, or coping potential is very low, motivational intensity should be low (Brehm, 1999;
Brehm & Self, 1989).
In one experiment testing these hypotheses, individuals were induced to believe
that they would or would not be able to engage in action that might resolve the anger-
evoking event (i.e., sign petitions to halt a university tuition increase; Harmon-Jones
et al., 2003). Results revealed that individuals who were angered and believed they could
engage in action had greater relative left frontal activity than individuals who were
angered and believed they could not engage in action (the tuition increase had already
been approved). Moreover, in the action-possible condition, greater relative left frontal
activity in response to the angering event correlated with more approach-related be-
havior (i.e., signing the petition and taking more petitions to have others sign). Other
studies have conceptually replicated this effect of anger and approach action possibility
on relative left frontal activity using pictorial stimuli to evoke anger (Harmon-Jones
et al., 2006).
These results could be interpreted to indicate that greater relative left frontal cor-
tical activity only occurs to anger evocations when individuals are given explicit
approach motivational opportunities. Other research suggests, however, that these
explicit approach motivational opportunities increase relative left frontal activity but
are not necessary for it to occur. That is, individuals who were exposed to pictures
that evoked anger and were given no explicit approach opportunities had increased
relative left frontal cortical activity if they were high in trait anger (Harmon-Jones,
2007). Thus, explicit opportunities for approach-motivated behavior are not neces-
sary to cause relative left frontal activity during anger. Anger can evoke approach
motivation without explicit approach motivational opportunities being immediately
present.
Other experiments have manipulated approach motivation independently of anger
and found that the approach motivation drives the increase in relative left frontal ac-
tivity. For example, one experiment used a manipulation of whole-body posture to ma-
nipulate approach motivation. When individuals are in a supine body position (lying flat
on their backs), they are likely to be less approach motivated. In this anger experiment,
individuals were interpersonally insulted while sitting upright or while in a supine pos-
ture. Results revealed that individuals who were insulted while in the supine posture
had lower relative left frontal cortical activity than individuals who were insulted while
sitting upright (and the latter condition had greater relative left frontal activity than an
ASYMMETRIC FRONTAL CORTICAL ACTIVITY 231
As noted, models of frontal asymmetry link approach motivation with greater relative
left frontal activity. Both the affective valence model and the motivational direction
model of frontal asymmetry associate withdrawal motivation with greater relative right
frontal activity. Despite evidence supporting this model, many studies have failed to
replicate the relationship between right frontal activity and withdrawal motivated traits
and states (Amodio et al., 2008; Berkman & Lieberman, 2010; Coan & Allen, 2003; Coan
et al., 2001; De Pascalis et al., 2013; Henriques & Davidson, 2000; Hewig et al., 2004,
2006; Jackson et al., 2003; Keune et al., 2012; Kline et al., 2000; Pizzagalli et al., 2005;
Quirin et al., 2013; Wacker et al., 2008; Wacker et al., 2010). This has led researchers to
question what could be causing the inconsistencies in the literature. Some suggest that
withdrawal motivation is a complex system that is difficult to accurately measure inde-
pendently of confounding variables (Amodio et al., 2008; Coan & Allen, 2004). Others
suggest that another system entirely accounts for greater right frontal asymmetry.
A model presented by Gable and colleagues (2018) suggested that right frontal activity
is more related to regulatory control than to withdrawal motivation. Regulatory con-
trol is thought to be carried out by the revised BIS (r-BIS; Gray & McNaughton, 2000),
which may act as a governing body over conflicts between the approach and withdrawal
systems. Activation of r-BIS leads to enhanced attention to, memory for, and detec-
tion of negatively valenced information, allowing it to manage conflicts by enhancing
aversion to one motivated behavior over the other (Heym et al., 2008). This can occur as
either the suppression of a behavioral response or an override of motivational impulses
(Aron et al., 2004, 2014; Carver & Connor-Smith, 2010; Hester & Garavan, 2004, 2009).
As such, r-BIS is related to effortful control, constraint, self-control, inhibition, conflict
monitoring, and error detection (Carver & Connor-Smith, 2010; Carver et al., 2008;
Derryberry & Rothbart, 1997; Gray & McNaughton, 2000; Kochanska & Knaack, 2003;
Nigg, 2006; Rothbart & Rueda, 2005). Low functioning r-BIS is thought to be related to
impulsive behavior, deficits in inhibitory control, and externalizing disorders such as
substance abuse (Enticott et al., 2006; Logan et al., 1997). Unusually high functioning
of r-BIS, on the other hand, may be related to passive avoidance, anxious inaction,
and internalizing disorders such as generalized anxiety disorder (Carver et al., 2008;
DeYoung, 2015; Eisenberg et al., 2004; Rothbart et al., 2004; Strack & Deutsch, 2004;
Valiente et al., 2003).
While much research has examined the relationship between asymmetric frontal cor-
tical activity and motivational systems, few studies have investigated the connections
between regulatory control (r-BIS) and asymmetric frontal activity (Gable et al., 2015;
Grimshaw & Carmel, 2014; Neal & Gable, 2016; Wacker et al., 2003). R-BIS acts as a
controlling agent over approach and withdrawal behaviors. The present model suggests
ASYMMETRIC FRONTAL CORTICAL ACTIVITY 233
that greater relative right frontal activity is indicative of greater r-BIS functioning
whereas reduced relative right frontal activity suggests reduced r-BIS functioning.
results suggest that impulsivity is associated with reduced right frontal activity, inde-
pendent of affective valence.
This study did not find a relationship between trait sensation seeking and frontal
activity. However, the type of sensation seeking scale used (UPPS-P measure of trait
sensation seeking) does not correlate well with other subscales of the UPPS-P, and
some researchers have suggested that it may reflect a construct other than impulsivity
(Simons et al., 2010). Santesso et al. (2008), however, used the Zuckerman Sensation
Seeking Scale to measure trait sensation seeking and found that it was related to greater
left frontal (reduced right frontal) activity.
Overall, these findings suggest that deficits in persistence and inhibiting behavior are
associated with reduced right frontal activity. Source localization of the relationship be-
tween r-BIS and frontal asymmetry suggests that the asymmetry is driven by reduced
activity in the right medial and lateral frontal areas, including areas of the right pre-
frontal cortex.
to express it, causes greater right frontal activity (Zinner et al., 2008). Although greater
right frontal activity may be associated with the motivation to withdraw in some anger
states, these results also support that greater right frontal activity is associated with ef-
fortful control stemming from emotional suppression of anger.
Functioning of r-BIS can also be connected to drug and alcohol reactivity. Greater
relative left frontal (reduced right frontal) activity has been connected to drug-cue
reactivity such as alcohol exposure (Myrick et al., 2004) and cocaine cravings (van
de Laar et al., 2004). These increases in left frontal activity in response to substance
cues are thought to be indicative of appetitive responses evoked from substance-
related stimuli (Carter & Tiffany, 1999). Hypoactivation of rBIS, however, can also
cause increased responsiveness toward alcohol. Mechin and colleagues (2016)
investigated whether it is trait impulsivity or trait approach motivation that drives
the increase in relative left frontal activity in response to alcohol-related cues. To
determine this, they had participants complete the UPPS-P Behavioral Impulsivity
Scale (Cyders & Smith, 2007; Whiteside et al., 2005), the BIS/BAS scales (Carver &
White, 1994), and questions about their drinking habits. They then collected EEG
data while participants viewed alcohol-related and neutral picture cues. Results
suggested that the reduction in right frontal activity toward alcohol cues was due
to trait impulsivity rather than trait approach motivation. The relationship between
impulsivity and alcohol picture presentation remained constant when controlling for
drinking behaviors and asymmetric frontal activity in response to neutral pictures.
These results suggest that r-BIS moderates asymmetric frontal activity to alcohol
cues while BAS does not.
Impulsivity is presumably related to relative right frontal cortical activity because im-
pulsivity is associated with reduced effortful control. Evidence demonstrating increases
in relative right frontal activity when individuals demonstrate greater effortful control
and increases in relative left frontal activity when individuals demonstrate greater im-
pulsivity could provide more compelling evidence for this model. Neal and Gable (2019)
used a Balloon Analogue Risk Task (BART) to test whether impulsive or controlled
behaviors influence asymmetric frontal activity. In this study, EEG recordings were
collected as participants performed the BART while simultaneously viewing alcohol
stimuli designed to enhance impulsive tendencies. Participants could win money by
successfully blowing up a virtual balloon. With each pump of the balloon, more money
could be won, but the likelihood of the balloon popping increased. Each trial ended with
either the participant cashing out on the balloon or the balloon popping.
EEG data collected during this task showed that asymmetry scores shifted throughout
alcohol trials. Frontal activity would shift to greater relative left frontal activity in the last
half of trials where the balloon popped (impulsive behavior) while activity would shift
to the right on the last half of trials where the participant cashed out (successful inhib-
ition). These shifts were localized to reduced activity in the rIFG and lIFG, respectively.
This suggests that increased right frontal activity is indicative of impulse control and
diminished right frontal activity leads to more impulsive behavior.
236 EDDIE HARMON-JONES, TAYLOR POPP, AND PHILIP A. GABLE
The right frontal cortex has also been found to play an important role in the r-BIS
functions of error detection, emotion regulation, and self-control. When an incor-
rect or inappropriate behavior is enacted, the error-related negativity (ERN) is evoked
in response. Individuals with greater levels of behavioral inhibition, anxiety, and
emotion regulation tend to have increased ERN amplitudes, suggesting that those
with higher functioning r-BIS show greater neural responses in response to conflict
monitoring (Amodio et al., 2008; Proudfit et al., 2013; Teper & Inzlicht, 2013). Greater
relative right frontal activity at baseline has been linked to increased ERN amplitudes,
while greater relative left frontal activity has been linked to reduced ERNs (Nash et al.,
2012). Taken together, these results suggest that greater relative right frontal activity
is related to greater r-BIS functioning in terms of conflict detection in response to
errors.
Recently, Lacey and colleagues (2020) further connected emotion regulation to
right frontal activity. In one experiment, participants listened to anxiety-inducing
and neutral sound clips and were told to either listen naturally or suppress their emo-
tional reactions to the clips. Participants recorded their level of effort when suppressing
their reactions and noted their affective experience on each trial. Results showed that,
when participants recorded higher levels of effortful control, right frontal activity was
increased. However, experience of negative emotion was not associated with this in-
crease in right frontal activity. In a second experiment, participants were shown nega-
tive and neutral pictures and told that looking at the images for a long time would earn
them money, while choosing to escape from looking at the pictures would earn them
no reward. In this study, right frontal activity was associated with looking at negative
pictures for a longer time during reward trials, but not during non-reward trials. Both
studies indicate that it was not the negative affect associated with the aversive stimuli,
but rather the effort to engage the negative stimuli that was related to an increase in right
frontal activity.
Schmeichel and colleagues (2016) evaluated the relationship between self-control
and asymmetrical frontal activity. In this study, participants either had their self-
control depleted or not, and then underwent EEG recording while viewing posi-
tive pictures. Among individuals who were relatively higher in trait BAS than BIS,
those with depleted (vs. non-depleted) self-control showed increased left frontal ac-
tivity in response to positive pictures. Among those with no relative difference in
BIS and BAS scores, those with depleted self-control showed decreased left frontal
activity in response to positive pictures. This suggests that when r-BIS is depleted,
those who generally exhibit more approach motivation show greater left frontal ac-
tivity in response to rewarding stimuli. An enhanced r-BIS, on the other hand, may
be able to lower approach motivation in those with generally hyperactive BAS and
hypoactive BIS.
Overall, these studies suggest that greater right frontal activity may be associated
with processes involving emotion regulation while lower right frontal activity may be
associated with reduced self-control and hindered error monitoring. Additionally, both
ASYMMETRIC FRONTAL CORTICAL ACTIVITY 237
situational contexts where control must be utilized and control-related personality traits
have been related to right frontal activity as measured with EEG.
Researchers have presented additional models to explain frontal asymmetry and its re-
lationship to emotional and motivational variables. The Bilateral BAS Model (Hewig
et al, 2004) and the Activation vs. Inhibition Model (Wacker et al., 2003) each propose
variants of how approach, withdrawal, and inhibition influences frontal asymmetry.
both types of reinforcement should activate the BAS, as both are a form of reward or
non-punishment. Participants then completed a go/no-go task in which no-go trials
were considered to be a form of passive avoidance, therefore activating the BIS. After the
task, participants were shown monetary feedback based on their performance. Results
showed that all participants showed greater bilateral frontal activation in response to
reinforcement trials—regardless of whether it was positive or negative reinforcement—
and that this effect was strongest in those with higher trait BAS scores. When these
results are considered with prior research suggesting that approach motivation aligns
with greater relative left frontal activity and withdrawal motivation aligns with greater
relative right frontal activity, the suggestion is that the two are subsystems of the
overarching BAS, once again supporting Hewig and colleagues’ bilateral BAS model.
Rodrigues and colleagues (2018) found further support for the bilateral BAS model
using a virtual reality paradigm. In this study, participants navigated a virtual T maze
in which they were to respond to various events. These events could be positive, nega-
tive, or neutral in valence. Positive trials encouraged participants to move toward the
event while negative trials encouraged withdrawal from the event. Approach-avoidance
conflict trials involved having a positive stimulus guarded by a negative stimulus and
approach-approach conflict trials involved choosing between two positive stimuli.
There were also two control trial types. In each trial, participants chose in which dir-
ection they would like to move while EEG was recorded. Left frontal activity increased
when participants chose an approach direction while right frontal activity increased
when participants chose a withdrawal direction. Further, bilateral frontal activity
increased during any choice in behavior relative to the choice to not move in any dir-
ection. The authors argue that these results support the role of frontal asymmetry in
behavioral approach or avoidance motivation, as well as the role of bilateral frontal acti-
vation in active behavior.
In an early study, Wacker and colleagues (2003) compared the viability of the BIS/
BAS, motivational direction, and valence models of frontal asymmetry. Participants
completed a mental image script task in which emotion (fear vs. anger) and motivation
(approach vs. withdrawal) were manipulated and the participants reported the degree
to which they agreed that the outcome of each task was the best option (degree of con-
flict). Results did not fall in line with the predictions of either the motivational direc-
tion or valence models of anterior asymmetry, but the degree of conflict experienced by
participants was positively correlated with relative right frontal activation. Assuming
that agreement ratings were a valid measure of BIS activation, these results support the
model of frontal asymmetry suggested by Wacker and colleagues.
Similar results were found in a later study by Wacker and colleagues (2008), which
directly compared the BIS/BAS model of anterior activation (BBMAA) to the mo-
tivational direction model. During a mental image script task similar to that used by
Wacker and colleagues (2003), participants showed greater relative right frontal acti-
vation in response to scripts designed to target BIS than to those targeting FFFS. Self-
reported measures of FFFS activation, conversely, were associated with greater relative
left frontal activation.
The relationship between trait BIS and frontal asymmetry was investigated by Wacker
and colleagues (2010) via a go/no-go task. Such a paradigm was used because no-go
tasks have been suggested to be a viable measure of the conflict and behavioral inhibition
functions of Wacker’s revised BIS. Consistent with this assumption, those participants
who had greater trait BIS showed greater relative right frontal activation in response to
no-go trials than to go trials.
Because frontal alpha asymmetry plays a role in the experience of motivation and
emotion, it follows that it would also factor into the experience of mental illness, par-
ticularly mood disorders. The psychological conditions most frequently investigated in
relation to frontal asymmetry are depression, anxiety, and bipolar disorder.
11.9.1 Depression
Depression is one of the most commonly studied conditions in relation to frontal asym-
metry. Depressive symptomology includes experiences such as decreased response to
reward, lack of positive affect, and greater tendencies toward withdrawal from triggering
activities. Davidson and colleagues (2002) pointed out that these symptoms can often
be described as deficiencies in approach motivation and hyperactive withdrawal motiv-
ation, and these trends are associated with decreased relative left frontal cortical activity
240 EDDIE HARMON-JONES, TAYLOR POPP, AND PHILIP A. GABLE
(greater relative right frontal activity). A number of studies have found a relationship
between depressive symptoms and reduced left-frontal activity during a resting state.
For instance, Schaffer and colleagues (1983) found a negative correlation between scores
on the Beck Depression Inventory (BDI) and relative left frontal activity. Subsequent
studies have found similar relationships between relative left frontal activity at rest
and self-reported (Diego et al., 2001) or clinically diagnosed depression (Henriques &
Davidson, 1990; Smith et al., 2018; Stewart et al., 2010). Henriques and Davidson (1990)
even found that participants who had previously been depressed showed lower relative
left frontal activity at rest compared to those who had never been depressed, and that
these patterns of asymmetry were comparable to those in participants experiencing
acute depressive symptoms. These results suggest that frontal asymmetry may be a state-
independent marker of depression.
It is of note that diminished relative left frontal activity in depressed individuals is
particularly robust in women compared to men (Stewart et al., 2010). Stewart and Allen
(2018) found evidence supporting this notion in a sample with no history of major de-
pressive disorder. Participants engaged in resting state EEG recordings before returning
one year later to report depressive symptomology during the worst two-week period
experienced throughout the one-year interim. It was found that women—but not
men—with lower relative left frontal activity at baseline reported greater degrees of de-
pressive symptomology during the interim period.
While resting data provides a means of studying the relationship between frontal
asymmetry and depression, some studies have found a null relationship between de-
pression and frontal asymmetry (Harmon-Jones et al., 2002; Metzger et al., 2004;
Tomarken & Davidson, 1994; McFarland et al., 2006). These null results have led
some researchers to argue that state-measures of frontal asymmetry may be more re-
liably associated with depression. For example, the capability model of frontal asym-
metry suggests that cortical activity measured during emotional challenges is more
indicative of predispositions toward psychopathology than cortical activity measured
at rest (Stewart et al., 2014). Lower relative left frontal activity during a facial emotion
task was found to be indicative of a history of major depressive disorder (Stewart et al.,
2011). Participants prone to depressive symptomology have also exhibited lower rela-
tive left frontal activity during tasks that evoke anger (Harmon-Jones et al., 2002) and
sadness (Nitschke et al., 2004). Those with early-onset depression also exhibited lower
relative left frontal activity during an approach-related reward paradigm (Shankman
et al., 2007).
Frontal asymmetry may act as more of a risk factor than an indicator of current de-
pression. In one study, Possel and colleagues (2008) measured resting frontal asym-
metry in adolescent boys and related it to depressive symptoms experienced throughout
the year following the EEG recording. Results suggested that lower baseline left frontal
activity was predictive of melancholic depressive symptoms and increased right frontal
activity was predictive of non-melancholic depressive symptoms one year later, but
depressive state was not predictive of baseline asymmetry. This suggests that frontal
asymmetry may be indicative of a predisposition to depression. Mitchell and Possel
(2012) conducted a similar study comparing depression and resting asymmetry in a
ASYMMETRIC FRONTAL CORTICAL ACTIVITY 241
nonclinical population and found that individuals with lower baseline left frontal ac-
tivity were more likely to develop depressive symptoms over the next year. Nusslock and
colleagues’ (2011) results echoed these findings, suggesting that cognitive vulnerability
to depression was both associated with lower relative left frontal activity at baseline and
predicted onset of depressive symptoms one year later.
In addition, frontal asymmetry has been found to be a predictor of treatment re-
sponse in depressed individuals. Greater relative left frontal activity prior to treatment
predicted more successful response to fluoxetine (Bruder et al., 2001), as well as to
escitalopram and sertraline in women (Arns et al., 2015).
11.9.2 Anxiety
Anxiety symptoms also appear to be related to greater right frontal activity. In a sample
of participants consisting of healthy controls, those with major depression in remis-
sion, those with acute depression without comorbid anxiety disorder, and those with
acute depression with a comorbid anxiety disorder, the only group difference observed
was greater relative right frontal activity in the group with both major depression and a
comorbid anxiety disorder compared to healthy controls (Feldmann et al., 2018). Other
findings suggested that those with both anxiety and depression show frontal asym-
metry patterns similar to those with depression alone (Mathersul et al., 2008). Nusslock
and colleagues (2018) found asymmetry patterns to be most similar between healthy
controls and those with comorbid depression and anxiety. In this study, women with a
history of childhood onset depression without anxiety diagnoses showed reduced left
frontal activity, which is consistent with past research on depression. However, women
with a history of childhood onset depression and with pathological levels of anxious ap-
prehension (in the form of generalized anxiety disorder, obsessive compulsive disorder,
or separation anxiety disorder) showed resting asymmetry patterns statistically indis-
tinguishable from healthy controls. These results highlight the role of comorbid depres-
sion and anxiety in its complex relationship with frontal asymmetry.
Similar results have been found in individuals with clinical diagnoses of anxiety
disorders. Adults with panic disorder, for example, showed increased right frontal ac-
tivity in both resting and anxiety-provoking contexts relative to controls (Wiedemann
et al., 1999). Individuals with social phobia also showed increased right frontal activity
compared to controls in response to anticipation of giving a speech (Davidson et al.,
2000). Additionally, participants with post-traumatic stress disorder showed greater
state-dependent right frontal activity when they were presented with trauma-relevant
stimuli (Meyer et al., 2015). When participants engaged in an emotional Stroop task
consisting of human faces depicting various emotions, those participants with greater
levels of trait anxiety showed increased right frontal activity in response to fearful
faces than did those with lower levels of trait anxiety (Avram et al., 2010). Similarly,
participants who scored higher in trait anxiety showed greater right frontal activity
during anxiety-provoking situations (Balconi & Pagani, 2014; Cole et al., 2012; Crost
et al., 2008).
242 EDDIE HARMON-JONES, TAYLOR POPP, AND PHILIP A. GABLE
Inconsistent patterns of frontal asymmetry have been found in children with anxiety
disorders. In a study comparing resting data of boys and girls aged 8 years and 11 years,
anxious girls aged 8 and 11 showed decreased relative left frontal activity at rest, while
their healthy control counterparts showed no significant frontal asymmetry at 8 years
and greater relative left frontal activity at 11 years. While these results are consistent with
research on adults with anxiety disorders, the young boys in the study showed incon-
sistent patterns. Anxious 8-year-old boys showed no significant frontal asymmetry and
anxious 11-year-old boys showed greater relative left frontal activity while healthy boys
showed greater relative right frontal activity (Baving et al.,, 2002). These results suggest
that patterns of frontal asymmetry in the context of anxiety are not consistent across
gender and age; further research is needed to better understand this relationship.
Inconsistent patterns have also been found when comparing various subtypes of
anxiety. For instance, at baseline, participants with generalized anxiety disorder and
increased worry show greater relative left frontal activity while those with high trait anx-
iety and low worry show lower relative left frontal activity (Smith et al., 2016). Crost and
colleagues (2008) also found that those with more anxiety show greater relative right
frontal activity in response to social threat while those higher in defensiveness show
greater relative left frontal activity in response to social threat. It has further been argued
that symptoms of anxious arousal (e.g., panic) are correlated with lower relative left
frontal activity and symptoms of anxious apprehension (e.g., worry) are correlated with
greater relative left frontal activity. As these results suggest, symptoms of anxiety have
complex relationships with patterns of frontal asymmetry.
11.9.3 Bipolar Disorder
Individuals with bipolar disorder experience heightened approach motivation and
hypersensitivity to goal-and reward-relevant cues during episodes of mania/hypomania
(Alloy & Abramson, 2010; Johnson, 2005; Urosevic et al., 2010). This trend is indicative
of increased BAS sensitivity in individuals with bipolar disorder. Indeed, self-reported
BAS sensitivity scores are higher in those with bipolar I disorder (Meyer et al., 2001;
Salavert et al., 2007), bipolar II disorder, and cyclothymia (Alloy et al., 2008), as well
as those prone to hypomanic symptomology (Meyer et al., 1999). Since increased BAS
sensitivity and approach motivation are positively correlated with relative left frontal
cortical activity, it is expected that this trend should be found in individuals with bipolar
disorder. Kano and colleagues (1992) collected resting EEG data from participants with
bipolar disorder and found greater left frontal cortical activation, suggesting that those
experiencing manic symptoms (as opposed to depressive symptoms) have greater rela-
tive left frontal activity at baseline.
Nusslock and colleagues (2012) also measured resting EEG frontal asymmetry in
individuals with cyclothymia and followed up approximately 5 years later to observe
changes in bipolar course. Those with greater levels of left frontal activity at base-
line had a greater likelihood of converting to more severe diagnoses over the interim
period (i.e., conversion from cyclothymia or bipolar II to bipolar I). Greater relative left
ASYMMETRIC FRONTAL CORTICAL ACTIVITY 243
frontal activity at baseline was also predictive of earlier age-of-onset of a first bipolar
spectrum episode, an indicator of severity of bipolar disorder. This relationship was
observed even when controlling for mood state and medication status at the time of
EEG recording.
An early study by Harmon-Jones and colleagues (2002) also found associations be-
tween bipolar disorder and frontal asymmetry in a task-based paradigm. In this study,
participants completed the General Behavior Inventory (GBI; Depue et al., 1989) to
assess potential risk for developing bipolar or depressive disorders. Then, EEG data was
collected while participants engaged in an anger-evoking task. The authors hypothesized
that individuals with hypomania/mania symptoms would have greater relative left
frontal activity when angered, based on prior research (Depue & Iacono, 1989). They
also hypothesized that the opposite pattern (a decrease in left frontal activity) would be
seen in participants with depressive symptoms. As predicted, those with hypomanic/
manic symptoms showed increased left frontal activity during the anger-inducing situ-
ation while those with depressive symptoms showed decreased left frontal activity. These
results suggest that the greater approach tendencies experienced by those with hypo-
manic/manic symptoms and the diminished approach tendencies experienced by those
with depressive symptoms manifest through variations in frontal asymmetry.
Harmon-Jones and colleagues (2008) tested the BAS dysregulation theory of bipolar
disorder by measuring frontal asymmetry in response to tasks of varying difficulty.
Participants had either a bipolar spectrum diagnosis or no major psychopathology.
During EEG data collection, participants engaged in an anagram task during which
they were given cues indicating the difficulty (i.e., easy, medium, or hard) of the up-
coming trial. They were also told whether they could receive money (win) or avoid
losing money (loss) by successfully completing the upcoming trial. Results indicated
that those with bipolar disorder had greater relative left frontal activity in preparation
for hard/win trials, while control participants showed a decrease in relative left frontal
activity in anticipation of the same trial type. Greater relative left frontal activity was
also correlated with self-reported hypomanic/manic experience during the task in
individuals with bipolar disorder. These results therefore provide evidence supporting
BAS dysregulation theory. They also suggest that increased relative left frontal ac-
tivity, which may be related to manic symptoms, can be triggered by more difficult and
rewarding stimuli.
11.11 CONCLUSION
The motivational model of frontal asymmetry suggests that relative left frontal activity
is associated with approach motivation while right frontal activity is associated with
withdrawal motivation. The former assumption of this model is supported by the litera-
ture, both through possible motivation confounds associated with the affective valence
model and through more direct manipulations of motivation. Anger—an emotion that
is both negative and approach motivated—is associated with greater left frontal activity
246 EDDIE HARMON-JONES, TAYLOR POPP, AND PHILIP A. GABLE
at the state and trait level. Further support stems from research connecting higher BAS
scores with greater baseline left frontal activity. The connection between BIS scores and
baseline right frontal activity, however, has been widely contested.
This inconsistency in the literature concerning withdrawal motivation and right
frontal has led to the development of the effortful control model of right frontal asym-
metry. The effortful control model suggests that right frontal activity is associated with
control and regulatory processing. Research supporting the effortful control model has
linked right frontal activity with increased anxious inaction, error detection, and emo-
tional control-related behaviors; right frontal activity is also negatively correlated with
state and trait impulsivity. Past work linking right frontal activity and withdrawal mo-
tivation may have activated effortful control, because withdrawal manipulations may
have co-activated effortful control to stay engaged with the aversive stimuli, but also be-
cause activation of r-BIS increases the cognitive load devoted to negative stimuli.
Psychopathologies influence patterns of frontal asymmetry. For instance, individuals
with depression show reduced levels of left frontal activity at rest and during emo-
tionally salient tasks. Reduced left frontal activity may be a risk factor for depression
in nonclinical populations. Those with anxiety generally show increased levels of right
frontal activity at rest and during tasks that elicit fear and/or anxiety; however, this re-
lationship is more complicated in children and when comparing subtypes of anxiety.
Those with bipolar disorder who experience manic symptoms exhibit greater left frontal
activity at rest and during anger-evoking or difficult rewarding tasks. These patterns
are consistent with the motivational direction and effortful control models of frontal
asymmetry.
While there have been failures to replicate some past frontal asymmetry findings,
these failures to replicate may be due to the complex nature of the systems underlying
frontal asymmetry. Phenomena such as affect and motivation change by the moment
and can be swayed by situational effects. Despite this, frontal asymmetry research has
a long and well-established history in EEG frequency research and continues to spark
new predictions, models, and discoveries.
REFERENCES
Alema, G., Rosadini, G., & Rossi, G. F. (1961). Preliminary experiments on the effects of the
intracarotid introduction of sodium amytal in parkinsonian syndromes. Bollettino della
Società Italiana di Biologia Sperimentale, 37, 1036–1037.
Allen, J. J. B., Harmon-Jones, E., & Cavender, J. H. (2001). Manipulation of frontal EEG asym-
metry through biofeedback alters self-reported emotional responses and facial EMG.
Psychophysiology, 38, 685–693.
Allen, J. J. B., Iacono, W. G., Depue, R. A., & Arbisi, P. (1993). Regional electroencephalographic
asymmetries in bipolar seasonal affective disorder before and after exposure to bright light.
Biological Psychiatry, 33, 642–646.
Alloy, L. B., Abramson, L. Y., Walshaw, P. D., Cogswell, A., Grandin, L. D., Hughes, M. E.,
... & Hogan, M. E. (2008). Behavioral approach system and behavioral inhibition system
ASYMMETRIC FRONTAL CORTICAL ACTIVITY 247
Brehm, J. W., & Self, E. A. (1989). The intensity of motivation. Annual Review of Psychology, 40,
109–131.
Bruder, G. E., Stewart, J. W., Tenke, C. E., McGrath, P. J., Leite, P., Bhattacharya, N., & Quitkin,
F. M. (2001). Electroencephalographic and perceptual asymmetry differences between
responders and nonresponders to an SSRI antidepressant. Biological psychiatry, 49(5),
416–425.
Carter, B. L. & Tiffany, S. T. (1999). Meta-analysis of cue-reactivity in addiction research.
Addiction, 94(3), 327–340. doi:10.1046/ j.1360-0443.1999.9433273.x
Carver, C. S. (2004). Negative affects deriving from the behavioral approach system. Emotion,
4, 3–22.
Carver, C. S. & Connor-Smith, J. (2010). Personality and coping. Annual Review of Psychology,
61, 679–704. doi:10.1146/annurev.psych.093008.100352
Carver, C. S. & Harmon-Jones, E. (2009). Anger is an approach-related affect: Evidence and
implications. Psychological Bulletin, 135, 183–204.
Carver, C. S. & White, T. L. (1994). Behavioral inhibition, behavioral activation, and affective
responses to impending reward and punishment: The BIS/BAS scales. Journal of Personality
and Social Psychology, 67, 319–333.
Carver, C. S., Johnson, S. L., & Joormann, J. (2008). Serotonergic function, two-mode models
of self-regulation, and vulnerability to depression: What depression has in common with
impulsive aggression. Psychological bulletin, 134(6), 912. doi:10.1037/a0013740
Coan, J. A. & Allen, J. J. (2003). Frontal EEG asymmetry and the behavioral activation and in-
hibition systems. Psychophysiology, 40, 106–114.
Coan, J. A. & Allen, J. J. (2004). Frontal EEG asymmetry as a moderator and mediator of
emotion. Biological Psychology, 67, 7–50.
Coan, J. A., Allen, J. J. B., & Harmon-Jones, E. (2001). Voluntary facial expression and hemi-
spheric asymmetry over the frontal cortex. Psychophysiology, 38, 912–925.
Cohen, J. (1977). Statistical power analysis for the behavioral sciences (rev. ed.). Academic Press.
Cohen, J. (1988). Statistical power analysis for the behavioral sciences (2nd ed.). Erlbaum.
Cole, C., Zapp, D. J., Nelson, S. K., & Perez-Edgar, K. (2012). Speech presentation cues mod-
erate frontal EEG asymmetry in socially withdrawn young adults. Brain and Cognition,
78(2), 156–162. doi:10.1016/j.bandc.2011.10.013
Cook, I. A., O’Hara, R., Uijtdehaage, S. H. J., Mandelkern, M., & Leuchter, A. F. (1998).
Assessing the accuracy of topographic EEG mapping for determining local brain function.
Electroencephalography and Clinical Neurophysiology, 107, 408–414.
Crost, N. W., Pauls, C. A., & Wacker, J. (2008). Defensiveness and anxiety predict frontal
EEG asymmetry only in specific situational contexts. Biological Psychology, 78(1), 43–52.
doi:10.1016/j. biopsycho.2007.12.008
Cyders, M. A., & Smith, G. T. (2007). Mood-based rash action and its components: Positive
and negative urgency. Personality and Individual Differences, 43(4), 839–850. doi:10.1016/
j.paid.2007. 02.008
Cyders, M. A., Zapolski, T. C., Combs, J. L., Settles, R. F., Fillmore, M. T., & Smith, G. T. (2010).
Experimental effect of positive urgency on negative outcomes from risk taking and on
increased alcohol consumption. Psychology of Addictive Behaviors, 24(3), 367–375.
Darwin, C. (1872/ 1965). The expressions of the emotions in man and animals. Oxford
University Press.
Davidson, R. J. & Fox, N. (1982). Asymmetric brain activity discriminates between positive and
negative affective stimuli in 10 month old infants. Science, 218, 1235–1237.
ASYMMETRIC FRONTAL CORTICAL ACTIVITY 249
Davidson, R. J., Chapman, J. P., Chapman, L. J., & Henriques, J. B. (1990). Asymmetric brain
electrical-activity discriminates between psychometrically-matched verbal and spatial cog-
nitive tasks. Psychophysiology, 27, 528–543.
Davidson, R. J., Marshall, J. R., Tomarken, A. J., & Henriques, J. B. (2000). While a phobic
waits: Regional brain electrical and autonomic activity in social phobics during antici-
pation of public speaking. Biological Psychiatry, 47(2), 85–95. https://doi.org/10.1016/
s0006-3223(99)00222-x
Davidson, R. J., Pizzagalli, D., Nitschke, J. B., & Putnam, K. (2002). Depression: Perspectives
from affective neuroscience. Annual Review of Psychology, 53, 545–574. doi:10.1146/annurev.
psych.53.100901.135148
De Pascalis, V., Cozzuto, G., Caprara, G. V., & Alessandri, G. (2013). Relations among EEG-
alpha asymmetry, BIS/BAS, and dispositional optimism. Biological Psychology, 94, 198–209.
Depue, R. A., Krauss, S., Spoont, M. R., & Arbisi, P. (1989). General Behavior Inventory identi-
fication of unipolar and bipolar affective conditions in a nonclinical university population.
Journal of Abnormal Psychology, 98(2), 117–126.
Depue, R.A. & Iacono, W.G. (1989). Neurobehavioral aspects of affective disorders. Annual
Review of Psychology, 40, 457–492.
Derryberry, D. & Rothbart, M. K. (1997). Reactive and effortful processes in the organ-
ization of temperament. Development and Psychopathology, 9(4), 633–652. doi:10.1017/
S095457949700 1375
DeYoung, C. G. (2015). Cybernetic big five theory. Journal of Research in Personality, 56, 33–58.
doi:10.1016/j.jrp.2014.07.004
Diego, M. A., Field, T., & Hernandez-Reif, M. (2001). CES-D depression scores are correlated
with frontal EEG alpha asymmetry. Depression and Anxiety, 13(1), 32–37.
Eisenberg, N., Spinrad, T. L., Fabes, R. A., Reiser, M., Cumberland, A., Shepard, S. A., ...
Thompson, M. (2004). The relations of effortful control and impulsivity to children’s resili-
ency and adjustment. Child Development, 75(1), 25–46. doi:10.1111/j.1467- 8624.2004.00652.x
Ekman, P. & Davidson, R. J. (1993). Voluntary smiling changes regional brain activity.
Psychological Science, 4, 342–345.
Ekman, P. & Friesen, W. V. (1975). Unmasking the face: A guide to recognizing emotions from fa-
cial clues. Prentice-Hall.
Elgavish, E., Halpern, D., Dikman, Z., & Allen, J. J. B. (2003). Does frontal EEG asymmetry
moderate or mediate responses to the international affective picture system (IAPS)?
Psychophysiology, 40, S38.
Enticott, P. G., Ogloff, J. R., & Bradshaw, J. L. (2006). Associations between laboratory measures
of executive inhibitory control and self-reported impulsivity. Personality and Individual
Differences, 41(2), 285–294. doi:10.1016/j.paid.2006.01.011
Feldmann, L., Piechaczek, C. E., Grünewald, B. D., Pehl, V., Bartling, J., Frey, M., ... &
Greimel, E. (2018). Resting frontal EEG asymmetry in adolescents with major depres-
sion: impact of disease state and comorbid anxiety disorder. Clinical Neurophysiology,
129(12), 2577–2585.
Funder, D. C. & Ozer, D. J. (2019). Evaluating effect size in psychological research: Sense and
nonsense. Advances in Methods and Practices in Psychological Science, 2(2), 156–168.
Gable, P. A. & Dreisbach, G. (2021). Approach motivation and positive affect. Current Opinion
in Behavioral Sciences, 39, 203–208.
Gable, P. A. & Harmon- Jones, E. (2008). Relative left frontal activation to appetitive
stimuli: Considering the role of individual differences. Psychophysiology, 45, 275–278.
250 EDDIE HARMON-JONES, TAYLOR POPP, AND PHILIP A. GABLE
Gable, P. A. & Poole, B. D. (2014). Influence of trait behavioral inhibition and behavioral
approach motivation systems on the LPP and frontal asymmetry to anger pictures. Social
Cognitive and Affective Neuroscience, 9, 182–190.
Gable, P. A., Mechin, N. C., Hicks, J. A., & Adams, D. L. (2015). Supervisory control system
and frontal asymmetry: Neurophysiological traits of emotion-based impulsivity. Social,
Cognitive, and Affective Neuroscience, 10(10), 1310–315.
Gable, P. A., Mechin, N., Hicks, J.A., & Adams, D.A. (2015). Positive urgency and left-frontal
asymmetry: Neurophysiological traits of emotion-based impulsivity. Social Cognitive and
Affective Neuroscience, 10, 1310–1315
Gable, P. A., Neal, L. B., & Threadgill, A. H. (2018). Regulatory behavior and frontal ac-
tivity: Considering the role of revised- BIS in relative right frontal asymmetry.
Psychophysiology, 55(1), e12910.
Gable, P. A., Poole, B. D., & Cook, M. S. (2013). Asymmetrical hemisphere activation enhances
global-local processing. Brain and Cognition, 83, 337–341.
Gainotti, G. (1972). Emotional behavior and hemispheric side of the lesion. Cortex, 8, 41–55.
Gasparrini, W. G., Satz, P., Heilman, K., & Coolidge, F. L. (1978). Hemispheric asymmetries
of affective processing as determined by the Minnesota Multiphasic Personality Inventory.
Journal of Neurology, Neurosurgery and Psychiatry, 41, 470–473.
Gianotti, L. R., Knoch, D., Faber, P. L., Lehmann, D., Pascual-Marqui, R. D., Diezi, C., ... & Fehr,
E. (2009). Tonic activity level in the right prefrontal cortex predicts individuals’ risk taking.
Psychological Science, 20, 33–38.
Goldstein, K. (1939). The organism: An holistic approach to biology, derived from pathological
data in man. American Books.
Gray, J. (1982). Précis of The neuropsychology of anxiety: An enquiry into the functions of the
septo-hippocampal system. Behavioral and Brain Sciences, 5(3), 469– 484. doi:10.1017/
S0140525X00013066
Gray, J. A. & McNaughton, N. (2000). The neuropsychology of anxiety: An enquiry into the
functions of the septo-hippocampal system (2nd ed.). Oxford University Press.
Grimshaw, G. M. & Carmel, D. (2014). An asymmetric inhibition model of hemi-
spheric differences in emotional processing. Frontiers in Psychology, 5, 489. doi:10.3389/
fpsyg.2014.00489
Gross, J. J. & Levenson, R. W. (1993). Emotional suppression: Physiology, self-report, and ex-
pressive behavior. Journal of Personality and Social Psychology, 64, 970–986. doi:10.1037/
0022- 3514.64.6.970
Hagemann, D., Hewig, J., Seifert, J., Naumann, E., & Bartussek, D. (2005). The latent state-
trait structure of resting EEG asymmetry: Replication and extension. Psychophysiology, 42,
740–752.
Hagemann, D., Naumann, E., Becker, G., Maier, S., & Bartussek, D. (1998). Frontal brain asym-
metry and affective style: A conceptual replication. Psychophysiology, 35, 372–388.
Hagemann, D., Naumann, E., Thayer, J. F., & Bartussek, D. (2002). Does resting EEG asym-
metry reflect a trait? An application of latent state-trait theory. Journal of Personality and
Social Psychology, 82, 619–641.
Harmon-Jones, C., Schmeichel, B. J., Mennitt, E., & Harmon-Jones, E. (2011). The expression
of determination: Similarities between anger and approach-related positive affect. Journal of
Personality and Social Psychology, 100, 172–181. doi:10.1037/a0020966
Harmon-Jones, E. (2003a). Clarifying the emotive functions of asymmetrical frontal cortical
activity. Psychophysiology, 40, 838–848.
ASYMMETRIC FRONTAL CORTICAL ACTIVITY 251
Harmon-Jones, E. (2003b). Anger and the behavioural approach system. Personality and
Individual Differences, 35, 995–1005.
Harmon-Jones, E. (2004). On the relationship of anterior brain activity and anger: Examining
the role of attitude toward anger. Cognition and Emotion, 18, 337–361.
Harmon-Jones, E. (2006). Unilateral right-hand contractions cause contralateral alpha power
suppression and approach motivational affective experience. Psychophysiology, 43, 598–603.
Harmon-Jones, E. (2007). Trait anger predicts relative left frontal cortical activation to anger-
inducing stimuli. International Journal of Psychophysiology, 66, 154–160.
Harmon-Jones, E., Abramson, L. Y., Nusslock, R., Sigelman, J. D., Urosevic, S., Turonie, L. D.,
Alloy, L. B., & Fearn, M. (2008). Effect of bipolar disorder on left frontal cortical responses to
goals differing in valence and task difficulty. Biological Psychiatry, 63, 693–698.
Harmon-Jones, E., Abramson, L. Y., Sigelman, J., Bohlig, A., Hogan, M. E., & Harmon-Jones,
C. (2002). Proneness to hypomania/mania or depression and asymmetric frontal cor-
tical responses to an anger-evoking event. Journal of Personality and Social Psychology, 82,
610–618.
Harmon-Jones, E. & Allen, J. J. B. (1997). Behavioral activation sensitivity and resting frontal
EEG asymmetry: Covariation of putative indicators related to risk for mood disorders.
Journal of Abnormal Psychology, 106, 159–163.
Harmon-Jones, E. & Allen, J. J. B. (1998). Anger and prefrontal brain activity: EEG asym-
metry consistent with approach motivation despite negative affective valence. Journal of
Personality and Social Psychology, 74, 1310–1316.
Harmon-Jones, E. & Gable, P. A. (2009). Neural activity underlying the effect of approach-
motivated positive affect on narrowed attention. Psychological Science, 20, 406–409.
Harmon-Jones, E. & Gable, P. A. (2018). On the role of asymmetric frontal cortical activity in
approach and withdrawal motivation: An updated review of the evidence. Psychophysiology,
55(1), e12879.
Harmon-Jones, E., Gable, P. A., & Price, T. F. (2011). Leaning embodies desire: Evidence that
leaning forward increases relative left frontal activation to appetitive stimuli. Biological
Psychology, 87, 311–313.
Harmon-Jones, E., Harmon-Jones, C., Abramson, L. Y., & Peterson, C. K. (2009). PANAS posi-
tive activation is associated with anger. Emotion, 9, 183–196.
Harmon-Jones, E., Harmon-Jones, C., Amodio, D. M., & Gable, P. A. (2011). Attitudes toward
emotions. Journal of Personality and Social Psychology, 101, 1332–1350. doi:10.1037/a0024951
Harmon-Jones, E., Harmon-Jones, C., Fearn, M., Sigelman, J. D., & Johnson, P. (2008).
Left frontal cortical activation and spreading of alternatives: Tests of the action-based
model of dissonance. Journal of Personality and Social Psychology, 94, 1–15. doi:10.1037/
0022-3514.94.1.1
Harmon-Jones, E., Harmon-Jones, C., & Price, T. F. (2013). What is approach motivation?
Emotion Review, 5, 291–295. doi:10.1177/1754073913477509
Harmon-Jones, E., Lueck, L., Fearn, M., & Harmon-Jones, C. (2006). The effect of personal
relevance and approach-related action expectation on relative left frontal cortical activity.
Psychological Science, 17, 434–440.
Harmon-Jones, E. & Peterson, C. K. (2008). Effect of trait and state approach motivation on
aggressive inclinations. Journal of Research in Personality, 42, 1381–1385.
Harmon-Jones, E. & Peterson, C. K. (2009). Supine body position reduces neural response to
anger evocation. Psychological Science, 20, 1209–1210. doi:10.1111/j.1467-9280.2009.02416.x
252 EDDIE HARMON-JONES, TAYLOR POPP, AND PHILIP A. GABLE
Harmon-Jones, E., Peterson, C. K., & Harris, C. R. (2009). Jealousy: Novel methods and neural
correlates. Emotion, 9, 113–117.
Harmon-Jones, E. & Sigelman, J. D. (2001). State anger and prefrontal brain activity: Evidence
that insult-related relative left-prefrontal activation is associated with experienced anger
and aggression. Journal of Personality and Social Psychology, 80, 797–803.
Harmon-Jones, E., Sigelman, J. D., Bohlig, A., & Harmon-Jones, C. (2003). Anger, coping, and
frontal cortical activity: The effect of coping potential on anger-induced left frontal activity.
Cognition and Emotion, 17, 1–24.
Harmon-Jones, E., Vaughn-Scott, K., Mohr, S., Sigelman, J., & Harmon-Jones, C. (2004). The
effect of manipulated sympathy and anger on left and right frontal cortical activity. Emotion,
4, 95–101.
Hellige, J. B. (1993). Hemispheric asymmetry: What’s right and what’s left. Harvard
University Press.
Henriques, J. B. & Davidson, R. J. (1990). Regional brain electrical asymmetries discriminate
between previously depressed and healthy control subjects. Journal of Abnormal Psychology,
99, 22–31.
Henriques, J. B. & Davidson, R. J. (2000). Decreased responsiveness to reward in depression.
Cognition & Emotion, 14(5), 711–724. doi:10.1080/02699930050117684
Hester, R. & Garavan, H. (2004). Executive dysfunction in cocaine addiction: Evidence for dis-
cordant frontal, cingulate, and cerebellar activity. The Journal of Neuroscience, 24(49), 11017–
11022. doi:10.1523/JNEUROSCI.3321-04.2004
Hester, R. & Garavan, H. (2009). Neural mechanisms underlying drug-related cue distrac-
tion in active cocaine users. Pharmacology Biochemistry and Behavior, 93(3), 270–277.
doi:10.1016/j. pbb.2008.12.009
Hewig, J., Hagemann, D., Seifert, J., Naumann, E., & Bartussek, D. (2004). On the selective re-
lation of frontal cortical activity and anger-out versus anger-control. Journal of Personality
and Social Psychology, 87, 926–939.
Hewig, J., Hagemann, D., Seifert, J., Naumann, E., & Bartussek, D. (2006). The relation of cor-
tical activity and BIS/BAS on the trait level. Biological Psychology, 71(1), 42–53. doi:10.1016/j.
biopsycho.2005.01.006
Heym, N., Ferguson, E., & Lawrence, C. (2008). An evaluation of the relationship be-
tween Gray’s revised RST and Eysenck’s PEN: Distinguishing BIS and FFFS in Carver and
White’s BIS/BAS scales. Personality and Individual Differences, 45(8), 709–7 15. doi:10.1016/
j.paid.2008.07.013
Hughes, D. M., Yates, M. J., Morton, E. E., & Smillie, L. D. (2015). Asymmetric frontal cortical
activity predicts effort expenditure for reward. Social Cognitive and Affective Neuroscience,
10, 1015–1019.
Jackson, D. C., Mueller, C. J., Dolski, I., Dalton, K. M., Nitschke, J. B., Urry, H. L., ... & Davidson,
R. J. (2003). Now you feel it, now you don’t: Frontal brain electrical asymmetry and indi-
vidual differences in emotion regulation. Psychological Science, 14(6), 612–617. doi:10.1046/
j.0956-7976.2003.psci_1473.x
Jacobs, G. D. & Snyder, D. (1996). Frontal brain asymmetry predicts affective style in men.
Behavioral Neuroscience, 110, 3–6.
Jensen-Campbell, L. A., Knack, J. M., Waldrip, A. M., & Campbell, S. D. (2007). Do big five per-
sonality traits associated with self-control influence the regulation of anger and aggression?
Journal of Research in Personality, 41, 403–424.
ASYMMETRIC FRONTAL CORTICAL ACTIVITY 253
Johnson, S. L. (2005). Mania and dysregulation in goal pursuit: A review. Clinical Psychology
Review, 25, 241–262.
Kamiya, J. (1979). Autoregulation of the EEG alpha rhythm: A program for the study of con-
sciousness. In S. A. E. Peper & M. Quinn (Eds.), Mind/body integration: Essential readings in
biofeedback (pp. 289–297). Plenum Press.
Kano, K., Nakamura, M., Matsuoka, T., Iida, H., & Nakajima, T. (1992). The topographical
features of EEGs in patients with affective disorders. Electroencephalography and Clinical
Neurophysiology, 83, 124–129.
Keune, P. M., Bostanov, V., Kotchoubey, B., & Hautzinger, M. (2012). Mindfulness versus ru-
mination and behavioral inhibition: A perspective from research on frontal brain asym-
metry. Personality and Individual Differences, 53(3), 323–328. doi:10.1016/j. paid.2012.03.034
Keune, P. M., Schönenberg, M., Wyckoff, S., Mayer, K., Riemann, S., Hautzinger, M., & Strehl,
U. (2011). Frontal alpha-asymmetry in adults with attention deficit hyperactivity dis-
order: Replication and specification. Biological Psychology, 87, 306–310.
Keune, P. M., van der Heiden, L., Várkuti, B., Konicar, L., Veit, R., & Birbaumer, N. (2012).
Prefrontal brain asymmetry and aggression in imprisoned violent offenders. Neuroscience
Letters, 515, 191–195.
King, J. A., Rosal, M. C., Ma, Y., Reed, G., Kelly, T., Stanek, III, E.J., & Ockene, I. S. (2000).
Sequence and seasonal effects of salivary cortisol. Behavioral Medicine, 26, 67–73.
Kline, J. P., Blackhart, G. C., Woodward, K. M., Williams, S. R., & Schwartz, G. E. (2000).
Anterior electroencephalographic asymmetry changes in elderly women in response
to a pleasant and an unpleasant odor. Biological Psychology, 52(3), 241–250. doi:10.1016/
S0301-0511(99)00046-0
Knoch, D., Gianotti, L. R., Baumgartner, T., & Fehr, E. (2010). A neural marker of costly pun-
ishment behavior. Psychological Science, 21(3), 337–342. doi:10.1177/0956797609360750
Kochanska, G. & Knaack, A. (2003). Effortful control as a personality characteristic of young
children: Antecedents, correlates, and consequences. Journal of Personality, 71(6), 1087–1112.
doi:10.1111/1467-6494.7106008
Kuper, N., Käckenmester, W., & Wacker, J. (2019). Resting frontal EEG asymmetry and person-
ality traits: A meta-analysis. European Journal of Personality, 33(2), 154–175.
Lacey, M. F., Neal, L. B., & Gable, P. A. (2020). Effortful control of motivation, not with-
drawal motivation, relates to greater right frontal asymmetry. International Journal of
Psychophysiology, 147, 18–25.
Lagerspetz, K. M. J. (1969). Aggression and aggressiveness in laboratory mice. In S. Garattini, &
E. B. Sigg (Eds.), Aggressive Behavior (pp. 77–85). Wiley.
Lazarus, R. S. (1991). Emotion and adaptation. Oxford University Press.
Logan, G. D., Schachar, R. J., & Tannock, R. (1997). Impulsivity and inhibitory control.
Psychological Science, 8(1), 60–64. doi:10.1111/j.1467-9280.1997.tb00545.x
Mathersul, D., Williams, L. M., Hopkinson, P. J., & Kemp, A. H. (2008). Investigating models of
affect: Relationships among EEG alpha asymmetry, depression, and anxiety. Emotion, 8(4),
560–572
McFarland, B. R., Shankman, S. A., Tenke, C. E., Bruder, G. E., & Klein, D. N. (2006).
Behavioral activation system deficits predict the six-month course of depression. Journal of
Affective Disorders, 91(2–3), 229–234.
Mechin, N., Gable, P. A., & Hicks, J. A. (2016). Frontal asymmetry and alcohol cue re-
activity: Influence of core personality systems. Psychophysiology, 53(8), 1224–1231. doi:10.1111/
psyp.12659
254 EDDIE HARMON-JONES, TAYLOR POPP, AND PHILIP A. GABLE
Metzger, L. J., Paige, S. R., Carson, M. A., Lasko, N. B., Paulus, L. A., Pitman, R. K., & Orr, S. P.
(2004). PTSD arousal and depression symptoms associated with increased right-sided par-
ietal EEG asymmetry. Journal of Abnormal Psychology, 113(2), 324–329.
Meyer, B., Johnson, S. L., & Carver, C. S. (1999). Exploring behavioral activation and inhibition
sensitivities among college students at risk for bipolar spectrum symptomatology. Journal of
Psychopathology and Behavioral Assessment, 21, 275–292.
Meyer, B., Johnson, S.L., & Winters, R. (2001). Responsiveness to threat and incentive in bi-
polar disorder: Relations of the BIS/BAS scales with symptoms. Journal of Psychopathology
and Behavioral Assessment, 23, 133–143.
Meyer, T., Smeets, T., Giesbrecht, T., Quaedflieg, C. W., Smulders, F. T., Meijer, E. H., &
Merckelbach, H. L. (2015). The role of frontal EEG asymmetry in post-traumatic stress dis-
order. Biological Psychology, 108, 62–77. doi:10.1016/j.biopsycho.2015.03.018
Mitchell, A. M. & Pössel, P. (2012). Frontal brain activity pattern predicts depression in adoles-
cent boys. Biological Psychology, 89(2), 525–527.
Mize, K. D. & Jones, N. A. (2012). Infant physiological and behavioral responses to loss of ma-
ternal attention to a social-rival. International Journal of Psychophysiology, 83(1), 16–23.
doi:10.1016/j.ijpsycho.2011.09.018
Murphy, F. C., Nimmo-Smith, I., & Lawrence, A. D. (2003). Functional neuroanatomy of
emotion: A meta-analysis. Cognitive, Affective, & Behavioral Neuroscience, 3, 207–233.
Myrick, H., Anton, R. F., Li, X., Henderson, S., Drobes, D., Voronin, K., & George, M. S. (2004).
Differential brain activity in alcoholics and social drinkers to alcohol cues: Relationship to
craving. Neuropsychopharmacology, 29(2), 393–402. doi:10.1038/sj.npp.1300295
Nash, K., Inzlicht, M., & McGregor, I. (2012). Approach-related left prefrontal EEG asymmetry
predicts muted error-related negativity. Biological Psychology, 91, 96–102.
Neal, L. B. & Gable, P. A. (2016). Neurophysiological markers of multiple facets of impulsivity.
Biological Psychology, 115, 64–68. doi:10.1016/j.biopsycho.2012.05.005
Neal, L. B. & Gable, P. A. (2019). Shifts in frontal asymmetry underlying impulsive and
controlled decision-making. Biological Psychology, 140, 28–34.
Nigg, J. T. (2006). Temperament and developmental psychopathology. Journal of Child
Psychology and Psychiatry, 47(3-4), 395–422. doi:10.1111/j.1469-7610.2006.01612.x
Nitschke, J. B., Heller, W., Etienne, M. A., & Miller, G. A. (2004). Prefrontal cortex activity
differentiates processes affecting memory in depression. Biological Psychology, 67(1–2),
125–143.
Nusslock, R., Harmon-Jones, E., Alloy, L. B., Urosevic, S., Goldstein, K., & Abramson, L. Y.
(2012). Elevated left mid-frontal cortical activity prospectively predicts conversion to bi-
polar I disorder. Journal of Abnormal Psychology, 121(3), 592–601. doi:10.1037/a0028973
Nusslock, R., Shackman, A. J., Harmon-Jones, E., Alloy, L. B., Coan, J. A., & Abramson, L. Y.
(2011). Cognitive vulnerability and frontal brain asymmetry: Common predictors of first
prospective depressive episode. Journal of Abnormal Psychology, 120, 497–503.
Nusslock, R., Shackman, A. J., McMenamin, B. W., Greischar, L. L., Davidson, R. J., & Kovacs,
M. (2018). Comorbid anxiety moderates the relationship between depression history and
prefrontal EEG asymmetry. Psychophysiology, 55(1), e12953.
Pascual-Marqui, R. D. (2002). Standardized low-resolution brain electromagnetic tomog-
raphy (sLORETA): Technical details. Methods and Findings in Experimental and Clinical
Pharmacology, 24 (Suppl D), 5–12.
Perria, P., Rosadini, G., & Rossi, G. F. (1961). Determination of side of cerebral dominance with
Amobarbital. Archives of Neurology, 4, 175–181.
ASYMMETRIC FRONTAL CORTICAL ACTIVITY 255
Rothbart, M. K., Ellis, L. K., & Posner, M. I. (2004). Temperament and self-regulation. In
K. D. Vohs & R. F. Baumeister (Eds.), Handbook of self-regulation: Research, theory, and
applications (pp. 284–299). Guilford Press.
Sackeim, H., Greenberg, M. S., Weimen, A. L., Gur, R. C., Hungerbuhler, J. P., & Geschwind,
N. (1982). Hemispheric asymmetry in the expression of positive and negative
emotions: Neurologic evidence. Archives of Neurology, 39, 210–218.
Salavert, J., Caseras, X., Torrubia, R., Furest, S., Arranz, B., Dueñas, R., & San, L. (2007). The
functioning of the Behavioral Activation and Inhibition Systems in bipolar I euthymic
patients and its influence in subsequent episodes over an eighteen-month period. Personality
and Individual Differences, 42(7), 1323–1331.
Santesso, D. L., Segalowitz, S. J., Ashbaugh, A. R., Antony, M. M., McCabe, R. E., & Schmidt,
L. A. (2008). Frontal EEG asymmetry and sensation seeking in young adults. Biological
Psychology, 78, 164–172.
Schaffer, C. E., Davidson, R. J., & Saron, C. (1983). Frontal and parietal electroencephalogram
asymmetry in depressed and nondepressed subjects. Biological Psychiatry, 18, 753–762.
Schiff, B. B. & Lamon, M. (1989). Inducing emotion by unilateral contraction of fa-
cial muscles: A new look at hemispheric specialization and the experience of emotion.
Neuropsychologia, 27, 923–935.
Schiff, B. B. & Lamon, M. (1994). Inducing emotion by unilateral contraction of hand muscles.
Cortex, 30, 247–254.
Schmeichel, B. J., Crowell, A., & Harmon-Jones, E. (2016). Exercising self-control increases rela-
tive left frontal cortical activation. Social Cognitive and Affective Neuroscience, 11, 282–288.
Schöne, B., Schomberg, J., Gruber, T., & Quirin, M. (2016). Event-related frontal alpha
asymmetries: Electrophysiological correlates of approach motivation. Experimental Brain
Research, 234(2), 559–567. doi:10.1007/s00221-015-4483-6
Schutter, D. J. L. G. (2009). Transcranial magnetic stimulation. In E. Harmon-Jones & J. S. Beer
(Eds.), Methods in social neuroscience (pp. 233–258). Guilford Press.
Schutter, D. J. L. G. & Harmon-Jones, E. (2013). The corpus callosum: A commissural road to
anger and aggression. Neuroscience & Biobehavioral Reviews, 37, 2481–2488. http://dx.doi.
org/10.1016/j.neubiorev.2013.07.013
Schutter, D. J. L. G., van Honk, J., d’Alfonso, A. A. L., Postma, A., & de Haan, E.H.F. (2001).
Effects of slow rTMS at the right dorsolateral prefrontal cortex on EEG asymmetry and
mood. Neuroreport, 12, 445–447.
Shankman, S. A., Klein, D. N., Tenke, C. E., & Bruder, G. E. (2007). Reward sensitivity in de-
pression: A biobehavioral study. Journal of Abnormal Psychology, 116(1), 95–104
Simons, J. S., Dvorak, R. D., Batien, B. D., & Wray, T. B. (2010). Event-level associations be-
tween affect, alcohol intoxication, and acute dependence symptoms: Effects of urgency,
self-control, and drinking experience. Addictive Behaviors, 35(12), 1045–1053. doi:10.1016/
j.addbeh.2010.07.001
Siniatchkin, M., Kropp, P., & Gerber, W-D. (2000). Neurofeedback—The significance of re-
inforcement and the search for an appropriate strategy for the success of self-regulation.
Applied Psychophysiology and Biofeedback, 25, 167–175.
Smith E. E., Cavanagh, J. F., & Allen, J. J. B. (2018). Intracranial source activity (eLORETA)
related to scalp-level asymmetry scores and depression status. Psychophysiology, 55(1), 119–
134. https://doi.org/10.1111/psyp.13019
Smith, E. E., Zambrano-Vazquez, L., & Allen, J. J. B. (2016). Patterns of alpha asymmetry in
those with elevated worry, trait anxiety, and obsessive-compulsive symptoms: A test of the
ASYMMETRIC FRONTAL CORTICAL ACTIVITY 257
worry and avoidance models of alpha asymmetry. Neuropsychologia, 85, 118–126. https://doi.
org/10.1016/j.neuropsychologia.2016.03.010
Smits, D. J. M. & Kuppens, P. (2005). The relations between anger, coping with anger, and
aggression, and the BIS/BAS system. Personality and Individual Differences, 39, 783–793.
Spielberger, C. D. (1988). State–trait anger expression inventory. Psychological Assessment
Resources.
Spielberger, C. D., Reheiser, E. C., & Sydeman, S. J. (1995). Measuring the experience, expres-
sion, and control of anger. In H. Kassinove (Ed.), Anger disorders: Definition, diagnosis, and
treatment (pp. 49–67). Taylor & Francis.
Stewart, J. L. & Allen, J. J. (2018). Resting frontal brain asymmetry is linked to future depressive
symptoms in women. Biological Psychology, 136, 161–167.
Stewart, J. L., Bismark, A. W., Towers, D. N., Coan, J. A., & Allen, J. J. B. (2010). Resting frontal
EEG asymmetry as an endophenotype for depression risk: Sex-specific patterns of frontal
brain asymmetry. Journal of Abnormal Psychology, 119(3), 502–512.
Stewart, J. L., Coan, J. A., Towers, D. N., & Allen, J. J. B. (2011). Frontal EEG asymmetry during
emotional challenge differentiates individuals with and without lifetime major depres-
sive disorder. Journal of Affective Disorders, 129(1), 167–174. https://doi.org/https://doi.org/
10.1016/j.jad.2010.08.029
Stewart, J. L., Coan, J. A., Towers, D. N., & Allen, J. J. B. (2014). Resting and task-elicited
prefrontal EEG alpha asymmetry in depression: Support for the capability model.
Psychophysiology, 51(5), 446–455. https://doi.org/10.1111/psyp.12191
Strack, F. & Deutsch, R. (2004). Reflective and impulsive determinants of social behavior.
Personality and Social Psychology Review, 8(3), 220–247. doi:10.1207/s15327957pspr0803_1
Sutton, S. K. & Davidson, R. J. (1997). Prefrontal brain asymmetry: A biological substrate of the
behavioral approach and inhibition systems. Psychological Science, 8, 204–210.
Teper, R. & Inzlicht, M. (2013). Meditation, mindfulness and executive control: The import-
ance of emotional acceptance and brain-based performance monitoring. Social, Cognitive,
and Affective Neuroscience, 8(1), 85–92. doi:10.1093/scan/nss045
Terzian, H. & Cecotto, C. (1959). Determination and study of hemisphere dominance by
means of intracarotid sodium amytal injection in man: II. Electroencephalographic effects.
Bollettino della Societa Italiana Sperimentale, 35, 1626–1630.
Thibodeau, R., Jorgensen, R. S., & Kim, S. (2006). Depression, anxiety, and resting frontal
EEG asymmetry: A meta-analytic review. Journal of Abnormal Psychology, 115(4), 715–729.
doi:10.1037/ 0021-843X.115.4.715.
Tomarken, A. J. & Davidson, R. J. (1994). Frontal brain activation in repressors and
nonrepressors. Journal of Abnormal Psychology, 103(2), 339– 349. doi:10.1037/
0021-843X.103.2.339
Tomarken, A. J., Davidson, R. J., Wheeler, R. E., & Doss, R. (1992). Individual differences in an-
terior brain asymmetry and fundamental dimensions of emotion. Journal of Personality and
Social Psychology, 62, 676–687.
Urosevic, S., Abramson, L. Y., Alloy, L. B., Nusslock, R., Harmon-Jones, E., Bender, R., &
Hogan, M. E. (2010). Increased rates of events that activate or deactivate the Behavioral
Approach System, but not events related to goal attainment, in bipolar spectrum disorders.
Journal of Abnormal Psychology, 119, 610–615.
Valiente, C., Eisenberg, N., Smith, C. L., Reiser, M., Fabes, R. A., Losoya, S., ... Murphy, B. C.
(2003). The relations of effortful control and reactive control to children’s externalizing
258 EDDIE HARMON-JONES, TAYLOR POPP, AND PHILIP A. GABLE
OSCI LL ATORY AC T I V I T Y I N
SENSORIMOTOR FU NC T I ON
12.1 Understanding
the sensorimotor system
SMA
M1 S1
Striatum
PMC
GPe
VL/VA GPi
Thal.
STN
Thalamus
Brainstem
Cerebellum
Afferent
Efferent Pathways
Pathways
Spinal
Cord
Figure 12.1 Schematic overview of major cortical and subcortical pathways that comprise the
sensorimotor system. Left: primary and secondary motor cortex project directly to the spinal
cord and via brainstem nuclei. Afferent proprioceptive information travels to primary somato-
sensory cortex via thalamus and enters the cerebellum directly. Right: The basal ganglia con-
sist of the striatum, internal and external pallidum (GPi, GPe), and subthalamic nucleus (STN).
Together with the thalamus they form circuits with various cortical regions, of which the ventral
lateral and ventral anterior thalamus (VL, VA) project to primary and secondary motor cortex.
cord that innervate axial muscles for the regulation of posture and balance, and a lateral
tract innervating distal muscles. The sensorimotor system has a bilateral organization,
with most corticospinal tracts crossing at the level of the medulla. This suggests that
movements of the right arm and leg are primarily controlled by the left hemisphere, and
vice versa. Movements and actions are intricately linked with perception and cognition;
therefore, the sensorimotor system must uphold strong ties with other circuits. As such,
the distinction between systems is not always clear and some sensorimotor regions have
been ascribed more cognitive roles.
As much as one would like to understand how a gymnast performs a twisting
somersault, the study of (supraspinal) neural control of movement in laboratory
settings is typically restricted to what is often referred to as “simple” uni-or bimanual
movements like finger tapping or isometric force generation. Electroencephalography
(EEG), magnetoencephalography (MEG), and electrocorticography (ECoG) are pre-
dominantly sensitive to synchronous synaptic transmission in cortical regions, while
electromyography (EMG) may provide insight into the activation of motoneurons
in the spinal cord via the spread of action potential currents that can be picked up
non-invasively with electrodes placed on the skin over the muscle belly. Increasing
use of deep brain stimulation (DBS) for the treatment of movement disorders allows
for invasively recording local field potentials (LFPs) from basal ganglia and thalamic
OSCILLATORY ACTIVITY IN SENSORIMOTOR FUNCTION 261
M1 STN STN
Relative
90 90 350 log power
Frequency [Hz]
70 70 0.4
250
0.2
50 50
150 0
30 30
–0.2
10 10 50
–0.4
–2 0 2 4 6 –2 0 2 4 6 –2 0 2 4 6
Time from movement onset [s]
Figure 12.2 Movement-related time-frequency spectra for primary motor cortex (M1) and
subthalamic nucleus (STN). Panels show the relative change in spectral power with respect to
a pre-movement baseline window. Participants performed a simple self-paced button press
movement with three fingers simultaneously at time point 0. Beta power starts to decrease be-
fore movement onset (ERD) and shows a clear post-movement rebound (ERS). Gamma power
increases around movement onset. In STN, another movement-related increase with smaller
amplitude can be observed in the range for high-frequency oscillations (HFOs). Note how time-
frequency modulations look very similar for M1 and STN, suggesting generic principles of infor-
mation processing.
Spectra are based on MEG and DBS-LFP recordings from
eleven Parkinson’s disease patients on dopaminergic medication taken from Litvak et al., 2012.
262 Bernadette C. M. van Wijk
Ever since the first recordings, researchers have divided EEG/MEG time series into fre-
quency bands that are presumed to have their own functional roles. This section focuses
on the properties of frequency bands that are thought to be central to sensorimotor
function.
and dominates the functional connectivity profile for sensorimotor cortex. For these
reasons, beta oscillations are considered most associated with motor control.
pyramidal cells but at much sparser rates (Lacey et al., 2014). The dependence of beta
oscillation amplitude on GABAA receptors has also been established through in vivo
pharmacological studies in humans. After administration of benzodiazepines, a class
of GABAergic drugs often prescribed as anticonvulsants, sedatives, or muscle relaxants,
spectral beta power in EEG/MEG appears enhanced (Baker & Baker, 2003; Hall et al.,
2010; Jensen et al., 2005). Along these lines, computational modelling work suggests
that stronger synaptic inputs between pyramidal cells in different layers might underlie
the beta suppression that is observed during movement (Bhatt et al., 2016).
Despite the prominence of beta band oscillations in sensorimotor function,
movement-related modulations only seem to encode general motor aspects. More
forceful movements induce stronger alpha/mu and beta ERD (Mima et al., 1999; Stančák
et al., 1997; Stančák & Pfurtscheller, 1996) as do movements that are performed with a
higher frequency (Toma et al., 2002) or involving more complex sequences (Hummel
et al., 2003; Manganotti et al., 1998). Movements with more muscle mass involved do not
influence ERD but rather lead to stronger beta ERS (Pfurtscheller et al., 1998; Stančák
et al., 2000). Movement duration has little effect on either ERD or ERS (Cassim et al.,
2000; Stančák & Pfurtscheller, 1996). It is however difficult to pinpoint ERD and ERS
patterns to specific motor parameters.
12.2.3 Prokinetic Gamma
To find a mechanistic link between beta oscillations and the control of movement tra-
jectory and muscle force, it can be useful to consider the behavior of individual neurons.
Pyramidal tract neurons (PTNs) in infragranular layers project to motoneurons of in-
dividual muscles and groups of muscles in the spinal cord, and appear to be function-
ally organized in small clusters (Asanuma et al., 1979). They only form a minority of
cells in motor cortex and form intricate connections with other neurons within and
between cortical modules that together encode movement patterns (Keller, 1993).
Microelectrode recordings in sensorimotor cortices of the macaque monkey revealed
spikes of individual neurons to be phase-locked to the LFP beta oscillation during time
periods when the oscillation is well pronounced (Baker et al., 1997; Denker et al., 2007;
Murthy & Fetz, 1996a). Although the phase locking of individual PTNs to the beta os-
cillation may only be weak, summation over a population of neurons can give a clearer
picture (Baker et al., 2003). We also know that the encoding of parameters such as hand
position, direction of motion, velocity, and force emerges from the joint firing rates of
a group of neurons that are individually tuned (Ashe & Georgopoulos, 1994; Fu et al.,
1995; Georgopoulos et al., 1986; Moran & Schwartz, 2017; Paninski et al., 2004). This is
termed population coding.
When movements are executed, phase locking of spikes with the beta oscillation
drops and spike rates strongly increase (Baker et al., 2001; Spinks et al., 2008). There
is a crude inverse correlation between spike rates and LFP beta power. Spike rates
reach frequencies above 30 Hz (Baker et al., 2001; Grammont and Riehle, 2003) and
OSCILLATORY ACTIVITY IN SENSORIMOTOR FUNCTION 265
may well correspond to the brief increases in gamma power (~30–100 Hz) that can
be observed in EEG and MEG recordings in a time window around movement onset
(Cheyne et al., 2008; Muthukumaraswamy, 2010; Ohara et al., 2001; Pfurtscheller et al.,
1993; Pfurtscheller & Neuper, 1992). Unlike the alpha/mu and beta ERD, the increase
in gamma only occurs in the hemisphere contralateral to the moving body part and
is somatotopically more focused (Crone et al., 1998a; Miller et al., 2007; Szurhaj et al.,
2005). Some studies report the gamma amplitude increase to vary with movement dir-
ection in invasive recordings (Leuthardt et al., 2004; Rickert et al., 2005). However, it
remains to be seen whether gamma power in EEG and MEG recordings has the same
specificity. Taken together, one may hypothesize that the beta oscillation inhibits, or
constrains, neuronal spiking. That is, a suppression of the beta rhythm can lead to an in-
crease in excitability of individual neurons that together encode movement parameters.
12.2.4 High-Frequency Oscillations
The latest addition to the repertoire of movement-related oscillations is a spectral
component with a clear peak in the 150–400 Hz frequency range, aptly coined high-
frequency oscillations (HFO). HFO have been observed in LFP recordings from DBS
electrodes in the STN and internal pallidum. They are of interest to sensorimotor
function as they show a characteristic increase in amplitude during movement (Foffani
et al., 2003; Litvak et al., 2012; López-Azcárate et al., 2010; Tan et al., 2013; Tsiokos et al.,
2013), like the gamma band. Their peak frequency is modulated by dopaminergic medi-
cation (López-Azcárate et al., 2010; Özkurt et al., 2011; van Wijk et al., 2016) and indica-
tive of tremor symptoms in patients with Parkinson’s disease (Hirschmann et al., 2016).
Yet, a similar movement-related HFO component in the cortex has not been reported
although this may simply be due to the poor signal-to-noise ratio of commonly used
EEG/MEG compared to more focal, invasive techniques.
Any understanding of sensorimotor function will not be complete if the information ex-
change between regions that comprise the sensorimotor system is ignored. Rhythmic
synchronization in the form of oscillations has been recognized as a means by which
neural populations selectively gate their sensitivity to input from other populations (Fries,
2005). The degree of synchronization between neural populations can be estimated by
computing coherence between time series as a measure of functional connectivity. This
measure can be understood as spectral counterpart of conventional correlation.
266 Bernadette C. M. van Wijk
12.3.1 Cortico-Spinal Coherence
Corticospinal coherence, also called corticomuscular coherence, refers to the syn-
chronization between oscillations in cortex and motoneuron activity in the spinal cord.
Firing of motoneurons yields motor unit action potentials that induce contraction of
muscle fibers. Repetitive firing is necessary in order to build up force. The rate at which
this occurs ranges from 6 Hz in rest to 35 Hz during forceful isometric contractions
and bursts of 80–120 Hz in case of rapid, ballistic movements (Freund, 1983). As such,
corticospinal coherence may reflect whether synchronized activity in cortex also
reaches the muscles.
Several studies report weak but significant beta-band coherence between the EMG of
hand or foot muscles and EEG over contralateral sensorimotor regions during sustained
isometric muscle contractions (Conway et al., 1995; Gross et al., 2000; Halliday et al.,
1998; Salenius et al., 1997; van Wijk et al., 2012). Coherence levels increase for muscle
contractions with higher force levels (Chakarov et al., 2009; Witte et al., 2007) and peak
frequencies shift into the gamma range during maximal force production (Brown et al.,
1998; Mima et al., 1999). Task-dependent coherence has also been found for the EEG
with different muscle groups, thereby forming functional synergies (Zandvoort et al.,
2019), or between EMG recordings of different muscles themselves (Boonstra et al.,
2016). While beta-band corticospinal coherence diminishes during movement (Baker
et al., 1997; Kilner et al., 2000), alpha-band coherence between active muscle groups
increases (Boonstra et al., 2009). Typically, no significant coherence can be observed in
rest unless a strong tremor is present (Hellwig et al., 2001).
Corticospinal beta coherence can be modulated by cognitive factors. The division of
attention during dual task performance decreases coherence levels (Johnson et al., 2011;
Kristeva-Feige et al., 2002; Safri et al., 2007), whereas task instructions to maintain
force output at a target level with high precision lead to increased coherence (Kristeva
et al., 2007). Using a pre-cued choice reaction time task, Van Wijk and colleagues
(2009) demonstrated that corticospinal coherence can be up-regulated in anticipation
of an upcoming movement decision. Key to their experimental paradigm was the pre-
activation of muscles in each trial via a precision grip before stimuli were displayed.
This led to a build-up of corticospinal coherence that allows for modulations to be ob-
servable. After presentation of a warning cue of the likely upcoming response hand,
corticospinal synchronization increased for the non-selected hand, while cortical beta
power decreased for the selected hand. These and other findings jointly suggest that
the role of beta oscillations to facilitate or inhibit movements extends to the level of the
spinal cord.
Does corticospinal coherence reflect mere entrainment of motoneurons by cortical
output? Feedback signals from muscle afferents ascend the spinal cord to somatosen-
sory cortex, which through strong connections with primary motor cortex, closes the
corticospinal loop. Riddle and Baker (2005) sought to tackle this question experimen-
tally by cooling subjects’ arms. Cooling caused an additional time delay (inferred from
phase-frequency regression) that was about twice the conduction time in one direction.
OSCILLATORY ACTIVITY IN SENSORIMOTOR FUNCTION 267
The authors therefore concluded that both ascending and descending pathways con-
tribute to the occurrence of corticospinal coherence. One may determine the direction-
ality of coupling by using measures such as Granger causality. This similarly revealed
significant contributions of both ascending and descending pathways (Witham
et al., 2011). It is worth noting that corticospinal coherence does not strictly follow
modulations of cortical power. Baker and Baker (2003) demonstrated this by recording
subjects before and after intake of diazepam, a benzodiazepine that enhances inhibitory
post-synaptic potentials via GABAA receptors. EEG beta power doubled in amplitude,
but corticospinal coherence was little altered. Interactions between cortex and spinal
cord might therefore be more complex than a simple one-way drive.
12.3.2 Cortico-Subcortical Coherence
With EEG and MEG being most sensitive to cortical sources, it is easy to forget the
contribution of more deeply located brain structures. Many of these are highly rele-
vant for sensorimotor and cognitive functions. DBS treatment allows for invasively
recording electrophysiological activity from subcortical structures in humans. The
subthalamic nucleus (STN) and internal pallidum (GPi) are primary targets for the
treatment of Parkinson’s disease and dystonia, the pedunculopontine nucleus (PPN)
for postural instability and gait freezing, and the ventrolateral thalamus (including
ventral intermediate nucleus) for tremor. It is possible, albeit challenging, to combine
LFP recordings from DBS electrodes with simultaneous EEG or MEG to study cortico-
subcortical interactions.
Early simultaneous LFP-MEG studies sought to map functional connectivity with
cortical regions across different frequency ranges during rest. Litvak and colleagues
(2011) and Hirschmann and colleagues (2011) showed converging evidence for the
presence of two spatially distinct and frequency-specific networks for the STN and
ipsilateral cortical areas: alpha-band coherence with temporoparietal cortex and
beta-band coherence with pre-motor cortex. Coherence values for these sources
were little altered with dopaminergic medication. Using directionality analysis, cor-
tical activity was found to drive STN activity for both frequency bands (Litvak et al.,
2011). Both research labs followed up on these studies by investigating movement-
related coherence but found more conflicting results. Litvak and colleagues (2012)
reported an increase in gamma band coherence during movement execution that was
further increased by dopaminergic mediation. Beta coherence increased during the
post-movement beta rebound period but was unaltered by medication. In a separate
study, they describe a reduction of alpha-band coherence during movement that was
more pronounced with medication (Oswal et al., 2013). By contrast, Hirschmann
and colleagues (2013) found that medication reduced beta-band coherence during
movement but not alpha.
Simultaneous LFP-MEG recordings identified functional networks also for other
subcortical structures. Activity in PPN appeared coherent with that of the brainstem
268 Bernadette C. M. van Wijk
Coherence
0.02
0.01
0
10 20 30 40 50
Right GPi
Frequency [Hz]
Figure 12.3 Beta band coherence between the right internal pallidum (GPi) and ipsilateral
sensorimotor cortex. Left: Topography with darker colors indicating stronger coherence values.
Right: Coherence is reduced during movement compared to rest.
Both panels are based on MEG and DBS-LFP recordings from eight dystonia patients taken from Van Wijk et al., 2017.
and cingulate cortex in the alpha band, and with medial cortical motor areas in the
beta band (Jha et al., 2017). For GPi, coherence was found with temporal cortex in the
theta band, cerebellum in the alpha band, and sensorimotor cortex in the beta band
(Neumann et al., 2015). This study was conducted in dystonia patients, for whom dis-
ease severity significantly correlated only with theta band coherence. Beta-band co-
herence was unrelated to symptoms but showed a movement-related reduction (van
Wijk et al., 2017; see Figure 12.3). The lack of clear alterations of cortico-subcortical co-
herence by medication or correlation with clinical symptoms suggests that this form of
functional coupling might be physiological rather than disease-related. Experimental
paradigms with more refined task designs will be needed to unravel their func-
tional roles.
12.3.3 Cortico-Cortical Coherence
Many motor tasks call upon the coordinated activation of multiple cortical regions.
Researchers have used coherence analysis with the aim to find traces of functional
coupling between EEG/MEG time series recorded at different locations. Some of
these findings confirmed inter-regional coupling patterns as one would expect from
postulated sensorimotor functions of individual brain regions. For example, the ob-
servation of higher coherence values between sensorimotor cortex and mesial pre-
motor areas for internally compared to externally paced movement is in line with the
presumed involvement of SMA in self-initiated movements (Gerloff et al., 1998; Serrien,
2008). Other findings are arguably more open for interpretation, like the increased level
of interhemispheric coherence between sensorimotor regions during more complex
unimanual and bimanual tasks (Gerloff et al., 1998; Gross et al., 2005; Manganotti et al.,
1998; Mima et al., 2000), or the observed gamma band coherence between cerebellar
hemispheres during bimanual finger tapping (Pollok et al., 2007). Notably, cortico-
cortical coherence can be used to reveal the integration of information from sensory
modalities during motor tasks. Classen and colleagues (1998) demonstrated that beta-
band coherence between motor and visual cortex is higher when participants make
OSCILLATORY ACTIVITY IN SENSORIMOTOR FUNCTION 269
use of visual information to perform a visuomotor tracking task compared to when the
visual stimulus is a mere distractor. Similarly, significant alpha-band coherence between
auditory cortices and the motor network can be found when finger tapping is paced by a
metronome (Pollok et al., 2005).
Coherence analysis between EEG/MEG time series is easily prone to volume con-
duction: multiple electrodes or sensors pick up the same neural activity leading to
an overestimation of functional connectivity. This prompted the development of
alternative connectivity measures like the imaginary part of coherency (Nolte et al.,
2004), where the real part of the cross-spectrum is ignored as it may contain volume
conduction artifacts. The authors applied this measure to EEG data of a unimanual
finger tapping task and demonstrated a weak but significant 20 Hz coupling from
contralateral to ipsilateral sensorimotor cortex in the time period before movement
onset and in the reverse direction after movement. Another measure is the phase
lag index (Stam et al., 2007), which eliminates instantaneous coupling by looking at
the asymmetry of the relative phase distribution. Hillebrand and colleagues (2012)
showed that this measure removes volume conduction effects that are still present
after projecting MEG data to source space and identified a strongly connected sen-
sorimotor network in the beta band. The disadvantage of these measures is that any
true coupling with zero time-lag is also ignored, therefore their estimates are on the
conservative side.
Invasive recordings reduce the problem of volume conduction if electrodes can
be placed sufficiently far apart. Murthy and Fetz (1996b) studied the occurrence of
synchronized oscillations in LFP signals from bilateral sensorimotor areas in the ma-
caque monkey. Oscillations occurred spontaneously at rest but only infrequently. They
appeared more often during exploratory arm movements with fine use of the fingers,
during which they were frequently synchronized between M1 and S1, and between bi-
lateral M1s. Both occurred at near zero time-lags. Oscillations, however, did not seem to
occur at consistent points in time related to the movements. Other studies did observe
a systematic increase in synchronous oscillations between M1, S1, and premotor areas
of the same hemisphere or between bilateral M1s around or before movement onset
in more standardized tasks (Cardoso de Oliveira et al., 2001; Ohara et al., 2001; Sanes
& Donoghue, 1993), therefore providing support for the involvement of oscillations in
functional coupling between cortical areas.
Weaver et al., 2009). Besides stimulation, the electrodes can be used to record LFPs,
though recordings are typically limited to a time window of about a week after elec-
trode implantation when wires are still externalized. Such recordings have revealed
elevated levels of beta oscillations in patients off medication during rest (Figure 12.4;
e.g., Brown et al., 2001; Levy et al., 2002; Priori et al., 2004), and some studies report less-
pronounced beta ERD during self-paced movement (Doyle et al., 2005a).
One of the challenges with invasive recordings in patients is to determine the degree
to which observed activities are physiological or disease-related. Acquiring the same
recordings from healthy subjects for comparison is not an option as they would need to
undergo the same surgical procedures. Still there are strong reasons to believe that ex-
cessive beta oscillations in the STN are a marker of Parkinsonism:
1. The amplitude of beta oscillations correlates with severity of bradykinesia and ri-
gidity as measured with the Unified Parkinson’s Disease Rating Scale (Neumann
et al., 2016; van Wijk et al., 2016).
2. The amplitude of beta oscillations decreases after dopaminergic medication or
DBS in parallel with clinical improvement (Kühn et al., 2009, 2008, 2006b; Ray
et al., 2008).
3. The amplitude of beta oscillations in STN and GPi is higher in patients with
Parkinson’s disease than in patients with dystonia (Piña-Fuentes et al., 2019).
4. Stimulation at 20 Hz instead of the clinically effective 130 Hz impedes motor per-
formance (Chen et al., 2007), therefore underscoring a causal role of excessive
beta oscillations in motor impairment.
While bradykinesia and rigidity are associated with beta oscillations, tremor typ-
ically manifests at frequencies below 12 Hz. Several forms of physiological and patho-
logical tremors with central or peripheral origins can be distinguished, all characterized
by involuntary rhythmic movements in one or more body parts (Deuschl et al., 2001;
McAuley & Marsden, 2000). Essential tremor is the most common movement disorder,
0 150 150 0
10 20 30 40 50 5 10 15 20 25 30 35 5 10 15 20 25 30 35
Frequency [Hz] Phase Freq [Hz] Phase Freq [Hz]
Figure 12.4 Abnormal subthalamic nucleus oscillatory activity in Parkinson’s disease.
Left: Spectral power in the beta frequency range is elevated when the patient is withdrawn from
dopaminergic medication. Right: Phase-amplitude coupling (PAC) between beta and high-
frequency oscillations is another marker of motor impairment.
Example of single subject DBS-LFP recordings taken from Van Wijk et al., 2016.
OSCILLATORY ACTIVITY IN SENSORIMOTOR FUNCTION 273
an action tremor that occurs with voluntary muscle contraction. Neural entrainment
between inferior olive, cerebellum, thalamus, and motor cortex is believed to underlie
its emergence (Raethjen and Deuschl, 2012). Activity at the tremor frequency within
this network has been found to be coherent with EMG (Hellwig et al., 2001; Schnitzler
et al., 2009). By contrast, Parkinsonian tremor presents during rest and may have its
origin in the basal ganglia (Deuschl et al., 2001). Both Parkinsonian and essential
tremor can be suppressed by 130-Hz DBS in the ventrolateral thalamus (Benabid et al.,
1996). Stimulation at the tremor frequency could either enhance or reduce tremor
amplitude, depending on the phase at which stimulation pulses are delivered (Cagnan
et al., 2013).
Dystonia is another hyperkinetic movement disorder that can be associated with
increased power for frequencies below the beta band in GPi (Chen et al., 2006; Liu et al.,
2008; Sharott et al., 2008; Silberstein et al., 2003). This arguably reflects diminished and
more irregular neuronal firing that is considered distinctive of the disorder (Hendrix
& Vitek, 2012). Clinical symptoms include twisting movements and abnormal posture
resulting from involuntary sustained and sometimes repetitive muscle contractions
(Fahn, 1988). Also in dystonia symptoms might be suppressed via DBS but improve-
ment may not become evident before several weeks or months of continuous stimula-
tion (Vidailhet et al., 2005). Barow et al. (2014) showed that 4–12 Hz pallidal activity is
reduced upon DBS with more immediate effects for phasic compared to tonic dystonia
subtypes.
Elevated levels of oscillatory activity seem to be a common feature of a number of
movement disorders. Effective treatments are frequently associated with a reduction of
this activity although causal relations are often difficult to establish.
Until now, averaging of spectral power within frequency bands or across long time
windows has been common practice in the field. Recently, dynamical properties of
oscillations have received more attention. For example, beta oscillations are not present
with constant amplitude but appear in short bursts of varying amplitude and duration
(Feingold et al., 2015; Murthy and Fetz, 1996b). In Parkinson’s disease, bursts with long
duration occur more frequently in the STN of patients with larger clinical impair-
ment and also have a higher amplitude than short bursts (Tinkhauser et al., 2017a).
Dopaminergic medication significantly reduces the number of long bursts (Tinkhauser
et al., 2017b). These new insights help to further constrain computational models, like
that proposed by Sherman and colleagues (2016) for S1, which explore the synaptic
mechanisms by which beta bursts emerge.
Another example is the notion that functions may arise from the interaction be-
tween frequency bands as opposed to individual frequency bands forming parallel
communication channels. Using recordings from rat motor cortex, Igarashi et al. (2013)
274 Bernadette C. M. van Wijk
demonstrated gamma oscillations comprised of a slow and fast component that were
both coupled with an ongoing theta rhythm. Slow gamma oscillations were pronounced
when rats were holding on to a lever and were phase-locked to peaks of the theta oscilla-
tion, whereas fast gamma oscillations emerged around the time of pulling the lever and
were phase-locked to a trough of the theta oscillation. These findings are reminiscent of
phase precession in hippocampal cortex where firing of place cells in the gamma range is
nested in the ongoing theta rhythm. With individual gamma cycles representing a par-
ticular location in space, these progressively shift forward to earlier phases of the theta
cycle as the rat navigates towards that location (O’Keefe & Recce, 1993). The nesting of
oscillations allows for the encoding of near and far locations, i.e. information that is not
contained in the individual frequencies. On the other hand, cross-frequency coupling
might also arise due to pathology. We have found the strength of phase-amplitude coup-
ling between beta and HFOs in the STN to correlate with the severity of bradykinesia
and rigidity symptoms in Parkinson’s disease (Figure 12.4; van Wijk et al., 2016).
Cross-frequency coupling could also be indicative of non-sinusoidal waveforms as
these would appear at higher frequencies in the spectrum (van Wijk, 2017). Cole et al.
(2017) demonstrated that the waveform shape of beta oscillations could indeed explain
beta-gamma phase-amplitude coupling patterns observed in M1 of Parkinson’s patients
(de Hemptinne et al., 2013). It is interesting that DBS reduces the asymmetry of the beta
oscillation waveform (de Hemptinne et al., 2015) as it hints at less-synchronous synaptic
input from the basal ganglia. A focus on the detailed time dynamics, cross-frequency
coupling, and waveform shape of oscillations holds great promise to further unravel
their role in sensorimotor function.
We move to interact with the external world in a meaningful way. Our actions depend
on goals and intentions, which in turn depend on the context we are in. Cognition and
motor control are highly intertwined. Several lines of research indeed suggest that sen-
sorimotor regions contribute beyond the classical view on motor control.
First of all, (pre-)motor cortex is already activated during the process of action selec-
tion instead of merely being informed on the final outcome that needs to be executed
(Cisek and Kalaska, 2005). Donner and colleagues (2009) demonstrated that lateralized
beta suppression and gamma increase in motor cortex could predict on a single trial
level which left/right choice participants were going to make during a perceptual de-
tection task. Instructions to emphasize decision speed versus accuracy decrease pre-
stimulus M1 beta levels, which might explain the upsurge of errors under speed stress
(Pastötter et al., 2012; Steinemann et al., 2018). Pre-decision activation of the motor
system is therefore likely to speed-up or even influence behavioral performance.
OSCILLATORY ACTIVITY IN SENSORIMOTOR FUNCTION 275
12.7 Conclusions
Oscillations are omnipresent in the sensorimotor system and show distinct modulations
with experimental tasks and pathology. In particular beta oscillations seem to have a
prominent role by means of their association with movement facilitation and inhib-
ition. Movement-related beta ERD and ERS are among the most robust time-frequency
276 Bernadette C. M. van Wijk
patterns in EEG/MEG studies. Intriguingly, the cortex is not alone in displaying these
characteristic modulations. Very similar movement-related beta ERD/ERS and gamma
ERS can be observed in LFP recordings from the STN (Alegre et al., 2005; Androulidakis
et al., 2007b; Cassidy et al., 2002; Kühn et al., 2004; Litvak et al., 2012), GPi (Brücke et al.,
2008; Talakoub et al., 2016; Tsang et al., 2012), and the ventral lateral thalamus (Brücke
et al., 2013; Klostermann et al., 2007; Paradiso et al., 2004). This implies that these
modulations are generic principles of information processing that might have different
functional meaning in different regions of the sensorimotor network.
Over the years, the main focus of the field at large has been on oscillations in contra-
lateral M1. While simple flexion/extension movements of the fingers indeed seem to
be fairly restricted to activation in the primary sensorimotor cortices, more complex
rhythmic or bimanual movement patterns recruit additional regions like ipsilateral M1,
SMA, premotor cortex, cerebellum, and also primary and secondary sensory and asso-
ciation cortices if the task involves strong visual or auditory components (Heinrichs-
Graham & Wilson, 2015; Houweling et al., 2008; Hummel et al., 2003; Pollok et al., 2005;
Rueda-Delgado et al., 2014). The contribution of subcortical structures is more difficult
to study in humans but should not be overlooked. Intriguingly, lesions to motor cortex
do not affect execution of a task-specific motor sequence in rats once the sequence
has been learned, suggesting reliance on subcortical controllers with projections to
the spinal cord (Kawai et al., 2015). Simultaneous MEG and DBS-LFP recordings have
revealed a number of spatially and spectrally distinct cortico-subcortical networks that
appear to be disease-unrelated. The functional relevance of these networks deserves fur-
ther exploration in future studies.
Although we have learned a great deal about the neural control of movement by
studying oscillations, several aspects remain unexplained. Important questions that are
still open are mostly of mechanistic nature. For example, how does beta ERD lead to
muscle activations? What causes beta ERD to start and initiate movement in the first
place? Why is desynchronization so widespread even for simple finger movement?
Why are movement-related modulations so similar across various parts of the motor
system? How are movement plans translated into motor commands? How are they
encoded? Oscillations reflect the summed activity of numerous neurons, from which
it might be difficult to infer details of individual muscle control especially with non-
invasive techniques such as EEG and MEG. Instead, they are more likely to reflect gen-
eral states of activation or deactivation. I believe there is still much more to gain from
studying oscillations through the combination of recordings techniques, the use of
advanced signal processing algorithms and the development of computational models.
Fortunately, grasping a cup of coffee is much easier than understanding how we do it.
REFERENCES
Adrian, E. D. & Matthews, B. H. C. (1934). The Berger rhythm: Potential changes from the oc-
cipital lobes in man. Brain, 57, 355–385. https://doi.org/10.1093/brain/57.4.355
OSCILLATORY ACTIVITY IN SENSORIMOTOR FUNCTION 277
Alayrangues, J., Torrecillos, F., Jahani, A., & Malfait, N. (2019). Error-related modulations of
the sensorimotor post-movement and foreperiod beta-band activities arise from distinct
neural substrates and do not reflect efferent signal processing. NeuroImage 184, 10–24.
https://doi.org/10.1016/j.neuroimage.2018.09.013
Alegre, M., Alonso-Frech, F., Rodríguez-Oroz, M. C., Guridi, J., Zamarbide, I., Valencia,
M., . . . Artieda, J. (2005). Movement-related changes in oscillatory activity in the human
subthalamic nucleus: Ipsilateral vs. contralateral movements. European Journal of
Neuroscience, 22, 2315–2324. https://doi.org/10.1111/j.1460-9568.2005.04409.x
Alegre, M., Labarga, A., Gurtubay, I., Iriarte, J., Malanda, A., & Artieda, J. (2003). Movement-
related changes in cortical oscillatory activity in ballistic, sustained and negative movements.
Experimental Brain Research, 148, 17–25. https://doi.org/10.1007/s00221-002-1255-x
Andres, F. G., Mima, T., Schulman, A. E., Dichgans, J., Hallett, M., & Gerloff, C. (1999).
Functional coupling of human cortical sensorimotor areas during bimanual skill acquisi-
tion. Brain, 122, 855–870. https://doi.org/10.1093/brain/122.5.855
Androulidakis, A. G., Doyle, L. M. F., Yarrow, K., Litvak, V., Gilbertson, T. P., & Brown, P.
(2007a). Anticipatory changes in beta synchrony in the human corticospinal system and
associated improvements in task performance. European Journal of Neuroscience, 25, 3758–
3765. https://doi.org/10.1111/j.1460-9568.2007.05620.x
Androulidakis, A. G., Kühn, A. A., Chen, C. C., Blomstedt, P., Kempf, F., Kupsch, A., . . . Brown,
P. (2007b). Dopaminergic therapy promotes lateralized motor activity in the subthalamic
area in Parkinson’s disease. Brain, 130, 457–68. https://doi.org/10.1093/brain/awl358
Aron, A. R. (2011). From reactive to proactive and selective control: Developing a richer model
for stopping inappropriate responses. Biological Psychiatry, 69, e55–e68. https://doi.org/
10.1016/j.biopsych.2010.07.024
Aron, A. R., Herz, D. M., Brown, P., Forstmann, B. U., & Zaghloul, K. (2016). Frontosubthalamic
circuits for control of action and cognition. Journal of Neuroscience, 36, 11489–11495. https://
doi.org/10.1523/JNEUROSCI.2348-16.2016
Asanuma, H., Zarzecki, P., Jankowska, E., Hongo, T., & Marcus, S. (1979). Projection of indi-
vidual pyramidal tract neurons to lumbar motor nuclei of the monkey. Experimental Brain
Research, 34, 73–89. https://doi.org/103742
Ashe, J. & Georgopoulos, A. P. (1994). Movement parameters and neural activity in motor
cortex and area 5. Cerebral Cortex, 4, 590–600. https://doi.org/10.1093/cercor/4.6.590
Avanzini, P., Fabbri-Destro, M., Dalla Volta, R., Daprati, E., Rizzolatti, G., & Cantalupo,
G. (2012). The dynamics of sensorimotor cortical oscillations during the observation of
hand movements: An EEG study. PLoS One, 7, e37534. https://doi.org/10.1371/journal.
pone.0037534
Babiloni, C., Babiloni, F., Carducci, F., Cincotti, F., Cocozza, G., Del Percio, C., Moretti, D. V.,
& Rossini, P. M. (2002). Human cortical electroencephalography (EEG) rhythms during the
observation of simple aimless movements: A high-resolution EEG study. NeuroImage, 17,
559–572. https://doi.org/10.1006/nimg.2002.1192
Baker, M. R. & Baker, S. N. (2003). The effect of diazepam on motor cortical oscillations and
corticomuscular coherence studied in man. Journal of Physiology, 546, 931–942. https://doi.
org/10.1113/jphysiol.2002.029553
Baker, S. N., Olivier, E., & Lemon, R. N. (1997). Coherent oscillations in monkey motor cortex
and hand muscle EMG show task-dependent modulation. Journal of Physiology, 501, 225–
241. https://doi.org/10.1111/j.1469-7793.1997.225bo.x
278 Bernadette C. M. van Wijk
Baker, S. N., Pinches, E. M., & Lemon, R. N. (2003). Synchronization in monkey motor cortex
during a precision grip task. II. Effect of oscillatory activity on corticospinal output. Journal
of Neurophysiology, 89, 1941–1953. https://doi.org/10.1152/jn.00832.2002
Baker, S. N., Spinks, R., Jackson, A., & Lemon, R. N. (2001). Synchronization in monkey motor
cortex during a precision grip task. I. Task-dependent modulation in single-unit synchrony.
Journal of Neurophysiology, 85, 869–885. https://doi.org/10.1152/jn.00832.2002
Barow, E., Neumann, W. J., Brücke, C., Huebl, J., Horn, A., Brown, P., . . . Kühn, A.A., 2014.
Deep brain stimulation suppresses pallidal low frequency activity in patients with phasic
dystonic movements. Brain, 137, 3012–3024. https://doi.org/10.1093/brain/awu258
Benabid, A. L., Pollak, P., Gao, D., Hoffmann, D., Limousin, P., Gay, E., Payen, I., & Benazzouz,
A. (1996). Chronic electrical stimulation of the ventralis intermedius nucleus of the thal-
amus as a treatment of movement disorders. Journal of Neurosurgery, 84, 203–214. https://
doi.org/10.3171/jns.1996.84.2.0203
Berger, H. (1929). Über das Elektroenkephalogramm des Menschen. Archiv für Psychiatrie und
Nervenkrankheiten, 87, 527–570.
Bhatt, M. B., Bowen, S., Rossiter, H. E., Dupont-Hadwen, J., Moran, R.J., Friston, K.J., Ward,
N.S., 2016. Computational modelling of movement- related beta-
oscillatory dynamics
in human motor cortex. Neuroimage 133, 224–232. https://doi.org/10.1016/j.neuroim
age.2016.02.078
Blankertz, B., Dornhege, G., Krauledat, M., Müller, K.-R., & Curio, G. (2007). The non-invasive
Berlin Brain–Computer Interface: Fast acquisition of effective performance in untrained
subjects. NeuroImage, 37, 539–550. https://doi.org/10.1016/j.neuroimage.2007.01.051
Boonstra, T. W., Daffertshofer, A., Breakspear, M., & Beek, P. J. (2007). Multivariate time–
frequency analysis of electromagnetic brain activity during bimanual motor learning.
NeuroImage, 36, 370–377. https://doi.org/10.1016/j.neuroimage.2007.03.012
Boonstra, T. W., Danna-Dos-Santos, A., Xie, H.-B., Roerdink, M., Stins, J. F., & Breakspear,
M. (2016). Muscle networks: Connectivity analysis of EMG activity during postural control.
Scientific Reports, 5, 17830. https://doi.org/10.1038/srep17830
Boonstra, T. W., van Wijk, B. C. M., Praamstra, P., & Daffertshofer, A. (2009). Corticomuscular
and bilateral EMG coherence reflect distinct aspects of neural synchronization. Neuroscience
Letters, 463, 17–21. https://doi.org/10.1016/j.neulet.2009.07.043
Brown, P., Oliviero, A., Mazzone, P., Insola, A., Tonali, P., & Di Lazzaro, V. (2001). Dopamine
dependency of oscillations between subthalamic nucleus and pallidum in Parkinson’s
disease. Journal of Neuroscience, 21, 1033–1038. https://doi.org/10.1523/JNEURO
SCI.21-03-01033.2001
Brown, P., Salenius, S., Rothwell, J. C., & Hari, R. (1998). Cortical correlate of the Piper
rhythm in humans. Journal of Neurophysiology, 80, 2911–2917. https://doi.org/10.1152/
jn.1998.80.6.2911
Brücke, C., Bock, A., Huebl, J., Krauss, J. K., Schönecker, T., Schneider, G. H., Brown, P., &
Kühn, A. A. (2013). Thalamic gamma oscillations correlate with reaction time in a go/no-go
task in patients with essential tremor. NeuroImage, 75, 36–45. https://doi.org/10.1016/j.neu
roimage.2013.02.038
Brücke, C., Kempf, F., Kupsch, A., Schneider, G. H., Krauss, J.K., Aziz, T., . . . Kühn, A. A.
(2008). Movement-related synchronization of gamma activity is lateralized in patients
with dystonia. European Journal of Neuroscience, 27, 2322–2329. https://doi.org/10.1111/
j.1460-9568.2008.06203.x
OSCILLATORY ACTIVITY IN SENSORIMOTOR FUNCTION 279
Cagnan, H., Brittain, J.-S., Little, S., Foltynie, T., Limousin, P., Zrinzo, L., . . . Brown, P. (2013).
Phase dependent modulation of tremor amplitude in essential tremor through thalamic
stimulation. Brain, 136, 3062–3075. https://doi.org/10.1093/brain/awt239
Cannon, E. N., Yoo, K. H., Vanderwert, R. E., Ferrari, P. F., Woodward, A.L., & Fox, N. A. (2014).
Action experience, more than observation, influences mu rhythm desynchronization. PLoS
One, 9, e92002. https://doi.org/10.1371/journal.pone.0092002
Cardoso de Oliveira, S., Gribova, A., Donchin, O., Bergman, H., & Vaadia, E. (2001). Neural
interactions between motor cortical hemispheres during bimanual and unimanual arm
movements. European Journal of Neuroscience, 14, 1881–1896. https://doi.org/10.1046/
j.0953-816X.2001.01801.x
Cassidy, M., Mazzone, P., Oliviero, A., Insola, A., Tonali, P., Di Lazzaro, V., & Brown, P. (2002).
Movement-related changes in synchronization in the human basal ganglia. Brain, 125, 1235–
1246. https://doi.org/10.1093/brain/awf135
Cassim, F., Monaca, C., Szurhaj, W., Bourriez, J.-L., Defebvre, L., Derambure, P., & Guieu,
J.-D. (2001). Does post-movement beta synchronization reflect an idling motor cortex?
Neuroreport, 12, 3859–3863. https://doi.org/10.1097/00001756-200112040-00051
Cassim, F., Szurhaj, W., Sediri, H., Devos, D., Bourriez, J.-L., Poirot, I., . . . Guieu, J.-D.
(2000). Brief and sustained movements: Differences in event-related (de)synchronization
(ERD/ERS) patterns. Clinical Neurophysiology, 111, 2032–2039. https://doi.org/10.1016/
S1388-2457(0 0)00455-7
Cattaneo, L. & Rizzolatti, G. (2009). The mirror neuron system. Archives of Neurology, 66, 557–
560. https://doi.org/10.1001/archneurol.2009.41
Chakarov, V., Naranjo, J.R., Schulte-Monting, J., Omlor, W., Huethe, F., & Kristeva, R. (2009).
Beta-range EEG-EMG coherence with isometric compensation for increasing modulated low-
level forces. Journal of Neurophysiology, 102, 1115–1120. https://doi.org/10.1152/jn.91095.2008
Chatrian, G. E., Petersen, M. C., & Lazarte, J. A. (1959). The blocking of the Rolandic wicket
rhythm and some central changes related to movement. Electroencephalography and Clinical
Neurophysiology, 11, 497–510. https://doi.org/10.1016/0013-4694(59)90048-3
Chen, C.C., Kühn, A.A., Hoffmann, K.T., Kupsch, A., Schneider, G.H., Trottenberg,
T., . . . Brown, P. (2006). Oscillatory pallidal local field potential activity correlates with in-
voluntary EMG in dystonia. Neurology, 66, 418–420. https://doi.org/10.1212/01.wnl.0000196
470.00165.7d
Chen, C. C., Litvak, V., Gilbertson, T., Kühn, A., Lu, C., Lee, S., . . . Hariz, M. (2007). Excessive
synchronization of basal ganglia neurons at 20 Hz slows movement in Parkinson’s disease.
Experimental Neurology, 205, 214–221. https://doi.org/10.1016/j.expneurol.2007.01.027
Chen, R., Yaseen, Z., Cohen, L. G., & Hallett, M. (1998). Time course of corticospinal excit-
ability in reaction time and self-paced movements. Annals of Neurology, 44, 317–325. https://
doi.org/10.1002/ana.410440306
Cheyne, D., Bells, S., Ferrari, P., Gaetz, W., & Bostan, A. C. (2008). Self-paced movements in-
duce high-frequency gamma oscillations in primary motor cortex. NeuroImage, 42, 332–342.
https://doi.org/10.1016/j.neuroimage.2008.04.178
Cheyne, D., Gaetz, W., Garnero, L., Lachaux, J.-P., Ducorps, A., Schwartz, D., & Varela, F. J.
(2003). Neuromagnetic imaging of cortical oscillations accompanying tactile stimulation.
Cognitive Brain Research, 17, 599–611. https://doi.org/10.1016/S0926-6410(03)00173-3
Cisek, P. & Kalaska, J. F. (2005). Neural correlates of reaching decisions in dorsal premotor
cortex: specification of multiple direction choices and final selection of action. Neuron, 45,
801–814. https://doi.org/10.1016/j.neuron.2005.01.027
280 Bernadette C. M. van Wijk
Classen, J., Gerloff, C., Honda, M., & Hallett, M. (1998) Integrative visuomotor behavior
is associated with interregionally coherent oscillations in the human brain. Journal of
Neurophysiology, 79, 1567–1573. https://doi.org/10.1152/jn.1998.79.3.1567
Cochin, S., Barthelemy, C., Lejeune, B., Roux, S., & Martineau, J. (1998). Perception of motion
and qEEG activity in human adults. Electroencephalography and Clinical Neurophysiology,
107, 287–295. https://doi.org/10.1016/S0013-4694(98)00071-6
Cole, S. R., van der Meij, R., Peterson, E. J., de Hemptinne, C., Starr, P. A., & Voytek, B. (2017).
Nonsinusoidal beta oscillations reflect cortical pathophysiology in Parkinson’s disease.
Journal of Neuroscience, 37, 2208–2216. https://doi.org/10.1523/JNEUROSCI.2208-16.2017
Conway, B. A., Halliday, D. M., Farmer, S. F., Shahani, U., Maas, P., Weir, A. I., & Rosenberg, J.
R. (1995). Synchronization between motor cortex and spinal motoneuronal pool during the
performance of a maintained motor task in man. Journal of Physiology, 489, 917–924. https://
doi.org/10.1113/jphysiol.1995.sp021104
Crone, N. E., Miglioretti, D. L., Gordon, B., & Lesser, R. P. (1998a). Functional mapping of
human sensorimotor cortex with electrocorticographic spectral analysis. II. Event-related
synchronization in the gamma band. Brain, 121, 2301–2315. https://doi.org/10.1093/brain/
121.12.2301
Crone, N. E., Miglioretti, D. L., Gordon, B., Sieracki, J. M., Wilson, M. T., Uematsu, S., & Lesser,
R. P. (1998b). Functional mapping of human sensorimotor cortex with electrocorticographic
spectral analysis. I. Alpha and beta event-related desynchronization. Brain, 121, 2271–2299.
https://doi.org/10.1093/brain/121.12.2271
de Hemptinne, C., Ryapolova-Webb, E. S., Air, E. L., Garcia, P. A., Miller, K. J., Ojemann, J.
G.,. . . Starr, P. A. (2013). Exaggerated phase-amplitude coupling in the primary motor cortex
in Parkinson disease. Proceedings of the National Academy of Sciences of the United States of
America, 110, 4780–4785. https://doi.org/10.1073/pnas.1214546110
de Hemptinne, C., Swann, N. C., Ostrem, J. L., Ryapolova-Webb, E. S., San Luciano, M.,
Galifianakis, N. B., & Starr, P. A. (2015). Therapeutic deep brain stimulation reduces cortical
phase-amplitude coupling in Parkinson’s disease. Nature Neuroscience, 18, 779–786. https://
doi.org/10.1038/nn.3997
Denker, M., Roux, S., Timme, M., Riehle, A., & Grün, S. (2007). Phase synchronization between
LFP and spiking activity in motor cortex during movement preparation. Neurocomputing,
70, 2096–2101. https://doi.org/10.1016/j.neucom.2006.10.088
Deuschl, G., Raethjen, J., Lindemann, M., & Krack, P. (2001). The pathophysiology of tremor.
Muscle Nerve, 24, 716–735. https://doi.org/10.1002/mus.1063
Deuschl, G., Schade-Brittinger, C., Krack, P., Volkmann, J., Schäfer, H., Bötzel, K., . . . Voges, J.
(2006). A randomized trial of deep-brain stimulation for Parkinson’s disease. New England
Journal of Medicine, 355, 896–908. https://doi.org/10.1056/nejmoa060281
Donner, T. H., Siegel, M., Fries, P., & Engel, A. K. (2009). Buildup of choice-predictive activity
in human motor cortex during perceptual decision making. Current Biology, 19, 1581–1585.
https://doi.org/10.1016/j.cub.2009.07.066
Doyle, L. M. F., Kühn, A. A., Hariz, M., Kupsch, A., Schneider, G. H., & Brown, P. (2005a).
Levodopa- induced modulation of subthalamic beta oscillations during self- paced
movements in patients with Parkinson’s disease. European Journal of Neuroscience,. 21, 1403–
1412. https://doi.org/10.1111/j.1460-9568.2005.03969.x
Doyle, L. M. F., Yarrow, K., & Brown, P. (2005b). Lateralization of event-related beta desyn-
chronization in the EEG during pre-cued reaction time tasks. Clinical Neurophysiology, 116,
1879–1888. https://doi.org/10.1016/j.clinph.2005.03.017
OSCILLATORY ACTIVITY IN SENSORIMOTOR FUNCTION 281
Erbil, N. & Ungan, P. (2007). Changes in the alpha and beta amplitudes of the central EEG
during the onset, continuation, and offset of long-duration repetitive hand movements.
Brain Research, 1169, 44–56. https://doi.org/10.1016/j.brainres.2007.07.014
Espenhahn, S., de Berker, A. O., van Wijk, B. C. M., Rossiter, H. E., & Ward, N. S. (2017).
Movement-related beta oscillations show high intra-individual reliability. NeuroImage, 147,
175–185. https://doi.org/10.1016/j.neuroimage.2016.12.025
Espenhahn, S., van Wijk, B. C. M., Rossiter, H. E., de Berker, A. O., Redman, N. D., Rondina,
J., Diedrichsen, J., & Ward, N. S. (2019). Cortical beta oscillations are associated with motor
performance following visuomotor learning. NeuroImage, 195, 340–353. https://doi.org/
10.1016/j.neuroimage.2019.03.079
Fahn, S. (1988). Concept and classification of dystonia. Advances in Neurology, 50, 1–8.
Feingold, J., Gibson, D. J., DePasquale, B., & Graybiel, A. M. (2015). Bursts of beta oscilla-
tion differentiate postperformance activity in the striatum and motor cortex of monkeys
performing movement tasks. Proceedings of the National Academy of Sciences of the United
States of America, 112, 13687–92. https://doi.org/10.1073/pnas.1517629112
Foffani, G., Priori, A., Egidi, M., Rampini, P., Tamma, F., Caputo, E., . . . Barbieri, S. (2003). 300-
Hz subthalamic oscillations in Parkinson’s disease. Brain, 126, 2153–2163. https://doi.org/
10.1093/brain/awg229
Freund, H. J. (1983). Motor unit and muscle activity in voluntary motor control. Physiological
Reviews, 63, 387–436. https://doi.org/10.1152/physrev.1983.63.2.387
Fries, P. (2005). A mechanism for cognitive dynamics: Neuronal communication through
neuronal coherence. Trends in Cognitive Science, 9, 474–480. https://doi.org/10.1016/
j.tics.2005.08.011
Fu, Q. G., Flament, D., Coltz, J. D., & Ebner, T. J. (1995). Temporal encoding of movement
kinematics in the discharge of primate primary motor and premotor neurons. Journal of
Neurophysiology, 73, 836–854. https://doi.org/10.1152/jn.1995.73.2.836
Gaetz, W., MacDonald, M., Cheyne, D., & Snead, O. C. (2010). Neuromagnetic imaging of
movement-related cortical oscillations in children and adults: Age predicts post-movement
beta rebound. NeuroImage, 51, 792–807. https://doi.org/10.1016/j.neuroimage.2010.01.077
Gastaut, H. (1952). Etude électrocorticographique de la réactivité des rythmes rolandiques.
Revue neurologique, 87, 176–82.
Georgopoulos, A., Schwartz, A., & Kettner, R. (1986). Neuronal population coding of
movement direction. Science, 233, 1416–1419. https://doi.org/10.1126/science.3749885
Gerloff, C., Richard, J., Hadley, J., Schulman, A.E., Honda, M., Hallett, M., 1998. Functional
coupling and regional activation of human cortical motor areas during simple, internally
paced and externally paced finger movements. Brain 121, 1513–1531. https://doi.org/10.1093/
brain/121.8.1513
Gilbertson, T., Lalo, E., Doyle, L., Lazzaro, V. Di, Cioni, B., & Brown, P. (2005). Existing motor
state is favored at the expense of new movement during 13–35 Hz oscillatory synchrony in
the human corticospinal system. Journal of Neuroscience, 25, 7771–7779. https://doi.org/
10.1523/JNEUROSCI.1762-05.2005
Grammont, F. & Riehle, A. (2003). Spike synchronization and firing rate in a population of
motor cortical neurons in relation to movement direction and reaction time. Biological
Cybernetics, 88, 360–373. https://doi.org/10.1007/s00422-002-0385-3
Gross, J., Pollok, B., Dirks, M., Timmermann, L., Butz, M., & Schnitzler, A. (2005). Task-
dependent oscillations during unimanual and bimanual movements in the human primary
282 Bernadette C. M. van Wijk
motor cortex and SMA studied with magnetoencephalography. NeuroImage, 26, 91–98.
https://doi.org/10.1016/j.neuroimage.2005.01.025
Gross, J., Tass, P. A., Salenius, S., Hari, R., Freund, H. J., & Schnitzler, A. (2000). Cortico-
muscular synchronization during isometric muscle contraction in humans as revealed
by magnetoencephalography. Journal of Physiology, 527, 623–631. https://doi.org/10.1111/
j.1469-7793.2000.00623.x
Hall, S. D., Barnes, G. R., Furlong, P. L., Seri, S., & Hillebrand, A. (2010). Neuronal network
pharmacodynamics of GABAergic modulation in the human cortex determined using
pharmaco-magnetoencephalography. Human Brain Mapping, 31, 581–594. https://doi.org/
10.1002/hbm.20889
Halliday, D. M., Conway, B. A., Farmer, S. F., and Rosenberg, J. R. (1998). Using electroen-
cephalography to study functional coupling between cortical activity and electromyograms
during voluntary contractions in humans. Neuroscience Letters, 241, 5–8. https://doi.org/
10.1016/S0304-3940(97)00964-6
Hari, R., Forss, N., Avikainen, S., Kirveskari, E., Salenius, S., & Rizzolatti, G. (1998). Activation
of human primary motor cortex during action observation: A neuromagnetic study.
Proceedings of the National Academy of Sciences of the United States of America, 95, 15061–
15065. https://doi.org/10.1073/pnas.95.25.15061
Heinrichs-Graham, E., McDermott, T. J., Mills, M. S., Wiesman, A. I., Wang, Y.-P., Stephen, J.
M., Calhoun, V. D., & Wilson, T. W. (2018). The lifespan trajectory of neural oscillatory ac-
tivity in the motor system. Developmental Cognitive Neuroscience, 30, 159–168. https://doi.
org/10.1016/j.dcn.2018.02.013
Heinrichs-Graham, E. & Wilson, T. W. (2015). Coding complexity in the human motor circuit.
Human Brain Mapping, 36, 5155–5167. https://doi.org/10.1002/hbm.23000
Hellwig, B., Häußler, S., Schelter, B., Lauk, M., Guschlbauer, B., Timmer, J., & Lücking, C. H.
(2001). Tremor-correlated cortical activity in essential tremor. The Lancet, 357, 519–523.
https://doi.org/10.1016/S0140-6736(0 0)04044-7
Hendrix, C. M. & Vitek, J. L. (2012). Toward a network model of dystonia. Annals
of the New York Academy of Sciences, 1265, 46–55. https://doi.org/10.1111/
j.1749-6632.2012.06692.x
Hillebrand, A., Barnes, G. R., Bosboom, J. L., Berendse, H. W., & Stam, C. J. (2012). Frequency-
dependent functional connectivity within resting-state networks: An atlas-based MEG
beamformer solution. NeuroImage, 59, 3909–3921. https://doi.org/10.1016/j.neuroim
age.2011.11.005
Hirschmann, J., Butz, M., Hartmann, C. J., Hoogenboom, N., Özkurt, T. E., Vesper, J., Wojtecki,
L., & Schnitzler, A. (2016). Parkinsonian rest tremor is associated with modulations of
subthalamic high-frequency oscillations. Movement Disorders, 31, 1551–1559. https://doi.org/
10.1002/mds.26663
Hirschmann, J., Özkurt, T. E., Butz, M., Homburger, M., Elben, S., Hartmann, C.
J., . . . Schnitzler, A. (2011). Distinct oscillatory STN-cortical loops revealed by simultaneous
MEG and local field potential recordings in patients with Parkinson’s disease. NeuroImage,
55, 1159–1168. https://doi.org/10.1016/j.neuroimage.2010.11.063
Hirschmann, J., Özkurt, T.E., Butz, M., Homburger, M., Elben, S., Hartmann, C.J., . . . Schnitzler,
A. (2013). Differential modulation of STN-cortical and cortico-muscular coherence by
movement and levodopa in Parkinson’s disease. NeuroImage, 68, 203–213. https://doi.org/
10.1016/j.neuroimage.2012.11.036
OSCILLATORY ACTIVITY IN SENSORIMOTOR FUNCTION 283
Houweling, S., Daffertshofer, A., van Dijk, B. W., & Beek, P. J. (2008). Neural changes induced
by learning a challenging perceptual-motor task. NeuroImage, 41, 1395–1407. https://doi.org/
10.1016/j.neuroimage.2008.03.023
Houweling, S., van Dijk, B. W., Beek, P. J., & Daffertshofer, A. (2010). Cortico-spinal syn-
chronization reflects changes in performance when learning a complex bimanual task.
NeuroImage, 49, 3269–3275. https://doi.org/10.1016/j.neuroimage.2009.11.017
Hummel, F., Kirsammer, R., & Gerloff, C. (2003). Ipsilateral cortical activation during finger
sequences of increasing complexity: representation of movement difficulty or memory
load? Clinical Neurophysiology, 114, 605–613. https://doi.org/10.1016/S1388-2457(02)00417-0
Huo, X., Wang, Y., Kotecha, R., Kirtman, E.G., Fujiwara, H., Hemasilpin, N., . . . Xiang, J.
(2011). High gamma oscillations of sensorimotor cortex during unilateral movement in the
developing brain: A MEG study. Brain Topography, 23, 375–384. https://doi.org/10.1007/s10
548-010-0151-0
Igarashi, J., Isomura, Y., Arai, K., Harukuni, R., & Fukai, T. (2013). A θ-γ oscillation code for
neuronal coordination during motor behavior. Journal of Neuroscience, 33, 18515–18530.
https://doi.org/10.1523/JNEUROSCI.2126-13.2013
Jasper, H. & Penfield, W. (1949). Electrocorticograms in man: Effect of voluntary
movement upon the electrical activity of the precentral gyrus. Archiv für Psychiatrie und
Nervenkrankheiten, 183, 163–174. https://doi.org/10.1007/BF01062488
Jasper, H. H. & Andrews, H. L. (1938). Electro-encephalography. III. Normal differentiation of
occipital and precentral regions in man. Archiv für Psychiatrie und Nervenkrankheiten, 39,
96–115.
Jeannerod, M. (2001). Neural simulation of action: A unifying mechanism for motor cogni-
tion. NeuroImage, S103–S109. https://doi.org/10.1006/nimg.2001.0832
Jensen, O., Goel, P., Kopell, N., Pohja, M., Hari, R., & Ermentrout, B. (2005). On the human
sensorimotor-cortex beta rhythm: Sources and modeling. NeuroImage, 26, 347–355. https://
doi.org/10.1016/j.neuroimage.2005.02.008
Jha, A., Litvak, V., Taulu, S., Thevathasan, W., Hyam, J.A., Foltynie, T., . . . Brown, P. (2017).
Functional connectivity of the pedunculopontine nucleus and surrounding region in
Parkinson’s disease. Cerebral Cortex, 27, 54–67. https://doi.org/10.1093/cercor/bhw340
Johnson, A. N., Wheaton, L. A., & Shinohara, M. (2011). Attenuation of corticomuscular co-
herence with additional motor or non-motor task. Clinical Neurophysiology, 122, 356–363.
https://doi.org/10.1016/j.clinph.2010.06.021
Kawai, R., Markman, T., Poddar, R., Ko, R., Fantana, A. L., Dhawale, A. K., Kampff, A.R., &
Ölveczky, B. P. (2015). Motor cortex is required for learning but not for executing a motor
skill. Neuron, 86, 800–812. https://doi.org/10.1016/j.neuron.2015.03.024
Keller, A. (1993). Intrinsic synaptic organization of the motor cortex. Cerebral Cortex, 3, 430–
441. https://doi.org/10.1093/cercor/3.5.430
Kilner, J. M., Baker, S. N., Salenius, S., Hari, R., & Lemon, R. N. (2000). Human cortical muscle
coherence is directly related to specific motor parameters. Journal of Neuroscience, 20, 8838–
8845. https://doi.org/10.1523/JNEUROSCI.20-23-08838.2000
Kilner, J. M., Marchant, J. L., & Frith, C. D. (2009). Relationship between activity in human pri-
mary motor cortex during action observation and the mirror neuron system. PLoS One, 4,
e4925. https://doi.org/10.1371/journal.pone.0004925
Klostermann, F., Nikulin, V. V., Kühn, A. A., Marzinzik, F., Wahl, M., Pogosyan, A., . . . Curio,
G. (2007). Task-related differential dynamics of EEG alpha-and beta-band synchronization
284 Bernadette C. M. van Wijk
Litvak, V., Jha, A., Eusebio, A., Oostenveld, R., Foltynie, T., Limousin, P., . . . Brown, P. (2011).
Resting oscillatory cortico-subthalamic connectivity in patients with Parkinson’s disease.
Brain, 134, 359–374. https://doi.org/10.1093/brain/awq332
Liu, X., Wang, S., Yianni, J., Nandi, D., Bain, P. G., Gregory, R., Stein, J. F., & Aziz, T. Z. (2008).
The sensory and motor representation of synchronized oscillations in the globus pallidus in
patients with primary dystonia. Brain, 131, 1562–1573. https://doi.org/10.1093/brain/awn083
López-Azcárate, J., Tainta, M., Rodríguez-Oroz, M. C., Valencia, M., González, R., Guridi,
J., . . . Alegre, M. (2010). Coupling between beta and high-frequency activity in the human
subthalamic nucleus may be a pathophysiological mechanism in Parkinson’s disease. Journal
of Neuroscience, 30, 6667–6677. https://doi.org/10.1523/JNEUROSCI.5459-09.2010
Manganotti, P., Gerloff, C., Toro, C., Katsuta, H., Sadato, N., Zhuang, P., Leocani, L., & Hallett,
M. (1998). Task-related coherence and task-related spectral power changes during sequen-
tial finger movements. Electroencephalography and Clinical Neurophysiology, 109, 50–62.
https://doi.org/10.1016/S0924-980X(97)00074-X
McAuley, J. H. & Marsden, C. D. (2000). Physiological and pathological tremors and rhythmic
central motor control. Brain, 123, 1545–1567. https://doi.org/10.1093/brain/123.8.1545
Miller, K. J., Leuthardt, E. C., Schalk, G., Rao, R. P. N., Anderson, N. R., Moran, D. W., Miller,
J. W., & Ojemann, J. G. (2007). Spectral changes in cortical surface potentials during motor
movement. Journal of Neuroscience, 27, 2424–2432. https://doi.org/10.1523/JNEURO
SCI.3886-06.2007
Mima, T., Matsuoka, T., & Hallett, M. (2000). Functional coupling of human right and left cor-
tical motor areas demonstrated with partial coherence analysis. Neuroscience Letters, 287,
93–96. https://doi.org/10.1016/S0304-3940(0 0)01165-4
Mima, T., Simpkins, N., Oluwatimilehin, T., & Hallett, M. (1999). Force level modulates human
cortical oscillatory activities. Neuroscience Letters, 275, 77–80. https://doi.org/10.1016/
S0304-3940(99)00734-X
Moisello, C., Blanco, D., Lin, J., Panday, P., Kelly, S.P., Quartarone, A., . . . Ghilardi, M. F. (2015).
Practice changes beta power at rest and its modulation during movement in healthy subjects
but not in patients with Parkinson’s disease. Brain and Behavior, 5, e00374. https://doi.org/
10.1002/brb3.374
Moran, D. W. & Schwartz, A. B. (2017). Motor cortical representation of speed and direc-
tion during reaching. Journal of Neurophysiology, 82, 2676–2692. https://doi.org/10.1152/
jn.1999.82.5.2676
Müller, K., Kass- Iliyya, F., & Reitz, M. (1997). Ontogeny of ipsilateral corticospinal
projections: A developmental study with transcranial magnetic stimulation. Annals of
Neurology, 42, 705–7 11. https://doi.org/10.1002/ana.410420506
Murthy, V. N. & Fetz, E. E. (1996a). Synchronization of neurons during local field potential
oscillations in sensorimotor cortex of awake monkeys. Journal of Neurophysiology, 76, 3968–
3982. https://doi.org/10.1152/jn.1996.76.6.3968
Murthy, V. N. & Fetz, E. E. (1996b). Oscillatory activity in sensorimotor cortex of awake
monkeys: synchronization of local field potentials and relation to behavior. Journal of
Neurophysiology, 76, 3949–3967. https://doi.org/10.1152/jn.1996.76.6.3949
Muthukumaraswamy, S. D. (2010). Functional properties of human primary motor cortex gamma
oscillations. Journal of Neurophysiology, 104, 2873–2885. https://doi.org/10.1152/jn.00607.2010
Muthukumaraswamy, S. D., Johnson, B. W., & McNair, N. A. (2004). Mu rhythm modula-
tion during observation of an object-directed grasp. Cognitive Brain Research, 19, 195–201.
https://doi.org/10.1016/j.cogbrainres.2003.12.001
286 Bernadette C. M. van Wijk
Neumann, W.-J., Degen, K., Schneider, G.-H., Brücke, C., Huebl, J., Brown, P., & Kühn, A. A. (2016).
Subthalamic synchronized oscillatory activity correlates with motor impairment in patients with
Parkinson’s disease. Movement Disorders, 31, 1748–1751. https://doi.org/10.1002/mds.26759
Neumann, W.-J., Jha, A., Bock, A., Huebl, J., Horn, A., Schneider, G., . . . Kühn, A. A. (2015).
Cortico-pallidal oscillatory connectivity in patients with dystonia. Brain, 138, 1894–1906.
https://doi.org/10.1093/brain/awv109
Neuper, C. & Pfurtscheller, G. (2001). Event-related dynamics of cortical rhythms: Frequency-
specific features and functional correlates. International Journal of Psychophysiology, 43, 41–
58. https://doi.org/10.1016/S0167-8760(01)00178-7
Nolte, G., Bai, O., Wheaton, L., Mari, Z., Vorbach, S., & Hallett, M. (2004). Identifying
true brain interaction from EEG data using the imaginary part of coherency. Clinical
Neurophysiology, 115, 2292–307. https://doi.org/10.1016/j.clinph.2004.04.029
O’Keefe, J. & Recce, M. L. (1993). Phase relationship between hippocampal place units and the
EEG theta rhythm. Hippocampus 3, 317–330. https://doi.org/10.1002/hipo.450030307
Ohara, S., Mima, T., Baba, K., Ikeda, A., Kunieda, T., Matsumoto, R., . . . Shibasaki, H. (2001).
Increased synchronization of cortical oscillatory activities between human supplemen-
tary motor and primary sensorimotor areas during voluntary movements. Journal of
Neuroscience, 21, 9377–9386. https://doi.org/10.1523/JNEUROSCI.21-23-09377.2001
Oswal, A., Brown, P., & Litvak, V. (2013). Movement related dynamics of subthalmo-cortical
alpha connectivity in Parkinson’s disease. NeuroImage, 70, 132–142. https://doi.org/10.1016/
j.neuroimage.2012.12.041
Özkurt, T. E., Butz, M., Homburger, M., Elben, S., Vesper, J., Wojtecki, L., & Schnitzler, A.
(2011). High-frequency oscillations in the subthalamic nucleus: A neurophysiological
marker of the motor state in Parkinson’s disease. Experimental Neurology, 229, 324–31.
https://doi.org/10.1016/j.expneurol.2011.02.015
Paninski, L., Fellows, M. R., Hatsopoulos, N. G., & Donoghue, J. P. (2004). Spatiotemporal
tuning of motor cortical neurons for hand position and velocity. Journal of Neurophysiology,
91, 515–532. https://doi.org/10.1152/jn.00587.2002
Paradiso, G., Cunic, D., Saint-Cyr, J. A., Hoque, T., Lozano, A. M., Lang, A. E., & Chen, R.
(2004). Involvement of human thalamus in the preparation of self-paced movement. Brain,
127, 2717–2731. https://doi.org/10.1093/brain/awh288
Pastötter, B., Berchtold, F., & Bäuml, K.-H. T. (2012). Oscillatory correlates of controlled speed-
accuracy tradeoff in a response-conflict task. Human Brain Mapping, 33, 1834–49. https://
doi.org/10.1002/hbm.21322
Pfurtscheller, G. & Lopes Da Silva, F. H. (1999). Event-related EEG/MEG synchronization and
desynchronization: Basic principles. Clinical Neurophysiology, 110, 1842–1857. https://doi.
org/10.1016/S1388-2457(99)00141-8
Pfurtscheller, G. & Neuper, C. (1992). Simultaneous EEG 10 Hz desynchronization and 40 Hz
synchronization during finger movements. Neuroreport, 3(12), 1057–1060. https://doi.org/
10.1097/00001756-199212000-00006
Pfurtscheller, G. & Neuper, C. (1997). Motor imagery activates primary sensorimotor area in
humans. Neuroscience Letters, 239, 65–8. https://doi.org/10.1016/s0304-3940(97)00889-6
Pfurtscheller, G., Neuper, C., & Kalcher, J. (1993). 40-Hz oscillations during motor behavior in
man. Neuroscience Letters, 164, 179–182. https://doi.org/10.1016/0304-3940(93)90886-P
Pfurtscheller, G., Stančák, A., & Neuper, C. (1996). Post-movement beta synchronization.
A correlate of an idling motor area? Electroencephalography and Clinical Neurophysiology,
98, 281–293. https://doi.org/10.1016/0013-4694(95)00258-8
OSCILLATORY ACTIVITY IN SENSORIMOTOR FUNCTION 287
Pfurtscheller, G., Zalaudek, K., & Neuper, C. (1998). Event-related beta synchronization after
wrist, finger and thumb movement. Electroencephalography and Clinical Neurophysiology,
109, 154–160. https://doi.org/10.1016/S0924-980X(97)00070-2
Piña-Fuentes, D., van Dijk, J. M. C., Drost, G., van Zijl, J. C., van Laar, T., Tijssen, M.
A. J., & Beudel, M. (2019). Direct comparison of oscillatory activity in the motor system
of Parkinson’s disease and dystonia: A review of the literature and meta-analysis. Clinical
Neurophysiology, 130, 917–924. https://doi.org/10.1016/j.clinph.2019.02.015
Pogosyan, A., Gaynor, L. D., Eusebio, A., & Brown, P. (2009). Boosting cortical activity at beta-
band frequencies slows movement in humans. Current Biology, 19, 1637–41. https://doi.org/
10.1016/j.cub.2009.07.074
Pollok, B., Butz, M., Gross, J., & Schnitzler, A. (2007). Intercerebellar coupling contributes
to bimanual coordination. Journal of Cognitive Neuroscience, 19, 704–7 19. https://doi.org/
10.1162/jocn.2007.19.4.704
Pollok, B., Gross, J., Müller, K., Aschersleben, G., & Schnitzler, A. (2005). The cerebral oscilla-
tory network associated with auditorily paced finger movements. NeuroImage, 24, 646–655.
https://doi.org/10.1016/j.neuroimage.2004.10.009
Priori, A., Foffani, G., Pesenti, A., Tamma, F., Bianchi, A., Pellegrini, M., . . . Villani, R. (2004).
Rhythm-specific pharmacological modulation of subthalamic activity in Parkinson’s dis-
ease. Experimental Neurology, 189, 369–379. https://doi.org/10.1016/j.expneurol.2004.06.001
Raethjen, J. & Deuschl, G. (2012). The oscillating central network of Essential tremor. Clinical
Neurophysiology, 123, 61–64. https://doi.org/10.1016/j.clinph.2011.09.024
Ray, N. J., Jenkinson, N., Wang, S., Holland, P., Brittain, J. S., Joint, C., Stein, J. F., & Aziz, T.
(2008). Local field potential beta activity in the subthalamic nucleus of patients with
Parkinson’s disease is associated with improvements in bradykinesia after dopamine and
deep brain stimulation. Experimental Neurology, 213, 108–113. https://doi.org/10.1016/
j.expneurol.2008.05.008
Rickert, J., de Oliveira, S. C., Vaadia, E., Aertsen, A., Rotter, S., & Mehring, C. (2005). Encoding
of movement direction in different frequency ranges of motor cortical local field potentials.
Journal of Neuroscience, 25, 8815–24. https://doi.org/10.1523/JNEUROSCI.0816-05.2005
Riddle, C. N. & Baker, S. N. (2005). Manipulation of peripheral neural feedback loops alters
human corticomuscular coherence. Journal of Physiology, 566, 625–639. https://doi.org/
10.1113/jphysiol.2005.089607
Ritter, P., Moosmann, M., & Villringer, A. (2009). Rolandic alpha and beta EEG rhythms’
strengths are inversely related to fMRI-BOLD signal in primary somatosensory and motor
cortex. Human Brain Mapping, 30, 1168–1187. https://doi.org/10.1002/hbm.20585
Rossiter, H. E., Davis, E. M., Clark, E. V, Boudrias, M.-H., & Ward, N. S. (2014). Beta oscillations
reflect changes in motor cortex inhibition in healthy ageing. NeuroImage, 91, 360–5. https://
doi.org/10.1016/j.neuroimage.2014.01.012
Rueda-Delgado, L. M., Heise, K. F., Daffertshofer, A., Mantini, D., & Swinnen, S. P. (2019). Age-
related differences in neural spectral power during motor learning. Neurobiology of Aging,
77, 44–57. https://doi.org/10.1016/j.neurobiolaging.2018.12.013
Rueda-Delgado, L. M., Solesio-Jofre, E., Serrien, D. J., Mantini, D., Daffertshofer, A., &
Swinnen, S. P. (2014). Understanding bimanual coordination across small time scales from
an electrophysiological perspective. Neuroscience & Biobehavioral Reviews, 47, 614–635.
https://doi.org/10.1016/j.neubiorev.2014.10.003
288 Bernadette C. M. van Wijk
Safri, N. M., Murayama, N., Hayashida, Y., & Igasaki, T. (2007). Effects of concurrent visual
tasks on cortico-muscular synchronization in humans. Brain Research, 1155, 81–92. https://
doi.org/10.1016/j.brainres.2007.04.052
Salenius, S., Portin, K., Kajola, M., Salmelin, R., & Hari, R. (1997). Cortical control of human
motoneuron firing during isometric contraction. Journal of Neurophysiology, 77, 3401–3405.
https://doi.org/10.1016/j.conb.2003.10.008
Salmelin, R., Hämäläinen, M., Kajola, M., & Hari, R. (1995). Functional segregation of
movement-related rhythmic activity in the human brain. NeuroImage, 2, 237–243. https://
doi.org/10.1006/nimg.1995.1031
Salmelin, R. & Hari, R. (1994). Spatiotemporal characteristics of sensorimotor neuromagnetic
rhythms related to thumb movement. Neuroscience, 60, 537–550. https://doi.org/10.1016/
0306-4522(94)90263-1
Sanes, J. N. & Donoghue, J. P. (1993). Oscillations in local field potentials of the primate motor
cortex during voluntary movement. Proceedings of the National Academy of Sciences of the
United States of America, 90, 4470–4474. https://doi.org/10.1073/pnas.90.10.4470
Schmiedt-Fehr, C., Mathes, B., Kedilaya, S., Krauss, J., & Başar-Eroğlu, C. (2016). Aging dif-
ferentially affects alpha and beta sensorimotor rhythms in a go/ go task. Clinical
no-
Neurophysiology, 127, 3234–42. https://doi.org/10.1016/j.clinph.2016.07.008
Schnitzler, A., Münks, C., Butz, M., Timmermann, L., & Gross, J. (2009). Synchronized
brain network associated with essential tremor as revealed by magnetoencephalography.
Movement Disorders, 24, 1629–1635. https://doi.org/10.1002/mds.22633
Schnitzler, A., Salenius, S., Salmelin, R., Jousmäki, V., & Hari, R. (1997). Involvement of pri-
mary motor cortex in motor imagery: A neuromagnetic study. NeuroImage, 6, 201–8. https://
doi.org/10.1006/nimg.1997.0286
Serrien, D. J. (2008). The neural dynamics of timed motor tasks: Evidence from a
synchronization–continuation paradigm. European Journal of Neuroscience, 27, 1553–1560.
https://doi.org/10.1111/j.1460-9568.2008.06110.x
Sharott, A., Grosse, P., Kühn, A.A., Salih, F., Engel, A.K., Kupsch, A., . . . Brown, P. (2008). Is the
synchronization between pallidal and muscle activity in primary dystonia due to peripheral
afferance or a motor drive? Brain, 131, 473–484. https://doi.org/10.1093/brain/awm324
Sherman, M. A., Lee, S., Law, R., Haegens, S., Thorn, C. A., Hämäläinen, M. S., Moore, C. I., &
Jones, S. R. (2016). Neural mechanisms of transient neocortical beta rhythms: Converging
evidence from humans, computational modeling, monkeys, and mice. Proceedings of the
National Academy of Sciences of the United States of America, 113, E4885–E4894. https://doi.
org/10.1073/pnas.1604135113
Silberstein, P., Kühn, A. A., Kupsch, A., Trottenberg, T., Krauss, J. K., Wöhrle, J.C., . . . Brown,
P. (2003). Patterning of globus pallidus local field potentials differs between Parkinson’s dis-
ease and dystonia. Brain, 126, 2597–608. https://doi.org/10.1093/brain/awg267
Spinks, R. L., Kraskov, A., Brochier, T., Umilta, M. A., & Lemon, R. N. (2008). Selectivity for
grasp in local field potential and single neuron activity recorded simultaneously from M1
and F5 in the awake macaque monkey. Journal of Neuroscience, 28, 10961–7 1. https://doi.org/
10.1523/JNEUROSCI.1956-08.2008
Stam, C. J., Nolte, G., & Daffertshofer, A. (2007). Phase lag index: Assessment of functional
connectivity from multi-channel EEG and MEG with diminished bias from common
sources. Human Brain Mapping, 28, 1178–1193. https://doi.org/10.1002/hbm.20346
OSCILLATORY ACTIVITY IN SENSORIMOTOR FUNCTION 289
Stančák, A., Feige, B., Lücking, C. H., & Kristeva-Feige, R. (2000). Oscillatory cortical ac-
tivity and movement- related potentials in proximal and distal movements. Clinical
Neurophysiology, 111, 636–650. https://doi.org/10.1016/S1388-2457(99)00310-7
Stančák, A. & Pfurtscheller, G. (1996). Mu-rhythm changes in brisk and slow self-paced finger
movements. Neuroreport, 7, 1161–1164. https://doi.org/10.1097/00001756-199604260-00013
Stančák, A., Riml, A., & Pfurtscheller, G. (1997). The effects of external load on movement-
related changes of the sensorimotor EEG rhythms. Electroencephalography and Clinical
Neurophysiology, 102, 495–504. https://doi.org/10.1016/S0013-4694(96)96623-0
Steinemann, N. A., O’Connell, R. G., & Kelly, S. P. (2018). Decisions are expedited through
multiple neural adjustments spanning the sensorimotor hierarchy. Nature Communications,
9, 3627. https://doi.org/10.1038/s41467-018-06117-0
Swann, N., Tandon, N., Canolty, R., Ellmore, T. M., McEvoy, L. K., Dreyer, S., DiSano, M., &
Aron, A. R. (2009). Intracranial EEG reveals a time-and frequency-specific role for the right
inferior frontal gyrus and primary motor cortex in stopping initiated responses. Journal of
Neuroscience, 29, 12675–12685. https://doi.org/10.1523/JNEUROSCI.3359-09.2009
Szurhaj, W., Bourriez, J. L., Kahane, P., Chauvel, P., Mauguière, F., & Derambure, P. (2005).
Intracerebral study of gamma rhythm reactivity in the sensorimotor cortex. European
Journal of Neuroscience, 21, 1223–1235. https://doi.org/10.1111/j.1460-9568.2005.03966.x
Talakoub, O., Neagu, B., Udupa, K., Tsang, E., Chen, R., Popovic, M. R., & Wong, W. (2016).
Time-course of coherence in the human basal ganglia during voluntary movements.
Scientific Reports, 6, 34930. https://doi.org/10.1038/srep34930
Tan, H., Jenkinson, N., & Brown, P. (2014). Dynamic neural correlates of motor error
monitoring and adaptation during trial-to-trial learning. Journal of Neuroscience, 34, 5678–
5688. https://doi.org/10.1523/JNEUROSCI.4739-13.2014
Tan, H., Pogosyan, A., Anzak, A., Ashkan, K., Bogdanovic, M., Green, A.L., . . . Brown, P.
(2013). Complementary roles of different oscillatory activities in the subthalamic nucleus in
coding motor effort in Parkinsonism. Experimental Neurology, 248, 187–195. https://doi.org/
10.1016/j.expneurol.2013.06.010
Tan, H., Wade, C., & Brown, P. (2016). Post-movement beta activity in sensorimotor cortex
indexes confidence in the estimations from internal models. Journal of Neuroscience, 36,
1516–1528. https://doi.org/10.1523/JNEUROSCI.3204-15.2016
Tinkhauser, G., Pogosyan, A., Little, S., Beudel, M., Herz, D. M., Tan, H., & Brown, P. (2017a).
The modulatory effect of adaptive deep brain stimulation on beta bursts in Parkinson’s dis-
ease. Brain, 140, 1053–1067. https://doi.org/10.1093/brain/awx010
Tinkhauser, G., Pogosyan, A., Tan, H., Herz, D. M., Kühn, A. A., & Brown, P. (2017b). Beta
burst dynamics in Parkinson’s disease OFF and ON dopaminergic medication. Brain, 140,
2968–2981. https://doi.org/10.1093/brain/awx252
Toma, K., Mima, T., Matsuoka, T., Gerloff, C., Ohnishi, T., Koshy, B., Andres, F., & Hallett, M.
(2002). Movement rate effect on activation and functional coupling of motor cortical areas.
Journal of Neurophysiology, 88, 3377–3385. https://doi.org/10.1152/jn.00281.2002
Trevarrow, M. P., Kurz, M. J., McDermott, T. J., Wiesman, A. I., Mills, M. S., Wang, Y.-
P., . . . Wilson, T. W. (2019). The developmental trajectory of sensorimotor cortical
oscillations. NeuroImage, 184, 455–461. https://doi.org/10.1016/j.neuroimage.2018.09.018
Tsang, E.W., Hamani, C., Moro, E., Mazzella, F., Lozano, A.M., Hodaie, M., Yeh, I.-J., & Chen, R.
(2012). Movement related potentials and oscillatory activities in the human internal globus
pallidus during voluntary movements. Journal of Neurology, Neurosurgery, and Psychiatry,
83, 91–7. https://doi.org/10.1136/jnnp.2011.243857
290 Bernadette C. M. van Wijk
Tsiokos, C., Hu, X., & Pouratian, N. (2013). 200–300 Hz movement modulated oscillations in
the internal globus pallidus of patients with Parkinson’s Disease. Neurobiological Disorders,
54, 464–474. https://doi.org/10.1016/j.nbd.2013.01.020
van Wijk, B. C. M. (2017). Is broadband gamma activity pathologically synchronized to the
beta rhythm in Parkinson’s disease? Journal of Neuroscience, 37, 9347–9349. https://doi.org/
10.1523/JNEUROSCI.2023-17.2017
van Wijk, B. C. M., Beek, P. J., & Daffertshofer, A. (2012). Differential modulations of ipsilateral
and contralateral beta (de)synchronization during unimanual force production. European
Journal of Neuroscience, 36, 2088–2097. https://doi.org/10.1111/j.1460-9568.2012.08122.x
van Wijk, B .C. M., Beudel, M., Jha, A., Oswal, A., Foltynie, T., Hariz, M. I., . . . Litvak, V. (2016).
Subthalamic nucleus phase-amplitude coupling correlates with motor impairment in Parkinson’s
disease. Clinical Neurophysiology, 127, 2010–2019. https://doi.org/10.1016/j.clinph.2016.01.015
van Wijk, B. C. M., Daffertshofer, A., Roach, N., & Praamstra, P. (2009) A role of beta oscilla-
tory synchrony in biasing response competition? Cerebral Cortex, 19, 1294–1302. https://doi.
org/10.1093/cercor/bhn174
van Wijk, B. C. M., Neumann, W.-J., Schneider, G.-H., Sander, T. H., Litvak, V., & Kühn, A. A.
(2017). Low-beta cortico-pallidal coherence decreases during movement and correlates with
overall reaction time. NeuroImage, 159, 1–8. https://doi.org/10.1016/j.neuroimage.2017.07.024
Vidailhet, M., Vercueil, L., Houeto, J.-L., Krystkowiak, P., Benabid, A.-L., Cornu, P., . . . Pollak,
P. (2005). Bilateral deep-brain stimulation of the globus pallidus in primary generalized dys-
tonia. New England Journal of Medicine, 352, 459–467. https://doi.org/10.1056/nejmoa042187
Weaver, F. M., Follett, K., Stern, M., Hur, K., Harris, C., Marks, . . . Huang, G.D.; CSP 468 Study
Group. (2009). Bilateral deep brain stimulation vs best medical therapy for patients with
advanced Parkinson disease: A randomized controlled trial. Journal of the American Medical
Association, 301, 63–73. https://doi.org/10.1001/jama.2008.929
Wessel, J. R., Ghahremani, A., Udupa, K., Saha, U., Kalia, S. K., Hodaie, M., . . . Chen, R. (2016).
Stop-related subthalamic beta activity indexes global motor suppression in Parkinson’s dis-
ease. Movement Disorders, 12, 1846–1853. https://doi.org/10.1002/mds.26732
Witham, C. L., Riddle, C. N., Baker, M. R., & Baker, S. N. (2011). Contributions of descending
and ascending pathways to corticomuscular coherence in humans. Journal of Physiology,
589, 3789–3800. https://doi.org/10.1113/jphysiol.2011.211045
Witte, M., Patino, L., Andrykiewicz, A., Hepp-Reymond, M.-C., & Kristeva, R. (2007).
Modulation of human corticomuscular beta- range coherence with low- level static
forces. European Journal of Neuroscience, 26, 3564–3570. https://doi.org/10.1111/
j.1460-9568.2007.05942.x
Wolpaw, J. R., Birbaumer, N., McFarland, D. J., Pfurtscheller, G., & Vaughan, T. M. (2002).
Brain–computer interfaces for communication and control. Clinical Neurophysiology, 113,
767–791. https://doi.org/10.1016/S1388-2457(02)00057-3
Yamawaki, N., Stanford, I. M., Hall, S. D., & Woodhall, G. L. (2008). Pharmacologically induced
and stimulus evoked rhythmic neuronal oscillatory activity in the primary motor cortex in
vitro. Neuroscience, 151, 386–395. https://doi.org/10.1016/j.neuroscience.2007.10.021
Zandvoort, C. S., van Dieën, J. H., Dominici, N., & Daffertshofer, A. (2019). The human sen-
sorimotor cortex fosters muscle synergies through cortico-synergy coherence. NeuroImage,
199, 30–37. https://doi.org/10.1016/j.neuroimage.2019.05.041
Pa rt I I I
CHAPTER 13
EEG FREQU E NC Y
DEVEL OPM EN T AC RO S S
INFANCY AND C H I L DH O OD
13.1 Introduction
Hans Berger (1929) published the first report of the electroencephalogram (EEG)
recorded on an adult scalp. After success with adult recordings, Berger began to record
electrical activity from the scalps of a wide age range of individuals and it was Berger’s
fifth report three years later (1932) that created the field of developmental neurophysi-
ology. Berger said in the earlier part of his fifth report that he had recorded EEG from
the scalps of a few children but had not yet recorded EEG from anyone under 5 years of
age. His rationale for avoiding younger children was that he was using needle electrodes
and did not want to use local anesthesia with such young individuals. Thus, for his initial
recordings with infants, Berger used silver foil electrodes on the forehead and occiput,
with each the size of the palm of his hand. On each section of silver foil was a piece of
flannel cloth soaked in saline. Another piece of silver foil was placed on top so that the
flannel would not dry out. He held this on the scalp with a rubber bandage (Berger, 1932).
Berger reported that he attempted to record EEG from six infants between 8 and
13 days old, but there were no characteristic alpha waves that he had come to expect
after doing so many adult and older child EEG recordings. He concluded that there was
no EEG in the first weeks after birth because the cerebral cortex had not yet assumed
its function. A 35-day-old infant (a “healthy and very strong boy”) was the youngest in
whom Berger could record EEG. The infant lay completely still and began to fall asleep
and there was some evidence of the posterior alpha activity that Berger had seen for the
past three years in older children and adults. Berger (1932) concluded in his fifth report
that one could see EEG in all healthy and strong children who are older than 2 months
294 KIMBERLY CUEVAS and MARTHA ANN BELL
of age and that by the age of 5 years, children show an EEG tracing comparable to that of
an adult.
The EEG recordings Berger accomplished from infants were highly intriguing. Infant
EEG had greater amplitude and cycled at a lower frequency than adult EEG. From
Berger’s time, researchers have assumed that EEG differences among infants, chil-
dren, and adults reflect differences in brain maturation. Longitudinal research studies,
including our own contributions (Bell & Fox, 1992; Cuevas & Bell, 2011), have verified
that position. Researchers continue to examine infant and young child EEG as a rela-
tively inexpensive, non-invasive way to study brain development.
This chapter provides an overview of the ontogeny of EEG frequency bands during
infancy and early childhood. We include infant correlates of adult frequency bands
labeled as delta, theta, alpha family, beta, and gamma. We focus most intensely on
the 6–9 Hz “infant alpha” frequency band, as it is the band that has received the most
attention in the developmental neurophysiology research literature and the frequency
band that we have examined so intensely in our own programs of research and our col-
laborative efforts. We organize the discussion of infant and child alpha around studies
focused on action-perception processes, executive processes, and affective processes.
The chapter ends with a brief section on broader impacts of infant and child EEG and
future directions.
This section focuses on the oscillatory rhythms that compose the waking EEG during
infancy and childhood. The primary frequency bands of normative EEG oscillatory ac-
tivity based on the adult literature are delta (1–4 Hz), theta (4–8 Hz), alpha (8–13 Hz),
beta (13–30 Hz), and gamma (36–44 Hz). These oscillations often occur at slower rates
during early development. However, throughout ontogeny, EEG amplitude is typically
inversely related to frequency, with greater amplitude exhibited for slower oscillations.
The functional properties of oscillatory rhythms are defined here in terms of event-
related changes in EEG power values (EEG reactivity) in response to external stimuli
and/or psychological (e.g., cognitive, affective) processing. Specifically, the term syn-
chronization refers to an increase in EEG power values relative to resting baseline,
whereas desynchronization refers to a decrease in EEG power values.
We provide an overview of what is known about each of these EEG frequency bands
during early development, including functional properties as well as age- related
changes in frequency and amplitude. In general, there is a decrease in the amount of low
frequency activity (1–5 Hz; i.e., delta, lower-theta) and an increase in intermediate fre-
quency oscillations (6–12 Hz; i.e., upper-theta; alpha) from infancy through childhood
(Gibbs & Knott, 1949; Hagne, 1968; Marshall et al., 2002). As there is no standardization
of EEG rhythms in the developmental neurophysiology literature as found in adolescent
EEG FREQUENCY DEVELOPMENT ACROSS INFANCY AND CHILDHOOD 295
and adult EEG work, EEG researchers have used multiple techniques to determine ap-
propriate frequency band definitions for infants and young children.
13.2.1 Delta Band
Large, slow wave oscillations in the delta band are typically investigated in the con-
text of sleep. Although oscillatory activity within the delta frequency band is present
in the waking EEG during early development, descriptions of this band have primarily
focused on decreases in delta activity and corresponding increases in higher frequency
oscillations as a function of age (Hagne, 1968; Ogawa et al., 1984). Although Lindsley
(1938) recorded from only occipital sites, he noted the presence of “type V waves” in
subjects from 3 months of age through adulthood, with the average frequency increasing
gradually from 0.68 to 1.08 Hz over this developmental period. The delta rhythm was
unaffected by stimulation and reached the 1 Hz oscillation rate by 3 years of age.
13.2.2 Theta Band
Oscillatory activity within the theta range is faster than the adjacent delta band. Theta
rhythms are prominent in the waking EEG during early development, but the theta fre-
quency range is slightly lower than that of the adult 4–8 Hz theta band. The adult litera-
ture has noted that emotional and cognitive processing is associated with increases in
theta power (e.g., Walter & Walter, 1949; see also Chapter 15). Theta oscillations have
been hypothesized to be linked to reward processing and active learning during early
development (Begus & Bonawitz, 2020). Visual perceptual-stimulation techniques
have also been used to “entrain” (elicit) neural oscillatory activity at a particular fre-
quency, with specific focus on the functional properties of the theta rhythm (Köster
et al., 2019). To our knowledge, however, there have been no systematic examinations of
theta’s functional properties throughout development. Here, we highlight some of the
characteristics of theta reactivity during infancy and childhood.
In the context of examining 7-to 12-month-olds’ alpha rhythms, Stroganova and
colleagues (Stroganova et al., 1999) identified 3.6–6 Hz activity with unique functional
properties. This band exhibited increases in theta power during “internally controlled
(anticipatory) attention” as infants were engaged in a peek-a-boo game (Orekhova
et al., 1999; Stroganova et al., 1998). Recent analysis of 6-to 12-month-olds’ sustained
attention, using heart-rate-defined phases of attention, has confirmed and extended
these findings. By 10–12 months of age, sustained attention during a video clip was
associated with theta (2–6 Hz) synchronization at multiple scalp sites, including frontal
pole, central, and parietal (Xie et al., 2018). Event-related theta synchronization during
attentional processes may also be related to infants’ subsequent memory and cognitive
abilities. For instance, increases in 11-month-olds’ frontal theta (3–5 Hz) power during
object exploration was associated with variability in infants’ future object recognition
296 KIMBERLY CUEVAS and MARTHA ANN BELL
(Begus et al., 2015). Likewise, measures of frontal theta (3–6 Hz) enhancement during
a dynamic non-social video at 6 months of age is associated with non-verbal cognitive
abilities at 9 months (Braithwaite et al., 2020). Together, these findings indicate that
theta synchronization is intricately linked to aspects of infant attention and memory;
however, the precise developmental onset and trajectory of event-related theta reactivity
are important open areas of inquiry.
Increases in theta power are also exhibited in response to positive and negative af-
fective processing during infancy (Futagi et al., 1998; Nikitina et al., 1985) as well as nu-
tritive sucking (Lehtonen et al., 2016; Paul et al., 1996). However, little is known about
age-related changes in theta reactivity beyond infancy. To this end, Orekhova and
colleagues (Orekhova et al., 2006) directly compared infants’ and young children’s theta
reactivity to “theta-provoking” stimuli—toy exploration and attention to social stimula-
tion (i.e., adult speech). Theta scalp topography differed as a function of age and stimulus
type; though, both 8-to 12-month-olds and 3-to 6-year-olds exhibited increases in
theta (3.6–5.6 Hz and 4–8 Hz, respectively) power in response to these stimuli. Thus, in
some affective contexts, the adult theta range is developmentally appropriate by early
childhood. Recent evidence also indicates changes in the topography of infants’ theta
(3–6 Hz) synchronization in response to social stimuli between 6 and 12 months of age
(Jones et al., 2015). In sum, theta synchronization is displayed in a variety of cognitive
and emotional settings early in development, but the precise nature of the scalp topog-
raphy of theta responsivity varies as a function of multiple factors, including age and
reactive context.
band (6–9 Hz) and its associations with affective and cognitive processes. Frontal alpha
activity is proposed to have unique neural generators by some, and to be associated with
central and/or posterior alpha rhythms by others (see Stroganova & Orekhova, 2007 for
discussion).
13.2.3.1 Posterior Alpha
Early developmental EEG investigations using visual quantification noted posterior os-
cillatory activity that had similar rhythmic qualities to the adult posterior alpha rhythm,
though the rate of oscillation was slower. Around 3–4 months of age, an occipital rhythm
with a frequency of 3–5 Hz has been identified in the waking EEG (Lindsley, 1938,
1939; Smith, 1938, 1941). It increases in both amplitude and frequency during the first
postnatal year (6–7 Hz), with the posterior alpha oscillation rate continuing to gradually
increase throughout childhood. The occipital alpha amplitude remains fairly stable until
3–4 years when there is a sharp drop followed by a gradual decline to adult levels by
15–16 years (Lindsley, 1939). The adult alpha frequency range is reached during middle
childhood, with mean occipital frequencies of 9 and 10 Hz at 8 and 11–16 years, respect-
ively (Lindsley, 1938, 1939; Smith, 1938, 1941). Subsequent investigations using frequency
analyzers and spectral analysis have confirmed these early findings (Hagne, 1968; Hagne
et al., 1973; Miskovic et al., 2015).
The functional properties of the posterior alpha rhythm were first noted during
early development by Lindsley (1938). Visual stimuli (i.e., bright lights in a dark room)
blocked occipital oscillatory activity in young participants, including infants who were
a “few months” of age. However, scalp recordings were limited to occipital sites and
detailed developmental analyses were not reported. Using spectral analysis, Stroganova
and colleagues (1999) confirmed and extended these early findings by comparing 7-to
12-month-olds’ posterior oscillatory activity during conditions of darkness (i.e., lights
off in the testing room) and quite visual attention (i.e., watching soap bubbles). These
conditions were designed to be developmentally appropriate analogues of the “eyes
closed versus eyes open” procedures used to identify posterior alpha reactivity with
older children and adults. Their work indicated that the 5.2–9.6 Hz band exhibited
prominent posterior oscillatory activity in the response to the “lights off ” condition,
and this activity was attenuated during visual attention. Importantly, these properties
were not displayed at anterior/central scalp locations, indicating similar functional
properties and scalp topography to the adult posterior alpha rhythm.
indicated that the central peak frequency rises from 6–7 Hz at 5 months to 9 Hz by
4 years of age (Marshall et al., 2002). These findings are consistent with Smith’s (1941)
original report of central alpha activity increasing to 8 Hz at 1.5 years, to 9 Hz at 3.5 years,
to 10 Hz at 10–14 years. Prior to reaching adult frequency ranges, the oscillatory rate of
the mu rhythm is faster than the posterior alpha rhythm.
Although the mu rhythm is attenuated during various conditions related to action
perception and production, the functional properties of this rhythm are most robust
and reliably identified via conditions of discrete motor movement. Currently, the most
systematic developmental examination of the functional properties of the central alpha
rhythm comes from a cross-sectional magnetoencephalography (MEG) study. Age-
related changes in the peak frequency of mu rhythm attenuation during a discrete hand
movement (i.e., squeezing a pipette) were found from infancy through childhood.
There was a linear increase in peak mu frequency from 2.75 to 8.25 Hz between 11 and
47 weeks of age (Berchicci et al., 2011). This was followed by a more gradual increase
in peak mu frequency during early childhood (2–5 years; average 8.5 Hz), which was
below the adult 10.2-Hz peak. These findings have been confirmed and extended by a
recent cross-sectional EEG study of mu rhythm attenuation during reaching-grasping
actions. Mu peaks were found at 7.39, 8.81, and 10.51 Hz at 12 months, 4 years, and 18–
21 years, respectively (Thorpe et al., 2016). A bourgeoning recent literature examining
the infant mu rhythm has found 6–9 Hz attenuation during action perception and/or
execution (Cuevas et al., 2014; Marshall & Meltzoff, 2011); a detailed description of the
EEG correlates of action-perception processes is provided later in the chapter.
13.2.4 Beta Band
Berger (1929) identified beta waves in conjunction with alpha waves in the adult
waking EEG, but as higher frequency oscillations. Similar to the central alpha rhythm,
the “precentral” beta rhythm is associated with motor processes (Jasper & Andrews,
1938; Jasper & Penfield, 1949). The adult literature has revealed that the beta rhythm is
attenuated during action execution, motor planning, and action perception (Jarvelainen
et al., 2004; Pfurtscheller et al., 1997). The beta frequency is often a harmonic of the mu
rhythm frequency. Although these rhythms have overlapping functional properties,
there are a variety of unique characteristics of each band (see Crone et al., 1998; Hari &
Salmelin, 1997).
Initial description of developmental changes in beta rhythm comes from Lindsley
(1938), who identified “type IV waves” from the waking EEG activity at occipital sites
(the only recording sites); these waves had mean frequencies approximately double the
alpha wave frequency throughout development. This oscillatory activity exhibited sub-
stantial increases in frequency throughout infancy and childhood: 7 Hz at 3 months, 11.6
Hz at 12 months, and 16 Hz at 4 years. The adult 20-Hz frequency was attained around 11–
12 years. The literature is mixed, however, on whether there are age-related increases or
decreases in beta band power (Gasser et al., 1988; Matsuura et al., 1985; Ogawa et al., 1984).
EEG FREQUENCY DEVELOPMENT ACROSS INFANCY AND CHILDHOOD 299
13.2.5 Gamma Band
Because of its low amplitude, high-frequency oscillations, systematic analysis of gamma
band activity via visual examination was not feasible (Lindsley, 1938). A variety of
broadband and sub-band definitions have been used to define the gamma oscillatory
activity, but the vast majority of the developmental and adult literature has focused on
oscillations near 40 Hz. Work with adults has revealed that increases in gamma band
power are associated with both motor and cognitive processes, including perceptual
binding and memory (see Başar, 2013, for review).
Early examination of the gamma (35–45 Hz) band with a large sample of 3-to 12-
year-olds (N =707), noted increased power between 3 and 4 years at all recorded sites
(frontal, central, occipital) with a distinct peak at frontal sites between 4 and 5 years of
age (Takano & Ogawa, 1998). Although occipital gamma power was stable after 4 years
of age, there were some age-related decreases in frontal and central gamma power be-
tween 4 and 11 years. A recent analysis of gamma (31–50 Hz) oscillations with a compara-
tively smaller sample of 3-to 38-year-olds (N =156) also identified age-related decreases
in resting gamma power; however, these changes were seen throughout the scalp
300 KIMBERLY CUEVAS and MARTHA ANN BELL
(Tierney et al., 2013). Tierney and colleagues proposed that decreases in gray matter—
specifically, synaptic density via synaptic pruning (Whitford et al., 2007)—may underlie
the widespread reductions in resting gamma power as a function of age. Additional
work is needed to characterize the development of gamma power during the first three
postnatal years. A recent large-scale longitudinal investigation between 2 to 6 postnatal
months (N =518) found gradual increases in gamma (30–50 Hz) power across all scalp
sites (Pivik et al., 2019). Pivik and colleagues hypothesize that ontogenetic changes in
GABAergic (γ‑aminobutyric acid) interneurons contribute to emerging gamma activity
(Paredes et al., 2016; Xu et al., 2011).
During infancy, resting gamma power varies as a function of internal and ex-
ternal factors, such as socioeconomic status, sex, and diet (Pivik et al., 2019; Tomalski
et al., 2013, but see Brito et al., 2016). Further, higher levels of resting gamma oscilla-
tory activity are positively associated with both concurrent and future measures of
early cognitive development, including language, memory, and executive function
(Benasich et al., 2008; Brito et al., 2013; Tarullo et al., 2017). Developmental studies
have also revealed changes in gamma band activity associated with ongoing cognitive
processing. An initial investigation of gamma’s involvement in perceptual binding
revealed bursts of gamma oscillatory activity at frontal areas during the perception
of illusory objects in 8-month-olds, but not 6-month-olds (Csibra et al., 2000). These
findings confirmed behavioral evidence that younger infants were unsuccessful at
perceiving illusory objects (Ghim, 1990). Subsequent work also shows increases in
6-month-olds’ gamma oscillatory activity over temporal areas in the context of ob-
ject occlusion events; presumably contributing infants’ ability to maintain object
representations (Kaufman et al., 2003, 2006). Gamma oscillations are also associated
with aspects of language learning during infancy, including recognition of familiar
objects with verbal labels (Gliga et al., 2010) and phonemic perceptual narrowing
(Ortiz-Mantilla et al., 2016).
Our primary focus in this section is on the infant and child frontal alpha and central mu
rhythms, as the 6–9 Hz alpha band has received the most attention in the developmental
literature. We noted at the beginning of this chapter that the developmental neurophysi-
ology literature has no standardization of EEG rhythms as found in adult EEG work. The
research we have reviewed thus far shows how individual research teams since Berger’s
time have worked toward defining specific frequency bands of interest, with clear indi-
cation that the traditional frequency band definitions used with adults (e.g., 8–13 Hz for
alpha) do not apply to studies of infant and young child EEG.
EEG FREQUENCY DEVELOPMENT ACROSS INFANCY AND CHILDHOOD 301
After six decades of developmental EEG research, the Society for Psychophysiological
Research published a set of recommendations for recording and analyzing EEG in re-
search contexts (Pivik et al., 1993). Those guidelines included two suggestions for EEG
researchers who focus on infants and young children. The first was wide band analysis,
potentially including all frequencies with evidence of power after doing spectral plots.
Although some researchers working with infant samples have used this approach in
studies of infant emotion (e.g., 3–13 Hz: Diego et al., 2006), the wide band approach was
never commonly adopted for use in infant and early childhood research. This may be
because of the assumption that there was the potential for the developmental EEG field
to eventually standardize frequency bands based on temporal, spatial, amplitude, and
functional characteristics much as the adult EEG field has done.
The second suggestion was that spectral plots be examined and frequency bands
determined that center around the peaks in the spectrum. This is the approach we used
with our first infant EEG data set (Bell & Fox, 1992). We recorded resting baseline EEG
from 13 infants from 7 to 12 months of age because we had hypotheses about relations
between the development of frontal EEG and performance on a classic infant task (i.e.,
Piaget’s A-not-B task). Spectral plots of the frontal EEG leads (F3, F4) were created for
each infant’s monthly EEG recording for a total of 91 frontal spectral plots. At each age
the plots revealed a dominant frequency around 6–9 Hz. More importantly, the plots
revealed month-to-month changes in the peak frequency for some infants. The spectral
plots from one infant in our Bell and Fox (1992) data set are shown in Figure 13.1. The
four spectral plots represent this infant’s EEG data at F3 and F4 from 8 until 11 months of
age. One can observe a peak frequency at 7 Hz in the EEG recording made at 8 months
of age and the sharing of peak frequency between 7 and 8 Hz at 9 months of age. In
the 10-month recording, peak frequency appears to be shifting toward 8 Hz, whereas at
11 months of age there is only one peak at 8 Hz. These spectral plots demonstrate a def-
inite shift in peak frequency on a month-to-month basis during infancy. They also show
a shift in this overall dominant frequency band from 5–9 Hz at 8 months to 6–10 Hz at
11 months for this one infant. After examining all 91 spectral plots, we determined that
6–9 Hz frequency band captured the dominant frequency in this sample of infants.
Our selection of the 6–9 Hz frequency band as dominant in our infant longitudinal
study was replicated and extended to early childhood by Marshall and colleagues (2002)
with a longitudinal data set that included recordings of baseline EEG from a group of
29 children at 5, 10, 14, 24, and 51 months of age. They reported a peak frequency in
the 6–9 Hz band that emerged across multiple scalp locations. Although not evident
during the earliest recordings at 5 months of age, the 6–9 Hz band was reliable across all
scalp locations by 10 months of age and continued to be the dominant frequency band
through the 51-month EEG recordings. Thus, with each of these two small longitudinal
datasets, the 6–9 Hz band contained infant and child peak frequency, much as adult
alpha band of 8–13 Hz contains the typical mature peak frequency at 10 Hz, although
there are individual differences in specific peak during adulthood (Klimesch, 1999).
This section highlights our 6–9 Hz alpha band research and selectively reviews other
work examining the alpha band’s relation to critical action-perception, cognitive,
302 KIMBERLY CUEVAS and MARTHA ANN BELL
8 months 9 months
F3 F3
F4 F4
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18
HERTZ HERTZ
10 months 11 months
F3 F3
F4 F4
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18
HERTZ HERTZ
Figure 13.1 Month-to-month spectral plots from one infant. Note the change in peak fre-
quency from 8 to 11 months. The x-axis is frequency in single hertz (Hz) bins and the y-axis is
EEG power (mean square microvolts).
From the Bell & Fox, 1992 dataset.
and affective processes during early development. Although 6–9 Hz oscillatory ac-
tivity has been linked to multiple aspects of early cognition, including future thinking
(Blankenship et al., 2018) and memory (Cuevas et al., 2012c), the most extensive area
of research has investigated emerging executive functions. Here, we provide a develop-
mental analysis of the characteristics and individual variations in alpha band activity in
association with action-perception, executive, and affective processes.
13.3.1 Action-Perception Processes
The functional properties of the mu and beta rhythms during motor movement were
established by early investigations with adults. A relatively recent area of inquiry in the
developmental literature has focused on these sensorimotor rhythms and their involve-
ment in action-perception processes. Specifically, there is evidence that the mu and beta
rhythms exhibit “neural mirroring” properties: oscillatory activity is attenuated during
both action execution and observation (Lepage & Théoret, 2006; Liao et al., 2015).
EEG FREQUENCY DEVELOPMENT ACROSS INFANCY AND CHILDHOOD 303
Characterizing the conditions under which the sensorimotor mu and beta rhythms are
reactive has been an area of substantial interest for developmental research as neural
mirroring systems have been proposed to be involved in action understanding as well as
broader socio-cognitive development (e.g., imitation; Marshall & Meltzoff, 2011).
Research with non- human primates has found evidence of neural mirroring
during the observation and execution of facial gestures in 1-to 7-day-old rhesus ma-
caque monkeys (Ferrari et al., 2012). Subsequent work has revealed that early experi-
ence with the mother, as compared to nursery-rearing, is associated with (1) greater mu
rhythm attenuation when observing facial gestures and (2) more frequent imitation of
these gestures (Vanderwert et al., 2015). These findings have been recently extended to
human infants, with 9-month-olds exhibiting attenuation of the mu rhythm during fa-
cial expression perception and production (Rayson et al., 2017). Further, variability in
9-month mu rhythm reactivity was related to their mothers’ tendency to mirror their
facial gestures at 2 months of age, with more frequent maternal mirroring at 2 months
of age being associated with greater action perception mu attenuation at age 9 months.
Together, these findings highlight the biopsychosocial underpinnings of emerging
action-perception processes.
Another line of research has examined the impact of children’s action experience on
the neural correlates of action perception. Recent evidence indicates that the effects of
early motor experience are evident shortly after birth. Newborn monkeys with more
frequent reaching attempts display greater beta rhythm reactivity when observing
goal-directed reaching movements (Festante et al., 2018). Likewise, 9-and 12-month-
old human infants with more advanced manual motor skills (e.g., reaching, grasping)
show enhanced mu rhythm attenuation during manual action perception (Cannon
et al., 2016; Upshaw et al., 2016; Yoo et al., 2016). Training paradigms have extended
these findings by controlling children’s short-term motor or perceptual experience with
novel tools/actions. In an action-sound perception paradigm, 10-month-olds exhibited
greater mu rhythm attenuation when hearing sounds associated with actions that they
had received motor (active training), as compared to observational (passive training),
experience (Gerson et al., 2015). A recent extension of this work to 3-to 6-year-olds has
revealed greater EEG indices of visual processing (occipital alpha attenuation) as chil-
dren observed tool-use actions that they had received active training (Bryant & Cuevas,
2019). However, there were no experience-related effects on sensorimotor rhythms,
indicating that their role in action understanding might vary depending on age and
context.
Other work has taken a broader perspective to consider whether individual
differences in motor system activation are related to aspects of socio-cognitive de-
velopment. For instance, in the context of a goal-directed reaching task, 7-month-
olds with greater mu rhythm attenuation were more likely to express the same goal
as an adult; thus, selecting the goal toy more often than the non-goal toy (Filippi
et al., 2016). These early socially-contingent behavioral and neural responses may
provide the foundation for more advanced socio-cognitive skills that emerge during
early childhood. Recent evidence indicates that 3-to 5-year-olds’ motor skills,
304 KIMBERLY CUEVAS and MARTHA ANN BELL
13.3.2 Executive Processes
Executive function (EF) is an umbrella term that captures a wide array of higher-order
cognitive processes—working memory, inhibitory control, cognitive flexibility—that
organize and coordinate behavior to support complex goal-directed actions. This con-
struct has been a major focus of developmental research as proficiency in executive pro-
cessing is associated with optimal behavioral and academic outcomes (see Diamond,
2013 for review). EF emerges during the first postnatal year, with substantial advances
throughout childhood and beyond (see Bell & Cuevas, 2016; Cuevas et al., 2018, for
reviews). Individual differences in the developmental progression of EF are linked to
biopsychosocial factors, including frontal lobe development, temperament, and co-
occurring socialization experiences. Here, we discuss our cross-sectional and longitu-
dinal work examining associations between EF and both resting-state and task-related
EEG measures throughout early development.
Resting-state (baseline) EEG Bell & Fox, 1992, 1997; Perone et al. 2018;*
is associated with EF (task MacNeill et al. 2018 Wolfe & Bell, 2004
performance)
Changes in EEG during executive Bell, 2001, 2002; Bell & Wolfe, 2007;
processing (baseline-to-task Bell & Wolfe, 2007; Swingler et al., 2011
related changes) Cuevas & Bell, 2011;
Cuevas et al., 2012a, 2012d
Changes in EEG during increased Cuevas et al., 2012e Broomell & Bell, 2017;
inhibitory demands Cuevas et al., 2016
Executive processing-related EEG Bell, 2001, 2002; 2012; Cuevas et al., 2016;
is associated with EF Cuevas et al., 2012a, 2012d Morasch & Bell, 2011;
Swingler et al., 2011;
Watson & Bell, 2013;
Wolfe & Bell, 2004, 2007a,
2007b
Resting-state EEG is associated Broomell et al. 2019; Cuevas et al., 2012b
with future EF Kraybill & Bell, 2013
13.3.3 Affective Processes
Developmental evidence suggests that resting state 6-9 Hz frontal EEG activation
patterns are associated with individual differences in general affective style from at least
4 or 5 months of age (e.g., Fox et al., 1992). Resting state frontal EEG patterns also dif-
ferentiate between infants born to women who experienced prenatal depression versus
nondepressed women as early one week after birth at the broader 3–9 Hz band (e.g.,
Diego et al., 2010). These types of EEG findings are based on the measurement of frontal
alpha-band EEG asymmetry, which highlights relative activation patterns between
homologous left and right frontal scalp electrodes. Specifically, right frontal asymmetry
reflects greater relative right frontal activation (i.e., lower EEG power values at right
hemisphere) and left frontal asymmetry reflects the opposite pattern.
Frontal EEG asymmetry is typically conceptualized as reflecting approach and
avoidance motivational tendencies across the lifespan (Coan & Allen, 2004; Harmon-
Jones & Allen, 1998). During infancy and early childhood, frontal EEG asymmetry
is also associated with temperament- based differences in emotion reactivity and
emotion regulation (Fox, 1994). Thus, when examining resting state (or baseline) EEG,
left frontal asymmetry is linked to approach motivation, positive emotions (plus the
approach emotion of anger; Harmon-Jones & Allen, 1998), and better emotion regula-
tion abilities. Right frontal asymmetry is associated with withdrawal motivation, nega-
tive emotions, and less skill at emotion regulation. Baseline frontal EEG asymmetry
patterns are conceptualized as trait or biomarker indicators of motivational and emotion
tendencies, whereas task-related frontal EEG asymmetry patterns are considered to be
state indicators of reactivity and regulation (Allen & Reznik, 2015; Coan & Allen, 2004;
Diaz & Bell, 2012).
Developmental research has examined links between frontal alpha (6–9 Hz) EEG
asymmetry and temperament, which is defined as biologically based individual
differences in emotion reactivity and emotion regulation (Rothbart & Bates, 2006).
Other work has used resting frontal EEG asymmetry in early development to pre-
dict later socioemotional outcomes. A third line of frontal EEG asymmetry research
308 KIMBERLY CUEVAS and MARTHA ANN BELL
includes parenting behaviors in examining both simple and complex links between
child trait asymmetry and later outcomes. We briefly discuss each of these three areas by
highlighting our research and selectively reviewing the work of others.
hemisphere frontal EEG power for girls, whereas early right hemisphere EEG power
predicted subsequent later change in left hemisphere frontal EEG power for boys. These
findings might account for reports of sex differences in temperament and related be-
havior problems (e.g., Else-Quest et al., 2006).
challenging and negative behaviors while playing with mothers had mothers with right
frontal EEG asymmetry after controlling for mother’s own negativity (Atzaba-Poria
et al., 2017). Thus, mother and child frontal EEG asymmetry patterns during dyadic
recordings are influenced by partner’s level of negativity.
Although there exist associations between frontal 6-9 Hz EEG asymmetry and some
aspect of emotion during infancy and early childhood (and beyond), the link between
frontal asymmetry and emotion is likely more than a simple association, as we have
described them here. It may be that frontal EEG asymmetry is either a mediator or a
moderator of emotion (Coan & Allen, 2004; Reznick & Allen, 2018). Two examples from
our own work highlight frontal asymmetry as a moderator. Resting frontal EEG asym-
metry was recorded in a large group of 5-month-old infants. Behavioral coding included
mothers’ responsive behaviors during free play with her infant, as well as infants’ regula-
tory behaviors and level of negativity during the arm restraint procedure where mother
faces her infant and prevents the infant from moving his/her arms. Maternal responsive
behaviors during free play with infants prior to the arm restraint procedure predicted
infant regulatory behaviors for infants with left frontal asymmetry but predicted infant
negativity for infants with right frontal asymmetry (Swingler et al., 2014). Thus, the same
maternal behavior is associated with different infant regulatory behaviors depending on
infant frontal EEG asymmetry pattern, suggesting an advantage for infants with both
left frontal EEG asymmetry and an environment with responsive parenting.
In this same large group of infants, we examined age 5-month frontal EEG asymmetry
as a moderator of the link between maternal behavior and child negative affect at age
24 months. Maternal sensitive and responsive behaviors at age 5 months were associated
with less toddler negative affect only for children who also had left frontal EEG asymmetry
at age 5 months. The same pattern held when examining maternal behaviors at 24 months
of age (Diaz et al., 2019). We interpreted this to mean that the level of toddler negative
affect is influenced by the child’s own neurophysiology as well as the parenting context.
In sum, frontal alpha EEG asymmetry, conceptualized as approach-avoidance mo-
tivation, has been examined in studies of temperament, socio-emotional outcomes, and
parenting. The underlying neurological basis for frontal EEG asymmetry is typically
focused on reciprocal metabolic connections between prefrontal cortex and the amyg-
dala. Activation of the right amygdala is associated with increased dispositional (i.e.,
temperament-based) negative affect (Abercrombie et al., 1998); conversely, individuals
with greater left frontal metabolic rate exhibit decreased amygdala activation and thus
decreased negative affect (Davidson, 2001). Although frontal EEG asymmetry does not
measure amygdala activity, it may be measuring the frontal cortical activation linked
with amygdala connectivity.
relative to the other are associated with emotion reactivity or emotion regulation. In
studies of cognition, increases in EEG power values are associated with higher levels
of performance on cognitive tasks. In the adult EEG literature, researchers have long
focused on the 8–13 Hz peak in the adult spectrum. It is commonly reported that alpha
activity (8–13 Hz) exhibits desynchronization (decreased power values) during increased
cortical processing (cognitive, affective), although there are some reports of alpha syn-
chronization (increased power values) during long-term memory tasks (Klimesch et al.,
1999). There are also suggestions that alpha power is related to an inhibitory filter that is
reflected in alpha synchronization (Klimesch, 2012).
Researchers have noted that adult theta activity (4–7 Hz) exhibits synchronization
during memory and attention tasks (e.g., Burgess & Gruzelier, 2000; Klimesch, 1999).
Thus, for the mature EEG signal, unique patterns of fluctuations in power levels at the
defined frequency bands are associated with different types of cognitive processing
(Cavanaugh, 2019). This type of information is lacking with respect to the EEG signal
recorded from infants and young children. Currently, it appears that infant and child
6–9 Hz alpha behaves like adult theta or adult alpha depending on the type of cogni-
tive processing, and definitely behaves like adult alpha during affective processing
(Anderson & Perone, 2018; Bell & Cuevas, 2012; Saby & Marshall, 2012). Systematic de-
velopmental research is needed to understand the neurophysiological and functional
significance of infant and child alpha at 6–9 Hz.
The brain’s electrical activity consists of multiple, co-occurring oscillatory rhythms that
underlie neural computation, communication, and transmission of information be-
tween brain networks—functions critical to simple and complex brain processes (Lopes
da Silva, 2013). Transient interactions among neural rhythms are critical for the coord-
ination of neural, cognitive, affective, and behavioral processes (Jensen & Colgin, 2007);
yet, holistic ontogenetic analysis is largely missing from the field. As shown throughout
this chapter, developmental analyses of EEG rhythms typically focus on a single oscil-
latory rhythm either cross-sectionally or at a particular age. These approaches preclude
integrative analysis of co-occurring oscillatory rhythms, including associated ontogen-
etic changes in neural activity.
The current review focused on individual frequency bands during early develop-
ment, with emphasis on the infant and child alpha 6–9 Hz band. We were among the
first to emphasize this particular frequency band, based on spectral plots of our ini-
tial infant longitudinal study (Figure 13.1; Bell & Fox, 1992). Stroganova and colleagues
(1999) were among the first to label 6–9 Hz as “alpha” based on their spectral plots of the
spatial and amplitude properties of this frequency band when infants were in a room
312 KIMBERLY CUEVAS and MARTHA ANN BELL
with lights turned off compared with quiet alert baseline. More recently, developmental
investigations of the central 6–9 Hz band have identified functional properties of the
mu rhythm during action perception and execution (see Cuevas et al., 2014, for review).
Furthermore, resting-state 6–9 Hz alpha rhythm measures are associated with con-
current and future affective, cognitive, and motor processes during infancy and early
childhood (see Anderson & Perone, 2018, for review). Despite the potential functional
significance of alpha (and gamma) resting-state activity, relatively little is known about
other co-occurring resting-state oscillatory rhythms, such as theta and beta frequency
bands. In sum, what is missing is systematic examination of the alpha family frequency
in conjunction with other co-occurring oscillatory rhythms to aid our understanding of
their frequency signatures, spatial topography, and functional significance, all within a
developmental time course.
Klimesch (2012) proposed that the alpha rhythm plays a central role in inhibition
and timing of brain activity, and thus may be importantly involved in activity at other
frequency bands. Accordingly, he proposed that for adult EEG, other frequency bands
are best defined by a frequency architecture focused on the alpha band and its activity.
Because alpha is a central process in modulating resting state and task behaviors in the
awake brain, the best way to understand frequency bands is by examining bands that
have a harmonic mean with alpha. The neurophysiology literature is missing this type of
functional thinking about alpha from a developmental point of view.
At the same time, there is increasing interest in how multiple oscillatory rhythms
interact to produce complex brain functions. Cross-frequency coupling and band
power ratios are example measures that permit more integrative analysis of resting-state
EEG as well as during cognitive, perceptual, affective, and motor processes. Their ap-
plication in the developmental literature has been limited to date (e.g., Brooker et al.,
2016; Perone et al., 2018; Stamoulis et al., 2015), but they provide a promising avenue
for understanding ontogenetic changes in neural oscillations and corresponding brain-
behavior associations (e.g., Brooker et al., 2021).
Finally, there have been recent initiatives aimed at enhancing the interpretability
and scientific rigor of pediatric EEG. One line of work has focused on factors related
to attrition and EEG data loss for developmental samples with recommendation for
the field (e.g., van der Velde & Junge, 2020). Concurrently, efforts to establish standard
automated, freely-available, data preprocessing pipelines designed specifically for pedi-
atric EEG (e.g., MADE and HAPPE) are promising in terms of increasing the amount of
useable data as well as improving reliability and cross-lab comparisons (Debnath et al.,
2020; Gabard-Durnam et al., 2018). Furthermore, recent evidence suggests that analysis
of infant and child oscillatory rhythms will be enhanced via techniques that extract the
periodic and aperiodic (1/f) components from the neural power spectra (Cellier et al.,
2021; Schaworonkow & Voytek, 2021). Even with the aforementioned strategies, analysis
of the psychometric properties of developmental EEG measures (e.g., longitudinal sta-
bility, reliability, validity) highlights important considerations for pediatric data (Anaya
et al., 2021; Vincent et al., 2021). At a time of pre-registration and open-access initiatives,
EEG FREQUENCY DEVELOPMENT ACROSS INFANCY AND CHILDHOOD 313
we are particularly excited about the widespread effects these methodological, theoret-
ical, and analytical approaches will have on the field.
Related Topics
References
Abercrombie, H. C., Schaefer, S. M., Larson, K. L., Oakes, T. R., Lindgre, K. A., . . . Davidson, R.
J. (1998). Metabolic rate in the right amygdala predicts negative affect in depressed patients.
NeuroReport, 9, 3301–3307.
Allen, J. J. B. & Reznik, S. J. (2015). Frontal EEG asymmetry as a promising marker of depression
vulnerability: Summary and methodological considerations. Current Opinion in Psychology,
4, 93–97. https://doi.org/10.1016/j.copsyc.2014.12.017
Anaya, B., Ostlund, B., LoBue, V., Buss, K., & Pérez-Edgar, K. (2021). Psychometric properties
of infant electroencephalography: Developmental stability, reliability, and construct val-
idity of frontal alpha asymmetry and delta-beta coupling. Developmental Psychobiology, 63,
e22178. https://doi.org/10.1002/dev.22178
Anderson, A. J. & Perone, S. (2018). Developmental change in the resting state electroenceph-
alogram: Insights into cognition and the brain. Brain and Cognition, 126, 40–52. https://doi.
org/10.1016/j.bandc.2018.08.001
Atzaba-Poria, N., Deater-Deckard, K., & Bell, M. A. (2017). Mother-child interaction: Links
between mother and child frontal electroencephalograph asymmetry and negative be-
havior. Child Development, 88, 544–554. doi:10.111/cdev.12583
Başar, E. (2013). A review of gamma oscillations in healthy subjects and in cognitive impair-
ment. International Journal of Psychophysiology, 90, 99–117. http://dx.doi.org/10.1016/j.ijpsy
cho.2013.07.005
314 KIMBERLY CUEVAS and MARTHA ANN BELL
Begus, K. & Bonawitz, E. (2020). The rhythm of learning: Theta oscillations as an index of
active learning in infancy. Developmental Cognitive Neuroscience, 45, 100810. https://doi.
org/10.1016/j.dcn.2020.100810
Begus, K., Southgate, V., & Gliga, T. (2015). Neural mechanisms of infant learning: differences
in frontal theta activity during object exploration modulate subsequent object recognition.
Biology Letters, 11, 20150041. http://dx.doi.org/10.1098/rsbl.2015.0041
Bell, M. A. (2001). Brain electrical activity associated with cognitive processing during a
looking version of the A-not-B task. Infancy, 2, 311–330.
Bell, M. A. (2002). Power changes in infant EEG frequency bands during a spatial working
memory task. Psychophysiology, 39, 450–458.
Bell, M. A. (2012). A psychobiological perspective on working memory performance at 8 months
of age. Child Development, 83, 251–265. http://dx.doi.org/10.1111/j.1467-8624.2011.01684.x
Bell, M. A. & Adams, S. E. (1999). Comparable performance on looking and reaching versions
of the A-not-B task at 8 months of age. Infant Behavior and Development, 22, 221–235.
Bell, M. A. & Cuevas, K. (2012). The use of EEG to study cognitive development: Issues and
practices. Journal of Cognition and Development, 13, 281–294. http://dx.doi.org/10.1080/
15248372.2012.691143
Bell, M. A. & Cuevas, K. (2016). Psychobiology of executive function in early development.
In J. A. Griffin, P. McCardle, & L. S. Freund (Eds.), Executive function in preschool-age chil-
dren: Integrating measurement, neurodevelopment, and translational research (pp. 157–179).
APA Press. http://dx.doi.org/10.1037/14797-008
Bell, M. A. & Fox, N. A. (1992). The relations between frontal brain electrical activity and cog-
nitive development during infancy. Child Development, 63, 1142–1163.
Bell, M. A. & Fox, N. A. (1997). Individual differences in object permanence performance at
8 months: Locomotor experience and brain electrical activity. Developmental Psychobiology,
31, 287–297.
Bell, M. A. & Wolfe, C. D. (2007). Changes in brain functioning from infancy to early
childhood: Evidence from EEG power and coherence during working memory tasks.
Developmental Neuropsychology, 31, 21–38.
Benasich, A. A., Gou, Z., Choudhury, N., & Harris, K. D. (2008). Early cognitive and language
skills are linked to resting frontal gamma power across the first 3 years. Behavioural Brain
Research, 195, 215–222. doi:10.1016/j.bbr.2008.08.049
Berchicci, M., Zhang, T., Romero, L., Peters, A., Annett, R., Teuscher, U., . . . Comani, S. (2011).
Development of mu rhythm in infants and preschool children. Developmental Neuroscience,
33, 130–143. doi:10.1159/000329095
Berger, H. (1929). Über das Elektrenkephalogramm des Menschen. Archiv für Psychiatrie und
Nervenkrankheiten, 87, 527–570.
Berger, H. (1932). Über das Elektroenzephalogramm des Menschen. Fünfter Bericht. Archiv für
Psychiatrie und Nervenkrankheiten, 98, 231–254.
Blankenship, T. L., Broomell, A. P. R., & Bell, M. A.(2018). Semantic future thinking and execu-
tive functions at age 4: The moderating role of frontal brain electrical activity. Developmental
Psychobiology, 60, 608–614. https://doi.org/10.1002/dev.21629
Bowman, L. C., Thorpe, S. G., Cannon, E. N., & Fox, N. A. (2017). Action mechanisms for social
cognition: Behavioral and neural correlates of developing Theory of Mind. Developmental
Science, 20, e12447. https://doi.org/10.1111/desc.12447
Braithwaite, E. K., Jones, E. J. H., Johnson, M. H., & Holmboe, K. (2020). Dynamic modu-
lation of frontal theta power predicts cognitive ability in infancy. Developmental Cognitive
Neuroscience, 45, 100818. https://doi.org/10.1016/j.dcn.2020.100818
EEG FREQUENCY DEVELOPMENT ACROSS INFANCY AND CHILDHOOD 315
Brito, N. H., Fifer, W. P., Myers, M. M., Elliott, A. J., & Noble, K. G. (2016). Associations
among family socioeconomic status, EEG power at birth, and cognitive skills during in-
fancy. Developmental Cognitive Neuroscience, 19, 144–151. http://dx.doi.org/10.1016/
j.dcn.2016.03.004
Brooker, R. J., Mistry-Patel, S., Kling, J. L., & Howe, H. A. (2021). Deriving within-person
estimates of delta-beta coupling: A novel measure for identifying individual differences in
emotion and neural function in childhood. Developmental Psychobiology, 63, e22172. https://
doi.org/10.1002/dev.22172
Brooker, R. J., Phelps, R. A., Davidson, R. J., & Goldsmith, H. H. (2016). Contextual
differences in delta beta coupling are associated with neuroendocrine reactivity in infants.
Developmental Psychobiology, 58, 406–418. doi:10.1002/dev.21381
Broomell, A. P. R. & Bell, M. A. (2017). Inclusion of a mixed condition makes the day/night
task more analogous to the adult Stroop. Developmental Neuropsychology, 42, 241–252.
doi:10.1080/87565641.2017.1309655
Broomell, A. P. R., Salva, J., & Bell, M. A. (2019). Infant electroencephalogram coherence and
toddler inhibition are associated with social responsiveness at age 4. Infancy, 24, 43–56.
doi:10.1111/infa.12273
Bryant, L. J. & Cuevas, K. (2019). Effects of active and observational experience on EEG activity
during early childhood. Psychophysiology, 56, e13360. https://doi.org/10.1111/psyp.13360
Burgess, A. P. & Gruzelier, J. H. (2000). Short duration power changes in the EEG during
recognition memory for words and faces. Psychophysiology, 37, 596–606. https://doi.org/
10.1111/1469-8986.3750596
Buss, K. A., Schumacher, J. R. M., Dolski, I., Kalin, N. H., Goldsmith, H. H., & Davidson R.
J. (2003). Right frontal brain activity, cortisol, and withdrawal behavior in 6-month-old
infants. Behavioral Neuroscience, 117, 11–20. doi:10.1037/0735-7044.117.1.11
Cannon, E. N., Simpson, E. A., Fox, N. A., Vanderwert, R. E., Woodward, A. L., & Ferrari, P. F.
(2016). Relations between infants’ emerging reach-grasp competence and event-related de-
synchronization in EEG. Developmental Science, 19, 50–62. https://doi.org/10.1111/desc.12295
Cavanaugh, J. F. (2019). Electrophysiology as a theoretical and methodological hub for the
neural sciences. Psychophysiology, 56, e13314. https://doi.org/10.1111/psyp.13314
Cellier, D., Riddle, J., Petersen, I., & Hwang, K. (2021). The development of theta and alpha
neural oscillations from ages 3 to 34 years. Developmental Cognitive Neuroscience, 50,
100969. https://doi.org/10.1016/j.dcn.2021.100969
Cheyne, D., Jobst, C., Tesan, G., Crain, S., & Johnson, B. (2014). Movement- related
neuromagnetic fields in preschool age children. Human Brain Mapping, 35, 4858–4875.
doi:10.1002/hbm.22518
Coan, J. A. & Allen, J. J. B. (2004). Frontal EEG asymmetry as a moderator and mediator of
emotion. Biological Psychology, 67, 7–50. https://doi.org/10.1016/j.biopsycho.2004.03.002
Crone, N. E., Miglioretti, D. L., Gordon, B., Sieracki, J. M., Wilson, M. T., Uematsu,
S., & Lesser, R. P. (1998). Functional mapping of human sensorimotor cortex with
electrocorticographic spectral analysis. I. Alpha and beta event-related desynchron-
ization. Brain, 121, 2271–2299.
Csibra, G., Davis, G., Spratling, M. W., & Johnson, M. H. (2000). Gamma oscillations and ob-
ject processing in the infant brain. Science, 290, 1582–1585. doi:10.1126/science.290.5496.1582
Cuevas, K. & Bell, M. A. (2010). Developmental progression of looking and reaching perform-
ance of the A-not-B task. Developmental Psychology, 46, 1363–1371. http://dx.doi.org/10.1037/
a0020185
316 KIMBERLY CUEVAS and MARTHA ANN BELL
Cuevas, K. & Bell, M. A. (2011). EEG and ECG from 5 to 10 months of age: Developmental
changes in baseline activation and cognitive processing during a working memory task.
International Journal of Psychophysiology, 80, 119–128. http://dx.doi.org/10.1016/j.ijpsy
cho.2011.02.009
Cuevas, K., Bell, M. A., Marcovitch, S., & Calkins, S. D. (2012a). EEG and heart rate measures of
working memory at 5 and 10 months of age. Developmental Psychology, 48, 907–917. http://
dx.doi.org/10.1037/a0026448
Cuevas, K., Calkins, S. D., & Bell, M. A. (2016). To Stroop or not to Stroop: Sex-related
differences in brain-behavior associations during early childhood. Psychophysiology, 53, 30–
40. http://dx.doi.org/10.1111/psyp.12464
Cuevas, K., Cannon, E. N., Yoo, K., & Fox, N. A. (2014). The infant EEG mu
rhythm: Methodological considerations and best practices. Developmental Review, 34, 26–
43. http://dx.doi.org/10.1016/j.dr.2013.12.001
Cuevas, K., Hubble, M., & Bell, M. A. (2012b). Early childhood predictors of post-kindergarten
executive function: Behavior, parent-report, and psychophysiology. Early Education and
Development, 23, 59–73. http://dx.doi.org/10.1080/10409289.2011.611441
Cuevas, K., Raj, V., & Bell, M. A. (2012c). A frequency band analysis of two-year-olds’ memory
processes. International Journal of Psychophysiology, 83, 315–322. http://dx.doi.org/10.1016/
j.ijpsycho.2011.11.009
Cuevas, K., Raj, V., & Bell, M. A. (2012d). Functional connectivity and infant spatial working
memory: A frequency band analysis. Psychophysiology, 49, 271–280. http://dx.doi.org/
10.1111/j.1469-8986.2011.01304.x
Cuevas, K., Rajan, V., & Bryant, L. J. (2018). Emergence of executive function in infancy. In S. A.
Wiebe & J. Karbach (Eds.), Executive function: Development across the life span (pp. 11–28).
Routledge.
Cuevas, K., Swingler, M. M., Bell, M. A., Marcovitch, S., & Calkins, S. D. (2012e). Measures
of frontal functioning and the emergence of inhibitory control processes at 10 months
of age. Developmental Cognitive Neuroscience, 2, 235–243. http://dx.doi.org/10.1016/
j.dcn.2012.01.002
Davidson, R. J. (2001). Toward a biology of personality and emotion. Annals of the New York
Academy of Sciences, 935, 191–207.
Davidson, R. J. & Fox, N. A. (1989). Frontal brain asymmetry predicts infants’ response to
maternal separation. Journal of Abnormal Psychology, 98, 127–131. https://doi.org/10.1037/
0021-843X.98.2.127
Dawson, G., Frey, K., Panagiotides, H., Osterling, J., & Hessl, D. (1997). Infants of depressed
mothers exhibit atypical frontal brain activity: A replication and extension of previous
findings. Journal of Child Psychology and Psychiatry, 38, 179–186. https://doi.org/10.1111/
j.1469-7610.1997.tb01852.x
Debnath R., Buzzell G. A., Morales S., Bowers M. E., Leach S. C., Fox N. A. (2020). The
Maryland analysis of developmental EEG (MADE) pipeline. Psychophysiology, 57, e13580.
https://doi.org/10.1111/psyp.13580
Diamond, A. (2013). Executive functions. Annual Review of Psychology, 64, 135–168. http://
dx.doi.org/10.1146/annurev-psych-113011-143750
Diaz, A. & Bell, M. A. (2012). Frontal EEG asymmetry and fear reactivity in different contexts
at 10 months. Developmental Psychobiology, 54, 536–545. doi:10.1002/dev.20612
Diaz, A., Swingler, M. M., Tan, L., Smith, C. L., Calkins, S. D., & Bell, M. A. (2019). Infant
frontal EEG asymmetry moderates the association between maternal behavior and toddler
EEG FREQUENCY DEVELOPMENT ACROSS INFANCY AND CHILDHOOD 317
Gartstein, M. A., Hancock, G. R., Potapova, N. V., Calkins, S. D., & Bell, M. A. (2019). Modeling
development of frontal electroencephalogram (EEG) asymmetry: Sex differences and links
with temperament. Developmental Science, 23, e12891. https://doi.org/10.1111/desc.12891
Gasser, T., Verleger, R., Bacher, P., & Sroka, L. (1988). Development of the EEG of school-age
children and adolescents. I. Analysis of band power. Electroencephalography and Clinical
Neurophysiology, 69, 91–99.
Gastaut, H., Dongier, M., & Courtois, G. (1954). On the significance of “wicket rhythms”
(“rythmes en arceau”) in psychosomatic medicine. Electroencephalography and Clinical
Neurophysiology, 6, 687.
Gatzke-Kopp, L. M. (2016). Diversity and representation: Key issues for psychophysiological
science. Psychophysiology, 53, 3–13. https://doi.org/10.1111/psyp.12566
Gerson, S. A., Bekkering, H., & Hunnius, S. (2015). Short-term motor training, but not obser-
vational training, alters neurocognitive mechanisms of action processing in infancy. Journal
of Cognitive Neuroscience, 27, 1207–1214. https://doi.org/10.1162/jocn_a_00774
Ghim, H.-R. (1990). Evidence for perceptual organization in infants: Perception of subjective
contours by young infants. Infant Behavior and Development, 13, 221–248. https://doi.org/
10.1016/0163-6383(90)90032-4
Gibbs, F. A. & Knott, J. R. (1949). Growth of the electrical activity of the cortex.
Electroencephalography and Clinical Neurophysiology, 1, 223–229.
Gliga, T., Volein, A., & Csibra, G. (2010). Verbal labels modulate perceptual object processing
in 1-year-old children. Journal of Cognitive Neuroscience, 22, 2781–2789.
Hagne, I. (1968). Development of the waking EEG in normal infants during the first year of life.
A longitudinal study including automatic frequency analysis. In P. Kellaway and I. Petersen
(Eds.), Clinical electroencephalography of children (pp. 97–118). Grune & Stratton.
Hagne, I., Persson, J., Magnusson, R., & Petersen, I. (1973). Spectral analysis via fast Fourier
transform of waking EEG in normal infants. In P. Kellaway and I. Petersen (Eds.),
Automation of clinical electroencephalography (pp. 103–143). Raven Press.
Hannesdottir, D. K., Doxie, J., Bell, M. A., Ollendick, T. H., & Wolfe, C. D. (2010). A longi-
tudinal study of emotion regulation and anxiety in middle childhood: Associations with
frontal EEG asymmetry in early childhood. Developmental Psychobiology, 52, 197–204.
doi:10.1002/dev.20425
Hari, R., & Salmelin, R. (1997). Human cortical oscillations: a neuromagnetic view through the
skull. Trends in Neurosciences, 20, 44–49.
Harmon-Jones, E. & Allen, J. J. B. (1998). Anger and frontal brain activity: EEG asymmetry con-
sistent with approach motivation despite negative affective valence. Journal of Personality
and Social Psychology, 74, 1310–1316. https://doi.org/10.1037/0022-3514.74.5.1310
Henriques, J. B. & Davidson, R. J. (1990). Regional brain electrical asymmetries discriminate
between previously depressed and healthy control subjects. Journal of Abnormal Psychology,
99, 22–31. https://doi.org/10.1037/0021-843X.99.1.22
Howarth, G. Z., Fettig, N. B., Curby, T. W., & Bell, M. A. (2016). Frontal electroencephalogram
asymmetry and temperament across infancy and early childhood: An exploration of sta-
bility and bidirectional relations. Child Development, 87, 465–476. doi:10.1111/cdev.12466
Jarvelainen, J., Schurmann, M., & Hari, R. (2004). Activation of the human primary
motor cortex during observation of tool use. NeuroImage, 23, 187– 192. doi:10.1016/
j.neuroimage.2004.06.010
Jasper, H. H. & Andrews, H. L. (1938). Electro-encephalography. III. Normal differenti-
ation of occipital and precentral regions in man. Archives of Neurology and Psychiatry, 39,
96–115.
EEG FREQUENCY DEVELOPMENT ACROSS INFANCY AND CHILDHOOD 319
Lusby, C. M., Goodman, S. H., Bell, M. A., & Newport, D. J. (2014). Electroencephalogram
patterns in infants of depressed mothers. Developmental Psychobiology, 56, 459–473.
doi:10.1002/dev.21112
MacNeill, L. A., Ram, N., Bell, M. A., Fox, N. A., & Perez-Edgar, K. (2018). Trajectories of
infants’ biobehavioral development: Timing and rate of A-not-B performance gains and
EEG maturation. Child Development, 89, 711–724. doi:10.1111/cdev.13022
Marshall, P. J., Bar-Haim, Y., & Fox, N. A. (2002). Development of the EEG from 5 months to
4 years of age. Clinical Neurophysiology, 113, 1199–1208.
Marshall, P. J. & Meltzoff, A. N. (2011). Neural mirroring systems: Exploring the EEG mu
rhythm in human infancy. Developmental Cognitive Neuroscience, 1, 110–123. https://doi.org/
10.1016/j.dcn.2010.09.001
Matsuura, M., Yamamoto, K., Fukuzawa, H., Okubo, Y., Uesugi, H., Moriiwa, M., . . . Shimazono,
Y. (1985). Age development and sex differences of various EEG elements in healthy chil-
dren and adults— Quantification by a computerized wave form recognition method.
Electroencephalography and Clinical Neurophysiology, 60, 394–406.
Meyer, M., Braukmann, R., Stapel, J. C., Bekkering, H., & Hunnius, S. (2016). Monitoring
others’ errors: The role of the motor system in early childhood and adulthood. British
Journal of Developmental Psychology, 34, 66–85. doi:10.1111/bjdp.12101
Miskovic, V., Ma, X., Chou, C.-A., Fan, M., Owens, M., Sayama, H., & Gibb, B. E. (2015).
Developmental changes in spontaneous electrocortical activity and network organization
from early to late childhood. NeuroImage, 118, 237–247. http://dx.doi.org/10.1016/j.neuroim
age.2015.06.013
Morasch, K. C. & Bell, M. A. (2011). The role of inhibitory control in behavioral and physio-
logical expressions of toddler executive function. Journal of Experimental Child Psychology,
108, 593–606. https://doi.org/10.1016/j.jecp.2010.07.003
Muthukumaraswamy, S. D. & Johnson, B. W. (2004). Changes in Rolandic mu rhythm
during observation of a precision grip. Psychophysiology, 41, 152– 156. doi:10.1046/
j.1469-8986.2003.00129.x
Nikitina, G. M., Posikera, I. N., & Stroganova, T. A. (1985). Central organization of emotional
reaction of infants during first year of life. Ontogenesis of the Brain, 4, 223–227.
Ogawa, T., Sugiyama, A., Ishiwa, S., Suzuki, M., Ishihara, T., & Sato, K. (1984). Ontogenic devel-
opment of autoregressive component waves of waking EEG in normal infants and children.
Brain & Development, 6, 289–303.
Orekhova, E. V., Stroganova, T. A., & Posikera, I. N. (1999). Theta synchronization during
sustained anticipatory attention in infants over the second half of the first year of life.
International Journal of Psychophysiology, 32, 151–172.
Orekhova, E. V., Stroganova, T. A., Posikera, I. N., & Elam, M. (2006). EEG theta rhythm in
infants and preschool children. Clinical Neurophysiology, 117, 1047– 1062. doi:10.1016/
j.clinph.2005.12.027
Ortiz-Mantilla, S., Hamalainen, J. A., Realpe-Bonilla, T., & Benasich, A. A. (2016). Oscillatory
dynamics underlying perceptual narrowing of native phoneme mapping from 6 to 12 months
of age. The Journal of Neuroscience, 36, 12095–12105. doi:10.1523/JNEUROSCI.1162-16.2016
Paredes, M. F., James, D., Gil-Perotin, S., Kim, H., Cotter, J. A., Ng, C., . . . Alvarez-Buylla, A.
(2016). Extensive migration of young neurons into the infant human frontal lobe. Science,
354, aaf7073. doi:10.1126/science.aaf7073
Paul, K., Dittrichova, J., & Papousek, H. (1996). Infant feeding behavior: Development in
patterns and motivation. Developmental Psychobiology, 29, 563–576.
EEG FREQUENCY DEVELOPMENT ACROSS INFANCY AND CHILDHOOD 321
Perone, S., Palanisamy, J., & Carlson, S. M. (2018). Age-related change in brain rhythms from
early to middle childhood: Links to executive function. Developmental Science, 21, e12691.
https://doi.org/10.1111/desc.12691
Pfurtscheller, G., Brunner, C., Schlogel, A., & Lopes da Silva, F. H. (2006). Mu rhythm (de)
synchronization and EEG single- trial classification of different motor imagery tasks.
NeuroImage, 31, 153–159. doi:10.1016/j.neuroimage.2005.12.003
Pfurtscheller, G., Stančák, A., & Edlinger, G. (1997). On the existence of different types of cen-
tral beta rhythms below 30 Hz. Electroencephalography and Clinical Neurophysiology, 102,
316–325.
Pivik R. T., Andres, A., Tennal, K. B., Gu, Y., Downs, H., Bellando, B. J. . . . Badger, T. M. (2019).
Resting gamma power during the postnatal critical period for GABAergic system develop-
ment is modulated by infant diet and sex. International Journal of Psychophysiology, 135, 73–
94. https://doi.org/10.1016/j.ijpsycho.2018.11.004
Pivik, R. T., Broughton, R. J., Coppola, R., Davidson, R. J., Fox, N. A., & Nuwer, M. R. (1993).
Guidelines for the recording and quantitative analysis of electroencephalographic activity
in research contexts. Psychophysiology, 30, 547–558. https://doi.org/10.1111/j.1469-8986.1993.
tb02081.x
Poole, K. L., Santesso, D. L., Van Lieshout, R. J., & Schmidt, L. A. (2018). Trajectories of frontal
brain activity and socio-emotional development in children. Developmental Psychobiology,
60, 353–363. doi:10.1002/dev.21620
Rayson, H., Bonaiuto, J. J., Ferrari, P. F., & Murray, L. (2017). Early maternal mirroring predicts
infant motor system activation during facial expression observation. Scientific Reports, 7,
11738. doi:10.1038/s41598-017-12097-w
Reznik, S. J. & Allen, J. J. B. (2018). Frontal asymmetry as a mediator and moderator of
emotion: An updated review. Psychophysiology, 55, e12965. https://doi.org/10.1111/psyp.12965
Rothbart, M. K., & Bates, J. E. (2006). Temperament. In W. Damon & R.M. Lerner (Series Eds.)
& N. Eisenberg (Vol. Ed.), Handbook of child psychology: Vol. 3. Social, emotional, and per-
sonality development, 6th ed. (pp. 99–166). Wiley.
Saby, J. N. & Marshall, P. J. (2012). The utility of EEG band power analysis in the study of in-
fancy and early childhood. Developmental Neuropsychology, 37, 253–273. https://doi.org/
10.1080/87565641.2011.614663
Schaworonkow, N. & Voytek, B. (2021). Longitudinal changes in aperiodic and periodic ac-
tivity in electrophysiological recordings in the first seven months of life. Developmental
Cognitive Neuroscience, 47, 100895. https://doi.org/10.1016/j.dcn.2020.100895
Short, S. J., Elison, J. T., Goldman, B. D., Styner, M., Gu, H., Connelly, M., ... Gilmore, J. H.
(2013). Associations between white matter microstructure and infants’ working memory.
NeuroImage, 64, 156–166. https://doi.org/10.1016/j.neuroimage.2012.09.021
Smith, C. L. & Bell, M. A. (2010). Stability in infant frontal asymmetry as a predictor of toddler-
hood internalizing and externalizing behaviors. Developmental Psychobiology, 52, 158–167.
doi:10.1002/dev.20427
Smith, C. L., Diaz, A., Day, K. L., & Bell, M. A. (2016). Infant frontal electroencephalogram
asymmetry and negative emotional reactivity as predictors of toddlerhood effortful control.
Journal of Experimental Child Psychology, 142, 262–273. doi:10.1016/j.jecp.2015.09.031
Smith, J. R. (1938). The electroencephalogram during normal infancy and childhood: II. The
nature of the growth of the alpha waves. The Journal of Genetic Psychology, 53, 455–469.
Smith, J. R. (1939). The “occipital” and “pre-central” alpha rhythms during the first two years.
The Journal of Psychology, 7, 223–226.
322 KIMBERLY CUEVAS and MARTHA ANN BELL
Smith, J. R. (1941). The frequency growth of the human alpha rhythms during normal infancy
and childhood. The Journal of Psychology, 11, 177–198.
Stamoulis, C., Vanderwert, R. E., Zeanah, C. H., Fox, N. A., & Nelson, C. A. (2015). Early psy-
chosocial neglect adversely impacts developmental trajectories of brain oscillations and their
interactions. Journal of Cognitive Neuroscience, 27, 2512–2528. doi:10.1162/jocn_a_00877
Stroganova, T. A. & Orekhova, E. V. (2007). EEG and infant states. In M. de Haan (Ed.), Infant
EEG and event-related potentials (pp. 251–287). Psychology Press.
Stroganova, T. A., Orekhova, E. V., & Posikera, I. N. (1998). Externally and internally controlled
attention in infants: An EEG study. International Journal of Psychophysiology, 30, 339–351.
Stroganova, T. A., Orekhova, E. V., & Posikera, I. N. (1999). EEG alpha rhythm in infants.
Clinical Neurophysiology, 110, 997–1012.
Swingler, M. M., Perry, N. B., Calkins, S. D., & Bell, M. A. (2014). Maternal sensitivity and in-
fant response to frustration: The moderating role of EEG asymmetry. Infant Behavior &
Development, 37, 523–535. doi:10.1016.j.infbeh.2014.06.010
Swingler, M. M., Willoughby, M. T., & Calkins, S. D. (2011). EEG power and coherence during
preschoolers’ performance of an executive function battery. Developmental Psychobiology,
53, 771–784. doi:10.1002/dev.20588
Takano, T. & Ogawa, T. (1998). Characterization of developmental changes in EEG-gamma
band activity during childhood using the autoregressive model. Acta Paediatrica Japonica,
40, 446–452.
Tarullo, A. R., Obradovic, J., Keehn, B., Rasheed, M. A., Siyal, S., Nelson, C. A., & Yousafzai,
A. K. (2017). Gamma power in rural Pakistani children: Links to executive function and
verbal ability. Developmental Cognitive Neuroscience, 26, 1–8. http://dx.doi.org/10.1016/
j.dcn.2017.03.007
Thatcher, R. W., Krause, P. J., & Hrybyk, M. (1986). Cortico- cortical associations and
EEG coherence: A two- compartmental model. Electroencephalography and Clinical
Neurophysiology, 64, 123–143. https://doi.org/10.1016/0013-4694(86)90107-0
Thorpe, S. G., Cannon, E. N., & Fox, N. A. (2016). Spectral and source structural development
of mu and alpha rhythms from infancy through adulthood. Clinical Neurophysiology, 127,
254–269. http://dx.doi.org/10.1016/j.clinph.2015.03.004
Tierney, A., Strait, D. L., O’Connell, S., & Kraus, N. (2013). Developmental changes in resting
gamma power from age three to adulthood. Clinical Neurophysiology, 124, 1040–1042. http://
dx.doi.org/10.1016/j.clinph.2012.09.023
Tomalski, P., Moore, D. G., Ribeiro, H., Axelsson, E. L., Murphy, E., Karmiloff-Smith,
A., . . . Kushnerenko, E. (2013). Socioeconomic status and functional brain development –
associations in early infancy. Developmental Science, 16, 676–687. doi:10.1111/desc.12079
Upshaw, M. B., Bernier, R. A., & Sommerville, J. A. (2016). Infants’ grip strength predicts
mu rhythm attenuation during observation of lifting actions with weighted blocks.
Developmental Science, 19, 195–207. doi:10.1111/desc.12308
van der Velde, B., & Junge, C. (2020). Limiting data loss in infant EEG: putting hunches to
the test. Developmental Cognitive Neuroscience, 45, 100809. https://doi.org/10.1016/
j.dcn.2020.100809
Vanderwert, R. E., Simpson, E. A., Paukner, A., Suomi, S. J., Fox, N. A., & Ferrari, P. F. (2015).
Early social experience affects neural activity to affiliative facial gestures in newborn
nonhuman primates. Developmental Neuroscience, 37, 243–252. doi:10.1159/000381538
Vincent, K. M., Xie, W., & Nelson, C. A. (2021). Using different methods for calculating
frontal alpha asymmetry to study its development from infancy to 3 years of age in a large
EEG FREQUENCY DEVELOPMENT ACROSS INFANCY AND CHILDHOOD 323
DEVEL OPMENTA L RE SE A RC H
ON TIME-F RE QU E NC Y
ACTIVIT Y IN A D OL E S C E NC E
AND EARLY ADU LT H O OD
The notion that any complex system depends critically on its history (D’Souza &
Karmiloff-Smith, 2016) is central to a developmental perspective, which holds that the
state of an organism at any moment in its lifetime reflects its history, or trajectory in
developmental time. A full understanding of a human characteristic is only possible
through a systematic examination of how it came to be. The time-frequency features
of an adult study participant’s response to target stimuli in a visual oddball task, for
instance, also reflect the current state of the dynamic interplay among genetic, bio-
logical, psychological, and external forces over time that have given rise to the psy-
chological and neural processes, motivational state, and personality characteristics
influencing activity recorded in the laboratory. Understanding the nature of this
interplay is critical to understanding the response, whether frontal theta or parietal
delta activity.
An approach informed by this type of perspective examines phenotypic variation
over time to ask questions about mechanisms underlying a psychological, behavioral,
or psychopathological process (Rutter, 1988). A developmental perspective on cogni-
tive control, for example, can help elucidate the mechanisms underlying it by examining
how time-frequency features related to cognitive control develop over the course of the
lifespan. Discontinuities or inflection points in this process might reflect important
transitions, which can be exploited to test hypotheses about causal mechanisms. A de-
velopmental approach thus has the potential to inform us about how time-frequency
features come about, as well as what those features are.
DEVELOPMENTAL RESEARCH ON TIME-FREQUENCY 325
14.1.1 Fine-Grained Analysis
Although a developmental perspective offers a unique view of time-frequency features,
developmental research on time-frequency features can also offer purely pragmatic
advantages. Time- frequency decompositions of electroencephalographic (EEG)
signals into moment-by-moment activity at different frequencies lend themselves to a
fine-grained analysis of brain activity, with accumbent advantages for developmental
research.
35
40
Age 11 Age 11
30
Age 14 20
25
Age 17 0 Age 24
20
V)
15
Age 20
10
Age 24
5
-5
-200 0 200 400 600 800 1000 1200 1400
Latency (ms)
Figure 14.1 Grand mean ERP waveforms across five assessment waves spanning
preadolescence to early adulthood. ERPs represent the average response to target stimuli at Pz
among subjects assessed longitudinally at target ages of 11, 14, 17, 20, and 24. ERPs for successive
assessment waves are progressively smaller in amplitude, with the ERP at age 11 being the largest
and the ERP at age 24 the smallest. The inset depicts the same ERPs for the age-11 and age-24
assessments only.
0 500 1000 0 500 1000 0 500 1000 0 500 1000 0 500 1000
6 6 6 6 6
4 4 4 4 4
2 2 2 2 2
0 500 1000 0 500 1000 0 500 1000 0 500 1000 0 500 1000
Frequency (Hz)
6 6 6 6 6
4 4 4 4 4
2 2 2 2 2
0 500 1000 0 500 1000 0 500 1000 0 500 1000 0 500 1000
6 6 6 6 6
4 4 4 4 4
2 2 2 2 2
0 500 1000 0 500 1000 0 500 1000 0 500 1000 0 500 1000
6 6 6 6 6
4 4 4 4 4
2 2 2 2 2
0 500 1000 0 500 1000 0 500 1000 0 500 1000 0 500 1000
Time (ms)
Figure 14.2 Component weights (loadings) across assessment waves. The figure consists
of false color maps (heatmaps) of weights at each time-frequency bin from PCA of time-
frequency energy for all target trials. PCA was conducted separately for each assessment wave.
Components were matched by congruence coefficients computed between loadings at successive
assessment waves.
across assessments are evident as well, including shifts in the latency of the maximum
amount of energy in Component 5 and a trend toward increasing compactness in time-
frequency space of Component 1.
Thus, a common set of time-frequency features characterizes the EEG data across de-
velopment from early adolescence to early adulthood. Yet at the same time the subtle
developmental change in ERP morphology across waves evident in Figure 14.1 occurred
through small changes in this common set of features. Investigating differences between
those components that shifted in time and/or frequency with development and those
that did not will contribute to a full understanding of the decision-making processes
elicited by a visual oddball task.
In the remainder of this chapter we summarize the main findings of research on devel-
opmental aspects of time-frequency activity, which highlights the importance of meth-
odological choices for interpreting developmental trends. We then discuss advantages
328 STEPHEN M. MALONE, JEREMY HARPER, and WILLIAM G. IACONO
This section reviews findings from selected studies that highlight the range of develop-
mental insights contributed by time-frequency research.
1 ITPC is also known as intertrial phase coherence or clustering and phase-locking factor. Because
time-frequency activity is often expressed as a complex number, it has both magnitude and phase. ITPC
is a measure of the degree to which the phase of the complex-valued signal at a given time-frequency bin
is consistent across trials.
330 STEPHEN M. MALONE, JEREMY HARPER, and WILLIAM G. IACONO
y
why this would be true because log ( y ) − log ( x ) = log . (See chapter 18 in Cohen,
x
2014, for a treatment of the issue of baseline correction in time-frequency analysis).
Several of the studies using baseline normalization also examined time-frequency
power in the pre-stimulus period (without baseline correction), which serves to help
reconcile the divergent findings. All report an inverse association between partici-
pant age and raw theta and delta power, which contrasts with the positive association
with post-stimulus power modulation but mirrors the inverse association between age
and power commonly reported in studies using raw baseline correction (e.g., Mathes
et al., 2016).
In summary, the phasic brain response to an experimental event (stimulus onset or
incorrect or inappropriate response) becomes increasingly aligned temporally with the
event, which is reflected in the positive association between age and ITPC commonly
observed. Although there are exceptions (Bowers et al., 2018; Crowley et al., 2014),
baseline subtraction most often results in finding overall reductions in power with age,
which is sometimes interpreted to reflect increased processing efficiency. However,
baseline normalization most often results in overall increases with age, which is not as
easily explained in terms of efficiency. In addition, time-frequency power in the pre-
stimulus baseline period also decreases. This suggests that pre-and post-stimulus power
both decrease with development, but at different rates. It may be that with development,
fewer neurons are recruited in response to an experimental event, leading to reduced
power overall, but children and adolescents become better able to maintain an appro-
priate task set, resulting in greater phase consistency in the EEG response and increased
post-stimulus power. Both interpretations are speculative but point to the fact that the
different methods of accounting for baseline activity can be complementary.
(a) T1 T2 T3 T4 T5 T6 T7 T8 T9
(b) T1 T2 T3 T4 T5 T6 T7 T8 T9
ages rather than at all of them. This design is illustrated in Figure 14.3, which depicts two
possible data collection plans for a hypothetical study of adolescents between the ages of
14 and 20 or older. (What we present here is purely hypothetical; see Brandmaier et al.,
2020 for guidelines for planned missingness in latent growth curve designs.) There are
three age cohorts in this example, with all participants assessed annually for a period
of two years. Subjects in the three respective cohorts were 14, 16, or 18 years old at the
initial assessment for the first design plan (Plan A) and 14, 17, and 20 years old at the
initial assessment for the second design plan. Plan A depicts a cohort-sequential de-
sign, in which participants’ mean age at the last assessment for each of the two youngest
age groups is the same as the mean age of participants in the next older age group at
their initial visit. Overlap in ages allows the researcher to compare cohorts. In Plan B,
participants in the three age cohorts are assessed at nonoverlapping ages, effectively
increasing the age range spanned by the sample. In both cases, the missing data are
missing by virtue of the experimental design, which makes missingness ignorable and
yields unbiased parameter estimation. Thus, they allow the researcher to characterize
developmental change across the entire age range of the sample despite the fact that each
subject is only observed three times.
14.5.2 Pubertal Stage
The ages of participants spanned middle childhood to late adolescence or
adulthood in several studies, which creates a potential confound between age and
pubertal status, and thus neuroendocrinological development. A chronological age
of 13, say, may not be equivalent for boys and girls with respect to the phenomenon
under study. Sample sizes may not be large enough for a careful analysis of effects
of puberty, but it should not be overlooked. Indeed, we found that incorporating
a measure of pubertal status in growth curve models eliminated apparent sex
differences in rates of change in time-frequency power in adolescence. In addition,
accounting for individual differences in pubertal status in this manner resulted
in superior correspondence between the observed data and predicted trajectories
of change in the amplitude of one particular time-frequency component (Malone
et al., 2021).
in mean levels of the outcome measure. Thus, these models are well suited for lon-
gitudinal analysis.2 An additional advantage of an LME framework over traditional
ANOVA models is that it uses all available data as long as missingness is ignorable.
Of course, one is not limited to linear models: a nonlinear or nonparametric model
may be appropriate for capturing growth that follows a well-specified form, such as ex-
ponential and logistic growth models. Models with a latent age or time basis can be an
attractive way to model nonlinearity in mean levels of an outcome; we illustrate the use
of such a model in the next section. Kernel regression is a form of locally weighted re-
gression that has been used to model change in time-frequency features (Chorlian et al.,
2015). Nonparametric regression can be useful when a simple parametric model is not ap-
propriate or as a preliminary step in choosing an appropriate model (Helwig, 2019). Thus,
a wide variety of models are available in regression analysis, whether the data are from a
longitudinal design or a cross-sectional one, which should prompt the researcher to think
about the most appropriate model of change for the data at hand and the question asked.
2
ANOVA models can now incorporate a variety of residual covariance structures, which overcomes
some of their limitations (Liu et al., 2012).
336 STEPHEN M. MALONE, JEREMY HARPER, and WILLIAM G. IACONO
trends in the data. The change point, or knot, in each regression was identified empir-
ically and ranged from 16 to 19.5. Developmental trajectories in time-frequency power
were characterized by an initial decrease in (baseline-corrected) raw power, followed by
a flatter trajectory in later adolescence and early adulthood. Although the rate of change
was smaller in magnitude for the second phase of these models, it was still significant
(and negative). Data for one representative component are plotted in Figure 14.4. Panel
(a) depicts the component structure across waves. This recapitulates the first row of
Figure 14.2 and indicates that the component structure became less dispersed in time-
frequency space with development. Panel (b) consists of violin plots representing the
distribution of scores across waves, separately for male and female participants. This
illustrates the overall trend for this component, its similarity for participants of both
Age 11
(a) (b)
6
6
4
Males
4
2
Component Score
Age 14 0
6
7.5
4
Females
2 5.0
6 11 14 17 20 24
4 Assessment Age
2
SEX Male Female
0 200 400 600 800 1000
Age 20
(c)
6 3
4
Component Score
2
2
0 200 400 600 800 1000
Age 24
6
1
4
2
Figure 14.4 Component scores from a PCA analysis of time-frequency energy plotted against
assessment age. Subjects are the same as in Figure 14.1. Data are for Component 1 in Figure 14.2,
chosen to illustrate key results. Panel A depicts the component weights. This is identical to the first
row of time-frequency weights in Figure 14.2. Panel B consists of violin plots of the distribution
of Component 1 scores across assessment waves, separately for the two sexes. A nonparametric
smoother (loess) represents the average developmental trend. Panel C depicts trajectories
implied by a piecewise linear model fit to these scores together with the observed mean scores at
each wave and 95% confidence intervals around them.
338 STEPHEN M. MALONE, JEREMY HARPER, and WILLIAM G. IACONO
sexes and a striking decrease in variance with development. Finally, Panel (c) depicts the
mean trajectory implied by our piecewise linear regression model along with observed
means and confidence intervals. The (nonlinear) model-implied trajectory accounts for
the observed data well.
That the magnitude of time-frequency activity could be modeled by piecewise linear
trajectories suggests that there were distinct developmental phases in this sample with
respect to the brain response to target stimuli in the oddball paradigm used. These
results also illustrate one of the advantages of LME models: only 485 participants had
data from all five assessments, yet all data were used in fitting models, including age-11
data from the approximately 200 ES twins, who were assessed at subsequent visits using
a different EEG recording system (described next) and as a result contributed only a
single observation in this investigation.
3
https://www.github.com/sjburwell/eeg_commander
DEVELOPMENTAL RESEARCH ON TIME-FREQUENCY 339
4 If there are T time points, T –2 parameters can be estimated; two parameters must be fixed to
identify these models. With only three time points, there is a single free parameter, but the number of
parameters estimated can increase substantially with more time points.
340 STEPHEN M. MALONE, JEREMY HARPER, and WILLIAM G. IACONO
wave, rather than assuming they are exactly the same, as SEM software often requires.
Because our interest is in individual trajectories we used a cluster-robust sandwich es-
timator via the COMPLEX samples option to derive appropriate standard errors given
that this sample comprises twins. In light of evidence of sex differences in level or shape
of trajectories of time-frequency activity (Chorlian et al., 2015; Malone et al., 2021), we
allowed for (estimated) sex differences in intercept and slope. Only the sex effect on the
intercept of theta ITPC was significant, p =.005, which is evident in the difference in
overall level between males and females in panel (b) of Figure 14.5. The rate of change
in all four measures could be considered roughly equivalent for males and females (p-
values ≥ .110). Loadings for the age-11 and age-17 data on the slope latent variable were
fixed to the average age at the age-11 and age-17 assessments, respectively, in order to
identify the model, whereas the slope loading on age-14 scores (or, strictly speaking,
the offset in years from the average age at this assessment) was estimated from the data.
This coefficient was positive for all four time-frequency measures, indicating greater
change by the age-14 assessment than assumed by a strictly linear age basis. Wald z
tests of the slope coefficient were significant, with the exception of theta total power,
indicating a meaningful departure from linearity. Intercept and slope were significantly
inversely correlated for both ITPC measures, indicating that those with higher levels
of ITPC at age 11 tended to show a slower rate of change with development. (See Table
14.1 for a close approximation to these correlations from the bivariate models examined
Tabled values are correlations between pairs of latent variables and their associated p-values for each
model: ITPC and power for delta and theta, respectively, and theta and delta power or ITPC, respectively.
Intercept1 and Slope1 refer to the first measure in each heading, whereas Intercept2 and Slope2 refer to the
second. For instance, in the Delta ITPC-Power model, Intercept1 is the delta ITPC intercept and Intercept2 is
for delta power. Age-specific residuals are the correlations between residual variance in the two measures,
independent of growth trajectories. Age-11 residuals for three of the four measures were constrained to 0 to
avoid negative estimates. Correlations involving these measures were therefore indeterminate.
342 STEPHEN M. MALONE, JEREMY HARPER, and WILLIAM G. IACONO
next.) These analyses thus indicate similarities and differences in trajectories of delta
and theta power and ITPC, as well as illustrate the potential of a latent time basis model
in representing a nonlinear trend, which in the present situation captured nonlinear
change in a particularly parsimonious model.
results of hierarchical clustering of the correlations. Correlations among theta and delta
measures segregated into separate respective clusters, with the former in the upper
right-hand corner of the correlation matrix and the latter in the lower left.5 Within the
theta cluster, measures of total power were particularly closely related, with correlations
ranging from .47 to .58, whereas ITPC at 11 and 14 formed a subcluster. By contrast,
delta total power and ITPC at 14 and total power at 17 tended to cluster together, and the
measures that formed a subcluster were the two age-11 measures. The largest correlation
coefficient was .58, between age-14 and age-17 theta power (between waves) as well as
between age-14 delta power and ITPC (within a wave). Age-17 ITPC for both theta and
delta were relatively independent of the other measures, with a median correlation of
.06 and .10, respectively. This pattern suggests a greater degree of shared variance among
delta power and ITPC than among theta power and ITPC, and relative independence
of ITPC measures at age 17. In addition, a shift in the pattern of associations around the
age-14 assessment is suggested.
To determine the degree to which change in one measure was associated with change
in another, we fit a bivariate parallel process model. The parallel process model allows
for the latent influences on each measure to be related to the latent influences on the
other measure. Specifically, the model includes correlations (covariances) between the
two intercepts and the two slopes as well as a regression of the slope of change in one
measure on the intercept of the other measure. The latter permits an assessment of the
degree to which the level of time-frequency ITPC, say, affects the rate of change in time-
frequency power. This model is illustrated in Figure 14.7.
We fit four such models, one for each pairing of measures. Modeling the trajectory
in each measure recapitulates the univariate analyses just described. This is evident in
Table 14.1, which presents the most relevant parameter estimates from these models,
in that the first two rows present correlations between the latent intercept and slope
characterizing growth in a given measure. These were significant only for delta and theta
ITPC. The parameters of greatest interest here are those that characterize relationships
between the two measures: the correlations between latent intercepts (ρITPIITPC in
Figure 14.7) and slopes (ρSTPSITPC) and the effect of one latent intercept on slope of the
other measure (γITPSITPC and γIITPCSTP). The former, cross-domain correlations appear
in the shaded rows in Table 14.1. Only the correlations between intercepts and slopes
for delta were significant, with results indicating that the greater the initial level of
delta ITPC power, the greater the initial level of delta ITPC and the greater the rate of
change in ITPC, the greater the rate of change in power. In addition, mean initial level
in delta power was inversely associated with the rate of change in ITPC (p =.008). This
would lead delta power and ITPC become somewhat more distinct with development,
consistent with the pattern of correlations in Figure 14.6. It is also consistent with the
5 Although PCA yields orthogonal components and varimax is an orthogonal rotation method,
any type of rotation of the PCA solution results in lack of orthogonality of scores, loadings or both,
depending on the form of normalization used (Jolliffe, 2002). Normalizing eigenvectors to unit length,
as we did, sacrifices independence of component scores while maintaining orthogonality of the loadings.
344 STEPHEN M. MALONE, JEREMY HARPER, and WILLIAM G. IACONO
Figure 14.7 Partial path diagram for a parallel process bivariate growth model. Latent variables
are represented by circles, whereas manifest variables are represented by squares. TP is time-
frequency total power, ITPC is intertrial phase coherence. Each is measured at three different
assessment waves in a longitudinal design. The left-hand panel focuses on the growth model for
total power (top) and ITPC (bottom). The mean level of each is modeled as a linear function of
a latent intercept (ITP and IITPC) and latent slope (STP and SITPC). The loadings of each observed
variable on its latent intercept are fixed at 1. The Greek letter a is used to represent the loadings on
the latent slope. In the latent age basis model used in the example described in the text, a0 is fixed
at 0 and a2 at 1, but a1 is estimated from the data. The coefficient corresponding to a1 thus gives the
proportion of total growth between 11 and 17 observed at age 14. At each assessment wave, the two
regression residuals covary (are correlated), which allows for age-specific influences common
to power and ITPC. These are drawn for ages 11 and 17 but not for age 14 so as to avoid an un-
necessarily crowded diagram. The right-hand panel “explodes” the center of the left-hand figure
to emphasize the relationships between the two latent growth processes. These consist of two
correlations, between the two latent intercepts and the two latent slopes, and two regression paths
predicting mean change in TP from initial level of ITPC and vice versa. The triangle containing
the number 1 is used to represent the mean level of intercept and slope for TP and ITPC, respect-
ively. In reality, these are intercepts, rather than means, because the model includes regressions of
the latent variables on sex (not shown). In both panels, unlabeled circular double-headed arrows
indicate residual variances.
correlations between the two intercepts or the two slopes were significant. All p-values
were greater than .360, although the ersatz effect size for the regression of delta ITPC on
rate of change in theta ITPC was somewhat larger than that of the significant effect in
the delta model. In addition, the correlations between paired intercepts and slopes were
not significant. Thus, the two processes in each model developed independently.
In broad terms, aspects of these results were anticipated by the pattern of correlations
in Figure 14.6 in that correlations between theta and delta measures were relatively
small in magnitude and total power and ITPC correlations tended to segregate for theta
compared to delta. However, most of the covariances between wave-specific residuals
were significant and positive for both the age-14 and age-17 assessment waves. That is,
in the absence of significant cross-measure regression effects for the most part, most of
the covariance between measures was due to age-specific influences. This was especially
true of the age-14 wave (and to a lesser extent the age-17 wave). Individual differences
in pubertal status may induce correlations between time-frequency measures that are
independent of developmental trajectories in these measures through hormonal effects
on brain organization, an inference not readily gleaned from the correlation matrix.
14.7.4 Conclusion
We do not by any means consider these results definitive. For one thing, the parallel pro-
cess model is only one possible approach, and it does not provide a direct assessment of
influences of one measure on change in the other. This requires a more complex model
with more time points (McArdle, 2009). Our results here are intended solely to illus-
trate some of the types of insights that longitudinal-developmental designs provide.
In both investigations, we found that a common set of dimensions of time-frequency
power, whether raw or dB-transformed, accounted well for time-frequency activity in
adolescence and, in the first investigation, into early adulthood. The consistency of these
dimensions across age in both instances was striking. Nevertheless, the particular con-
figuration of these dimensions shifted in subtle ways with age. For instance, congruence
coefficients between successive assessment waves for the component depicted in Figure
14.4 were all 0.95 or greater, indicating very close congruence between one wave and the
next. Nevertheless, the coefficient for the same component between age-11 and age-24
waves was 0.77, indicating much less congruence over the course of adolescence. Thus,
developmental change occurred through a gradual shift in time-frequency space of the
locus of activity.
In our original investigation, we found that piecewise linear regression models
accounted well for mean levels of the data, which suggests distinct phases charac-
terize the development of time-frequency activity in the rotated heads oddball task.
The timing of the transition between phases varied among components, which helps
to illuminate age-related differences in the nature of the brain response. In our second,
more preliminary investigation, our ability to replicate this finding was hamstrung by
the fact that only three data points were available. Nevertheless, we found evidence of
346 STEPHEN M. MALONE, JEREMY HARPER, and WILLIAM G. IACONO
Research conducted to date has yielded several consistent general findings. The ma-
jority of studies have reported consistent associations between age and time-frequency
power and phase-locking (ITPC). The association with ITPC is positive, such that
phase-locking increases with age. As discussed, the nature of the association with power
tends to be negative in studies using raw power or baseline subtraction but positive
in studies examining modulation of power, such as by means of a dB transform, what
Makeig and colleagues have called event-related spectral perturbations (ERSP; Makeig,
1993). These general trends are interpreted to mean that time-frequency responses
to stimuli and events in an experimental setting become more mature in some sense.
However, the consistency of these trends across paradigms presents something of an
interpretive challenge for researchers. If more “efficient” responses are observed in a go/
no-go task among older participants, does this reflect a change in the processes involved
in withholding a response specifically or a developmental change in information pro-
cessing more generally? Comparing age-related trends across a variety of tasks differing
in processing and response demands will be necessary to begin answering this question.
A developmental perspective prompts a shift in the types of questions one asks. For
instance, finding age-related differences in mean levels of a time-frequency phenom-
enon is really the first step of inquiry into the phenomenon (Rutter, 1988). This is in
part because age differences involve other differences among individuals as well that are
correlated with age. Observing age differences also suggests lines of follow-up research.
If younger adolescents differ from older ones on a measure such as time-frequency
power, how is this related to other time-frequency or performance measures (e.g., Liu
et al., 2014; Papenberg et al., 2013)? At what point in development do these differences
disappear? Do age-related differences in time-frequency features related to cognitive
control in one task, such as increased theta power, predict similar age-related differences
in those same features in another task? Failure to find such differences might call into
question whether these features truly reflect cognitive-control processes. It might also
indicate that the developmental course of these time-frequency features varies because
of differences in the cognitive control processes required by tasks that appear similar
on the surface. Control of interference in the attentional field by flanking stimuli in an
Eriksen flanker tasks is likely somewhat different from control of response selection
or inhibition in a go/no-go task. Yet at the same time, the similarities between these
two forms of control suggest that the time-frequency activity elicited by each should be
related and the relationship between them can be traced longitudinally within subjects.
DEVELOPMENTAL RESEARCH ON TIME-FREQUENCY 347
REFERENCES
Basar, E. (1980). EEG-brain dynamics: Relation between EEG and brain evoked potentials.
Amsterdam.
Begleiter, H., Porjesz, B., Bihari, B., & Kissin, B. (1984). Event-related brain potentials in boys at
risk for alcoholism. Science, 225, 1493–1496.
Bender, S., Weisbrod, M., Bornfleth, H., Resch, F., & Oelkers-Ax, R. (2005). How do children
prepare to react? Imaging maturation of motor preparation and stimulus anticipation by late
contingent negative variation. NeuroImage, 27, 737–752.
Bernat, E. M., Malone, S. M., Williams, W. J., Patrick, C. J., & Iacono, W. G. (2007).
Decomposing delta, theta, and alpha time–frequency ERP activity from a visual oddball
task using PCA. International Journal of Psychophysiology, 64, 62–74.
Bernat, E. M., Williams, W. J., & Gehring, W. J. (2005). Decomposing ERP time-frequency en-
ergy using PCA. Clinical Neurophysiology, 116, 1314–1334.
348 STEPHEN M. MALONE, JEREMY HARPER, and WILLIAM G. IACONO
Bowers, M. E., Buzzell, G. A., Bernat, E. M., Fox, N. A., & Barker, T. V. (2018). Time-frequency
approaches to investigating changes in feedback processing during childhood and adoles-
cence. Psychophysiology, 55, e13208.
Brandmaier, A. M., Ghisletta, P., & Oertzen, T. V. (2020). Optimal planned missing data de-
sign for linear latent growth curve models. Behavioral Research Methods, 52, 1445–1458.
doi:10.3758/s13428-019-01325-y
Burwell, S. J., Malone, S. M., Bernat, E. M., & Iacono, W. G. (2014). Does electroencephalogram
phase variability account for reduced P3 brain potential in externalizing disorders? Clinical
Neurophysiology, 125, 2007–2015.
Carlson, S. R. & Iacono, W. G. (2008). Deviant P300 amplitude development in males is
associated with paternal externalizing psychopathology. Journal of Abnormal Psychology,
117, 910–923. https://doi.org/10.1037/a0013443
Casey, B. J., Jones, R. M., & Hare, T. A. (2008). The adolescent brain. Annals of the New York
Academy of Sciences, 1124, 111–126. https://doi.org/10.1196/annals.1440.010
Chorlian, D. B., Rangaswamy, M., Manz, N., Kamarajan, C., Pandey, A. K., Edenberg,
H., . . . Porjesz, B. (2015). Gender modulates the development of theta event related
oscillations in adolescents and young adults. Behavioral Brain Research, 292, 342–352.
doi:10.1016/j.bbr.2015.06.020
Chorlian, D. B., Rangaswamy, M., Manz, N., Meyers, J. L., Kang, S. J., Kamarajan, C., . . . Porjesz,
B. (2017). Genetic correlates of the development of theta event related oscillations in
adolescents and young adults. International Journal of Psychophysiology, 115, 24–39.
doi:10.1016/j.ijpsycho.2016.11.007
Cohen, M. X. (2014). Analyzing neural time series data: theory and practice: MIT press.
Crowley, M. J., van Noordt, S. J. R., Wu, J., Hommer, R. E., South, M., Fearon, R. M. P., & Mayes,
L. C. (2014). Reward feedback processing in children and adolescents: medial frontal theta
oscillations. Brain and Cognition, 89, 79–89.
Curran, P. J., & Bauer, D. J. (2011). The disaggregation of within-person and between-person
effects in longitudinal models of change. Annual Review of Psychology, 62, 583–619.
doi:10.1146/annurev.psych.093008.100356
D’Souza, D. & Karmiloff-Smith, A. (2016). Why a developmental perspective is critical for
understanding human cognition. Behavioral and Brain Sciences, 39, e122. doi:10.1017/
S0140525X15001569
DuPuis, D., Ram, N., Willner, C. J., Karalunas, S., Segalowitz, S. J., & Gatzke-Kopp, L. M. (2015).
Implications of ongoing neural development for the measurement of the error-related nega-
tivity in childhood. Developmental Science, 18, 452–468.
Ehlers, J., Struber, D., & Başar-Eroğlu, C. (2016). Multistable perception in children: Prefrontal
delta oscillations in the developing brain. International Journal of Psychophysiology, 103, 129–
134. doi:10.1016/j.ijpsycho.2015.02.013
Ferrer, E., Salthouse, T. A., Stewart, W. F., & Schwartz, B. S. (2004). Modeling age and retest
processes in longitudinal studies of cognitive abilities. Psychology and Aging, 19, 243–259.
Grimm, K. J., Ram, N., & Hamagami, F. (2011). Nonlinear growth curves in developmental re-
search. Child Development, 82, 1357–1371. doi:10.1111/j.1467-8624.2011.01630.x
Hallquist, M. N. & Wiley, J. F. (2018). Mplus Automation: An R package for facilitating
large-scale latent variable analyses in Mplus. Structural Equation Modeling, 48(2), 1–18.
doi:10.1080/10705511.2017.1402334
DEVELOPMENTAL RESEARCH ON TIME-FREQUENCY 349
Harper, J., Malone, S. M., & Iacono, W. G. (2019). Target-related parietal P3 and medial frontal
theta index the genetic risk for problematic substance use. Psychophysiology, 56(8), e13383.
doi:10.1111/psyp.13383
Helwig, N. E. (2019). Robust nonparametric tests of general linear model coefficients: A com-
parison of permutation methods and test statistics. NeuroImage, 201, 116030. doi:10.1016/
j.neuroimage.2019.116030
Hill, S. Y. & Shen, S. (2002). Neurodevelopmental patterns of visual P3b in association with
familial risk for alcohol dependence and childhood diagnosis. Biological Psychiatry, 51,
621–631.
Hwang, K., Ghuman, A. S., Manoach, D. S., Jones, S. R., & Luna, B. (2016). Frontal pre-
paratory neural oscillations associated with cognitive control: A developmental study
comparing young adults and adolescents. NeuroImage, 136, 139– 148. doi:10.1016/
j.neuroimage.2016.05.017
Iacono, W. G., Carlson, S. R., Taylor, J., Elkins, I. J., & McGue, M. (1999). Behavioral disinhib-
ition and the development of substance use disorders: Findings from the Minnesota Twin
Family Study. Development and Psychopathology, 11, 869–900.
Iacono, W. G. & Malone, S. M. (2011). Developmental endophenotypes: Indexing genetic
risk for substance abuse with the P300 brain event‐related potential. Child Development
Perspectives, 5, 239–247.
Iacono, W. G., Malone, S. M., & Vrieze, S. I. (2017). Endophenotype best practices. International
Journal of Psychophysiology, 111, 115–144. doi:10.1016/j.ijpsycho.2016.07.516
Jolliffe, I. T. (2002). Principal component analysis. Springer.
Keyes, M. A., Malone, S. M., Elkins, I. J., Legrand, L. N., McGue, M., & Iacono, W. G. (2009).
The enrichment study of the Minnesota twin family study: Increasing the yield of twin
families at high risk for externalizing psychopathology. Twin Research and Human Genetics,
12, 489.
Kolev, V. & Yordanova, J. (1997). Analysis of phase-locking is informative for studying event-
related EEG activity. Biological Cybernetics, 76, 229–235. doi:10.1007/s004220050335
Little, R. J. & Rubin, D. B. (2002). Statistical analysis with missing data (2nd ed.). Wiley.
Liu, S., Rovine, M. J., & Molenaar, P. (2012). Selecting a linear mixed model for longitudinal
data: Repeated measures analysis of variance, covariance pattern model, and growth curve
approaches. Psychological Methods, 17, 15.
Liu, Z.-X., Woltering, S., & Lewis, M. D. (2014). Developmental change in EEG theta activity in
the medial prefrontal cortex during response control. NeuroImage, 85, 873–887.
Luciana, M., & Collins, P. F. (2012). Incentive motivation, cognitive control, and the adolescent
brain: Is it time for a paradigm shift? Child Development Perspectives, 6, 392–399. doi:10.1111/
j.1750-8606.2012.00252.x
Makeig, S. (1993). Auditory event-related dynamics of the EEG spectrum and effects of ex-
posure to tones. Electroencephalography and Clinical Neurophysiology, 86, 283–293.
doi:10.1016/0013-4694(93)90110-h
Malone, S. M., Harper, J., Bernat, E. M., & Iacono, W. G. (2021). Stability and change in the de-
velopment of time-frequency activity in adolescence and early adulthood. Preprint [online]
https://doi.org/10.31219/osf.io/8kdh
Mathes, B., Khalaidovski, K., Wienke, A. S., Schmiedt-Fehr, C., & Basar-Eroglu, C. (2016).
Maturation of the P3 and concurrent oscillatory processes during adolescence. Clinical
Neurophysiology, 127, 2599–2609.
350 STEPHEN M. MALONE, JEREMY HARPER, and WILLIAM G. IACONO
McArdle, J. J. (2009). Latent variable modeling of differences and changes with longitudinal
data. Annual Review of Psychology, 60, 577–605.
Mišić, B., Mills, T., Vakorin, V. A., Taylor, M. J., & McIntosh, A. R. (2014). Developmental trajec-
tory of face processing revealed by integrative dynamics. Journal of Cognitive Neuroscience,
26, 2416–2430.
Muthén, L. K., & Muthén, B. O. (1998-2012). Mplus User’s Guide (7th ed.). Muthén & Muthén.
Papenberg, G., Hämmerer, D., Müller, V., Lindenberger, U., & Li, S.-C. (2013). Lower theta
inter-trial phase coherence during performance monitoring is related to higher reaction
time variability: a lifespan study. NeuroImage, 83, 912–920.
R Development Team. (2019). R: A Language and Environment for Statistical Computing. R
Foundation for Statistical Computing. Retrieved from https://www.R-project.org/
Rutter, M. (1988). Epidemiological approaches to developmental psychopathology. Archives of
General Psychiatry, 45, 486–495. doi:10.1001/archpsyc.1988.01800290106013
Sander, M. C., Werkle-Bergner, M., & Lindenberger, U. (2012). Amplitude modulations
and inter- trial phase stability of alpha- oscillations differentially reflect working
memory constraints across the lifespan. NeuroImage, 59, 646– 654. doi:10.1016/
j.neuroimage.2011.06.092
Schneider, J. M., Abel, A. D., Ogiela, D. A., McCord, C., & Maguire, M. J. (2018). Developmental
differences in the neural oscillations underlying auditory sentence processing in children
and adults. Brain and Language, 186, 17-25.
Schneider, J. M., Abel, A. D., Ogiela, D. A., Middleton, A. E., & Maguire, M. J. (2016).
Developmental differences in beta and theta power during sentence processing.
Developmental Cognitive Neuroscience, 19, 19–-30.
Somerville, L. H. & Casey, B. J. (2010). Developmental neurobiology of cognitive control
and motivational systems. Current Opinion in Neurobiology, 20, 236–241. doi:10.1016/
j.conb.2010.01.006
Stiles, J. & Jernigan, T. L. (2010). The basics of brain development. Neuropsychology Review, 20,
327–348. doi:10.1007/s11065-010-9148-4
Tallon-Baudry, C., Bertrand, O., Delpuech, C., & Pernier, J. (1996). Stimulus specificity
of phase-locked and non-phase-locked 40 Hz visual responses in human. Journal of
Neuroscience, 16, 4240–4249.
Tucker, L. R. (1951). A method for synthesis of factor analysis studies (Vol. No. PRS-984).
Educational Testing Service.
Uhlhaas, P. J., Roux, F., Singer, W., Haenschel, C., Sireteanu, R., & Rodriguez, E. (2009). The
development of neural synchrony reflects late maturation and restructuring of functional
networks in humans. Proceedings of the National Academy of Sciences of the United States of
America, 106, 9866–9871.
Wienke, A. S., Basar-Eroglu, C., Schmiedt-Fehr, C., & Mathes, B. (2018). Novelty N2-P3a
complex and theta oscillations reflect improving neural coordination within frontal brain
networks during adolescence. Frontiers in Behavioral Neuroscience, 12, 218.
Williams, W. J. (2001). Reduced interference time- frequency distributions: Scaled
decompositions and interpretations. In L. Debnath (Ed.), Wavelet transforms and time-
frequency signal analysis (pp. 381–417). Birkhauser.
Wilson, S., Haroian, K., Iacono, W. G., Krueger, R. F., Lee, J. J., Luciana, M., . . . Vrieze, S. (2019).
Minnesota Center for Twin and Family Research. Twin Research and Human Genetics, 22(6),
746–752. doi:10.1017/thg.2019.107
DEVELOPMENTAL RESEARCH ON TIME-FREQUENCY 351
Yordanova, J., & Kolev, V. (1996). Brain theta response predicts P300 latency in children.
Neuroreport, 8, 277–280. doi:10.1097/00001756-199612200-00055
Yordanova, J. & Kolev, V. (1998). Developmental changes in the theta response system: a single-
sweep analysis. Journal of Psychophysiology, 12, 113–126.
Yordanova, J., & Kolev, V. (2008). Event-related brain oscillations in normal development. In L.
A. Schmidt & S. J. Segalowitz (Eds.), Developmental psychophysiology: Theory, systems, and
methods (pp. 15–68). Cambridge University Press.
CHAPTER 15
15.1 Introduction
when the visual cortex is deprived of sensory input (i.e., eyes closed), yet instantly dis-
appear when the thalamocortical coupling is restored and the thalamus starts to transfer
signals from our eyes to the visual cortex again (i.e., eyes open). This observation had
led researchers to propose that alpha waves represent a so-called cortical idling rhythm
that is inversely related to processing activity (Adrian & Matthews, 1934). In addition to
the posterior areas, activity in the alpha band is also present in the more anterior regions
and involved in motor, cognitive, and emotion-related functions (e.g., Hoptman &
Davidson, 1998: Harmon-Jones & Gable, 2018).
Furthermore, EEG frequencies of 13 Hz and higher (beta) have been proposed to re-
flect synchronized field potentials in more local cortico-cortical circuits (Lubar, 1997).
Beta activity can be registered across the cerebral cortex including the more anterior
regions of the cerebral cortex (Engel & Fries, 2010). The association between beta ac-
tivity and an anterior distribution concurs with the proposed posterior-anterior gra-
dient model in which the natural peak EEG frequency increases from the posterior
perceptual to the anterior action dedicated areas of the cerebral cortex (Rosanova et al.,
2009). In sum, neural rhythms are preserved across the mammalian species, and they
provide a signature of functionally dedicated brain regions and circuits.
The ratio of frequency bands is the relative contribution of neural rhythms to charac-
terize the interdependency between these neural signals that can be recorded from the
scalp. For example, scalp-recorded resting state frontocentral theta activity can at least
in part be explained by input from signals coming from the subcortical system, whereas
beta activity represents endogenous inhibitory activity that is more cortical of origin
(Schutter & van Honk, 2005). Hence, the ratio between theta and beta power arguably
reflects the balance between bottom-up excitatory subcortical signals and the cortical
signals associated with top-down regulation and attentional control (Schutte et al.,
2017). In this conceptual framework, a high theta-beta power ratio indicates a dominant
bottom-up subcortical input over top-down cortical regulation, whereas a low theta-
beta power ratio implies a dominant top-down cortical regulation over the bottom-up
subcortical system. The general idea behind the theta-beta power ratio is that the ratio
captures the interaction of more motivation-driven processes (theta) on the one hand,
and attention-and cognitive-related processes (beta) on the other.
20.0
40.0
0
1 sec 0 10 20 30 40 50 Hz
beta: 13–30 Hz μV2
μV 1.2
–40.0
–20.0
20.0
40.0
0
1 sec 0 10 20 30 40 50 Hz
Figure 15.1 Example of band-pass filtered scalp recorded theta and beta oscillations and spec-
tral power distributions.
The replicability of these findings together with results showing that the pharma-
cological agent methylphenidate (Ritalin) improves ADHD symptoms paralleled
by a lowering of theta activity and decrease in the theta-beta power ratio (Arns et al.,
2018) has led to the suggestion that the resting state theta-beta power ratio may be a
biomarker for ADHD and can be used in diagnostics. In one of the first studies, three
experiments were reported that aimed to replicate previous findings, examine criterion-
related validity of EEG on inattention, and determine test-retest reliability of the EEG
(Monastra et al., 2001).
In the first experiment, earlier findings of significantly higher theta-beta power
ratios in patients aged between 6–20 years with inattentive and hyperactive-combined
ADHD as compared to controls, were replicated. The second experiment showed that
the theta-beta power ratio derived attentional index could reliably differentiate between
ADHD and non-ADHD participants (sensitivity: 90%, specificity: 94%). In the third
experiment a high test-retest correlation coefficient (0.96) of the theta-beta power ratio
was found in a sample of fifty-five volunteers that underwent an EEG session on two
occasions thirty days apart. Comparable sensitivity (87%) and specificity (94%) scores
356 DENNIS J. L. G. SCHUTTER and J. LEON KENEMANS
in identifying ADHD within a clinical sample (n =159) were reported by another multi-
center validation study (Snyder et al., 2008). However, the theta-beta power ratio could
not reliably differentiate between ADHD and comorbidities or alternative diagnoses.
Based on these findings the authors advised the use of EEG as a complementary measure
in the clinical evaluation of ADHD (Snyder et al., 2008).
This conclusion follows up results from the first meta-analysis by Snyder & Hall
(2006), who investigated the presence of specific EEG traits in ADHD and evaluated
nine high quality studies (n =1498) in which the primary analysis of interest concerned
a group comparison between ADHD patient’s diagnosis according to the criteria
of the Diagnostic and Statistical Manual of Mental Disorders and controls on resting
state theta and beta activity. Standardized mean effect size (Glass’ delta) and 95% confi-
dence intervals (CI) were calculated. Results showed significantly higher theta (pooled
effect size: 1.31; 95% CI: 1.14–1.48) and lower beta activity (pooled effect size: −0.51; 95%
CI: −0.65 to −0.35) in patients compared to controls.
Consistent with these observations a strong significant effect was observed for an
increase of theta-beta power ratio in ADHD patients versus controls (pooled effect
size: 3.08; 95% CI: 2.90 ̶ 3.26). A second fixed effects model meta-analysis published
in 2013 reported a notably lower overall effect size, Hedges’ d: 0.75; 95% CI: 0.57 ̶ 0.93,
favoring increased theta-beta power ratio in children and adolescents with ADHD as
compared to controls (Arns et al., 2013).1 Furthermore, considerable heterogeneity in
effect sizes across studies was observed, indicating that the variance of the effect size
across the included studies was greater than to be expected from sampling error. Even
though the effect size was still large, subsequent analyses showed no strong support for
the applicability of the theta-beta power ratio in diagnostics (Arns et al., 2013).
Despite the inconclusive scientific evidence, in July 2013, the U.S. Food and
Drug Administration (FDA) approved the use of the Neuropsychiatric EEG-Based
Assessment Aid Health for the assessment of ADHD based on the theta-beta power
ratio (FDA, 2013). To further examine this, we updated the meta-analysis by Arns
et al. (2013) and included additional studies that reported group differences between
theta-beta power ratios of children and adolescent with and without an ADHD diag-
nosis. In total, eighteen studies that were published in peer-reviewed English scientific
journals including 2425 participants (1600 patients and 825 controls) were analysed.
The standardized effect size (Hedges’ delta) and variance were calculated for each study
(Hedges & Olkin, 1985).
Based on the assumption that, in addition to sampling error, effect sizes contain a
true random component, a random effects model was performed. Results showed an
estimated mean effect size of 0.92 (95% CI: 0.40–1.44) (Figure 15.2).
1
This meta-analysis included nine studies and used different selection criteria compared to the earlier
meta-analysis. As a result, only two of the nine studies used in the original meta-analysis by Snyder and
colleagues (2008) were considered eligible. In addition, this study-based effect size calculation used the
pooled standard deviation of the ADHD and control groups instead of exclusively using the SD of the
control group in computing the Glass’ delta.
THETA-BETA POWER RATIO 357
The test for study heterogeneity was not significant, Qtotal =12.95, p =0.74. Rosenberg’s
fail-safe number analysis indicates that 2415 studies with null findings (p< 0.05) are
required to change the significant meta-analytic result into a non-significant finding.
In addition, Orwin’s method shows that 68 studies are needed to reduce the mean effect
size of 0.92 to a small effect size of 0.2. As neither method accounts for sample size or
variance of the studies, these fail-safe number results should be interpreted with caution.
While the mean effect size observed in the current meta-analysis is larger than the effect
size reported in 2013, the overlapping 95% CIs show that the difference between the
effect sizes is not statistically significant at the p=0.05 level. The wide 95% CI interval
indicates considerable uncertainty in terms of the true value of the effect size and the use
of the theta-beta power ratio for diagnosing ADHD in children and adolescents remains
unclear (Arns et al., 2013). Several authors have proposed that adding cognitive, emotive
and genetic information to the theta-beta power ratio may have substantial diagnostic
as well as prognostic value (Williams et al., 2010).
It is suggested that determining theta activity relative to the individual alpha peak-
frequency may contribute to the diagnostic value of the resting state theta-beta power
ratio (Saad et al., 2018). This suggestion adds to the idea of multiple theta rhythms and
that the dependency on conventional fixed frequency bands may be problematic in the
classification of ADHD as well as for gaining a deeper understanding on the relation be-
tween EEG ratios and their functional meaning.
Determining the so- called transition- frequency reflects the true physiological
boundary between theta and alpha activity, and may present a physiological plausible
358 DENNIS J. L. G. SCHUTTER and J. LEON KENEMANS
alternative as this frequency marks the intersection of resting state with active EEG
(Saad et al., 2018: Klimesch, 1999). Perhaps the individualized peak alpha frequency may
also be useful to indicate the transition from alpha to beta.
Personalized approaches like these may be able to control for individual differences in
EEG make-up unrelated to ADHD symptoms and improve the specificity (true negative
rate) and sensitivity (true positive rate) of the theta-beta power ratio in diagnostics.
Even although the concept of resting state theta-beta power ratio as a biomarker for
diagnostic purposes is promising, further research is needed to further develop this
concept. In addition, higher resting state theta-beta power ratio in ADHD children has
been suggested to be indicative for treatment outcome (Loo et al., 1999: Clarke et al.,
2002b). Support for this idea comes from results showing that good as compared to poor
responders to methylphenidate have higher resting state theta-beta power ratio prior to
treatment (Clarke et al., 2002a). This fits the observation that stimulant medication can
increase beta activity in children with ADHD, particularly in frontal cortical regions, in
addition to improving symptoms related to attentional focus and response inhibition
(Loo et al., 2004).
Together, the normalization of dopaminergic activity in the subcortical circuits as
indicated by a reduction in theta activity, and the increase in beta activity, may explain
the lowering of the power ratio in response to methylphenidate. Indeed, oscillatory beta
activity is in turn linked to dopamine levels at sites of cortical input to subcortical areas
including the basal ganglia (Jenkinson & Brown, 2011). While the resting state theta-
beta power ratio may be a positive response predictor, a multi-center, prospective open-
label trial found that lower frontal alpha (9 Hz) peak frequency, but not the resting state
theta-beta ratio in male adolescent ADHD patients, is a predictor for non-response to
methylphenidate (Arns et al., 2018). This negative predictor response was found to be
independent from age, medication dosage, and baseline symptom severity. The lower
individual anterior alpha peak is suggestive of a maturational lag in the structural de-
velopment of the frontal cortex (Whitford et al., 2007). Together these findings stress
the importance of considering both functional (e.g., theta-beta power ratio) and struc-
tural aspects (e.g., alpha peak) of brain organization for understanding the efficacy of
treatments. More recent work points towards the direction that EEG measures may be
more useful biomarkers of ADHD outcome or treatment response rather than diagnosis
(Sari Gokten et al., 2019: Loo et al., 2018). Finally, the different findings leave open the
possibility that ADHD symptoms can have several neuro-etiological origins and should
be evaluated within an environmental context.
In addition to research on the clinical features and applicability of the theta-beta power ratio
in ADHD, basic neuroscientific research has also made considerable progress in identifying
THETA-BETA POWER RATIO 359
cognitive and affective processes associated with the theta-beta power ratio. This line of re-
search provides insights not only into the specific symptomatology of ADHD, but also in
understanding the neurocognitive processes within the general population.
Experimental human brain research has shown that the prefrontal cortex (PFC) plays
a prominent role in the regulation of attention and motivational tendencies, and other
executive control functions (Corbetta et al., 2008).
On the other hand, for theta specifically, based on combined fMRI and
electrocorticography, significant associations have been reported between theta power
and BOLD activity in “intrinsic” resting-state networks, including the default-mode
network (DMN, the “idling network”) and a network involved in episodic and working
memory (Hacker et al., 2017). Furthermore, the CNS stimulant nicotine reduces both
theta power (Knott et al., 2007), as well DMN activity (Hahn et al., 2007). Nicotine-
induced reduction in DMN activity correlates positively with nicotine- induced
improvements in task performance (Hahn et al., 2007). These data support an interpret-
ation that theta power indexes idling and/or hypo-arousal.
In addition, patients responding to the CNS stimulant methylphenidate dem-
onstrate reductions in the theta-beta power ratio (Isiten et al., 2017). Notably, the
methylphenidate-related change was driven by an increase in beta, rather than a reduc-
tion in theta, power. If indeed beta power is more closely related to cerebral cortical
activity, then the reductions in ADHD symptoms may be due to improved executive
control functions. Raclopride positron emission tomography (PET) showed that me-
thylphenidate (dose: 0.8 ± 0.11 mg/kg) potentiates dopaminergic activity by blocking
dopamine transporters causing an increase of synaptically available DA in the human
brain (especially the striatum, Volkow et al., 2001). The findings were replicated at lower
dosages which are more commonly prescribed in daily practice (Gottlieb, 2001).
As methylphenidate also increases noradrenergic activity (Kuczenski & Segal,
2001), catecholaminergic dysfunctions is argued to contribute to elevated theta-beta
power ratios in individuals with ADHD who respond to methylphenidate (Isiten
et al., 2017).
Based on the proposed neural generators of theta and beta oscillations, in conjunc-
tion with the dopaminergic dysfunctions described earlier, a neuroanatomical model
has been developed to explain the motivational, cognitive, and motoric symptoms
of ADHD. The model is a midbrain ventral tegmental area (VTA)-centered frame-
work that includes both the mesolimbic and mesocortical dopaminergic pathway. The
mesolimbic pathway refers to the connection between the VTA and nucleus accumbens
(ventral striatum), implicated in reward motivation and subcortical motor planning
routines. In addition, the VTA extensively projects to the septo-hippocampal complex
and amygdala indicative of a complex subcortical circuit dedicated to instrumental
learning and memory processes (Hefco et al., 2003).
The connection between the VTA and PFC constitutes the backbone of the
mesocortical pathway. In particular, the VTA projections to the medial frontal cortex
and ACC provide a neural basis for monitoring and top-down regulatory functions (Di
Michele et al., 2005).
360 DENNIS J. L. G. SCHUTTER and J. LEON KENEMANS
Next to the septo-hippocampal complex, the ACC is another brain area that operates
in the theta frequency range (Asada et al., 1999). It is suggested that together these
regions make up a circuit dedicated to monitoring and goal-directed motor planning,
whereas the ventral and dorsolateral parts of the FC take part in executive functions
including working memory, attentional control, and top-down regulation.
The functionality of the latter anterior cortical regions is suggested to involve beta
activity. From this viewpoint, the theta-beta power ratio may provide a global index
for cortico-subcortical interactions wherein slow wave (theta) activity represents the
subcortical bottom up and fast wave (beta) activity in cortical top-down processes
(Schutter et al., 2006). In particular, the low tonic dopaminergic activity in the cortico-
subcortical circuits and subsequently elevated theta-beta power ratio could perhaps
provide a neural marker for attention, impulsivity, reward sensitivity and motor activity
in individuals with and without ADHD (Heilman et al., 1991: Robbins, 2000). A more
specific formulation of this idea holds that a reduction in dopamine level is associated
with disinhibition of the ACC causing an increase in spontaneous theta oscillations
and higher responsivity (e.g., the frontal-related negativity (FRN)) to external feedback
(Holroyd & Coles, 2002).
Thus, high theta relative to low beta activity may reflect a state of heightened subcor-
tical drive in conjunction with cortical hypoarousal. Furthermore, the ACC is among
the prime neural generators of theta oscillations and may a take central position between
the subcortical and prefrontal brain areas. This idea is further supported by research
showing that the ACC and the anterior portions of the medial PFC can exert direct
top-down control over the VTA via monosynaptic excitatory projections (Holroyd and
Yeung, 2012: Elston et al., 2019). Increased 4 Hz (theta) oscillatory signaling between the
ACC and VTA has been demonstrated during anticipatory decision making, feedback
processing, and behavioral flexibility (Elston et al., 2019).
Studies on the spectral density of local field potentials in the ACC and VTA show that
while the ACC shows peaks between 3–5 Hz, the VTA emits signals between 3–5 Hz and
7–9 Hz (Elston et al., 2019). The frequency specificity of theta activity across regions
suggests the earlier proposed existence of multiple theta rhythms that each have a dis-
tinct functional role in behavior. Rhythmic excitability allows for temporal windows of
coordinated spike timing to exchange signals between distal regions (Buzsáki & Watson,
2012: Cavanagh et al., 2012). Further evidence for frequency specificity was provided by
a study showing that administering exogenous oscillatory field potentials at 4 Hz (low
theta) increases working memory capacity, while at 7 Hz (high theta) working memory
performance deteriorates (Wolinski et al., 2008). The latter idea also fits the proposed
role of theta activity carrying information encoded in gamma oscillations (>30 Hz) from
the hippocampus to the cerebral cortex in memory formation (Jensen and Lisman, 1996).
Taken together the studies imply the existence of multiple functionally inde-
pendent theta activity related to (1) arousal, (2) motivation, (3) cognitive flexibility, and
(4) memory-related processes.
THETA-BETA POWER RATIO 361
Studying the frequency specific of theta and beta may therefore be particularly inform-
ative about whether a higher theta-beta power ratio, for example, reflects drowsiness as
indicated by a broad increase in the theta EEG spectrum or heightened motivation as
reflected by increased information transfer between the ACC and VTA as reflected by
4 Hz oscillations. As the conventional calculations of theta-beta power ratio are based
on the averaging of power values in the 4–7 Hz (theta) and 13–30 Hz (beta) frequency
range, the study of specific theta, as well as beta oscillatory, components may be relevant
for understanding the meaning of the power ratio within a given neural and behavioral
context.
Several psychological studies have examined the proposed interrelations between theta-
beta power ratio and cognitive-affective processes in healthy volunteers.
On the basis of the proposed relation between the theta-beta power ratio and im-
pulsivity in ADHD, the predictive value of resting state theta-beta power ratio on
risky decision making was investigated in a sample of healthy volunteers (Schutter
& van Honk, 2005). Risky decision making was measured with the Iowa gambling
task. In the Iowa gambling task, players are instructed to try to gain as much money
as possible by drawing selections from a choice of four decks. Two of the decks are
disadvantageous, producing immediate large rewards, but on the long term cause
an overall loss due to even higher punishments. The other two decks are advan-
tageous in the sense that immediate rewards are modest but more consistent and
punishment rate is low. Decisions to choose from the decks become motivated by
the reward and punishment schedules associated with the advantageous and disad-
vantageous decks.
Results showed that higher resting state theta-beta power ratio was associated
with more risky decisions. These findings can be understood in terms of increased
reward sensitivity and/or lower punishment sensitivity. Importantly, participants
did show a cumulative increase in advantageous decision making during the
course of the task. Even though volunteers acquired knowledge about the reward-
punishment contingency, subjects with high theta-beta power ratio were none-
theless more inclined to take risky decisions as compared to subjects with a lower
theta-beta power ratio.
The association between theta-beta power ratio and disadvantageous decision
making was replicated in a follow-up experiment (Massar et al., 2014) which also
explored the specific contributions of reward and punishment sensitivity to the
theta-beta power ratio by testing how well subjects were able to learn associations be-
tween symbols that signal a high probability of winning money and low probability
362 DENNIS J. L. G. SCHUTTER and J. LEON KENEMANS
of winning nothing (reward learning), and symbols that signal a high probability of
losing money and low probability of losing nothing (punishment learning). Results
showed that the association between theta-beta power ratio and associative learning
was driven by the positive correlation between theta oscillations and reward learning.
This finding fits with the proposed role of theta oscillations in the dopaminergic
mesolimbic pathway.
Results of an earlier study demonstrated that the nature of the relation between
theta-beta power ratio and risk-taking involved feedback-related processing in the
ACC (Massar et al., 2012). More specifically, a higher theta-beta power ratio was found
to correlate with reduced feedback processing as reflected in lower FRN amplitudes.
Additionally, this association varied as a function of punishment sensitivity as evaluated
by the behavioral inhibition system questionnaire. The data suggest a bias toward mo-
tivationally driven behavior as a result of suboptimal action monitoring, leading to
top-down executive control being subsequently less involved due to reduced ACC
signaling. This effect becomes more prominent when subjects score higher on punish-
ment sensitivity and arguably already experience higher default levels of uncertainty.
In accordance, a meta-analysis has shown that midfrontal theta activity is linked to
ACC activity during action-outcome uncertainty and cognitive control (Cavanagh &
Shackman, 2015).
Notably, much like children with an anxiety disorder, children with ADHD display
an intolerance of uncertainty (Gramszlo et al., 2018). Indeed, if the ACC is central to
the integration of bottom-up feedback signals originating from midbrain structures
and top-down executive functioning and control, disturbances in the ACC may sub-
serve the lack of coordinated and goal-directed behaviors in ADHD. Furthermore,
increased theta-beta power ratios may signify compromised ACC-related feed-
back processes that contribute to subjective states of uncertainty. Uncertainty is
experienced when there is a lack of subjectively perceived contingency between action
and outcome.
Thorndike’s law of effect dictates that reward and punishment-related feedback
signals are responsible for shaping situation appropriate action-outcome contingencies.
In other words, the structural occurrence of rewards and punishment following a
specific action leads to the formation of stronger action-outcome links (Seth, 2013).
Experimentally, uncertainty can be manipulated by introducing unpredictable changes
in reward-punishment contingencies. Earlier findings indicate that particularly in un-
predictable environments, a situation-dependent balance between the exploitation
of acquired knowledge and exploration of new options is the best behavioral strategy
to maximize profit (Daw et al., 2006). In the event of a change in reward-punishment
schemes, individuals have to relearn the contingencies to adjust decision making ac-
cordingly (Clark et al., 2004: Bechara et al., 1997). The presence of theta activity is
associated with the experience of conflict and reward sensitivity, so it is reasonable to
assume that high theta-beta power ratio is associated with suboptimal reversal learning
THETA-BETA POWER RATIO 363
ratio and executive control in which under certain (less complex?) circumstances
higher ratios can have a positive effect on performance. This implies that during
high cognitive load and motivational value the positive relations turn into negative
correlations. These seemingly paradoxical results highlight the possibility that, de-
pending on the operational and conceptual perspective adopted by researchers and
theorists, high theta activity, and therefore a high theta-beta power ratio, can reflect
cortical hypo-arousal and increased cortical drive, as well as enhanced working-
memory function and cognitive control. This line of argumentation concurs with
the proposed existence of different functions of theta waves. Finally, a recent study
highlighted prestimulus high theta during the task as an indicator of drowsiness,
but also stimulus-elicited theta specifically in high-conflict conditions, supposedly
reflecting enhanced demands for cognitive control (Canales-Johnson et al., 2020; see
also Cooper et al., 2019).
self-reported attentional control in healthy young adults (Putman et al., 2010) has
prompted a series of scientific studies that further investigated the latter association
in the context of stress and anxiety (Putman et al., 2014). High levels of stress and
anxiety have a negative impact on central executive functioning and performance.
For example, cognitive performance anxiety (CPA) is the phenomenon in which anx-
iety/uncertainty about one’s cognitive competence causes lower and even impaired
performance.
The interaction between the theta-beta power ratio and CPA-related stress on per-
formance was examined in healthy participants who performed a stress induction
or control task. In addition to replicating the earlier inverse association between
the theta-beta power ratio and self-reported attentional control, lower levels of the
ratio were found to be protective for the negative effects of CPA. These finding can be
explained by the inverse relation between theta-beta power ratio and attentional con-
trol (Putman et al., 2010).
In a subsequent study, a pictorial dot-probe paradigm was deployed to further
investigate a role for the theta-beta power ratio in attention resource allocation to
emotionally arousing stimuli. Healthy volunteers were instructed to fixate on the
center of a screen and to locate a black dot that appeared either on the left or right
side immediately following the simultaneous presentation of two stimuli in the left
and right periphery of the visual field. Statistical analyses revealed that subjects with
a relatively high theta-beta power ratio directed attention towards mildly arousing
and avoided highly arousing negative pictures. Consequently, subjects with low
theta-beta power ratios displayed more attention towards high arousing negative
material. Even though correlational findings do not allow inferences on the direc-
tionality of statistical associations, the authors interpret these findings as further
support for the idea that an increased theta-beta power ratio reflects an enhanced
ability to regulate the impact of negative emotional information (Angelidis
et al., 2018).
An independent follow- up experiment replicated this finding and also found
that lower self-reported trait attentional control predicted more attention to mild as
compared to high arousing negative stimuli (Van Son et al., 2018). These findings concur
with earlier research that administered a go/no-go task in healthy volunteers. In this
particular study subjects were instructed to respond (go) or inhibit responses (no-go)
to fearful and happy facial expressions (Putman et al., 2010). In line with the ADHD
literature (Metin et al., 2012), a significant correlation was found between response in-
hibition and the theta-beta power ratio (Putman et al., 2010). More specifically, stronger
response inhibition to the fearful relative to the happy facial expressions was predicted
by lower theta-beta power ratios (Putman et al., 2010).
Furthermore, evidence was found for a positive relation between the theta-beta
power ratio and the drive scale of the behavioral activation system (Gray, 1985).
The drive scale is related to the brain’s seeking system, which promotes energized
366 DENNIS J. L. G. SCHUTTER and J. LEON KENEMANS
appetitive and approach-related behaviors that include exploration, foraging and fu-
ture anticipation of rewards (Panksepp, 1998, 2012). Neuroanatomically, the seeking
system corresponds to the mesolimbic dopaminergic pathway and provides input
to the higher cortical areas and promotes reward-driven actions. Within this frame-
work, theta activity may be a correlate of the seeking system that feeds information
into the PFC. This information is then used in attentional control and regulatory
processes, arguably reflected by the beta activity component of the ratio, to shape
context-appropriate behaviors. More recently, temperamental shyness was found to
correlate with higher levels of social anxiety in children with large baseline-to-task
decreases in frontal theta-beta power ratio in the anticipation of having to give a
speech (Poole et al., 2021). These findings were interpreted as increased attentional
control and emotion regulation during the experience of stress in shy children (Poole
et al., 2021). Observations that resting state the theta-beta power ratio is positively
associated with distraction tendency and not re-appraisal strategies (Kobayashi et al.,
2020) concur with the proposed link between theta-beta power ratio and attentional
control (Putman et al., 2014).
Due to the low spatial resolution of scalp-recorded electric field potentials, localizing
their neural generators is challenging. However, interleaving resting state func-
tional magnetic resonance imaging with resting state EEG recording provides a way
to examine the relations between different cortical and subcortical brain regions and
the theta-beta power ratio. Recently, associations between the theta-beta power ratio
and the brain’s executive control network (ECN) were explored in a sample of healthy
volunteers (Van Son et al., 2019b). The ECN consists of the dorsolateral PFC, ACC and
parietal cortex, and governs top-down regulation and attentional control processes
(Corbetta et al., 2008).
Reductions in attentional control and brain activity were investigated during mind
wandering. Mind wandering refers to transient changes of attentional focus from a task
to non-task related mental states and is accompanied by an increase in the theta-beta
power ratio (Braboszcz & Delorme, 2011: Van Son et al., 2019a). As compared to an at-
tentional focus condition, the reduction in ECN functional connectivity was correlated
to an increase in theta-beta power ratio during mind wandering. Even though this
finding does not imply that the ECN is the generator of theta activity, the result-concurs
with the positive relation between mind wandering and ADHD symptomatology
(Seli et al., 2015) as well as with aberrant functional connectivity of the executive con-
trol network in ADHD subjects (McCarthy et al., 2013). The results not only provide
a functional neuro-anatomical substrate for the theta-beta power ratio, but also fur-
ther strengthens construct validity of the ratio as a functional signature of the balance
between bottom-up motivational and top-down cognitive aspects of normal and ab-
normal human behavior. Figure 15.3 depicts a functional neuro-anatomic framework
of the theta-beta power ratio which represents a synthesis of the concepts and empir-
ical work reviewed in this chapter.
THETA-BETA POWER RATIO 367
DLPFC
beta
MPFC
ACC
VTA
theta
SH NA
While resting state EEG is robust and has a high test-retest reliability within a time
frame of one week (Angelidis et al., 2018), longer inter-assessment intervals and meth-
odological factors can nonetheless provide a source of error variance which can nega-
tively affect the reliability of the theta-beta power ratio. Results can among other factors
be influenced by wrong electrode placement, artifact contamination, drowsiness,
changes in medication use, and comparison with suboptimal control groups. The lack of
a standardized EEG protocol could make comparisons between studies and groups less
reliable. So, in addition to differences already mentioned, differences in, for example,
applied reference method and varying epoch and Fast Fourier Transform window
lengths can yield heterogenous results. Also, age has been found to influence the relation
368 DENNIS J. L. G. SCHUTTER and J. LEON KENEMANS
between the theta-beta power ratio and ADHD (Clarke et al., 2001a). One line of evi-
dence comes from work showing that the ratio is able to differentiate between ADHD in
children, but not in adults (Markovska-Simoska & Jordanova, 2017).
It should furthermore be noted that whereas scientific studies are well suited to pre-
dict and demonstrate effects on the group level, specifying these effects to the individual
level is more problematic. Finally, there is a difference between statistical significance
and clinical relevance, as in that a reliable systematic finding is too small to have a mean-
ingful contribution to the diagnosis, prognosis and/or treatment of a disorder. In sum,
the relation between available scientific evidence and clinical usefulness of EEG in
ADHD and other conditions remains a topic of research.
In addition to the theta-beta power ratio, only a limited number of studies have looked
at the functional relevance of other ratios. For example, an increase in parietal delta-
beta power ratio has been positively associated with spontaneous emotion regulation
as shown by lower ratings of discomfort after the offset of negative pictures (Tortella-
Feliu et al., 2014). These results suggest that the posterior delta-beta power ratio reflects
an electrophysiological dispositional feature for faster automatic recovery of unpleasant
emotional responses (Tortella-Feliu et al., 2014). These findings are in agreement
with a study that found a positive associations between parietal delta-beta waves and
self-reported attentional control in a group of healthy volunteers (Morillas-Romero
et al., 2015). In another study both the parietal delta-beta and frontal theta-beta power
ratio were correlated to risky decision making (Schutter & van Honk, 2005). A study
investigating ADHD subtypes in children by analyzing their EEG found that inatten-
tive type of ADHD was associated with increased delta-beta power ratios as compared
to matched controls (Aldemir et al., 2018). Although the underlying mechanism
remains unclear, cognitive impairments in neurological populations go accompanied
by increases in slow wave oscillations, including delta waves, in conjunction with a de-
crease in alpha and beta wave activity (Klass & Brenner, 1995). Perhaps, relative higher
delta power as a result of lower alpha activity is linked to reduced perceptual sensitivity
to external stimuli that results in less attentional resource allocation and in-depth pro-
cessing. Additionally, higher delta relative to alpha oscillations has been shown to suc-
cessfully discriminate between patients with acute ischemic stroke and healthy controls
(Finnigan et al., 2016). Moreover, in a group of twenty patients with ischemic stroke, a
significant inverse association was found between the delta-alpha EEG ratio assessed at
approximately 72 hours poststroke onset and cognitive performance after 3.5 months
(Schleiger et al., 2014). Similar to what was mentioned earlier, the ratio between slow
delta and alpha (and beta) oscillations may be a sign of cortical (dys)function and/
THETA-BETA POWER RATIO 369
Following the first observation of elevated theta-beta power ratios in children with
ADHD, many studies have replicated this finding. In 2013 the FDA formally approved
the use of the theta-beta power ratio as a criterion for the diagnosis of ADHD. However,
recent meta-analyses suggest that the theta-beta power ratio does not reliably differen-
tiate between individuals with and without ADHD. Recently, psychological research in
healthy volunteers has established reliable associations between the theta-beta power
ratio and indices of motivation and executive functioning. Consistent with observations
in individuals with ADHD, higher theta-beta power ratios are associated with reduced
top-down attentional control paralleled by approach-related motivation and seeking
behavior. These associations are conceptualized as a cortico-limbic balance wherein
beta activity stands for cortical top-down executive control and theta activity represents
the bottom up limbic motivational incentives. Together with the high test-retest reli-
ability coefficients (Angelidis et al., 2016), the theta-beta power ratio can be considered
a trait marker for capturing the reciprocal relationship between cognitive control and
motivational tendencies, and may have complementary value in diagnosis and treating
ADHD and/or related disorders.
REFERENCES
Adrian, E. D. & Matthews, B. H. (1934). The Berger rhythm: Potential changes from the oc-
cipital lobes in man. Brain, 57, 355–385.
Aldemir, R., Demirci, E., Per, H., Canpolat, M., Özmen, S., & Tokmakçi, M. (2018).
Investigation of attention deficit hyperactivity disorder (ADHD) sub-types in children via
EEG frequency domain analysis. International Journal of Neuroscience, 128, 349–336.
Angelidis, A., van der Does, W., Schakel, L., & Putman, P. (2016). Frontal EEG theta/beta
ratio as an electrophysiological marker for attentional control and its test-retest reliability.
Biological Psychology, 121, 49–52
2
Strictly speaking, an oscillation refers to rhythmic electrical activity in one specific frequency (e.g., 5
Hz) and not to a range of frequencies (e.g., theta: 4–7 Hz)
370 DENNIS J. L. G. SCHUTTER and J. LEON KENEMANS
Angelidis, A., Hagenaars, M., van Son, D., van der Does, W., & Putman, P. (2018). Do not look
away! Spontaneous frontal EEG theta/beta ratio as a marker for cognitive control over
attention to mild and high threat. Biological Psychology, 135, 8–17.
Arns, M., Conners, C. K., & Kraemer, H. C. (2013). A decade of EEG theta/beta ratio research
in ADHD: A meta-analysis. Journal of Attention Disorders, 17, 374–383.
Arns, M., Vollebregt, M. A., Palmer, D., Spooner, C., Gordon, E., Cohn, M., . . . Buitelaar, J. K.
(2018). Electroencephalographic biomarkers as predictors of methylphenidate response in
attention-deficit/hyperactivity disorder. European Neuropsychopharmacology, 28, 881–891.
Asada, H., Fukuda, Y., Tsunoda, S., Yamaguchi, M., & Tonoike, M. (1999). Frontal midline
theta rhythms reflect alternative activation of prefrontal cortex and anterior cingulate cortex
in humans. Neuroscience Letters, 274, 29–32.
Barry, R. J., Clarke, A. R., McCarthy, R., Selikowitz, M., Rushby, J. A., & Ploskova, E.
(2004). EEG differences in children as a function of resting-state arousal level. Clinical
Neurophysiology, 115, 402–408.
Bechara, A., Damasio, H., Tranel, D., & Damasio, A. R. (1997). Deciding advantageously before
knowing the advantageous strategy. Science, 275, 1293–1295.
Braboszcz, C. & Delorme, A. (2011). Lost in thoughts: Neural markers of low alertness during
mind wandering. Neuroimage, 54, 3040–3047.
Buyck, I. & Wiersema, J. R. (2015). Electroencephalographic activity before and after cog-
nitive effort in children with attention deficit/hyperactivity disorder. Clinical EEG and
Neuroscience, 46, 88–93.
Buzsaki, G. & Watson, B. O. (2012). Brain rhythms and neural syntax: Implications for effi-
cient coding of cognitive content and neuropsychiatric disease. Dialogues in Clinical
Neuroscience, 14, 345–367.
Canales-Johnson, A., Beerendonk, L., Blain, S., Kitaoka, S., Ezquerro-Nassar, A., Nuiten,
S., . . . Bekinschtein, T. A. (2020). Decreased alertness reconfigures cognitive control
networks. Journal of Neuroscience, 40, 7142–7 154.
Cavanagh, J. F., Zambrano-Vazquez, L., & Allen, J. J. (2012). Theta lingua franca: A common
mid-frontal substrate for action monitoring processes. Psychophysiology, 49, 220–238.
Cavanagh, J. F. & Shackman, A. J. (2015). Frontal midline theta reflects anxiety and cognitive
control: Meta-analytic evidence. Journal of Physiology, 109, 3–15.
Clarke, A. R., Barry, R. J., McCarthy, R., & Selikowitz, M. (2001a). EEG-defined subtypes
of children with attention-deficit/hyperactivity disorder. Clinical Neurophysiology, 112,
2098–2105.
Clarke, A. R., Barry, R. J., McCarthy, R., & Selikowitz, M. (2001b). Age and sex effects in the
EEG: Development of the normal child. Clinical Neurophysiology, 112, 806–814.
Clarke, A. R., Barry, R. J., McCarthy, R., Selikowitz, M., & Croft, R. J. (2002a). EEG differences
between good and poor responders to methylphenidate in boys with the inattentive type of
attention-deficit/hyperactivity disorder. Clinical Neurophysiology, 113, 1191–1198.
Clarke, A. R., Barry, R. J., Bond, D., McCarthy, R., & Selikowitz, M. (2002b). Effects of stimu-
lant medications on the EEG of children with attention-deficit/hyperactivity disorder.
Psychopharmacology, 164, 277–284.
Clark, L., Cools, R., & Robbins, T. W. (2004). The neuropsychology of ventral prefrontal
cortex: Decision-making and reversal learning. Brain and Cognition, 55, 41–53.
Cooper, P.S., Karayanidis, F., McKewen, M., McLellan-Hall, S., Wong, A.S., Skippen, P., &
Cavanagh, J. F. (2019). Frontal theta predicts specific cognitive control-induced behavioural
changes beyond general reaction time slowing. NeuroImage, 189, 130–140.
THETA-BETA POWER RATIO 371
Corbetta, M., Patel, G., & Shulman, G. L. (2008). The reorienting system of the human
brain: from environment to theory of mind. Neuron, 58, 306–324.
Daw, N. D., O’Doherty, D. P., Seymour, B., & Dolan R. J. (2006). Cortical substrates for explora-
tory decision making. Nature, 441, 876–879.
Di Michele, F., Prichep, L., John, E. R., & Chabot, R. J. (2005). The neurophysiology of attention-
deficit/hyperactivity disorder. International Journal of Psychophysiology, 58, 81–93.
Elston, T. W., Croy, E., & Bilkey, D. K. (2019). Communication between the anterior cingu-
late cortex and ventral tegmental area during a cost-benefit reversal task. Cell Reports, 26,
2353–2361.
Engel, A. K. & Fries, P. (2010). Beta-band oscillations: Signalling the status quo? Current
Opinion in Neurobiology, 20, 156–165.
Finnigan, S., Wong, A., & Read, S. (2016). Defining abnormal slow EEG activity in acute is-
chaemic stroke: Delta/alpha ratio as an optimal QEEG index. Clinical Neurophysiology, 127,
1452–1459.
Food and Drug Administration (2013). De novo classification request for Neuropsychiatric
EEG-Based Assessment Aid for ADHD (NEBA) System. Retrieved from http://www.acc
essdata.fda.gov/cdrh_docs/reviews/K1127 11.pdf
Gottlieb, S. (2001). Methylphenidate works by increasing dopamine levels. British Medical
Journal, 322, 259.
González Castro, P., Alvarez Pérez, L., Núñez Pérez, J. C., González-Pienda García, J. A.,
Alvarez García, D., & Muñiz Fernández, J. (2010). Cortical activation and attentional control
in ADAH subtypes. International Journal of Clinical and Health Psychology, 10, 23–39.
Gramszlo, C., Fogleman, N. D., Rosen, P. J., & Woodruff-Borden, J. (2018). Intolerance of
uncertainty in children with attention- hyperactivity disorder. Attention Deficit
deficit/
Hyperactivity Disorders, 10, 189–197.
Gray, J. A. (1982). The neuropsychology of anxiety: An enquiry into the septo-hippocampal system.
Oxford University Press.
Gray, J. A. (1985). Emotional behaviour and the limbic system. Advances in Psychosomatic
Medicine, 13, 1–25.
Guyton, A. C. (1976). Organ physiology: Structure and function of the nervous system. W.B. Saunders.
Hacker, C. D., Snyder, A. Z., Pahwa, M., Corbetta, M., & Leuthardt, E. C. (2017). Frequency-specific
electrophysiologic correlates of resting state fMRI networks. NeuroImage, 149, 446–457.
Hahn, B., Ross, T. J., Yang, Y., Kim, I., Huestis, M. A., & Stein, E. A. (2007). Nicotine enhances
visuospatial attention by deactivating areas of the resting brain default network. Journal of
Neuroscience, 27, 3477–3489.
Halawa, I. F., El Sayed, B. B., Amin, O. R., Meguid, N. A., & Abdel Kader, A. A. (2017). Frontal
theta/beta ratio changes during TOVA in Egyptian ADHD children. Neurosciences (Riyadh),
22, 287–291.
Harmon-Jones, E. & Gable P. A. (2018). On the role of asymmetric frontal cortical activity in
approach and withdrawal motivation: An updated review of the evidence. Psychophysiology
[Epub], 55. doi:10.1111/psyp.12879.
Hedges, L. V. & Olkin, I. (1985). Statistical methods for meta-analysis. Academic Press.
Hoptman, M. J. & Davidson, R. J. (1998). Baseline EEG asymmetries and performance on
neuropsychological tasks. Neuropsychologia, 36, 1343–1353.
Hefco, V., Yamada, K., Hefco, A., Hritcu, L., Tiron, A., & Nabeshima, T. (2003). Role of the
mesotelencephalic dopamine system in learning and memory processes in the rat. European
Journal of Pharmacology, 475, 55–60.
372 DENNIS J. L. G. SCHUTTER and J. LEON KENEMANS
Heilman, K. M., Voeller, K. K., & Nadeau, S. E. (1991). A possible pathophysiologic substrate of
attention deficit hyperactivity disorder. Journal of Child Neurology, 6, S76–81.
Holroyd, C. B. & Coles, M. G. H. (2002). The neural basis of human error processing: re-
inforcement learning, dopamine, and the error-related negativity. Psychological Review, 109,
679–709.
Holroyd, C. B. & Yeung, N. (2012). Motivation of extended behaviors by anterior cingulate
cortex. Trends in Cognitive Sciences, 16, 122–128.
Isiten, H. N., Cebi, M., Sutcubasi Kaya, B., Metin, B., & Tarhan, N. (2017). Medication effects on
EEG biomarkers in attention-deficit/hyperactivity disorder. Clinical EEG and Neuroscience,
48, 246–250.
Jarrett, M. A., Gable, P. A., Rondon, A. T., Neal, L. B., Price, H. F., & Hilton, D. C. (2017). An
EEG study of children with and without ADHD symptoms: Between-group differences
and associations with sluggish cognitive tempo symptoms. Journal of Attention Disorders, 1,
1087054717723986.
Jasper, H., Solomon, P., & Bradley, C. (1938) Electroencephalographic analyses of behaviour
problem children. American Journal of Psychiatry, 95, 641–658.
Jenkinson, N. & Brown, P. (2011). New insights into the relationship between dopamine, beta
oscillations and motor function. Trends in Neurosciences, 34, 611–618.
Jensen, O. & Lisman, J.E. (1996). Novel lists of 72 known items can be reliably stored in an
oscillatory short-term memory network: Interaction with long-term memory. Learning &
Memory, 3, 257–263.
Karakaş, S. & Barry, R. J. (2017). A brief historical perspective on the advent of brain oscillations
in the biological and psychological disciplines. Neuroscience and Biobehavioral Reviews, 75,
335–347
Klass, D. W. & Brenner, R. P. (1995). Electroencephalography of the elderly. Journal of Clinical
Neurophysiology, 12, 116–131.
Klimesch, W. (1999). EEG alpha and thetaoscillations reflect cognitive and memory perform-
ance: A review and analysis. Brain Research Reviews, 29, 169–195.
Knott, V. J. & Fisher, D. J. (2007). Naltrexone alteration of the nicotine-induced EEG and
mood activation response in tobacco-deprived cigarette smokers. Experimental and Clinical
Psychopharmacology, 15, 368–381.
Kobayashi, R., Honda, T., Hashimoto, J., Kashihara, S., Iwasa, Y., Yamamoto, K., . . . Nakao, T.
(2020). Resting-state theta/beta ratio is associated with distraction but not with reappraisal.
Biological Psychology, 155, 107942.
Kuczenski, R. & Segal, D.S. (2001). Locomotor effects of acute and repeated threshold doses of
amphetamine and methylphenidate: relative roles of dopamine and norepinephrine. Journal
of Pharmacology and Experimental Therapeutics, 296, 876–83.
Lansbergen, M. M., Schutter, D. J., & Kenemans, J. L. (2007). Subjective impulsivity and base-
line EEG in relation to stopping performance. Brain Research, 1148, 161–169.
Lansbergen, M. M., Arns, M., van Dongen-Boomsma, M., Spronk, D., & Buitelaar, J. K.
(2011). The increase in theta/beta ratio on resting-state EEG in boys with attention-
hyperactivity disorder is mediated by slow alpha peak frequency. Progress in
deficit/
Neuropsychopharmacology & Biological Psychiatry, 35, 47–52.
Loo, S. K., Teale, P. D., & Reite, M. L. (1999). EEG correlates of methylphenidate response
among children with ADHD: A preliminary report. Biological Psychiatry, 45, 1657–1660.
THETA-BETA POWER RATIO 373
Loo, S. K., Hopfer, C., Teale, P. D., & Reite, M. L. (2004). EEG correlates of methylphenidate
response in ADHD: Association with cognitive and behavioral measures. Journal of Clinical
Neurophysiology, 21, 457–464.
Loo, S. K. & Makeig, S. (2012). Clinical utility of EEG in attention-deficit/hyperactivity dis-
order: a research update. Neurotherapeutics, 9, 569–587.
Loo, S. K., McGough, J. J., McCracken, J. T., & Smalley, S. L. (2018). Parsing heterogeneity in
attention- deficit hyperactivity disorder using EEG- based subgroups. Journal of Child
Psychology and Psychiatry, 59, 223–231.
Lubar, J. F. (1991). Discourse on the development of EEG diagnostics and biofeedback for
attention-deficit/hyperactivity disorders. Biofeedback and Self-Regulation, 16, 201–225.
Lubar, J. F. (1997). Neocortical dynamics: Implications for understanding the role of
neurofeedback and related techniques for the enhancement of attention. Applied
Psychophysiology and Biofeedback, 22, 111–126.
Mann C., Lubar, J., Zimmerman, A., Miller, C., & Muenchen, R. (1992). Quantitative analysis
of EEG in boys with attention deficit hyperactivity disorder: Controlled study with clinical
implications. Pediatric Neurology, 8, 30–36.
Markovska-Simoska, S. & Pop-Jordanova, N. (2017). Quantitative EEG in children and adults
with attention deficit hyperactivity disorder: Comparison of absolute and relative power
spectra and theta/beta ratio. Clinical EEG and Neuroscience, 48, 20–32.
Massar, S. A., Rossi, V., Schutter, D. J., Kenemans, J. L. (2012). Baseline EEG theta/beta ratio
and punishment sensitivity as biomarkers for feedback-related negativity (FRN) and risk-
taking. Clinical Neurophysiology, 123, 1958–1965.
Massar, S. A., Kenemans, J. L., & Schutter, D. J. (2014). Resting-state EEG theta activity and risk
learning: sensitivity to reward or punishment? International Journal of Psychophysiology, 91,
172–177.
McCarthy, H., Skokauskas, N., Mulligan, A., Donohoe, G., Mullins, D., Kelly,
J., . . . Frold, T. (2013). Attention network hypoconnectivity with default and affective net-
work hyperconnectivity in adults diagnosed with attention-deficit/hyperactivity disorder in
childhood. JAMA Psychiatry, 70, 1329–1337.
Metin, B., Roeyers, H., Wiersema, J.R., van der Meere, J., & Sonuga-Barke, E. (2012). A meta-
analytic study of event rate effects on go/no-go performance in attention-deficit/hyper-
activity disorder. Biological Psychiatry, 72, 990–996.
Miyake, A., Friedman, N. P., Emerson, M. J., Witzki, A. H., Howerter, A., & Wager, T. D. (2000).
The unity and diversity of executive functions and their contributions to complex “frontal
lobe” tasks: A latent variable analysis. Cognitive Psychology, 41, 49–100.
Monastra, V. J., Lubar, J. F., Linden, M., van Deusen, P., Green, G., Wing, W., et al. (1999).
Assessing attention deficit hyperactivity disorder via quantitative electroencephalog-
raphy: An initial validation study. Neuropsychology, 13, 424–433.
Monastra, V. J., Lubar, J. F., & Linden, M. (2001). The development of a quantitative
electroencephalographic scanning process for attention deficit- hyperactivity dis-
order: Reliability and validity studies. Neuropsychology, 15, 136–144.
Morillas-Romero, A., Tortella-Feliu, M., Bornas, X., & Putman, P. (2015). Spontaneous EEG
theta/beta ratio and delta-beta coupling in relation to attentional network functioning
and self-reported attentional control. Cognitive Affective and Behavioral Neuroscience, 15,
598–560.
374 DENNIS J. L. G. SCHUTTER and J. LEON KENEMANS
Nazari, M. A., Wallois, F., Aarabi, A., & Berquin, P. (2011). Dynamic changes in quantitative
electroencephalogram during continuous performance test in children with attention-
deficit/hyperactivity disorder. International Journal of Psychophysiology, 81, 230–236.
Ogrim, G., Kropotov, J., & Hestad, K. (2012). The quantitative EEG theta/beta ratio in attention
deficit/hyperactivity disorder and normal controls: Sensitivity, specificity, and behavioral
correlates. Psychiatry Research, 198, 482–488.
Panksepp, J. (1998). Affective neuroscience: The foundations of human and animal emotions.
Oxford University Press.
Panksepp J. & Biven, L. (2012). The archeology of mind: Neuroevolutionary origins of human
emotions. W.W. Norton & Company.
Poole, K. L., Hassan, R., & Schmidt, L. A. (2021). Temperamental shyness, frontal EEG theta/
beta ratio, and social anxiety in children. Child Development, 92, 2006–2019.
Putman, P., van Peer, J., Maimari, I., & van der Werff, S. (2010). EEG theta/beta ratio in relation
to fear-modulated response-inhibition, attentional control, and affective traits. Biological
Psychology, 83, 73–78.
Putman, P., Verkuil, B., Arias-Garcia, E., Pantazi, I., & van Schie, C. (2014). EEG theta/beta
ratio as a potential biomarker for attentional control and resilience against deleterious
effects of stress on attention. Cognitive Affective and Behavioral Neuroscience, 14, 782–791.
Rahman, M. M., Shukla, A., & Chattarji, S. (2018). Extinction recall of fear memories formed
before stress is not affected despite higher theta activity in the amygdala. Elife, 7, e35450.
Robbins, T. W. (2000). Chemical neuromodulation of frontal-executive functions in humans
and other animals. Experimental Brain Research, 133, 130–138.
Rosanova, M., Casali, A., Bellina, V., Resta, F., Mariotti, M., & Massimini, M. (2009). Natural
frequencies of human corticothalamic circuits. Journal of Neuroscience, 29, 7679–7685.
Saad, J. F., Kohn, M. R., Clarke, S., Lagopoulos, J., & Hermens, D. F. (2018). Is the theta/beta
EEG marker for ADHD inherently flawed? Journal of Attention Disorders, 22, 815–826.
Satterfield, J. H. & Cantwell, D. P. (1974). Proceedings: CNS function and response to methyl-
phenidate in hyperactive children. Psychopharmacology Bulletin, 10, 36–38.
Satterfield, J. H. & Dawson, M. E. (1971). Electrodermal correlates of hyperactivity in children.
Psychophysiology, 8, 191-197.
Schleiger, E., Sheikh, N., Rowland, T., Wong, A., Read, S., & Finnigan, S. (2014). Frontal
EEG delta/alpha ratio and screening for post-stroke cognitive deficits: the power of four
electrodes. International Journal of Psychophysiology, 94, 19–24.
Schutte, I., Kenemans, J. L., & Schutter, D. J. (2017). Resting-state theta/beta EEG ratio is
associated with reward-and punishment-related reversal learning. Cognitive Affective and
Behavioral Neuroscience, 17, 754–763.
Schutter, D. J. & Knyazev, G. G. (2012). Cross-frequency coupling of brain oscillations in
studying motivation and emotion. Motivation and Emotion, 36, 46–54.
Schutter, D. J. & van Honk, J. (2005). Electrophysiological ratio markers for the balance be-
tween reward and punishment. Cognitive Brain Research, 24, 685–690.
Schutter, D. J., Leitner, C., Kenemans, J. L., & van Honk, J. (2006). Electrophysiological
correlates of cortico-subcortical interaction: A cross-frequency spectral EEG analysis.
Clinical Neurophysiology, 117, 381–387.
Schutter, D. J., de Weijer, A. D., Meuwese, J. D., Morgan, B., & van Honk, J. (2008). Interrelations
between motivational stance, cortical excitability, and the frontal electroencephalogram
asymmetry of emotion: A transcranial magnetic stimulation study. Human Brain Mapping,
29, 574–80.
THETA-BETA POWER RATIO 375
Sari Gokten, E., Tulay, E.E., Beser, B., Elagoz Yukse, M., Arikan, K., Tarhan, N., & Metin, B .
(2019). Predictive value of slow and fast EEG oscillations for methylphenidate response in
ADHD. Clinical EEG and Neuroscience, 50, 332–338.
Seli, P., Smallwood, J., Cheyne, J. A., & Smilek, D. (2015). On the relation of mind wandering
and ADHD symptomatology. Psychonomic Bulletin & Review, 22, 629–636.
Seth, A. K. (2013). Interoceptive inference, emotion, and the embodied self. Trends in Cognitive
Sciences, 17, 565–573.
Shi, T., Li, X., Song, J., Zhao, N., Sun, C., et al. (2012). EEG characteristics and visual cogni-
tive function of children with attention deficit hyperactivity disorder (ADHD). Brain &
Development, 34, 806–811.
Snyder, S. M. & Hall, J. R. (2006). A meta-analysis of quantitative EEG power associated with
attention-deficit hyperactivity disorder. Journal of Clinical Neurophysiology, 23, 440–455.
Snyder, S. M., Quintana, H., Sexson, S. B., Knott, P., Haque, A. F. M., & Reynolds, D. A. (2008).
Blinded, multi-center validation of EEG and rating scales in identifying ADHD within a
clinical sample. Psychiatry Research, 159, 346–358.
Sohn, H., Kim, I., Lee, W., Peterson, B.S., Hong, H., et al. (2010). Linear and non-linear EEG
analysis of adolescents with attention-deficit/hyperactivity disorder during a cognitive task.
Clinical Neurophysiology, 121, 1863–1870.
Tortella-Feliu, M., Morillas-Romero, A., Balle, M., Llabrés, J., Bornas, X., & Putman, P. (2014).
Spontaneous EEG activity and spontaneous emotion regulation. International Journal of
Psychophysiology, 94, 365–372.
Van Dongen-Boomsma, M., Lansbergen, M. M., Bekker, E. M., Sandra Kooij, J. J., van der
Molen, M., Kenemans, J. L., & Buitelaar, J. K. (2010). Relation between res-ting EEG to cog-
nitive performance and clinical symptoms in adults with attention-deficit/hyperactivity dis-
order. Neuroscience Letters, 469, 102–106.
Van Son, D., Angelidis, A., Hagenaars, M. A., van der Does, W., & Putman, P. (2018). Early and
late dot-probe attentional bias to mild and high threat pictures: Relations with EEG theta/
beta ratio, self-reported trait attentional control, and trait anxiety. Psychophysiology, 55,
e13274.
Van Son, D., de Blasio, F. M., Fogarty, J. S., Angelidis, A., Barry, R. J., & Putman, P. (2019a).
Frontal EEG theta/beta ratio during mind wandering episodes. Biological Psychology,
140, 19–27.
Van Son, D., de Rover, M., De Blasio, F. M, van der Does, W., Barry, R. J., & Putman, P. (2019b).
Electroencephalography theta/beta ratio covaries with mind wandering and functional connect-
ivity in the executive control network. Annals of the New York Academy of Sciences, 1452, 52–64.
Volkow, N. D., Wang, G., Fowler, J. S., Logan, J., Gerasimov, M., Maynard, L., . . . Franceschi, D.
(2001). Therapeutic doses of oral methylphenidate significantly increase extracellular dopa-
mine in the human brain. Journal of Neuroscience, 21, RC121.
Whitford, T. J., Rennie, C. J., Grieve, S. M., Clark, C. R., Gordon, E., & Williams, L. M. (2007).
Brain maturation in adolescence: concurrent changes in neuroanatomy and neurophysi-
ology. Human Brain Mapping, 28, 228–237.
Williams, L. M., Hermens, D. F., Thein, T., Clark, C. R., Cooper, N. J., Clarke, S. D., . . . Kohn,
M. R. (2010). Using brain based cognitive measures to support clinical decisions in ADHD.
Pediatric Neurology, 42, 118–126.
Williams, L. M., Tsang, T. W., Clarke, S., & Kohn, M. (2010). An ‘integrative neuroscience’
perspective on ADHD: linking cognition, emotion, brain and genetic measures with
implications for clinical support. Expert Review of Neurotherapeutics, 10, 1607–1621.
376 DENNIS J. L. G. SCHUTTER and J. LEON KENEMANS
Wischnewski, M., Zerr, P., & Schutter, D. J. (2016). Effects of theta transcranial alternating
current stimulation over the frontal cortex on reversal learning. Brain Stimulation, 9,
705–7 11.
Wischnewski, M., Joergensen M. L., Compen, B., & Schutter, D. J. (2020). Frontal beta
transcranial alternating current stimulation improves reversal learning. Cerebral Cortex,
30(, 3286–3295.
Wolinski, N., Cooper, N., Sauseng, P., & Romei, V. (2018). The speed of parietal theta frequency
drives visuospatial working memory capacity. PLoS Biology, 16, e2005348.
Zhang. D. W., Li, H., Wu, Z., Zhao, Q., Song, Y., Liu, L., . . . Sun, L. (2019). Electroencephalogram
Theta/Beta Ratio and Spectral Power Correlates of Executive Functions in Children and
Adolescents With AD/HD. Journal of Attention Disorders, 23, 721–732.
CHAPTER 16
C ORTICAL S OU RC E
L O CALIZ ATION I N E E G
F REQU ENCY A NA LYSI S
16.1 Introduction
to Source Localization
EEG signals recorded via electrodes placed on the scalp represent the postsynaptic
potentials generated by mass synchronized pyramidal neurons perpendicular to the
cortical surface. Source localization or source analysis is the activity that aims to iden-
tify the underlying cortical generators or sources of the EEG potentials measured on the
scalp. The sources of the EEG signals may be modeled as electrical dipoles with both dir-
ectionality and amplitude. The identification of the location of dipoles is often obscured
because the current generated by these sources spreads in all directions in the brain and
is smeared by the skull. However, recent advances made in improving the techniques
for source localization have substantially increased the spatial resolution of this method
and led to the tendency towards using it as a brain imaging tool (Michel & Murray, 2012).
Source localization consists of procedures creating a forward (Hallez et al., 2007) and
an inverse (Grech et al., 2008) model. The forward model is created with a head model
and electrodes and describes the effect of an electrical source inside the head on the
EEG. The inverse model is the “inverse” of the forward model and is used with the EEG
recording to compute the location and amplitude of the sources that generated the given
EEG. The present chapter provides an overview of some widely adopted source ana-
lysis techniques and their applications to EEG frequency analysis. The importance of
using realistic head models for source analysis is also discussed. In the last section of the
chapter, we present examples for using source analysis as a neuroimaging tool in EEG
frequency analysis.
378 WANZE XIE and JOHN E. RICHARDS
There are two major approaches to source localization. Equivalent current dipole (ECD)
models use a limited number of electrical dipoles that are computed with the forward
model to explain the distribution of the EEG on the scalp. Distributed source models use
a large set of dipole locations distributed across the brain and are calculated using the in-
verse model and the observed EEG to generate the amplitude of the current density (He
et al., 2018; Michel et al., 2004). This section presents both models and describes how
these models are used for source analysis.
number and location of dipoles based on other neuroimaging data, such as functional
magnetic resonance imaging (fMRI) and positron emission tomography (PET) (Agam
et al., 2011; Foxe et al., 2003). For example, a meta-analysis of previous fMRI reports was
conducted by Foxe and colleagues (2003) to find out the brain regions that were consist-
ently activated in the attention task used by the authors. The location of these regions
was used as a guide in their source analysis, such that the MRI coordinates of these
areas were used as the seeds for the dipoles in the forward solution, with the assumption
that the task-related EEG activity was generated in these areas. A priori assumptions
can also be made based on theoretical rationale and findigns from previous research,
such as using the fusiform and occipital face area (FFA; OFA) as the potential sources
for the N170 component (Gao et al., 2019), the occipital and parietal regions for alpha
oscillations (Xie et al., 2017), and the pre-central cortex for the mu rhythm (Thorpe
et al., 2016). Gao and colleagues (2019) conducted a thorough literature review of the
cortical source locations of the N170 ERP component found by previous EEG studies
and the 3D coordinates of the FFA identified by previous fMRI research, and then used
these locations as a priori defined regions of interest (ROIs) in their cortical source lo-
calization. Section 2.3 provides further information about how to define a priori ROIs
for source localization.
Thus, additional constraints must be imposed to obtain unique and well-posed linear
inverse solutions (Grech et al., 2008; He et al., 2018).
The solution to the underdetermined construction of the inverse spatial filter is to
constrain the solution by mathematical or quantitative procedures (Grech et al., 2008;
Michel & He, 2012). The following paragraphs introduce a couple of widely adopted
solutions in distributed source modeling: minimum norm estimation (MNE), low-
resolution electromagnetic tomography (LORETA), standardized low- resolution
electromagnetic tomography (sLORETA), and exact low-resolution electromagnetic
tomography (eLORETA). Pascual-Marqui and colleagues have made great contribution
to the development of the “LORETA family” in the past two decades (Sherlin, 2009).
A number of additional solutions have been developed and utilized by the EEG re-
search community, such as weighted MNE, shrinking LORETA FOCUSS (SLF), local
autoregressive average, and beamforming techniques. Detailed description of these
methods can be found elsewhere (Grech et al., 2008; Green & McDonald, 2009; Hallez
et al., 2007; He et al., 2018).
The first introduced algorithm or constraint to the inverse solution in distributed
source modeling was the minimum norm least- squares inverse (Hämäläinen &
Ilmoniemi, 1994), often called the MNE. This algorithm minimizes the least-square
error of the estimated inverse solution X in the equation described earlier, which means
it results in an inverse solution with the lowest overall intensity (Hauk, 2004). The in-
verse solution obtained from MNE tends to be biased towards weak and superficial
sources (Michel et al., 2004). To overcome the problem of this surface-restricted MNE
algorithm, several methods have been developed that keep the basic mathematical
relations of the MNE but alter its characteristics to resolve its weaknesses. For example,
an early modification was to use a depth-weighting matrix in the formula to account for
the MNE’s bias toward superficial surfaces.
The LORETA algorithm was devised to add to the MNE by the additional constraint
to smooth the sources with a Laplacian filter (Pascual-Marqui et al., 1994). The LORETA
method was also designed to solve the “surface source” issue of MNE, such that it favors
solutions with strong activation of a large number of sources and punishes solutions
with small number of surface sources. The Laplacian of the sources is a measure of spa-
tial roughness. Minimizing the Laplacian of sources leads to lower resolution in the
source space, and thus the solution of LORETA algorithm is smoother than MNE and
weighted MNE algorithms.
Pascual-Marqui (2002) developed the more widely used inverse algorithm sLORETA,
which uses the current density estimate given by the MNE solution and weights the so-
lution by the standardized values of the current density estimate. This partially alleviates
the emphasis on superficial sources found in the MNE by enhancing the smaller deeper
sources with their lower standard deviation. It was found that sLORETA results in fewer
errors in reconstructed sources than MNE (Pascual-Marqui, 2002). The source local-
ization accuracy has been compared using weighted MNE, LORETA, and sLORETA
by Monte-Carlo analysis of data with different noise levels and sources with different
depth in the brain (Grech et al., 2008). Grech and colleagues found that the sLORETA
CORTICAL SOURCE LOCALIZATION IN EEG FREQUENCY ANALYSIS 381
algorithm had the lowest level of localization error and fewest ghost sources, which are
sources estimated from the inverse solution but not actually present in the simulated
data. Therefore, sLORETA outperforms the other methods regarding the accuracy of
source localization using simulated data with different levels of artifacts.
The eLORETA algorithm is another member of the “MNE family” that is also built
upon the linear weighted MNE inverse solution and has been regarded as improvement
over the previously developed LORETA and sLORETA algorithms (Pascual-Marqui,
2007). The eLORETA algorithm provides exact localization with zero error (i.e., no
localization bias) to the inverse solution even in the presence of measurement and
structured noise in the data. Assuming there is an activated dipole in the brain with an
arbitrary orientation and known magnitude, the scalp EEG potentials generated by this
dipole can be simulated. Applying the eLORETA algorithm as the inverse solution to
the scalp measurement would give reconstructed current density fields. The “zero error
localization” property of eLORETA means that there is zero distance between the actual
point-test source and the reconstructed source with the absolute maximum amplitude.
It should be noted this property has not been achieved by previously published discrete
linear distributed source modeling algorithms (Pascual-Marqui, 2007).
Jatoi and colleagues (2014) compared the source localization accuracy between
sLORETA and eLORETA for EEG data collected in an experiment with visual stimuli.
The eLORETA algorithm was found to have enhanced ability to suppress less-significant
sources in the brain compared to sLORETA. Jatoi and colleagues also found that
eLORETA gave less localization error and clearer (less blurry) images as compared to
sLORETA. As a member of the LORETA family, the moments in neighboring neuronal
sources are also highly correlated for eLORETA.
2008; Michel et al., 2004). For example, using the CDR approach, all potential source
locations (e.g., brain volumes) are simultaneously estimated and the relative magni-
tude of the reconstructed source activity, also called the current source density (CSD),
provides the putative locations of the sources. The source space of the distributed source
models may be used to constrain the solution to physiologically reasonable positions.
For example, since the generation of the EEG occurs in the postsynaptic potentials of
the columnar pyramidal cells, the dipoles may be limited to gray matter (GM) and the
dipole directions fixed as perpendicular to the cortical surface. These characteristics
have led to a growing tendency in the past two decades to replace ECD optimization
methods with distributed source modeling.
A recent study by Gao and colleagues (2019) investigated the cortical sources of the
N170 face-sensitive ERP component. Individual structural MRI, fMRI, and EEG data
were collected. The anatomical information of the head and brain was used to create
individual realistic head models for source localization. The face-sensitive regions
identified during the fMRI scan were used as the a priori defined ROIs in the analyses
of the reconstructed source activity of the N170 component. This procedure reduces
the family-wise error rate that would otherwise be caused by multiple comparisons or
analyses across all source volumes in the brain.
The forward model that is used in both equivalent current dipole models and
distributed source models and is created with a head model. The head model in
EEG source localization describes media inside the head with their relative elec-
trical conductivity. The head model is the link between the source volumes and
the electrical potentials on the scalp in source analysis. The head models for
EEG source localization have progressed from spherical models with only two or
three compartments to realistic models with all different tissues of the head being
segmented and assigned for their own conductivity values. This section introduces
the importance of using realistic head models in EEG source analysis and its usage
in empirical research. Since this chapter mostly focuses on realistic head models, an
overview of other types of head models can be found elsewhere (Grech et al., 2008;
Hallez et al., 2007).
The creation of the FEM model used to be much more computational demanding
than the BEM model (Michel et al., 2004), but the computational efficiency of computers
has now been substantially improved, which makes the FEM model more applic-
able in source analysis. In a recent paper, Vorwerk and colleagues (2018) introduced
a MATLAB-based pipeline, the “Fieldtrip-SimBio” pipeline, that aims to reduce the
workload for the generation of multicompartment FEM models. This pipeline has been
added to the existing package of functions to create FEM models in the Fieldtrip toolbox
(Oostenveld et al., 2011), which allows for an easy construction and application of a FEM
model for source analysis (Vorwerk et al., 2018).
Many studies have found that it is important to model the inhomogeneous distri-
bution of brain and head tissues (Cho et al., 2015; Vorwerk et al., 2014). Vorwerk and
colleagues (2014) comprehensively investigated the influence of modeling versus not
modeling certain conductive tissues of the head on the EEG and MEG forward so-
lution. They tested the effect of hierarchically adding new materials to a basic three-
compartment head model (brain, skull, scalp) on the accuracy of source localization.
Specifically, they tested the effect of the addition of realistic distribution of CSF, GM,
and WM, differentiation of the skull spongiosa and compacta, and anisotropic WM
conductivities on source localization. The largest increase in the accuracy of signal
topography and amplitude from model to model came with the addition of CSF to the
three-compartment model, followed by lesser (but still appreciable) increases when
adding GM and WM and the anisotropic WM conductivities into the head models. The
lack of the GM/WM distinction, or CSF, in the BEM compartment models, would re-
sult in considerable errors in the reconstructed source activities, which concurs with
Cho and colleagues’ (2015) study, which showed that the distinction between these brain
tissues in source localization is important for analyzing EEG and MEG functional con-
nectivity between cortical regions. Taken together, these findings highlight the import-
ance of modeling the head conductive compartments as realistically as possible in order
to increase the accuracy of source localization.
By contrast, a few other studies have compared the difference between using BEM
and FEM models for source localization but found no substantial difference in the
reconstructed source activation (Michel & He, 2012). For example, Birot and colleagues
(2014) compared the source localization accuracy with LORETA using BEM and FEM
models with 38 epileptic patients. These patients underwent scalp EEG and intracranial
EEG recordings and subsequent resection surgery. The results of this study implicated
that using the BEM and FEM models did not cause a significant difference in the ac-
curacy of source localization (Birot et al., 2014). This finding suggests that the choice of a
head model may not be a crucial factor for source localization using distributed localiza-
tion algorithm for some applications.
brain-to-skull conductivity ratio (Acar & Makeig, 2013). There is a consensus that the
human skull has a much lower conductivity value than the other tissues inside the head.
However, there is dispute about the correct skull conductivity value or the skull-to-brain
conductivity ratio to use for creating head models (Michel & He, 2012). The ratio of 1/
80 is commonly specified as the skull-to-brain conductivity ratio in head models (ei-
ther spherical or realistic), with their conductivity values being 0.0042 and 0.33 S/m
(Siemens per meter, reciprocal of ohm Ω), respectively. The skull-to-brain conductivity
ratio of 1/80 originated from studies that measured the electrical properties of brain and
head tissues (Gabriel et al., 1996; Rush & Driscoll, 1969). This ratio has been set up as
the default in several source localization toolboxes, such as brain electrical source ana-
lysis (BESA; Megis Software GmbH, Gräfelfing, Germany) and Fieldtrip (Oostenveld
et al., 2011), and widely used in empirical research and simulated studies (Ryynanen
et al., 2006).
More recent studies have suggested that the traditional skull to brain conductivity
ratio of 1/80 might be misestimated. A few studies have been conducted to estimate
or measure the conductivity of human skull, and findings have shown that the correct
skull-to-brain conductivity ratio may be between 1/15 to 1/30, which is much higher than
the traditional ratio, and using a lower skull-to-brain conductivity ratio (e.g., 1/80) may
result in shallower source locations (Acar et al., 2016; Acar & Makeig, 2013; Dannhauer
et al., 2011; Lai et al., 2005; Oostendorp et al., 2000). Therefore, the traditional ratio of
skull-to-brain conductivity may need to be adjusted when creating head models for cor-
tical source localization.
Bone conductivities are age dependent, with infants and children having much
higher skull conductivity values compared to adults (Odabaee et al., 2014). This
means that using the adult skull conductivity values to create head models for infant
and child participants could lead to inaccurate source localization results (Reynolds
& Richards, 2009). To this end, some source localization software, for example,
BESA, provides references for the skull-to-brain conductivity ratio for children and
adolescents at different ages. Recent studies have also started to adopt age-appropriate
skull conductivity values for source localization on infant EEG data (Xie, Jensen,
et al., 2019).
Human brain oscillatory activity is generated by neural tissue in the central nervous
system spontaneously or in response to stimuli in the environment. The significance
of these oscillatory signals in sensory-cognitive processes has become increasingly
evident. Using source localization in EEG (time) frequency analysis has made great
contributions to our understanding of the cortical mechanisms of cognitive processes
associated with oscillatory activities in various frequency bands. Using source localiza-
tion of EEG signals to obtain reconstructed cortical activities also allows us to investi-
gate functional connectivity withing and between brain networks during cognitive tasks
and resting-state. The current section reviews the application of source localization
techniques in EEG frequency analysis and how this method has provided insights into
brain functions and network organizations.
et al., 2007), sustained attention and vigilance (Kim et al., 2017), error monitoring and
conflict processing (Cohen, 2011; Cohen & Ridderinkhof, 2013; Yeung et al., 2007),
working memory (Maurer et al., 2015; Michels et al., 2010; Michels et al., 2008; Onton
et al., 2005), motion observation and execution (Nystrom et al., 2011; Ritter et al., 2009;
Thorpe et al., 2016). There is a large body of literature on the developmental origin of
the relation between these cognitive processes and corresponding EEG oscillations (Bell
& Cuevas, 2012). EEG source localization has also made it possible to investigate the
cortical sources of EEG oscillatory activities associated with cognitive performance in
infants and children (see Section 16.4.4).
The source localization techniques discussed in the previous sections open the av-
enue to unravel the crucial role of EEG oscillatory activity in human brain functions.
For instance, Sauseng and colleagues (2007) investigated the function of the theta-
band oscillations in sustained attention and executive control. They examined how
EEG theta-band activity changed as a function of task difficulty and memory load. The
participants were required to execute complex sequential finger movements requiring
different levels of memory load, such as execution of a previously trained simple se-
quence, an overlearned complex sequence, or a novel complex sequence. Local frontal-
midline theta activity on the scalp was found to be associated with the general level of
cognitive demand, with the highest power found for the most demanding condition.
This theta activity was reconstructed to the source space using the LORETA method.
Specifically, the CSD was reconstructed for more than 2000 cortical voxels across the
brain using distributed source modeling, separately for different cognitive conditions.
The source activity was then compared voxel by voxel between the conditions, and mul-
tiple comparisons were statistically controlled. The authors found that the CSD in the
anterior cingulate gyrus was significantly different between the conditions, suggesting
that theta-band oscillations generated by this region may reflect the activation of an at-
tentional system responsible for allocating cognitive resources (Sauseng et al., 2007).
Theta-band activity generated by the frontal cortex also plays an important role
in error monitoring (Cohen, 2011). Cohen found that EEG theta-band activity in the
frontal electrodes was associated with error-monitoring behaviors in adults. The author
subsequently conducted source localization of the theta-band activity using equiva-
lent current dipole modeling with subject-specific BEM models created with individual
MRIs. One dipole that explained the most amount of variance of the scalp potentials was
estimated per subject. The author found that the dipoles in the medial prefrontal and
anterior cingulate cortexes best explained the variance in the frontal theta activity across
subjects, suggesting that these areas are the potential sources of the distribution of the
theta activity on the scalp.
EEG oscillations generated by the frontal area has been found to play an important
role in maintaining vigilance (sustained attention) during cognitive tasks. Kim and
colleagues (2017) investigated which brain areas are engaged in controlling the mainten-
ance of vigilance using source analysis of EEG signals. Distributed source modeling with
the sLORETA algorithm was conducted to obtain smooth source activation in different
frequency bands across the entire cortex. The authors found that the activity in the left
388 WANZE XIE and JOHN E. RICHARDS
prefrontal cortex was significantly correlated with vigilance variation in the delta, beta,
and gamma bands, which suggests the involvement of this brain region in maintaining
vigilance. To sum up, these empirical studies demonstrated how using EEG source ana-
lysis in frequency analysis could advance our understanding of the neural mechanisms
underlying different cognitive processes.
seconds following the demand for glucose via oxygenated hemoglobin. This restricts
fMRI connectivity analyses to “seconds” resolution and prohibits this technique from
studying dynamic changes in neural oscillations in sub-second frequency ranges. By
contrast, EEG and MEG directly measure synchronized neuronal activity with msec
time resolution and thus can measure dynamic changes in neural oscillations at the
resolution of the neural system (e.g., sub-seconds resolution).
The examination of EEG functional connectivity between electrodes on the scalp
has been used to study synchronized fluctuations in various rhythms that may reflect
the dynamic coordination and network structure inside the brain (Bastos & Schoffelen,
2016; Stam, 2005). A variety of methods have been developed to evaluate the “correl-
ation or synchronization” between the signals from two electrodes, such as the correl-
ation between power envelope (Hipp et al., 2012), coherence analysis (Bowyer, 2016),
and phase synchronization (Stam et al., 2007). In addition, previous work studying EEG
functional connectivity on the scalp level with infants and children provides insights
into the developmental origin of integrated brain networks (see Chu-Shore et al., 2011
for a review). For example, Cuevas and colleagues (2012) investigated the association be-
tween infants’ working memory and inhibitory control performance in cognitive tasks
and the coherence between frontal electrodes in the alpha band. The authors found that
frontal alpha coherence was associated with inhibitory control processes in 10-month-
old infants, which suggests that the cooperation between brain regions already plays an
important role in cognitive performance in infancy (Cuevas et al., 2012).
The volume conduction/field spread issue of surface EEG and the effect of the ref-
erence electrode on phase synchronization and correlation values impede our under-
standing of the underlying neural mechanisms from the EEG functional connectivity
results on the scalp (Bastos & Schoffelen, 2016; Guevara et al., 2005; Schiff, 2005).
Specifically, this problem means that the activity of a cortical source can be “visible” at
multiple electrodes due to the field smearing on the scalp. This will cause spurious cor-
relation or synchronization values between electrodes. Consequently, the interpretation
of the scalp-level connectivity requires considerable caution because two “connected
or synchronized” electrodes may not reflect the brain regions that are functionally
connected.
Applying cortical source reconstruction methods to EEG data could substantially
mitigate the consequence of volume conduction. Brain functional connectivity using
EEG could be done by first estimating the sources of the surface signals, deriving time-
domain activity of the cortical sources via the time-domain activity of the surface ac-
tivity, and then doing functional connectivity in the source-space by calculating the
correlation between amplitudes (or power envelope) of times-series in cortical sources
or phase synchronization in different frequency bands after Fourier transformation of
the time-series in cortical sources into the frequency domain. For example, Xie and
colleagues (2018) developed a pipeline to estimate functional connectivity in the cor-
tical source space to investigate the function of different brain networks of awake infants
(Figure 16.1). This pipeline includes cortical source reconstruction of EEG recordings
with age- appropriate MRIs, parcellation of the source activity into brain ROIs,
390 WANZE XIE and JOHN E. RICHARDS
Infant FEM
model
Source Reconstruction
20 seed-based ROIs
Connectivity
analysis
wPLI
Connectivity between
the 20 seed-based ROIs
Figure 16.1 The pipeline for EEG functional connectivity analysis in the source space used by
Xie et al. (2018).
CORTICAL SOURCE LOCALIZATION IN EEG FREQUENCY ANALYSIS 391
associated with attenuated connectivity in the alpha band within the dorsal attention
network and the DMN, as well as distinct network organization and efficiency indicated
by graph theory measures (Xie et al., 2018). This work highlights the possibility of
using source-space functional connectivity analysis to study the development of brain
networks in early childhood. The following paragraphs introduce more recent advances
that have been made in combining EEG source localization and functional brain con-
nectivity analysis in the frequency domain.
EEG source localization can be used to detect brain functional connectivity during
tasks. The functional connectivity between frontal and parietal cortices plays an
important role in executive control. Sauseng and colleagues (2007) found a more
distributed pattern of functional connectivity in the theta band between frontal
and parietal regions when participants executed novel compared to memorized
figure movement sequences. In a more recent study, Cohen and Ridderinkhof (2013)
investigated the functional coupling between EEG signals in different frequency bands
during spatial conflict processing. Inter-regional connectivity in the source space
was found to differ between the congruent and incongruent conditions in a Simon
task. Congruent trials induced stronger coupling between frontal theta and par-
ietal alpha and gamma power, while incongruent trials induced stronger coupling of
the theta power between the medial and lateral frontal regions. Studies also show that
the inter-regional phase synchrony in low-and high-frequency bands is associated
with top-down control of visual and auditory spatial attention (Doesburg et al., 2009;
Doesburg et al., 2012). For example, Doesburg and colleagues (2009) found increased
phase synchronization in the alpha band between the visual cortex and parietal regions
contralateral to the attended visual hemifield and decreased phase synchronization be-
tween those brain regions ipsilateral to the attended visual hemifield. The Cohen and
Doesburg studies used the beamforming technique for cortical source reconstruction.
The beamforming techniques are not covered here and have been reviewed elsewhere
(Green & McDonald, 2009).
The functional connectivity analysis in the source space of spontaneous EEG during
resting-state has provided insights into the structure and functioning of brain networks,
as well as the development of brain networks over childhood. Liu and colleagues (2017)
detected large-scale brain networks in human adults using high-density EEG recordings
along with the eLORETA technique and realistic head models for source analysis. ICA
analysis of source reconstructed power envelops was conducted to extract functionally
connected brain regions (networks) in different frequency bands. Their results showed
that the brain networks identified with resting-state EEG data (e.g., the DMN, attention
and language networks) were comparable to the brain networks obtained from prior
resting-state fMRI research (Liu et al., 2017). Mantini and colleagues’ (2007) simultan-
eous EEG-fMRI data were collected during resting-state to unravel the direct relation-
ship between functional connectivity measured by the two metrics. Six major brain
networks (e.g., the DMN, auditory, visual, and attention networks) were identified from
the BOLD signals using ICA. The fluctuations in each of these fMRI-defined resting-
state networks were found to be correlated with EEG power in different frequency bands
392 WANZE XIE and JOHN E. RICHARDS
(e.g., delta, theta, alpha, beta, and gamma). This study was one of the first attempts to ex-
plore the link between EEG brain rhythms and low-frequency coherent fluctuations of
the BOLD signal during resting-state.
sources observed during infant sustained attention. Specifically, the authors examined
whether infant sustained attention, defined with the infant heart-rate attention model
(Richards, 2003), was accompanied by distinct EEG rhythmic activities in the theta,
alpha and beta bands in attention-related cortical areas. Cortical source reconstruc-
tion of EEG power in different frequency bands was conducted using the eLORETA
method. Realistic head models were created for each participant using individual MRIs.
Infant sustained attention was found to be accompanied by increased theta power in the
orbitofrontal and ventral temporal areas and decreased alpha power in brain regions
within the default mode network (DMN). The relation between infant sustained
attention and EEG power was not found among infants aged 6 and 8 months, but the re-
lation emerged at age 10 months and became well established by age 12 months (Figure
16.2). This study established the a connection between infant sustained attention and
cortical oscillatory activity in the theta and alpha bands.
50 50
40
40 40
30
30 30
20
20 20
SO
SO
SO
SA
SA
SA
AT
AT
AT
10 10
10
0 0 0
10 Mos 12 Mos 10 Mos 12 Mos 10 Mos 12 Mos
Figure 16.2 Development of the effect of sustained attention on infant theta source activation.
(A) 3D displays for the difference in CDR amplitude between sustained attention and attention
termination separately for the four ages. Age-appropriate average MRI templates were used for
the display for each age. Sustained attention effect was primarily shown in the temporal pole,
orbital frontal, and ventral temporal regions, especially for 10 and 12 months. (B) Bar graphs
for the average CDR amplitudes in these brain networks for 10 and 12 months. Error bars repre-
sent SEMs.
Adapted from Xie et al., 2017.
394 WANZE XIE and JOHN E. RICHARDS
16.5 Conclusion
This chapter introduces EEG source localization and its usage in (time) frequency
analysis. EEG source localization offers superb temporal resolution and increasingly
improved spatial resolution. Methodological progress helped to improve the accuracy
of source localization, for example, techniques developed to constrain the inverse
solutions in distributed source reconstruction. The importance of using realistic head
models and age-appropriate conductivity values for source localization is also increas-
ingly recognized by the field, especially for research with pediatric populations. These
advances extended the application of source localization in clinical diagnosis and sci-
entific research. Studies using EEG source localization and frequency analysis provide
importance insights into the functional significance of oscillatory activities in different
frequency rhythms in cognitive processes, as well as the maturation of functional brain
networks over development.
REFERENCES
Acar, A. Z., Acar, C. E., & Makeig, S. (2016). Simultaneous head tissue conductivity
and EEG source location estimation. Neuroimage, 124(Pt A), 168– 180. doi:10.1016/
j.neuroimage.2015.08.032
CORTICAL SOURCE LOCALIZATION IN EEG FREQUENCY ANALYSIS 395
Acar, A. Z., & Makeig, S. (2013). Effects of forward model errors on EEG source localization.
Brain Topography, 26(3), 378–396. doi:10.1007/s10548-012-0274-6
Agam, Y., Hamalainen, M. S., Lee, A. K., Dyckman, K. A., Friedman, J. S., Isom, M., . . . Manoach,
D. S. (2011). Multimodal neuroimaging dissociates hemodynamic and electrophysiological
correlates of error processing. Proceedings of the National Academy of Sciences of the United
States of America, 108(42), 17556–17561. doi:10.1073/pnas.1103475108
Bastos, A. M. & Schoffelen, J. M. (2016). A tutorial review of functional connectivity ana-
lysis methods and their interpretational pitfalls. Frontiers in Systems Neuroscience, 9, 175.
doi:10.3389/fnsys.2015.00175
Bathelt, J., O’Reilly, H., Clayden, J. D., Cross, J. H., & de Haan, M. (2013). Functional brain net-
work organisation of children between 2 and 5 years derived from reconstructed activity
of cortical sources of high-density EEG recordings. Neuroimage, 82, 595-604. doi:10.1016/
j.neuroimage.2013.06.003
Bell, M. A., & Cuevas, K. (2012). Using EEG to study cognitive development: issues and practices.
Journal of Cognition and Development, 13(3), 281–294. doi:10.1080/15248372.2012.691143
Birot, G., Spinelli, L., Vulliemoz, S., Megevand, P., Brunet, D., Seeck, M., & Michel, C.
M. (2014). Head model and electrical source imaging: a study of 38 epileptic patients.
Neuroimage: Clinical, 5, 77–83. doi:10.1016/j.nicl.2014.06.005
Bowyer, S. M. (2016). Coherence a measure of the brain networks: past and present.
Neuropsychiatric Electrophysiology, 2(1), 1. doi:10.1186/s40810-015-0015-7
Brodbeck, V., Spinelli, L., Lascano, A. M., Wissmeier, M., Vargas, M. I., Vulliemoz, S., . . . Seeck,
M. (2011). Electroencephalographic source imaging: a prospective study of 152 operated epi-
leptic patients. Brain, 134(Pt 10), 2887–2897. doi:10.1093/brain/awr243
Cho, J. H., Vorwerk, J., Wolters, C. H., & Knosche, T. R. (2015). Influence of the head model
on EEG and MEG source connectivity analyses. Neuroimage, 110, 60–77. doi:10.1016/
j.neuroimage.2015.01.043
Chu-Shore, C. J., Kramer, M. A., Bianchi, M. T., Caviness, V. S., & Cash, S. S. (2011). Network
analysis: Applications for the developing brain. Journal of Child Neurology, 26(4), 488–500.
doi:10.1177/0883073810385345
Cohen, M. X. (2011). Error-related medial frontal theta activity predicts cingulate-related
structural connectivity. Neuroimage, 55(3), 1373–1383. doi:10.1016/j.neuroimage.2010.12.072
Cohen, M. X. & Ridderinkhof, K. R. (2013). EEG source reconstruction reveals frontal-
parietal dynamics of spatial conflict processing. PLoS One, 8(2), e57293. doi:10.1371/journal.
pone.0057293
Cuevas, K., Swingler, M. M., Bell, M. A., Marcovitch, S., & Calkins, S. D. (2012). Measures of
frontal functioning and the emergence of inhibitory control processes at 10 months of age.
Developmental and Cognitive Neuroscience, 2(2), 235–243. doi:10.1016/j.dcn.2012.01.002
Dannhauer, M., Lanfer, B., Wolters, C. H., & Knosche, T. R. (2011). Modeling of the human skull
in EEG source analysis. Human Brain Mapping, 32(9), 1383–1399. doi:10.1002/hbm.21114
Doesburg, S. M., Green, J. J., McDonald, J. J., & Ward, L. M. (2009). From local inhibition
to long- range integration: A functional dissociation of alpha- band synchronization
across cortical scales in visuospatial attention. Brain Research, 1303, 97–110. doi:10.1016/
j.brainres.2009.09.069
Doesburg, S. M., Green, J. J., McDonald, J. J., & Ward, L. M. (2012). Theta modulation of inter-
regional gamma synchronization during auditory attention control. Brain Research, 1431,
77–85. doi:10.1016/j.brainres.2011.11.005
396 WANZE XIE and JOHN E. RICHARDS
Foxe, J. J., McCourt, M. E., & Javitt, D. C. (2003). Right hemisphere control of visuospatial
attention: Line-bisection judgments evaluated with high-density electrical mapping and
source analysis. Neuroimage, 19(3), 710–726.
Gabriel, C., Gabriel, S., & Corthout, E. (1996). The dielectric properties of biological tissues.
1. Literature survey. Physics in Medicine and Biology, 41(11), 2231–2249. doi:10.1088/0031-
9155/41/11/001
Gao, C., Conte, S., Richards, J. E., Xie, W., & Hanayik, T. (2019). The neural sources of
N170: Understanding timing of activation in face-selective areas. Psychophysiology, 56(6),
e13336. doi:10.1111/psyp.13336
Grech, R., Cassar, T., Muscat, J., Camilleri, K. P., Fabri, S. G., Zervakis, M., . . . Vanrumste,
B. (2008). Review on solving the inverse problem in EEG source analysis. Journal of
Neuroengineering and Rehabilitation, 5, 25. doi:10.1186/1743-0003-5-25
Green, J. J. & McDonald, J. J. (2009). A practical guide to beamformer source reconstruc-
tion for EEG. In T. C. Handy (Ed.), Brain signal analysis: Advances in neuroelectric and
neuromagnetic methods (pp. 76–98). MIT Press
Guevara, R., Velazquez, J. L., Nenadovic, V., Wennberg, R., Senjanovic, G., & Dominguez,
L. G. (2005). Phase synchronization measurements using electroencephalographic
recordings: what can we really say about neuronal synchrony? Neuroinformatics, 3(4), 301–
314. doi:10.1385/NI:3:4:301
Guy, M. W., Zieber, N., & Richards, J. E. (2016). The cortical development of specialized face
processing in infancy. Child Development, 87(5), 1581–1600. doi:10.1111/cdev.12543
Hallez, H., Vanrumste, B., Grech, R., Muscat, J., De Clercq, W., Vergult, A., . . . Lemahieu,
I. (2007). Review on solving the forward problem in EEG source analysis. Journal of
Neuroengineering and Rehabilitation, 4, 46. doi:10.1186/1743-0003-4-46
Hämäläinen, J. A., Ortiz-Mantilla, S., & Benasich, A. A. (2011). Source localization of event-
related potentials to pitch change mapped onto age-appropriate MRIs at 6 months of age.
Neuroimage, 54(3), 1910–1918. doi:10.1016/j.neuroimage.2010.10.016
Hämäläinen, M. S. & Ilmoniemi, R. J. (1994). Interpreting magnetic fields of the brain: Minimum
norm estimates. Medical & Biological Engineering & Computing, 32(1), 35–42.
Hauk, O. (2004). Keep it simple: a case for using classical minimum norm estima-
tion in the analysis of EEG and MEG data. Neuroimage, 21(4), 1612–1621. doi:10.1016/
j.neuroimage.2003.12.018
He, B., Musha, T., Okamoto, Y., Homma, S., Nakajima, Y., & Sato, T. (1987). Electric dipole
tracing in the brain by means of the boundary element method and its accuracy. IEEE
Transactions on Biomedical Engineering, 34(6), 406–414.
He, B., Sohrabpour, A., Brown, E., & Liu, Z. (2018). Electrophysiological source imaging: A
noninvasive window to brain dynamics. Annual Review of Biomedical Engineering, 20, 171–
196. doi:10.1146/annurev-bioeng-062117-120853
Hipp, J. F., Hawellek, D. J., Corbetta, M., Siegel, M., & Engel, A. K. (2012). Large-scale cortical
correlation structure of spontaneous oscillatory activity. Nature Neuroscience, 15(6), 884–
890. doi:10.1038/nn.3101
Jatoi, M. A., Kamel, N., Malik, A. S., & Faye, I. (2014). EEG-based brain source localization
comparison of sLORETA and eLORETA. Australasian Physical & Engineering Sciences in
Medicine, 37(4), 713–721. doi:10.1007/s13246-014-0308-3
Jones, S. R., Kerr, C. E., Wan, Q., Pritchett, D. L., Hamalainen, M., & Moore, C. I. (2010).
Cued spatial attention drives functionally relevant modulation of the mu rhythm in
CORTICAL SOURCE LOCALIZATION IN EEG FREQUENCY ANALYSIS 397
Nystrom, P. (2008). The infant mirror neuron system studied with high density EEG. Social
Neuroscience, 3(3-4), 334–347. doi:10.1080/17470910701563665
Nystrom, P., Ljunghammar, T., Rosander, K., & von Hofsten, C. (2011). Using mu rhythm de-
synchronization to measure mirror neuron activity in infants. Developmental Science, 14(2),
327–335.
Odabaee, M., Tokariev, A., Layeghy, S., Mesbah, M., Colditz, P. B., Ramon, C., & Vanhatalo, S.
(2014). Neonatal EEG at scalp is focal and implies high skull conductivity in realistic neo-
natal head models. Neuroimage, 96, 73–80. doi:10.1016/j.neuroimage.2014.04.007
Onton, J., Delorme, A., & Makeig, S. (2005). Frontal midline EEG dynamics during working
memory. Neuroimage, 27(2), 341–356. doi:10.1016/j.neuroimage.2005.04.014
Oostendorp, T. F., Delbeke, J., & Stegeman, D. F. (2000). The conductivity of the human
skull: Results of in vivo and in vitro measurements. IEEE Transactions on Biomedical
Engineering, 47(11), 1487–1492. doi:10.1109/Tbme.2000.880100
Oostenveld, R., Fries, P., Maris, E., & Schoffelen, J. M. (2011). FieldTrip: Open source software
for advanced analysis of MEG, EEG, and invasive electrophysiological data. Computational
Intelligence and Neuroscience, 2011, 156869. doi:10.1155/2011/156869
Ortiz-Mantilla, S., Hamalainen, J. A., & Benasich, A. A. (2012). Time course of ERP generators
to syllables in infants: A source localization study using age-appropriate brain templates.
Neuroimage, 59(4), 3275–3287. doi:10.1016/j.neuroimage.2011.11.048
Pascual-Marqui, R. D. (2002). Standardized low-resolution brain electromagnetic tomog-
raphy (sLORETA): technical details. Methods and Findings in Experimental and Clinical
Pharmacology, 24, 5–12.
Pascual-Marqui, R. D. (2007). Discrete, 3D distributed, linear imaging methods of electric
neuronal activity. Part 1: Exact, zero error localization. https://arxiv.org/abs/0710.3341.
Pascual-Marqui, R. D., Michel, C. M., & Lehmann, D. (1994). Low-resolution electromagnetic
tomography -A new method for localizing electrical activity in the brain. International
Journal of Psychophysiology, 18, 49–65.
Phan, T. V., Smeets, D., Talcott, J. B., & Vandermosten, M. (2018). Processing of structural
neuroimaging data in young children: Bridging the gap between current practice and
state-of-the-art methods. Developmental Cognitive Neuroscience, 33, 206–223. doi:10.1016/
j.dcn.2017.08.009
Phillips, C., Rugg, M. D., & Friston, K. J. (2002a). Anatomically informed basis functions for
EEG source localization: Combining functional and anatomical constraints. Neuroimage,
16(3 Pt 1), 678–695.
Phillips, C., Rugg, M. D., & Fristont, K. J. (2002b). Systematic regularization of linear inverse
solutions of the EEG source localization problem. Neuroimage, 17(1), 287–301.
Reynolds, G. D. & Richards, J. E. (2005). Familiarization, attention, and recognition memory
in infancy: An event-related potential and cortical source localization study. Developmental
Psychology, 41(4), 598–615. doi:10.1037/0012-1649.41.4.598
Reynolds, G. D. & Richards, J. E. (2009). Cortical source localization of infant cognition.
Developmental Neuropsychology, 34(3), 312–329. doi:10.1080/87565640902801890
Richards, J. E. (2003). Cortical sources of event-related potentials in the prosaccade and
antisaccade task. Psychophysiology, 40(6), 878–894.
Richards, J. E., Sanchez, C., Phillips-Meek, M., & Xie, W. (2016). A database of age-appropriate
average MRI templates. Neuroimage, 124, 1254–1259. doi:10.1016/j.neuroimage.2015.04.055
CORTICAL SOURCE LOCALIZATION IN EEG FREQUENCY ANALYSIS 399
Richards, J. E. X. & Xie, W. (2015). Brains for all the ages: Structural neurodevelopment in
infants and children from a life-span perspective. In J. Benson (Ed.), Advances in child devel-
opment and behavior (Vol. 48, pp. 1–52). Elsevier.
Ritter, P., Moosmann, M., & Villringer, A. (2009). Rolandic alpha and beta EEG rhythms’
strengths are inversely related to fMRI-BOLD signal in primary somatosensory and motor
cortex. Human Brain Mapping, 30(4), 1168–1187. doi:10.1002/hbm.20585
Rush, S. & Driscoll, D. A. (1969). EEG electrode sensitivity -an application of reciprocity. IEEE
Transactions on Biomedical Engineering, 16(1), 15-&. doi:10.1109/Tbme.1969.4502598
Ryynanen, O. R., Hyttinen, J. A., & Malmivuo, J. A. (2006). Effect of measurement noise and
electrode density on the spatial resolution of cortical potential distribution with different re-
sistivity values for the skull. IEEE Transactions on Biomedical Engineering, 53(9), 1851–1858.
doi:10.1109/TBME.2006.873744
Sauseng, P., Hoppe, J., Klimesch, W., Gerloff, C., & Hummel, F. C. (2007). Dissociation of
sustained attention from central executive functions: local activity and interregional con-
nectivity in the theta range. European Journal of Neuroscience, 25(2), 587–593. doi:10.1111/
j.1460-9568.2006.05286.x
Scherg, M. (1992). Functional imaging and localization of electromagnetic brain activity. Brain
Topography, 5(2), 103–111.
Scherg, M., Bast, T., & Berg, P. (1999). Multiple source analysis of interictal spikes: goals,
requirements, and clinical value. Journal of Clinical Neurophysiology, 16(3), 214–224.
Schiff, S. J. (2005). Dangerous phase. Neuroinformatics, 3(4), 315–318. doi:10.1385/NI:3:4:315
Sherlin, L. (2009). Diagnosing and treating brain function through the use of low resolution
electromagnetic tomography (LORETA). In T. Budzynski, H. K. Budzynski, J. Evans, & A.
Abarbanel (Eds.), Introduction to quantitative EEG and neurofeedback: Advanced theory and
applications (2nd ed.), (pp. 83–102). Academic Press.
Stam, C. J. (2005). Nonlinear dynamical analysis of EEG and MEG: Review of an emerging
field. Clinical Neurophysiology, 116(10), 2266–2301. doi:10.1016/j.clinph.2005.06.011
Stam, C. J., Nolte, G., & Daffertshofer, A. (2007). Phase lag index: Assessment of functional
connectivity from multi channel EEG and MEG with diminished bias from common
sources. Human Brain Mapping, 28(11), 1178–1193. doi:10.1002/hbm.20346
Thorpe, S. G., Cannon, E. N., & Fox, N. A. (2016). Spectral and source structural development
of mu and alpha rhythms from infancy through adulthood. Clinical Neurophysiology, 127(1),
254–269. doi:10.1016/j.clinph.2015.03.004
Vinck, M., Oostenveld, R., van Wingerden, M., Battaglia, F., & Pennartz, C. M. (2011). An
improved index of phase-synchronization for electrophysiological data in the presence of
volume-conduction, noise and sample-size bias. Neuroimage, 55(4), 1548–1565. doi:10.1016/
j.neuroimage.2011.01.055
Vorwerk, J., Cho, J. H., Rampp, S., Hamer, H., Knosche, T. R., & Wolters, C. H. (2014). A guide-
line for head volume conductor modeling in EEG and MEG. Neuroimage, 100, 590–607.
doi:10.1016/j.neuroimage.2014.06.040
Vorwerk, J., Oostenveld, R., Piastra, M. C., Magyari, L., & Wolters, C. H. (2018). The FieldTrip-
SimBio pipeline for EEG forward solutions. Biomedical Engineering Online, 17, 37.
doi:10.1186/s12938-018-0463-y
Wendel, K., Vaisanen, J., Seemann, G., Hyttinen, J., & Malmivuo, J. (2010). The influence of
age and skull conductivity on surface and subdermal bipolar EEG leads. Computational
Intelligence and Neuroscience, 2010, 397272. doi:10.1155/2010/397272
400 WANZE XIE and JOHN E. RICHARDS
Xie, W., Jensen, K. G., Wade, M., Kumar, S., Westerlund, A., Kakon, S. H., ... Nelson, C. A.
(2019). Child growth predicts brain functional connectivity and future cognitive outcomes
in urban Bangladeshi children exposed to early adversities. BMC Medicine, 17, art. no. 199.
https://doi.org/10.1186/s12916-019-1431-5
Xie, W., Mallin, B. M., & Richards, J. E. (2017). Development of infant sustained attention and
its relation to EEG oscillations: An EEG and cortical source analysis study. Developmental
Science, 21, e12562. doi:10.1111/desc.12562
Xie, W., Mallin, B. M., & Richards, J. E. (2018). Development of brain functional connectivity
and its relation to infant sustained attention in the first year of life. Developmental Science,
22, e12703. doi:10.1111/desc.12703
Xie, W., McCormick, S. A., Westerlund, A., Bowman, L. C., & Nelson, C. A. (2018). Neural
correlates of facial emotion processing in infancy. Developmental Science, 22, e12758.
doi:10.1111/desc.12758
Xie, W., Jensen, S. K. G., Wade, M., Kumar, S., Westerlund, A., Kakon, S. H., Haque, R., Petri, W.
A., & Nelson, C. A. (2019). Growth faltering is associated with altered brain functional con-
nectivity and cognitive outcomes in urban Bangladeshi children exposed to early adversity.
BMC Med, 17(1), 199. https://doi.org/10.1186/s12916-019-1431-5
Yang, L., Liu, Z., & He, B. (2010). EEG-fMRI reciprocal functional neuroimaging. Clinical
Neurophysiology, 121(8), 1240–1250. doi:10.1016/j.clinph.2010.02.153
Yang, L., Wilke, C., Brinkmann, B., Worrell, G. A., & He, B. (2011). Dynamic imaging of
ictal oscillations using non-invasive high-resolution EEG. Neuroimage, 56(4), 1908–1917.
doi:10.1016/j.neuroimage.2011.03.043
Yeung, N., Bogacz, R., Holroyd, C. B., Nieuwenhuis, S., & Cohen, J. D. (2007). Theta phase
resetting and the error- related negativity. Psychophysiology, 44(1), 39– 49. doi:10.1111/
j.1469-8986.2006.00482.x
CHAPTER 17
FREQU E NC Y
CHARACTERISTI C S OF SL E E P
17.1 Introduction
Our life unfolds around the 24-hour clock, structured by the periodically alternating
episodes of sleep and wakefulness. This alternation is paralleled by a series of changes
in the pattern of EEG frequencies and the level of consciousness. Sleep is often regarded
as a state of inactivity and complete relaxation, securing a general brain recovery from
prior wakefulness. Nonetheless, the magnitude of variation in the pattern of EEG
frequencies in sleep are net higher compared to wakefulness, suggesting that sleep is not
a homogeneous physiological state. Indeed, sleep involves multiple transiently recurrent
stages characterized by specific oscillatory phenomena, often utterly absent from the
waking EEG (e.g. sleep spindles, slow oscillations). The rich diversity of neural rhythms
associated with total or partial lack of consciousness in non-rapid (NREM) or rapid eye
movement sleep (REM), respectively, conceives sleep as a unique time window allowing
for a deep insight into the functional neuro-architecture of the brain. As a result of the
remarkable advances in the acquisition and quantitative analysis techniques of EEG
data experienced in the last decades, as well as the continuously growing interest in sleep
research across the globe, our understanding of the nature and functional significance
of sleep dependent brain oscillatory phenomena has considerably evolved. There is a
wealth of evidence supporting a tight link between features of sleep-dependent EEG ac-
tivity and waking experience and cognitive performance. What is more, sleep EEG is
a promising source of reliable biomarkers of the disease process in neurodegenerative
disorders affecting the brain. Indeed, while sleep EEG is highly sensitive to extrinsic
factors (e.g., pharmacological agents, physical and electric stimulation), it is remarkably
modulated by intrinsic factors as well (e.g., age, sex, genetic polymorphisms, mental
health, neurological disorders). Prominent sleep dependent brain oscillations—sleep
402 ALPÁR S. LÁZÁR, ZSOLT I. LÁZÁR, AND RÓBERT BÓDIZS
slow waves and sleep spindles—are central to the cognitive and physiological function
of sleep and therefore have become the target of intervention studies aimed at boosting
sleep quality and cognitive performance by those sleep oscillations. This holds the
promise of a new line of future treatments aiming to slow down cognitive decline
associated with aging or various neurological conditions, as well as to help the rehabili-
tation process in acquired brain injury. The current chapter attempts to provide an over-
view of the most characteristic EEG rhythms and their complex coalescence/interaction
in making up human sleep. The chapter aims at summarizing the state of the art without
being oblivious of the historical groundwork.
The cyclic alternation of sleep and wakefulness follows a regular pattern governed by
several biological timekeeping systems that modulate a series of physiological processes,
including brain activity and sleep propensity, and thus govern the timing, duration, and
quality of sleep.
17.2.1 Circadian Rhythms
Circadian rhythms allow organisms to adapt more effectively to environmental changes
associated with the rotation of the earth (e.g., light, temperature, food availability, pre-
dation risk) and are generated by a molecular clockwork present across most cells of the
body. These individual clocks are synchronized by the master circadian clock located in
the suprachiasmatic nucleus (SCN) of the hypothalamus, which is entrained to the 24-h
light period through the retinohypothalamic tract (Mohawk et al., 2012). The circadian
clock has a substantial impact on the organism. Its effects range from the modulation
of gene expressions and protein synthesis (Mohawk et al., 2012) through a variety of
physiological processes, including hormonal changes, immune functions (Kizaki et al.,
2015), brain arousal and neural rhythms (Cajochen et al., 2013; Dijk & Czeisler, 1995;
Lazar et al., 2015), to direct effects on cognitive performance, including learning and
memory (Santhi et al., 2016; Schmidt et al., 2007; Smarr et al., 2014; Wright et al., 2012).
Given the diurnal nature of wake-sleep periodicity, it is not surprising that circadian
rhythms are at the heart of sleep regulation.
17.2.2 Sleep Homeostasis
The other major biological time-related system involved in sleep regulation is the so-
called sleep homeostasis or process H (originally called process S), which governs
FREQUENCY CHARACTERISTICS OF SLEEP 403
Figure 17.1 Sleep-wake cycles generated by the two-process model (Borbely, 1982).
Thick line represents homeostatic pressure, pressure building up during wake and decaying in sleep following an
exponential dynamic; thin lines represent the circadian, C(t), modulated thresholds. White areas mark wake episodes,
shaded areas mark sleep periods. Switches between wakefulness and sleep occur at intersection points between the
homeostatic pressure and the circadian, C(t), modulated thresholds (thin lines).
Used parameters: T =24 hours, α=-0.45, μ =1, a =0.1, χw =18.2 hours, χs =4.2 hours, H0–=0.17, H+0 =0.6.
the balance between the duration and intensity of wakefulness and sleep. The longer
one stays awake and the more stimulation the brain is exposed to, the greater sleep
pressure is being accumulated, as reflected by increased slow oscillatory activity in the
following sleep episode. Process H, however, does not track time awake or asleep in a
linear manner, but builds up along an asymptotic function (see the thick lines within
non-shaded regions in Figure 17.1), and following sleep onset it decreases exponen-
tially (see the thick lines within shaded regions in Figure 17.1). Sleep homeostasis is
closely linked to cellular metabolic-energetic processes as well as synaptic plasticity,
and the most commonly used EEG marker for sleep homeostasis is delta or slow wave
activity (SWA) ranging between 0.5 and 4 Hz in sleep (discussed more later). Sleep
homeostasis is also under a strong age-dependent regulation, presenting a mono-
tonically decreasing upper asymptote with aging, and is altered in depression and
neurodegenerative disorders.
17.2.4 Ultradian Cycles
Besides circadian and homeostatic regulation, sleep is cyclical in itself—periods of
NREM sleep alternate with REM along an ultradian rhythm in all mammals and pre-
sumably birds as well. In contrast with the largely universal period of the circadian
rhythm, which is clearly determined by alternating light-dark cycles on Earth, the ul-
tradian sleep cycles seem to be allometrically (i.e., in a body size-dependent manner)
regulated and are mainly a function of the metabolism ratio of the species. The average
(modal) duration of a complete NREM-REM cycle is around 90 minutes in adult
human subjects repeating four to six times over a typical full night sleep, but the dur-
ation of a complete cycle in laboratory rats is around 11 minutes (Kaiser, 2013; Stupfel &
Pavely, 1990; Savage & West, 2007). The time structure of the consecutive sleep cycles
is complex and non-random. Cyclic variations of various forms of sleep EEG micro-
structural phenomena, usually associated with autonomic activations, emerge with
infraslow periodicities (Halász & Bódizs, 2013). At a finer time scale, there are several
EEG frequencies that are more-or-less specific for either the NREM or the REM phase of
sleep. According to American Academy of Sleep Medicine (AASM) we can distinguish
non-rapid eye movement sleep (N or NREM) that consists of three stages: N1 (NREM
1 or NREM stage 1), N2 (NREM 2 or NREM stage 2), and N3 (slow wave sleep, SWS), as
well as rapid eye movement sleep (REM). As soon as we initiate sleep, the EEG, elec-
tromyographic (EMG), and electro-oculographic (EOG) activity goes through marked
changes compared to wakefulness (see Table 17.1). The transition from lighter (N1 or
NREM 1) to slow wave sleep (SWS) is associated with gradually decreasing muscle ac-
tivity, heart and breathing rate, and body temperature. Moreover, EEG becomes grad-
ually dominated by slower and slower frequency oscillations (in NREM 1 by theta, in
NREM 2 by mixed frequencies including K-complexes and sleep spindles) until the en-
tire brain will present continuous slow oscillations and delta activity (slow wave activity,
SWA) in SWS. However, during REM, also called paradoxical sleep, the brain’s electric
and metabolic activity presents a significant increase as reflected in higher frequency,
low amplitude mixed EEG frequencies resembling waking brain activity. REM is also
accompanied by bursts of rapid eye movements, increased heart and breathing rates
paralleled by muscle atonia. This is the stage of sleep when most dreams are reported.
The scoring of sleep stages is based on conventions and, beyond characteristic EEG
features, it requires other polygraphic readouts such as electrooculography (EOG) and
electromyography (EMG) (Table 17.1 and Figure 17.2).
17.2.5 Infradian Rhythms
The circadian rhythm is far from being the slowest envelope of sleep rhythmicity. There
are rhythms with periods significantly longer than 24 hours (infradian). Several studies
suggest that circatrigintan (~30 days) periodicity in coherence with menstrual cycles
FREQUENCY CHARACTERISTICS OF SLEEP 405
N1 (NREM 1) Low-amplitude, mixed frequency The chin EMG amplitude Slow-rolling eye
activity in the range 4–7 Hz is variable, often lower movements (conjugate,
and attenuated alpha activity than during wake. reasonably regular,
(<50% of the scoring epoch). sinusoidal eye
Vertex sharp waves. movements with an
initial deflection that
usually lasts >500 ms).
N2 (NREM 2) The presence of one or more K- The chin EMG is of Usually, eye movement
complexes unassociated with variable amplitude, but is not noticed during
arousal. usually lower than in N2.
One or more sleep spindles. wake.
N3 (SWS) SWS =slow wave sleep. The chin EMG is of Usually, eye movements
Dominant slow wave activity: EEG variable amplitude, are not noticed during
waves of 0.5–2 Hz and >75 often lower than in N3.
μV peak-to-peak amplitude, stage N2, sometimes as
measured over the frontal low as in REM sleep.
regions, referenced to the
contralateral ear or mastoid
(F4–M1, F3–M2).
≥20% of a staging epoch consists
of slow wave activity.
REM Low-amplitude, mixed-frequency Low-amplitude chin Conjugated,
EEG activity without K- EMG—baseline EMG episodic, sharply
complexes or sleep spindles. activity, usually is at peaked, irregular eye
Sawtooth waves may be present the lowest level of all movements.
as well. recording.
of females is prevalent and well measurable in human sleep EEG (Baker & Lee, 2018).
Interestingly, a circatrigintan (circalunar) periodicity (~29.5 days) of sleep-dependent
EEG activity has also been reported (Cajochen et al., 2013). According to this study, the
lunar cycle had a significant impact on the frequencies between 0.5 and 16 Hz. More
specifically, around the full moon, delta activity during NREM sleep decreased by 30%,
while spindles also exhibited a marked reduction compared to the days around the new
moon. This is intriguing given the role of delta waves and sleep spindles in the cognitive
and physiological function of sleep (discussed later). Although there are other studies
confirming an effect of lunar cycle on sleep duration and quality (Roosli et al., 2006;
406 ALPÁR S. LÁZÁR, ZSOLT I. LÁZÁR, AND RÓBERT BÓDIZS
Turanyi et al., 2014), these findings are still controversial due to multiple subsequent
studies failing to reproduce similar results (Cordi et al., 2014).
Details on these periodicities, rhythms, and regulatory mechanisms are discussed
across the rest of the chapter. Table 17.2 provides an overview of the relevant frequencies.
(continued)
408 ALPÁR S. LÁZÁR, ZSOLT I. LÁZÁR, AND RÓBERT BÓDIZS
FFT, Fast Fourier Transformation; PPA, Period Amplitude Analysis; *Evident allometry, human data
provided
Before providing a detailed analysis of specific sleep EEG oscillations, we present some
core, defining features of the global picture of sleep EEG, like overall EEG amplitude,
the amplitude vs. frequency relationship, power law scaling behavior of EEG spectra,
H-exponent, and fractality.
EEG amplitude reaches its maxima during SWS (N3 or NREM stages 3&4 according
to the Rechtschaffen and Kales criteria) in physiological circumstances (Agnew et al.,
1967). Sleep EEG total power values and/or amplitudes are characterized by age-related
decreases from childhood through the older ages (Astrom & Trojaborg, 1992). Moreover,
women are characterized by higher overall sleep EEG power values as compared to
men (Carrier et al., 2001; Bódizs et al., 2021). The age-related decrease in EEG ampli-
tude power in children/adolescents may reflect a combination of non-neuronal factors,
like the increase in skull thickness, and neural maturation (decrease in overall synaptic
density). Women vs. men differences in overall sleep EEG amplitude are thought to
mainly reflect differences in skull thickness (Dijk et al., 1989).
EEG rhythms with lower frequency predominate in NREM sleep Stage 2 and SWS as
compared to REM sleep and wakefulness. The logarithm of sleep EEG amplitude has
been shown to be a linear function (negative correlate) of the logarithm of frequency in
NREM sleep of human volunteers (Feinberg et al., 1984). The steepness of the negative
relationship between log amplitude and log frequency is higher in young as compared to
elder subjects. The power law scaling behavior of the sleep EEG spectra is an expression
FREQUENCY CHARACTERISTICS OF SLEEP 409
of the same phenomenon. EEG spectra can in fact be well approximated by a power law
decay function:
Pf = C × f α
Peak power (PPeak) at frequency f equals 1 if there is no peak and is larger than 1 if there
is a spectral peak at that frequency. It has to be noted, that PPeak(f) is a whitened power
measure, because it is characterized by roughly equal power along the frequency axis and
is thus statistically independent from the spectral slope (α), which constitutes the colored
noise part of the spectrum, characterized by a power-law type decrease along the fre-
quency scale. PPeak(f) is also normalized in terms of amplitude differences as the term C
is also partialed out from this function. By using these approaches, 191 spectral measures
were effectively reduced to 4, which were efficient in characterizing known age-effects, sex-
differences, and cognitive correlates of NREM sleep EEG (Figure 17.3; Bódizs et al., 2021).
Power law dependence is the hallmark of scale-free phenomena, such as fractals.
Global sleep EEG features can be characterized by concepts related to self-similarity
and quantified by a number of closely related measures (Ma et al., 2018; Stam, 2005).
These include the correlation dimension (Grassberger & Procaccia, 1983), the scaling
410 ALPÁR S. LÁZÁR, ZSOLT I. LÁZÁR, AND RÓBERT BÓDIZS
WAKE
REM
NREM1
NREM2
SWS
25 F3
Frequency (Hz) Frequency (Hz)
20
15
10
5
0
25 P3
20
15
10
5
0
0 1 2 3 4 5 6 7
Sleep time (hours)
Figure 17.3 Evolution of EEG power spectral density during one night’s sleep of a young adult
measured over the left frontal (F3) and parietal (P3) brain regions. Warmer colors represent
higher power (logarithmic scale). Temporal resolution is 10 seconds/pixel, each representing
the average of the short-time Fourier transforms of four segments of 4-s long, Hanning tapered
windows with 50% overlap.
exponent emerging from detrended fluctuation analysis (Gu & Zhou, 2006), and the
Hurst exponent, H, which quantifies the smoothness of the signal (Hurst et al., 1965).
That is, a higher H, 0.5 < H < 1, indicates that the changes in the EEG voltage tend to
“go in the same direction”, or maintain the direction of the deflection for longer periods
(the signal is relatively smooth). In this case, there are positive correlations between the
successive increments in voltage. If 0 < H <0.5, the overall correlation between successive
increments is negative. The H-exponent is closely related to the exponent, α, of the
power spectra as follows: α = 2H + 1. Moreover, the H-exponent correlates with band-
limited relative spectral power in a sleep stage-and EEG derivation-specific manner.
The H-exponent of human sleep EEG is regularly larger than 0.5, which means that the
sleep EEG (both NREM and REM) is characterized by long-term positive autocorrel-
ation. As expected, the H values have been reported higher during the deepest parts of
SWS compared to NREM2 and REM across all EEG derivations (Weiss et al., 2009).
In summary, sleep EEG spectra follows a power law function that is the state-specific
features are best described by multiple frequency bands and their ratio, rather than
single band measures. In addition, the sleep EEG has fractal structure which should be
considered when characterizing different sleep-waking states in terms of EEG criteria.
electrodes without Au coating and specific EEG gels are required in order to appropri-
ately detect these oscillations (Vanhatalo et al., 2005). Another class of periodic phe-
nomena in the ISO frequency range contribute multiplicatively to the signal, manifesting
as modulations of the higher frequency components (Parrino et al., 2014, Lecci et al.,
2017, Lazar et al.,2019). Evidence suggests that ISO modulates >0.5 Hz EEG activity, as
well as interictal epileptic activity during sleep (Vanhatalo et al., 2004). Long lasting (1–5
s) occipital negative transients of 200–700 µV are seen in the DC-EEG records of pre-
term infants, which is probably an early form of ISO during sleep. Delta bursts have
been seen to be embedded in these transients (Vanhatalo et al., 2002). The power of ISO
correlates positively with the fMRI BOLD signal intensity of several subcortical regions,
including the cerebellum, thalamus, and the basal ganglia. In contrast, paramedian
heteromodal cortical BOLD signal intensities correlate negatively with ISO during sleep
(Picchioni et al., 2011). The ISO is a global modulation of neural excitability. The gener-
ation of ISO is hypothesized to reflect the long lasting hyperpolarizations of thalamo-
cortical cells, which might be due to the opening of some type of inwardly rectifying K
+ channels. The latter are hypothesized to be opened by the activation of adenosine A1
receptors, while the endogenous ligand, adenosine, results from the degradation of ad-
enosine triphosphate (ATP) after its release from glial cells (Hughes et al., 2011).
sleep EEG slow wave activity (0.5–4.5 Hz) was shown to be characterized by a 20–30 s
periodicity in human volunteers (Achermann & Borbely, 1997), which might indicate
a CAP-like fluctuation of slow wave amplitude (involving A1 and/or A2 phases). Sleep
spindle activity (see below) was shown to be organized in rhythmic wave sequences
according to multiple frequencies. One of these frequencies falls in the ISO range
and implies a sleep spindle periodicity of 50–100 s (0.01–0.02 Hz) in both mice and
humans (Lazar et al., 2019; Lecci et al., 2017). Using the spectrum of spectra approach
(calculating the power of the multivariate time series obtained from the short-time
Fourier transform for the signal), it was shown that the ISO (0.01–0.02 Hz) of the sigma
(10–15 Hz) power (frequency range corresponding to sleep spindles) is coordinated
with cardiac oscillations and composed of two phases: one with enhanced and one with
decreased environmental alertness (Lecci et al., 2017). Moreover, the intensity of ISO
in the sigma power predominate in NREM 2 and over the posterior regions in humans,
and is positively associated with declarative memory recall performance following sleep
(Lecci et al., 2017). Recently, ISO in sleep spindle activity were also quantified. The novel
method computes the infrapower of an on/off binary square signal corresponding with
individually detected sleep spindles. This infrapower is contrasted with that from a com-
putation where the same sleep spindles and inter-spindle lapses are randomly reshuffled
so that large-scale temporal structures are suppressed. The method validated using
surrogate data strictly addresses the large-scale temporal structures in the sequence of
spindle events, does not manifest any obvious biases and is extensible to other phasic
phenomena such as slow waves. The study showed that ISO were strongest in the fast
sigma band (13–15 Hz). Further, a robust infra peak of individually detected fast sleep
spindles was identified. This intensity and frequency of this peak were modulated by
topography, sleep stage, and sleep history. The power of ISO in fast sleep spindles is most
prominent in the centro-parieto-occipital brain regions, left hemisphere, and second
half of the night, independent of the number of sleep spindles. The frequency of ISO
is higher over the frontal and temporal brain regions and in the first half of the night
(Lazar et al., 2019).
Slow EEG waves are among the most well established neurophysiological features
of sleep (Figure 17.2). Their presence was apparent in the very first all-night sleep
EEG records (Loomis et al., 1937). The most prominent type of slow wave activity in
the sleep EEG is the slow oscillation (SO; 0.1–1 Hz). The first comprehensive descrip-
tion and experimental analysis of the sleep SO was given by Mircea Steriade and his
colleagues (Steriade, Contreras et al., 1993; Steriade et al., 1993a, 1993b). Cycles of the SO
are characterized by a more-or-less regular sequence of neuronal silence/disfacilitation,
characterized by a surface negative deflection (down state) and a burst-like neuronal
firing, taking the form of a surface positive component (up state). Down states are
FREQUENCY CHARACTERISTICS OF SLEEP 413
Delta waves (activity) are medium-sized EEG waves of 1–4 Hz, prevailing in the NREM
phase of sleep (maximal delta is present in SWS). The term delta activity (the spectral
414 ALPÁR S. LÁZÁR, ZSOLT I. LÁZÁR, AND RÓBERT BÓDIZS
Perhaps one of the most persistent focuses of electrophysiological research is the phe-
nomenon of hippocampal rhythmic slow activity (RSA), or the so-called hippocampal
theta rhythms during exploratory behavior and REM sleep of different mammalian
species. However, the term theta is confusing as there is evidence for clear allometry
in the frequency of hippocampal RSA. Animals with larger brains are characterized
by lower frequency hippocampal RSA (Blumberg, 1989), which means that human
hippocampal RSA measured in REM sleep falls in the delta (1.5–3 Hz), but not in the
theta, frequency range (Bodizs et al., 2001). The confusion in the literature is further
increased by the fact that there are indeed clear theta frequency oscillations during
the REM phases of sleep in humans, but these oscillations are of frontomedial/an-
terior cingular origin (frontal midline theta), and not hippocampal, as it was implicitly
assumed.
1998). There is preliminary evidence for the anterior cingulate origin of the frontally
recorded theta rhythm in REM sleep (Nishida et al., 2004; Uchida et al., 2003). It has
been shown that REM sleep EEG frontal theta activity is involved in emotional memory
consolidation (Goldstein & Walker, 2014) and the incorporation of recent waking-life
experiences in dreams (Eichenlaub et al., 2018). In addition, resilience against PTSD is
indexed by higher REM sleep EEG theta activity (Cowdin et al., 2014). More recently a
decrease in REM-dependent theta activity has been identified as a marker of looming
neurodegeneration in pre-manifest and manifest patients carrying the mutant gene for
Huntington’s disease (Lazar, Panin, et al., 2015; Piano et al., 2017).
Sawtooth waves are 1.5–5 Hz (mean frequency: 2.5 Hz) notched triangular waves that
occur in bursts with a mean duration of 7 s (range 2–26 s) in the frontocentral region
of REM sleep EEG records (Figure 17.4). The bursts have the appearance of teeth on
a saw (Pearl et al., 2002). Evidence suggests the lower level of sawtooth wave activity
characterizes the first REM sleep episode (Pearl et al., 2002), as well as the early emer-
gence of sawtooth waves during the NREM-REM transition (Sato et al., 1997).
When compared to presleep quiet wakefulness, the level of EEG alpha activity is
attenuated during sleep. Various frequency limits for defining sleep EEG alpha ac-
tivity are seen in the literature. The major issue is the differentiation of alpha from
sleep spindles (Section 17.11). Realistic frequency boundaries in healthy adults are 7–10,
7–11, or 8–11 Hz. However, sleep spindles (as well as waking alpha) are slower in pre-
pubertal children, thus the above criteria cannot be automatically transferred to all ages.
Individual alpha peak frequency in the presleep period is lower than the spindle peak
frequencies; thus presleep alpha peaks should be used as individual anchors for a delib-
erate and individualized frequency determination.
Continuous or intermittent alpha waves are most frequently considered as signs of
(micro)arousals. Transient increases in alpha power are parts of the A2 and A3 types
of microarousals embedded in CAP sequences, which are characterized by mixed low
and high frequency and exclusively high frequency EEG components, respectively,
indicating neural activation (Parrino et al., 2014). Moreover, ongoing posterior alpha ac-
tivity can be considered an index of covert waking brain activity during sleep, indicating
shallower/more perturbated sleep and higher reactivity to external stimuli (McKinney
et al., 2011). A peculiar mix of alpha activity with sleep EEG delta waves is termed alpha-
delta sleep. Alpha-delta sleep consists of epochs made up by delta wave-dominated
FREQUENCY CHARACTERISTICS OF SLEEP 417
F3
NREM 1
C3
P3
O1
F3
C3
NREM 2
P3
O1
Sawtooth waves
F3
C3
REM
P3
20 μV
O1
1 sec
Figure 17.4 EEG samples with characteristic EEG oscillations including Vertex sharp
transients in NREM 1, positive occipital sharp transients (POST) in NREM 2 and Sawtooth waves
in REM sleep.
segments (5–20%) mixed and filled up with large amplitude alpha waves (Hauri &
Hawkins, 1973). Fibromyalgia syndrome (Branco et al., 1994), major depressive disorder
(Jaimchariyatam et al., 2011) and psychological features, like insecure attachment (Sloan
et al., 2007), are associated with more alpha-delta waves during sleep.
418 ALPÁR S. LÁZÁR, ZSOLT I. LÁZÁR, AND RÓBERT BÓDIZS
The issue of alpha activity in REM sleep is somewhat more complex, however.
Although, posterior alpha is a reliable measure of instantaneous sleep depth and re-
sponsiveness in REM sleep (McKinney et al., 2011), alpha activity is higher and more
synchronized in tonic (no actual rapid eye movements), as compared to phasic (actual
bursts of rapid eye movements) REM episodes (Simor et al., 2018; Simor et al., 2016). In
addition, there is evidence of increased high alpha frequencies (10–14 Hz, which are not
sleep spindles) in the REM sleep records of subjects with frequent nightmares (Simor
et al., 2013).
Sharp negative waves with a maximum at the vertex are frequently seen during the ini-
tiation of sleep, most frequently in NREM Stage 1 sleep (Figure 17.4). Although these
transients, called vertex sharp waves, resemble some interictal epileptic discharges,
these sharp waves are physiological and do not indicate the presence of any epileptic
processes, even if a complete lateralization of the phenomenon is seen (Brenton &
Mytinger, 2015). Vertex sharp waves and N300 components of averaged EEG responses
to stimuli with different modalities have been shown to have common sources (Colrain
et al., 2000).
Positive occipital sharp transients of sleep are also parts of the EEG picture of
shallower phases of sleep, however, unlike vertex sharp waves, they are often seen in
NREM Stage 2 sleep as well (Figure 17.4). The first thirty minutes of NREM sleep is the
period of preferred emergence of positive occipital sharp transients (Egawa et al., 1983).
The occipital maximum is hypothesized to be related to some activations in the visual
areas of the brain.
Putative sleep spindle-like waveforms were first described by Hans Berger, the founder
of human EEG, as oscillatory episodes with a 72-ms wavelength (equaling 13.89 Hz)
in a female schizophrenic patient in (rectally induced) tribromoethanol narcosis
(Berger, 1933). Soon thereafter, Alfred Lee Loomis described and named the sleep
spindles as being very regular 14-Hz bursts, lasting 1–1.5 seconds, adding that the amp-
litude increases and decreases steadily before and after reaching its maximum, respect-
ively (Loomis et al., 1935) (Figure 17.2). In the next decades, sleep spindles were mainly
considered as hallmarks of Stage 2 sleep, but experimental studies clearly indicated
FREQUENCY CHARACTERISTICS OF SLEEP 419
constituents of sleep spindles (single cycles) is most often tacitly considered as quasi-
sinusoidal. However, there are important specifications in this regard: sleep spindles
have been shown to reflect both subthreshold depolarizations of apical dendrites
(smooth surface negative waves) and apical dendritic depolarization (sharp negative-
positive waves) (Urakami et al., 2012). Sleep spindles with unipolar sharp peaks have
been registered near the parahippocampal region in epilepsy patients undergoing
presurgical examination with foramen ovale electrodes. Sharp peaks have been
associated with nested (phase-coupled) 80–140 Hz ripples in these patients (Clemens
et al., 2011). Sleep spindles with unusual sharp negative peaks have also been observed in
infancy, up to 2 years of age (Fois, 1961). However, there is no direct evidence indicating
that the above findings reflect enhanced neuronal firing during parahippocampal and
cortical sleep spindles in epilepsy patients and infants, respectively.
theory posits that sleep slow oscillations preferentially consolidate the stronger memory
traces, which also leads to the extinction of weak memories. Sleep spindles are thought
to contribute to a slow but reliable consolidation of the multiple competing memories
(Wei et al., 2018).
Historically, beta waves have been defined as EEG frequencies above 13 Hz (Kozelka &
Pedley, 1990). However, it is now clear that sleep spindles constitute a specific phenom-
enon, thus spindles and beta oscillations have to be defined separately in spite of the
clear overlap in the frequency of these phenomena. Consequently, frequencies higher
than sleep spindles, but lower than gamma (16–30 Hz), can be considered as beta ac-
tivity in NREM sleep, whereas a somewhat broader frequency is acceptable in states
without sleep spindling, like REM sleep and wakefulness (13–30 Hz). NREM sleep beta
EEG activity with frontocentral maxima is an integral part of sleep initiation, especially
in children (Kellaway & Fox, 1952). Moreover, the up states of the SO are associated with
an increase in beta/gamma activity (up to 80 Hz), indicating the entrainment of higher
frequency oscillations by SO commonly referred to as grouping effect (Neske, 2016). SO/
delta and beta activity fluctuate reciprocally across NREM and REM sleep indicating
that an inverse relationship between the two activities (Uchida et al., 1992). Beta activity
is evidently higher in REM as compared to NREM sleep. The anterior cingulate and the
dorsolateral prefrontal cortices are characterized by prominent and coherent beta ac-
tivity during REM sleep in humans (Vijayan et al., 2017).
Several reports suggest that enhanced beta/gamma EEG power (activity) is an index
of cortical arousal, or disturbed sleep. Most of the reports indicate that sleep EEG (both
NREM and REM) records of insomnia patients are characterized by heightened beta/
gamma power as compared to records of subjects who sleep well (Perlis, Merica et al.,
2001; Perlis, Smith et al., 2001).
Gamma oscillations are usually defined as 30–80 Hz waves indicating arousal and
(local) activation of the cortex. If gamma oscillations are coherent, they are related to
perceptual binding. As mentioned previously, there are several reports indicating
similar behavior of beta and gamma oscillations during sleep (see also those indicating
enhanced gamma in patients suffering from primary and secondary insomnia: Perlis,
Merica, et al., 2001; Perlis, Smith, et al., 2001). The specificity of gamma waves is seen
in cases in which consciousness is part of the story. Thus, frontal gamma EEG power
and coherence is increased in periods of lucid dreaming, as compared to non-lucid
dreaming (dream lucidity indicates the state in which the dreamer is aware of the
dream-like nature of the ongoing mental activity; see Voss et al., 2009). In addition, beta
and gamma activity are differentially expressed in tonic and phasic REM sleep: tonic
424 ALPÁR S. LÁZÁR, ZSOLT I. LÁZÁR, AND RÓBERT BÓDIZS
REM is characterized by increased beta, whereas phasic REM by increased gamma ac-
tivity (Simor et al., 2016).
Ripples are usually recorded in the invasive epilepsy monitoring setting. Slow (80–140
Hz) and fast (250–500 Hz) ripples are considered physiological and pathological (epi-
leptic enhancement) signals, respectively. The hippocampal formation and the medial
temporal lobe is the most frequent, but not exclusive origin of ripple oscillations in
all wake-sleep states. It has been shown that hippocampal ripples are nested in spe-
cific phases of sleep spindle oscillations, resulting in a fine-tuned coupling of the two
oscillations in human subjects (Clemens et al., 2011; Staresina et al., 2015). This hierarch-
ical nesting (SO-sleep spindle-ripple) is thought to play a crucial role in off-line memory
consolidation processes involving the reactivation of transient memory traces during
sleep. Recent investigations indicate that ripples can be recorded in the scalp EEG of
children as well. Mean frequency has been reported to be 102 Hz, mean duration 70
ms, and mean root mean square amplitude 0.95 µV (Mooij et al., 2017). Scalp-recorded
ripples have been associated with several sleep EEG transients (Mooij et al., 2018).
17.14 Conclusion
REFERENCES
Achermann, P. & Borbely, A. A. (1997). Low-frequency (< 1 Hz) oscillations in the human sleep
electroencephalogram. Neuroscience, 81(1), 213–222.
Agnew, H. W., Jr., Parker, J. C., Webb, W. B., & Williams, R. L. (1967). Amplitude measurement
of the sleep electroencephalogram. Electroencephalography and Clinical Neurophysiology,
22(1), 84–86.
FREQUENCY CHARACTERISTICS OF SLEEP 425
Destexhe, A., Contreras, D., & Steriade, M. (1998). Mechanisms underlying the synchronizing
action of corticothalamic feedback through inhibition of thalamic relay cells. Journal of
Neurophysiology, 79(2), 999–1016.
De Gennaro, L., Marzano, C., Fratello, F., Moroni, F., Pellicciari, M. C., Ferlazzo, F., . . . Rossini,
P. M. (2008). The electroencephalographic fingerprint of sleep is genetically determined: A
twin study. Annals of Neurology, 64(4), 455–460.
de Zambotti, M., Willoughby, A. R., Sassoon, S. A., Colrain, I. M., & Baker, F. C. (2015).
Menstrual cycle-related variation in physiological sleep in women in the early menopausal
transition. The Journal of Clinical Endocrinology and Metabolism, 100(8), 2918–2926.
Dijk, D. J., Beersma, D. G., & Bloem, G. M. (1989). Sex differences in the sleep EEG of young
adults: Visual scoring and spectral analysis. Sleep, 12(6), 500–507.
Dijk, D. J. & Czeisler, C. A. (1995). Contribution of the circadian pacemaker and the sleep
homeostat to sleep propensity, sleep structure, electroencephalographic slow waves, and
sleep spindle activity in humans. Journal of Neuroscience, 15(5 Pt 1), 3526–3538.
Dijk, D. J. (1999). Circadian variation of EEG power spectra in NREM and REM sleep in
humans: dissociation from body temperature. Journal of Sleep Research, 8(3), 189–195
Egawa, I., Yoshino, K., & Hishikawa, Y. (1983). Positive occipital sharp transients in the human
sleep EEG. Psychiatry and Clinical Neuroscience, 37(1), 57–65.
Eichenlaub, J. B., van Rijn, E., Gaskell, M. G., Lewis, P. A., Maby, E., Malinowski, J. E., . . . Blagrove,
M. (2018). Incorporation of recent waking-life experiences in dreams correlates with frontal
theta activity in REM sleep. Social Cognitive and Affective Neuroscience, 13(6), 637–647.
Evans, B. M. & Richardson, N. E. (1995). Demonstration pf a 3–5-s periodicity between the
spindle bursts in NREM sleep in man. Journal of Sleep Research, 4(3), 196–197.
Feinberg, I., March, J. D., Fein, G., & Aminoff, M. J. (1984). Log amplitude is a linear
function of log frequency in NREM sleep EEG of young and elderly normal subjects.
Electroencephalography and Clinical Neurophysiology, 58(2), 158–160.
Fell, J., Elfadil, H., Roschke, J., Burr, W., Klaver, P., Elger, C. E., & Fernandez, G. (2002). Human
scalp recorded sigma activity is modulated by slow EEG oscillations during deep sleep.
International Journal of Neuroscience, 112(7), 893–900.
Fois, A. (1961). The electroencephalogram of the normal child. Charles C Thomas.
Frank, M. G. & Cantera, R. (2014). Sleep, clocks, and synaptic plasticity. Trends in Neuroscience,
37(9), 491–501.
Freeman, W. J. & Zhai, J. (2009). Simulated power spectral density (PSD) of background elec-
trocorticogram (ECoG). Cognitive Neurodynamics, 3(1), 97–103.
Geiger, A., Huber, R., Kurth, S., Ringli, M., Jenni, O. G., & Achermann, P. (2011). The sleep EEG
as a marker of intellectual ability in school age children. Sleep, 34(2), 181–189.
Gibbs, E. A. & Gibbs, E. L. (1958). Atlas of electroencephalography (2nd ed.). Addison Wesley.
A glossary of terms most commonly used by clinical electroencephalographers. (1974).
Electroencephalography and Clinical Neurophysiology, 37(5), 538–548.
Godbout, R., Bergeron, C., Limoges, E., Stip, E., & Mottron, L. (2000). A laboratory study of
sleep in Asperger’s syndrome. Neuroreport, 11(1), 127–130.
Goldstein, A. N. & Walker, M. P. (2014). The role of sleep in emotional brain function. Annual
Review of Clinical Psychology, 10, 679–708.
Grassberger, P. & Procaccia, I. (1983). Characterization of strange attractors. Physical Review
Letters, 50(5), 346–349.
Gu, G.-F., & Zhou, W.-X. (2006). Detrended fluctuation analysis for fractals and multifractals
in higher dimensions. Physical Review E, 74(6), 061104.
428 ALPÁR S. LÁZÁR, ZSOLT I. LÁZÁR, AND RÓBERT BÓDIZS
Halasz, P. (2005). K-complex, a reactive EEG graphoelement of NREM sleep: An old chap in a
new garment. Sleep Med Rev, 9(5), 391–412.
Halász, P. & Bódizs, R. (2013). Dynamic structure of NREM sleep. Springer.
Hauri, P. & Hawkins, D. R. (1973). Alpha-delta sleep. Electroencephalography and Clinical
Neurophysiology, 34(3), 233–237.
Hoedlmoser, K., Heib, D. P., Roell, J., Peigneux, P., Sadeh, A., Gruber, G., & Schabus, M. (2014).
Slow sleep spindle activity, declarative memory, and general cognitive abilities in children.
Sleep, 37(9), 1501–1512.
Hori, T., Sugita, Y., Koga, E., Shirakawa, S., Inoue, K., Uchida, S., . . . Fukuda, N. (2001).
Proposed supplements and amendments to ‘A manual of standardized terminology,
techniques and scoring system for sleep stages of human subjects’, the Rechtschaffen & Kales
(1968) standard. Psychiatry and Clinical Neuroscience, 55(3), 305–310.
Huber, R., Esser, S. K., Ferrarelli, F., Massimini, M., Peterson, M. J., & Tononi, G. (2007). TMS-
induced cortical potentiation during wakefulness locally increases slow wave activity during
sleep. PLoS One, 2(3), e276.
Huber, R., Ghilardi, M. F., Massimini, M., Ferrarelli, F., Riedner, B. A., Peterson, M. J., &
Tononi, G. (2006). Arm immobilization causes cortical plastic changes and locally decreases
sleep slow wave activity. Nature Neuroscience, 9(9), 1169–1176.
Huber, R., Ghilardi, M. F., Massimini, M., & Tononi, G. (2004). Local sleep and learning.
Nature, 430(6995), 78–81.
Hughes, S. W., Lorincz, M. L., Parri, H. R., & Crunelli, V. (2011). Infraslow (<0.1 Hz) oscillations
in thalamic relay nuclei basic mechanisms and significance to health and disease states.
Progress in Brain Research, 193, 145–162.
Huguenard, J. R. & McCormick, D. A. (2007). Thalamic synchrony and dynamic regulation of
global forebrain oscillations. Trends in Neuroscience, 30(7), 350–356.
Hurst, H. E., Black, R. P., & Simaika, Y. M. (1965). Long-term storage, an experimental study.
Constable.
Iber, C. & American Academy of Sleep Medicine. (2007). The AASM manual for the scoring
of sleep and associated events: Rules, terminology and technical specifications. American
Academy of Sleep Medicine.
Inanaga, K. (1998). Frontal midline theta rhythm and mental activity. Psychiatry and Clinical
Neuroscience, 52(6), 555–566.
Jaimchariyatam, N., Rodriguez, C. L., & Budur, K. (2011). Prevalence and correlates of alpha-
delta sleep in major depressive disorders. Innovations in Clinical Neuroscience, 8(7), 35–49.
Jankel, W. R. & Niedermeyer, E. (1985). Sleep spindles. Journal of Clinical Neurophysiology,
2(1), 1–35.
Kaiser, D. (2013). Infralow frequencies and ultradian rhythms. Seminars in Pediatric Neurology,
20(4), 242–245.
Kane, N., Acharya, J., Benickzy, S., Caboclo, L., Finnigan, S., Kaplan, P. W., . . . van Putten,
M. J. A. M. (2017). A revised glossary of terms most commonly used by clinical
electroencephalographers and updated proposal for the report format of the EEG findings.
Revision 2017. Clinical Neurophysiology Practice, 2, 170–185.
Kattler, H., Dijk, D. J., & Borbely, A. A. (1994). Effect of unilateral somatosensory stimulation
prior to sleep on the sleep EEG in humans. Journal of Sleep Research, 3(3), 159–164.
Kellaway, P. & Fox, B. J. (1952). Electroencephalographic diagnosis of cerebral pathology
in infants during sleep: I. Rationale, technique, and the characteristics of normal sleep in
infants. The Journal of Pediatrics, 41(3), 262–287.
FREQUENCY CHARACTERISTICS OF SLEEP 429
Kizaki, T., Sato, S., Shirato, K., Sakurai, T., Ogasawara, J., Izawa, T., . . . Ohno, H. (2015). Effect of
circadian rhythm on clinical and pathophysiological conditions and inflammation. Critical
Reviews in Immunology, 35(4), 261–275.
Knoblauch, V., Martens, W., Wirz-Justice, A., Krauchi, K., & Cajochen, C. (2003). Regional
differences in the circadian modulation of human sleep spindle characteristics. European
Journal of Neuroscience, 18(1), 155–163.
Knoblauch, V., Martens, W. L., Wirz-Justice, A., & Cajochen, C. (2003). Human sleep spindle
characteristics after sleep deprivation. Clinical Neurophysiology, 114(12), 2258–2267.
Knoblauch, V., Munch, M., Blatter, K., Martens, W. L., Schroder, C., Schnitzler, C., . . . Cajochen,
C. (2005). Age-related changes in the circadian modulation of sleep-spindle frequency
during nap sleep. Sleep, 28(9), 1093–1101.
Kozelka, J. W. & Pedley, T. A. (1990). Beta and mu rhythms. Journal of Clinical Neurophysiology,
7(2), 191–207.
Kubicki, S., Meyer, C., & Rohmel, J. (1986). [The 4 second sleep spindle periodicity]. EEG-
EMG Zeitschrift fur Elektroenzephalogrphie, Elektromyographie und Verwandte Gebiete,
17(2), 55–61.
Latreille, V., Carrier, J., Lafortune, M., Postuma, R. B., Bertrand, J. A., Panisset, M., . . . Gagnon,
J. F. (2015). Sleep spindles in Parkinson’s disease may predict the development of dementia.
Neurobiology of Aging, 36(2), 1083–1090.
Lazar, A. S., Lazar, Z. I., & Dijk, D. J. (2015). Circadian regulation of slow waves in human
sleep: Topographical aspects. Neuroimage, 116, 123–134.
Lazar, A. S., Panin, F., Goodman, A. O., Lazic, S. E., Lazar, Z. I., Mason, S. L., . . . Barker, R. A.
(2015). Sleep deficits but no metabolic deficits in premanifest Huntington’s disease. Annals of
Neurology, 78(4), 630–648.
Lázár, Z. I., Dijk, D. J., & Lázár, A. S. (2019). Infraslow oscillations in human sleep spindle ac-
tivity. Journal of Neuroscience Methods, 316, 22–34. .
Lecci, S., Fernandez, L. M., Weber, F. D., Cardis, R., Chatton, J. Y., Born, J., & Luthi, A. (2017).
Coordinated infraslow neural and cardiac oscillations mark fragility and offline periods in
mammalian sleep. Science Advances, 3(2), e1602026.
Lee, J., Kim, D., & Shin, H.-S. (2004). Lack of delta waves and sleep disturbances during non-
rapid eye movement sleep in mice lacking alpha1G-subunit of T-type calcium channels.
Proceedings of the National Academy of Sciences of the United States of America, 101(52),
18195–18199.
Limoges, E., Mottron, L., Bolduc, C., Berthiaume, C., & Godbout, R. (2005). Atypical sleep
architecture and the autism phenotype. Brain, 128(Pt 5), 1049–1061.
Loomis, A. L., Harvey, E. N., & Hobart, G. (1935). Potential rhythms of the cerebral cortex
during sleep. Science, 81(2111), 597–598.
Loomis, A. L., Harvey, E. N., & Hobart, G. A. (1937). Cerebral states during sleep, as studied by
human brain potentials. Journal of Experimental Psychology, 21(2), 127–144.
Lustenberger, C., Boyle, M. R., Alagapan, S., Mellin, J. M., Vaughn, B. V., & Frohlich, F. (2016).
Feedback-controlled transcranial alternating current stimulation reveals a functional role of
sleep spindles in motor memory consolidation. Current Biology, 26(16), 2127–2136.
Lustenberger, C., Maric, A., Durr, R., Achermann, P., & Huber, R. (2012). Triangular relation-
ship between sleep spindle activity, general cognitive ability and the efficiency of declarative
learning. PLoS One, 7(11), e49561.
Ma, Y., Shi, W., Peng, C. K., & Yang, A. C. (2018). Nonlinear dynamical analysis of sleep elec-
troencephalography using fractal and entropy approaches. Sleep Medicine Review, 37, 85–93.
430 ALPÁR S. LÁZÁR, ZSOLT I. LÁZÁR, AND RÓBERT BÓDIZS
Mak- McCully, R. A., Rolland, M., Sargsyan, A., Gonzalez, C., Magnin, M., Chauvel,
P., . . . Halgren, E. (2017). Coordination of cortical and thalamic activity during non-REM
sleep in humans. Nature Communications, 8, 15499.
Massimini, M., Huber, R., Ferrarelli, F., Hill, S., & Tononi, G. (2004). The sleep slow oscillation
as a traveling wave. Journal of Neuroscience, 24(31), 6862–6870.
McKinney, S. M., Dang-Vu, T. T., Buxton, O. M., Solet, J. M., & Ellenbogen, J. M. (2011). Covert
waking brain activity reveals instantaneous sleep depth. PLoS One, 6(3), e17351.
Mohawk, J. A., Green, C. B., & Takahashi, J. S. (2012). Central and peripheral circadian clocks
in mammals. Annual Review of Neuroscience, 35, 445–462.
Molle, M., Bergmann, T. O., Marshall, L., & Born, J. (2011). Fast and slow spindles during the
sleep slow oscillation: Disparate coalescence and engagement in memory processing. Sleep,
34(10), 1411–1421.
Molle, M., Marshall, L., Gais, S., & Born, J. (2002). Grouping of spindle activity during slow
oscillations in human non-rapid eye movement sleep. Journal of Neuroscience, 22(24),
10941–10947.
Mooij, A. H., Frauscher, B., Goemans, S. A. M., Huiskamp, G. J. M., Braun, K. P. J., & Zijlmans,
M. (2018). Ripples in scalp EEGs of children: Co-occurrence with sleep-specific transients
and occurrence across sleep stages. Sleep, 41(11), zsy169-zsy169.
Mooij, A. H., Raijmann, R. C. M. A., Jansen, F. E., Braun, K. P. J., & Zijlmans, M. (2017).
Physiological ripples (± 100 Hz) in spike-free scalp EEGs of children with and without epi-
lepsy. Brain Topography, 30(6), 739–746.
Moroni, F., Nobili, L., De Carli, F., Massimini, M., Francione, S., Marzano, C., . . . Ferrara, M.
(2012). Slow EEG rhythms and inter-hemispheric synchronization across sleep and wakeful-
ness in the human hippocampus. Neuroimage, 60(1), 497–504.
Neske, G. T. (2016). The slow oscillation in cortical and thalamic networks: Mechanisms and
functions. Frontiers in Neural Circuits, 9, 88).
Nishida, M., Hirai, N., Miwakeichi, F., Maehara, T., Kawai, K., Shimizu, H., & Uchida, S.
(2004). Theta oscillation in the human anterior cingulate cortex during all-night sleep: An
electrocorticographic study. Neuroscience Research, 50(3), 331–341.
Noachtar, S., Binnie, C., Ebersole, J., Mauguiere, F., Sakamoto, A., & Westmoreland, B. (1999).
A glossary of terms most commonly used by clinical electroencephalographers and pro-
posal for the report form for the EEG findings. The International Federation of Clinical
Neurophysiology. Electroencephalography and Clinical Neurophysiology Supplement,
52, 21–41.
Parrino, L., Grassi, A., & Milioli, G. (2014). Cyclic alternating pattern in polysomnography: What
is it and what does it mean? Current Opinion in Pulmonary Medicine, 20(6), 533–541.
Pearl, P. L., LaFleur, B. J., Reigle, S. C., Rich, A. S., Freeman, A. A., McCutchen, C., & Sato,
S. (2002). Sawtooth wave density analysis during REM sleep in normal volunteers. Sleep
Medicine, 3(3), 255–258.
Pereda, E., Gamundi, A., Rial, R., & Gonzalez, J. (1998). Non-linear behaviour of human
EEG: Fractal exponent versus correlation dimension in awake and sleep stages. Neuroscience
Letters, 250(2), 91–94.
Perlis, M. L., Merica, H., Smith, M. T., & Giles, D. E. (2001). Beta EEG activity and insomnia.
Sleep Medicine Review, 5(5), 363–374.
Perlis, M. L., Smith, M. T., Andrews, P. J., Orff, H., & Giles, D. E. (2001). Beta/gamma EEG ac-
tivity in patients with primary and secondary insomnia and good sleeper controls. Sleep,
24(1), 110–117.
FREQUENCY CHARACTERISTICS OF SLEEP 431
Pesonen, A.-K., Ujma, P., Halonen, R., Räikkönen, K., & Kuula, L. (2019). The associations
between spindle characteristics and cognitive ability in a large adolescent birth cohort.
Intelligence, 72, 13–19.
Piano, C., Mazzucchi, E., Bentivoglio, A. R., Losurdo, A., Calandra Buonaura, G., Imperatori,
C., . . . Della Marca, G. (2017). Wake and sleep EEG in patients with Huntington disease: An
eLORETA study and review of the literature. Clinical EEG and Neuroscience, 48(1), 60–7 1.
Picchioni, D., Horovitz, S. G., Fukunaga, M., Carr, W. S., Meltzer, J. A., Balkin, T. J., . . . Braun, A.
R. (2011). Infraslow EEG oscillations organize large-scale cortical-subcortical interactions
during sleep: a combined EEG/fMRI study. Brain Research, 1374, 63–72.
Plante, D. T., Goldstein, M. R., Landsness, E. C., Peterson, M. J., Riedner, B. A., Ferrarelli,
F., . . . Benca, R. M. (2013). Topographic and sex-related differences in sleep spindles in major
depressive disorder: A high-density EEG investigation. Journal of Affective Disorders, 146(1),
120–125.
Potari, A., Ujma, P. P., Konrad, B. N., Genzel, L., Simor, P., Kormendi, J., . . . Bodizs, R.
(2017). Age-related changes in sleep EEG are attenuated in highly intelligent individuals.
Neuroimage, 146, 554–560.
Rechtschaffen, A. & Kales, A. (1968). A manual of standardized terminology, techniques and
scoring system for sleep stages of human subjects. University of California Los Angeles Brain
Information Service.
NINDB Neurological Information Network (US), Bethesda, Md., U. S. National Institute of
Neurological Diseases and Blindness, Neurological Information Network.
Ritter, P. S., Schwabedal, J., Brandt, M., Schrempf, W., Brezan, F., Krupka, A., . . . Nikitin, E.
(2018). Sleep spindles in bipolar disorder -a comparison to healthy control subjects. Acta
Psychiatrica Scandinavica, 138(2), 163–172.
Roosli, M., Juni, P., Braun-Fahrlander, C., Brinkhof, M. W., Low, N., & Egger, M. (2006).
Sleepless night, the moon is bright: Longitudinal study of lunar phase and sleep. Journal of
Sleep Research, 15(2), 149–153.
Santhi, N., Lazar, A. S., McCabe, P. J., Lo, J. C., Groeger, J. A., & Dijk, D. J. (2016). Sex differences
in the circadian regulation of sleep and waking cognition in humans. Proceedings of the
National Academy of Sciences of the United States of America, 113(19), E2730–2739.
Sato, S., McCutchen, C., Graham, B., Freeman, A., von Albertini-Carletti, I., & Alling, D.
W. (1997). Relationship between muscle tone changes, sawtooth waves and rapid eye
movements during sleep. Electroencephalography and Clinical Neurophysiology, 103(6),
627–632.
Savage, V. M. & West, G. B. (2007). A quantitative, theoretical framework for understanding
mammalian sleep. Proceedings of the National Academy of Sciences of the United States of
America, 104(3), 1051.
Schabus, M., Hodlmoser, K., Gruber, G., Sauter, C., Anderer, P., Klosch, G., . . . Zeitlhofer, J.
(2006). Sleep spindle-related activity in the human EEG and its relation to general cognitive
and learning abilities. European Journal of Neuroscience, 23(7), 1738–1746.
Schilling, C., Schlipf, M., Spietzack, S., Rausch, F., Eisenacher, S., Englisch,
S., . . . Schredl, M. (2017). Fast sleep spindle reduction in schizophrenia and healthy first-
degree relatives: Association with impaired cognitive function and potential intermediate
phenotype. European Archives of Psychiatry and Clinical Neuroscience, 267(3), 213–224.
Schmidt, C., Collette, F., Cajochen, C., & Peigneux, P. (2007). A time to think: Circadian
rhythms in human cognition. Cognitive Neuropsychology, 24(7), 755–789. doi:10.1080/
02643290701754158
432 ALPÁR S. LÁZÁR, ZSOLT I. LÁZÁR, AND RÓBERT BÓDIZS
Schonwald, S. V., Carvalho, D. Z., Dellagustin, G., de Santa-Helena, E. L., & Gerhardt, G. J.
(2011). Quantifying chirp in sleep spindles. Journal of Neuroscience Methods, 197(1), 158–164.
Shimizu, R. E., Connolly, P. M., Cellini, N., Armstrong, D. M., Hernandez, L. T., Estrada,
R., . . . Simons, S. B. (2018). Closed-loop targeted memory reactivation during sleep improves
spatial navigation. Frontiers in Human Neuroscience, 12, 28.
Silber, M. H., Ancoli-Israel, S., Bonnet, M. H., Chokroverty, S., Grigg-Damberger, M. M.,
Hirshkowitz, M., . . . Iber, C. (2007). The visual scoring of sleep in adults. Journal of Clinical
Sleep Medicine, 3(2), 121–131.
Simor, P., Gombos, F., Blaskovich, B., & Bodizs, R. (2018). Long-range alpha and beta and
short-range gamma EEG synchronization distinguishes phasic and tonic REM periods.
Sleep, 41(3), zsx210. https://doi.org/10.1093/sleep/zsx210
Simor, P., Gombos, F., Szakadat, S., Sandor, P., & Bodizs, R. (2016). EEG spectral power in
phasic and tonic REM sleep: different patterns in young adults and children. Journal of Sleep
Research, 25(3), 269–277.
Simor, P., Horváth, K., Ujma, P. P., Gombos, F., & Bódizs, R. (2013). Fluctuations between
sleep and wakefulness: Wake-like features indicated by increased EEG alpha power during
different sleep stages in nightmare disorder. Biological Psychology, 94(3), 592–600.
Skeldon, A. C., Dijk, D. J., & Derks, G. (2014). Mathematical models for sleep-wake dy-
namics: Comparison of the two-process model and a mutual inhibition neuronal model.
PLoS One, 9(8), e103877.
Sloan, E. P., Maunder, R. G., Hunter, J. J., & Moldofsky, H. (2007). Insecure attachment is
associated with the alpha-EEG anomaly during sleep. Biopsychosocial Medicine, 1, 20.
Smarr, B. L., Jennings, K. J., Driscoll, J. R., & Kriegsfeld, L. J. (2014). A time to remember: The
role of circadian clocks in learning and memory. Behavioral Neuroscience, 128(3), 283–303.
Spieweg, I., Sanders, S., & Kubicki, S. (1992). Periodische Entladungen von Schlafspindeln
unter Plazebo und Zopiclone. [Periodical discharges of sleep spindles under placebo and
Zopiclone]. Klinische Neurophysiologie, 23(04), 215–220.
Spinosa, M. J. & Garzon, E. (2007). Sleep spindles: Validated concepts and breakthroughs.
Journal of Epilepsy and Clinical Neurophysiology, 13, 179–182.
Stam, C. J. (2005). Nonlinear dynamical analysis of EEG and MEG: Review of an emerging
field. Clinical Neurophysiology, 116(10), 2266–2301.
Staresina, B. P., Bergmann, T. O., Bonnefond, M., van der Meij, R., Jensen, O., Deuker, L., . . . Fell,
J. (2015). Hierarchical nesting of slow oscillations, spindles and ripples in the human hippo-
campus during sleep. Nature Neuroscience, 18(11), 1679–1686.
Steriade, M. (2000). Corticothalamic resonance, states of vigilance and mentation.
Neuroscience, 101(2), 243–276.
Steriade, M., & Amzica, F. (1998). Coalescence of sleep rhythms and their chronology in
corticothalamic networks. Sleep Research Online, 1(1), 1–10.
Steriade, M., Contreras, D., Curro Dossi, R., & Nunez, A. (1993). The slow (< 1 Hz) oscillation
in reticular thalamic and thalamocortical neurons: Scenario of sleep rhythm generation in
interacting thalamic and neocortical networks. Journal of Neuroscience, 13(8), 3284–3299.
Steriade, M., Nunez, A., & Amzica, F. (1993a). Intracellular analysis of relations between the
slow (< 1 Hz) neocortical oscillation and other sleep rhythms of the electroencephalogram.
Journal of Neuroscience, 13(8), 3266–3283.
Steriade, M., Nunez, A., & Amzica, F. (1993b). A novel slow (< 1 Hz) oscillation of neocortical
neurons in vivo: Depolarizing and hyperpolarizing components. Journal of Neuroscience,
13(8), 3252–3265.
Stupfel, M., & Pavely, A. (1990). Ultradian, circahoral and circadian structures in endothermic
vertebrates and humans. Comparative Biochemistry & Physiology A: Physiology, 96(1), 1–11.
FREQUENCY CHARACTERISTICS OF SLEEP 433
Tessier, S., Lambert, A., Chicoine, M., Scherzer, P., Soulieres, I., & Godbout, R. (2015).
Intelligence measures and stage 2 sleep in typically-developing and autistic children.
International Journal of Psychophysiology, 97(1), 58–65.
Tononi, G., & Cirelli, C. (2014). Sleep and the price of plasticity: From synaptic and cellular
homeostasis to memory consolidation and integration. Neuron, 81(1), 12–34.
Turanyi, C. Z., Ronai, K. Z., Zoller, R., Veber, O., Czira, M. E., Ujszaszi, A., . . . Novak, M. (2014).
Association between lunar phase and sleep characteristics. Sleep Medicine, 15(11), 1411–1416.
Uchida, S., Maehara, T., Hirai, N., Kawai, K., & Shimizu, H. (2003). Theta oscillation in the an-
terior cingulate and beta-1 oscillation in the medial temporal cortices: A human case report.
Journal of Clinical Neuroscience, 10(3), 371–374.
Uchida, S., Maloney, T., & Feinberg, I. (1992). Beta (20–28 Hz) and delta (0.3–3 Hz) EEGs oscil-
late reciprocally across NREM and REM sleep. Sleep, 15(4), 352–358.
Ujma, P. P., Halasz, P., Simor, P., Fabo, D., & Ferri, R. (2018). Increased cortical involvement and
synchronization during CAP A1 slow waves. Brain Structure & Function, 223(8), 3531–3542.
Ujma, P. P., Sandor, P., Szakadat, S., Gombos, F., & Bodizs, R. (2016). Sleep spindles and in-
telligence in early childhood-developmental and trait-dependent aspects. Developmental
Psychology, 52(12), 2118–2129.
Urakami, Y., Ioannides, A. A., & Kostopoulos, G. K. (2012). Sleep spindles –as a Biomarker of
brain function and plasticity. In I. Ajeena (Ed.), Advances in clinical neurophysiology (pp.
73–108). IntechOpen.
Vanhatalo, S., Palva, J. M., Holmes, M. D., Miller, J. W., Voipio, J., & Kaila, K. (2004). Infraslow
oscillations modulate excitability and interictal epileptic activity in the human cortex during
sleep. Proceedings of the National Academy of Sciences of the United States of America, 101(14),
5053–5057.
Vanhatalo, S., Tallgren, P., Andersson, S., Sainio, K., Voipio, J., & Kaila, K. (2002). DC-EEG
discloses prominent, very slow activity patterns during sleep in preterm infants. Clinical
Neurophysiology, 113(11), 1822–1825.
Vanhatalo, S., Voipio, J., & Kaila, K. (2005). Full-band EEG (FbEEG): An emerging standard in
electroencephalography. Clinical Neurophysiology, 116(1), 1–8.
Vijayan, S., Lepage, K. Q., Kopell, N. J., & Cash, S. S. (2017). Frontal beta-theta network during
REM sleep. eLife, 6, e18894.
Voss, U., Holzmann, R., Tuin, I., & Hobson, J. A. (2009). Lucid dreaming: a state of conscious-
ness with features of both waking and non-lucid dreaming. Sleep, 32(9), 1191–1200.
Vyazovskiy, V. V., Olcese, U., Lazimy, Y. M., Faraguna, U., Esser, S. K., Williams, J. C., . . . Tononi,
G. (2009). Cortical firing and sleep homeostasis. Neuron, 63(6), 865–878.
Wei, Y., Krishnan, G. P., Komarov, M., & Bazhenov, M. (2018). Differential roles of sleep
spindles and sleep slow oscillations in memory consolidation. PLoS Computational Biology,
14(7), e1006322
Weiss, B., Clemens, Z., Bodizs, R., Vago, Z., & Halasz, P. (2009). Spatio-temporal analysis of
monofractal and multifractal properties of the human sleep EEG. Journal of Neuroscience
Methods, 185(1), 116–124.
Werk, C. M., Harbour, V. L., & Chapman, C. A. (2005). Induction of long-term potentiation leads
to increased reliability of evoked neocortical spindles in vivo. Neuroscience, 131(4), 793–800.
Wright, K. P., Lowry, C. A., & Lebourgeois, M. K. (2012). Circadian and wakefulness-sleep
modulation of cognition in humans. Frontiers in Molecular Neuroscience, 5, 50.
Zygierewicz, J., Blinowska, K. J., Durka, P. J., Szelenberger, W., Niemcewicz, S., & Androsiuk, W.
(1999). High resolution study of sleep spindles. Clinical Neurophysiology, 110(12), 2136–2147.
CHAPTER 18
A REVIEW OF OS C I L L ATORY
BRAIN DY NA MI C S I N
SCHIZOPH RE NIA
KEVIN M. SPENCER
18.1 Introduction
The idea that synchronous neural activity in the gamma band mediated information pro-
cessing in the brain arrived at a time in which schizophrenia was being conceptualized
as a disorder not just of malfunctioning of particular brain regions, but of connectivity
in both local microcircuits (McGlashan & Hoffman, 2000), and large-scale, distributed
networks (e.g., Friston & Frith, 1995). Gamma oscillations had been proposed to be
the primary means by which brain networks were functionally connected (e.g., Singer,
1999). When taken together with the idea that schizophrenia was a “splitting of the
mind”, characterized by a “disintegration of thought and personality” (Bleuler, 1911/
436 KEVIN M. SPENCER
18.2.1.1 ASSR Studies
The ASSR is the most commonly studied gamma oscillation in schizophrenia,
and deficits in the power and/or phase locking factor (PLF; Tallon-Baudry et al.,
1996) aspects of this oscillation in individuals with schizophrenia have been consistently
replicated (reviewed in O’Donnell et al., 2013; for meta-analysis see Thuné et al., 2016).
Following Kwon and colleagues’ (1999) seminal report, others have reported gamma
(40 Hz, sometimes also 30 Hz) ASSR power and/or PLF deficits in chronic patients
A REVIEW OF OSCILLATORY BRAIN DYNAMICS IN SCHIZOPHRENIA 437
(Brenner et al., 2003; Hong et al., 2004; Light et al., 2006; Teale et al., 2008; Krishnan
et al., 2009: Spencer et al., 2009; Hamm et al., 2011; Tsuchimoto et al., 2011: Edgar et al.,
2014; Hirano et al., 2015; Zhou et al., 2018).
However, an important factor to consider in schizophrenia research is that the
results of studies conducted with chronic patients, that is, patients who have had
schizophrenia for some time, could be confounded to some degree by the duration
of the patients’ illness and their exposure to medications, especially antipsychotic
drugs (e.g., Leung et al., 2011). Therefore, it is important to test whether patients who
have never been medicated, first-episode patients (who were recently diagnosed with
schizophrenia), and unmedicated individuals within the schizophrenia spectrum
show the same abnormalities as chronic patients. In general, these studies have found
the same pattern of gamma-frequency ASSR deficits as in chronic schizophrenia. The
gamma ASSR deficit has been reported in ultra-high-risk individuals (Tada et al.,
2016), first-degree relatives (Hong et al., 2004; Puvvada et al., 2018; Rass et al., 2012),
schizotypal personality disorder (Brenner et al., 2003; but not observed in Rass et al.,
2012), first-episode schizophrenia (Spencer et al., 2008a; Tada et al., 2016), early-onset
psychosis (Wilson et al., 2008), schizoaffective disorder (Zhou et al., 2018), and bipolar
I disorder (Isomura et al., 2016; Spencer et al., 2008a; Zhou et al., 2018). One study
also found the gamma ASSR deficit in bipolar I disorder without psychosis (Parker
et al., 2019), which raises the question of whether this deficit is particular to psychosis.
Additional studies of the ASSR in non-psychotic bipolar disorder need to be done to
confirm this result.
Thus, the deficit in gamma ASSR generation appears to affect individuals with psych-
osis in general, soon after the first episode of psychosis. The gamma ASSR deficit is even
present in first degree relatives of individuals with schizophrenia, who share genetic (and
possibly environmental) risk factors with persons with schizophrenia, even though they
are not psychotic and have not received antipsychotic medication. A limited amount of
evidence suggests that the gamma ASSR deficit is also present in individuals at high risk
of developing schizophrenia. As more studies take place with this population, it is im-
portant to ask whether the degree of the gamma ASSR deficit is predictive of the prob-
ability of their conversion to psychosis.
While the gamma ASSR is clearly sensitive to neural circuit abnormalities in psych-
osis, it is not so clearly associated with particular psychotic symptoms. Given the na-
ture of the ASSR as an auditory sensory response, and the hypothesized relationships
between gamma oscillations and perception, hallucination symptoms may be expected
to be correlated with gamma ASSR measures. Indeed, some studies have found
correlations between gamma ASSR measures and hallucination symptoms (Mulert
et al., 2011; Spencer et al., 2008a,2009; Zhou et al., 2018), although the direction of these
correlations is not consistent across studies.
The ASSR has become the most widely-studied gamma oscillation in schizophrenia
research because of several factors:
2. It is easy to analyze, as the frequency is known a priori and the power of the ASSR
can be measured with basic spectral analysis methods that are commonly found in
EEG analysis software.
3. The gamma ASSR deficit in schizophrenia appears to be independent of whether
or not the stimuli are attended (e.g., Hamm et al., 2015).
4. The basic set of neural generators of the ASSR are known (e.g., Herdman et al.,
2002, 2003).
5. The ASSR exhibits good test-retest reliability (Legget et al., 2017; McFadden et al.,
2014; Tan et al., 2015).
For these reasons, the ASSR is being extensively used in translational neuroscience
studies of animal models relevant to schizophrenia. While rodents do not appear to
have a clearly defined resonance frequency as humans, they do exhibit gamma ASSR
deficits when administered drugs that are commonly used to model aspects of psych-
osis, such as N-methyl-D-aspartate receptor (NMDAR) antagonists (e.g., Leishman
et al., 2015; Sivarao et al., 2016; Sullivan et al., 2015). Therefore, the ASSR is a promising
tool for translational neuroscience research.
18.2.1.2 EAGBR Studies
The next most-studied gamma oscillation in schizophrenia is the EAGBR, owing
to the extensive use of the auditory oddball and sensory gating paradigms in schizo-
phrenia ERP research, from which the EAGBR can be readily measured (typically in
the responses to the standard stimuli). In general, the evoked power and PLF aspects of
the EAGBR are decreased in the schizophrenia spectrum. EAGBR deficits in chronic
schizophrenia patients have been reported in auditory oddball (e.g., Hall et al., 2011a;
Lenz et al., 2011; Oribe et al., 2019; Roach & Mathalon, 2008), sensory gating (e.g., Hall
et al., 2011b; Popov et al., 2011), and tone discrimination (e.g., Leicht et al., 2010) tasks,
although some studies have failed to find deficits (Brenner et al., 2009; Spencer et al.,
2008b). In first-episode schizophrenia patients, various studies all reported reduced
evoked power and PLF (Leicht et al., 2015; Oribe et al., 2019; Taylor et al., 2013), although
Gallinat and colleagues (2004) did not find EAGBR deficits in a sample of unmedicated
patients. Studies of clinical high-risk individuals are mixed: Perez and colleagues (2013)
found that EAGBR evoked power was reduced but PLF did not differ compared with
healthy individuals, while Oribe and colleagues (2019) did not find any reductions in
clinical high-risk individuals.
Three studies of the EAGBR have provided some insight into the genetic basis of audi-
tory gamma oscillation deficits in schizophrenia. Leicht and colleagues (2011) found
EAGBR power and PLF deficits in both schizophrenia patients and their first-degree
relatives in a tone discrimination task, pointing to the EAGBR deficit being an inher-
itable trait. Hall and colleagues (2011a) examined the EAGBR from an oddball task in
a sample of schizophrenia patients and their monozygotic twins who were concordant
or discordant for schizophrenia, and compared them to a sample of healthy twin pairs.
Hall and colleagues (2011a) found EAGBR power and PLF deficits in the patients, with
A REVIEW OF OSCILLATORY BRAIN DYNAMICS IN SCHIZOPHRENIA 439
deficits to lesser degrees in the well co-twins of discordant pairs. These results were
clearly indicative of a strong genetic factor in the EAGBR deficits. But in contrast to
those studies, Hall and colleagues (2011b) did not find an EAGBR power deficit in the
unaffected first-degree relatives of schizophrenia patients in data from a sensory gating
task, which suggests that there is no heritable component of the EAGBR deficit in
schizophrenia. One possible explanation for these discrepant findings is that attention
plays a role in producing the EAGBR deficit, as the standard sensory gating task used by
Hall and colleagues (2011b) was passive (unlike the active tasks in Hall et al., 2011a and
Leicht et al., 2011, which found deficits). Thus, the EAGBR deficit in the schizophrenia
spectrum and its genetic basis could involve a major contribution from reduced top-
down attentional modulation of the EAGBR, rather than a dysfunctional EAGBR per se.
Further work is needed to test this hypothesis.
18.2.1.3 Summary
The ASSR and EAGBR studies provide strong evidence for impaired generation of
stimulus-evoked gamma oscillations in the auditory cortex in schizophrenia, although
the ASSR results as a whole are stronger than the EAGBR findings. A natural question
to ask is whether the gamma ASSR and EAGBR deficits are correlated. Roach and
colleagues (2013) found that 40-Hz ASSR PLF and EAGBR PLF were indeed correlated
within both patient and controls groups. Thus, auditory-evoked gamma oscillations
may provide biomarkers of cortical circuit abnormalities that are heritable and present
across the schizophrenia spectrum.
However, for long-duration stimuli, a transient evoked offset oscillation is also found
with similar characteristics as the early evoked oscillation.
These three visual oscillations have been investigated in schizophrenia. Spencer and
colleagues (2003, 2004, 2008b) found reduced PLF of the early visual evoked gamma
442 KEVIN M. SPENCER
oscillation in chronic schizophrenia patients, but in other studies this oscillation did
not differ from healthy controls (Ghorashi & Spencer, 2015; Spencer & Ghorashi,
2014). Grent-‘t-Jong and colleagues (2020) reported reduced PLF of the early evoked
gamma oscillation in first-episode patients and clinical high-risk individuals. Grent-
‘t-Jong and colleagues (2016, 2020) showed that the power of the visual induced oscil-
lation was reduced in schizophrenia patients during visual stimulation with a moving
grating stimulus, and during stimulation with upright and inverted Mooney face
stimuli (Grützner et al., 2013; Sun et al., 2013). And, the power of the offset response was
also reduced overall in schizophrenia patients compared to controls in Grützner and
colleagues (2013) and Sun and colleagues (2013). Thus, several studies have found some
impairments in gamma oscillations elicited by visual stimuli in schizophrenia, but the
causes for the discrepancies among studies are not clear.
As in the auditory system, steady-state stimulation has been used in the visual system
to probe the integrity of its circuitry in schizophrenia, although with little consistency
in the stimulation frequencies tested among studies. Using 1-Hz stimulation, Jin and
colleagues (2000) found that schizophrenia patients had reduced visual steady-state re-
sponse (VSSR) power at harmonics in the alpha band (10–12 Hz) compared to healthy
controls. Clementz and colleagues (2004) found that the temporal dynamics of the VSSR
to stimulation at 6.4 Hz were abnormal in schizophrenia, showing a delayed buildup and
a longer decay time. Similarly, using 12.5-Hz stimulation, Ethridge and colleagues (2011)
found a decline in VSSR power during the stimulation period in persons with schizo-
phrenia compared to controls. Krishnan and colleagues (2005) investigated VSSRs
at stimulation frequencies from 4–40 Hz and found reduced VSSR power in patients
for beta and gamma frequency stimulation. But in contrast, Riečanský and colleagues
(2010) found increased PLF of the initial transient phase of the VSSR for 40 Hz stimula-
tion in patients.
In sum, there is evidence for deficits in visual gamma generation in schizophrenia,
although these deficits may not be as general as with auditory gamma oscillations. There
is a substantial degree of inconsistency in the early visual-evoked oscillation and gamma
VSSR findings, while visual induced gamma power seems to be consistently decreased in
schizophrenia (although relatively few studies have investigated this oscillation). More
experiments assessing different types of stimuli, tasks, and multiple VSSR stimulation
frequencies are needed to better understand visual gamma generation in schizophrenia.
2005). Since most of the evidence for neural circuit abnormalities in schizophrenia
comes from studies of PFC microcircuitry, deficits should be found gamma oscillations
elicited in tasks that engage PFC circuits. Several studies have indeed reported gamma
oscillations deficits in working memory and executive control tasks in schizophrenia.
One notable series of studies employed the “preparing to overcome prepotency”
(POP) task, in which a rule governing the spatial correspondence between a response
cue and the manual response is maintained in working memory. On low cognitive
control trials, the response cue matches the response hand, while on high cognitive
control trials, the response cue signals the opposite response hand. In the first study
reporting strong evidence for PFC gamma deficits in working memory in schizo-
phrenia, Cho and colleagues (2006) examined gamma power in chronic schizophrenia
patients and controls performing the POP task, finding that increased cognitive control
requirements elicited higher PFC gamma power during the delay period at electrodes
over the PFC in controls but not in patients, who made more errors in the high con-
trol condition. Gamma power during the late part of the delay was correlated with ac-
curacy in the controls, and negatively correlated with disorganization symptoms in the
patients. In a study comparing medicated and unmedicated first-episode patients with
controls in the POP task, Minzenberg and colleagues (2010) found a similar pattern
of cognitive control-related gamma deficits at frontal electrodes in both groups of
patients, indicating that the gamma deficit was not related to antipsychotic medication,
and was present early in the course of the disorder. In healthy individuals, cognitive-
control-related gamma in the POP task was correlated with gamma-amino-butyric acid
(GABA) neurotransmission, but not in schizophrenia patients (Frankle et al., 2015).
And Minzenberg and colleagues (2015) showed that the administration of modafinil, a
drug that improves cognition through its actions on the dopamine and norepinephrine
systems, enhanced performance on the POP task and the associated cognitive control-
related gamma power in schizophrenia patients.
Using a version of the Sternberg paradigm, Haenschel and colleagues (2009) provided
further support for the findings of impaired gamma oscillations associated with working
memory maintenance in schizophrenia. Participants had to encode from one to three
objects in working memory and then maintain them for 12 seconds, until a probe
item was presented. Hence, this task enabled the dissociation of processes involved in
working memory encoding, rehearsal, and retrieval. The study found that their early-
onset schizophrenia patients had worse performance than controls as memory load
(number of objects) increased and showed deficits in stimulus-evoked oscillations in
several bands in the encoding phase. In the late delay period, induced gamma power at
frontal sites in the controls peaked in the three-item load condition, whereas induced
gamma power peaked for the patients in the two-item condition. This kind of abnormal
memory load/response function has been observed in functional neuroimaging studies
(e.g., Jansma et al., 2004). The patients also had reduced induced theta and gamma
power during the retrieval phase compared to controls.
These studies demonstrate that gamma activity likely originating in the PFC involved
in working memory and executive control processes is impaired in schizophrenia, and
444 KEVIN M. SPENCER
point to avenues for restoring cognitive function to normal levels. This area of research
makes the best link between studies of neural microcircuits and oscillatory activity
related to particular cognitive functions in schizophrenia.
Single TMS pulses evoke EEG oscillations in the cortex underneath the stimulated
scalp site. Evidence suggests that the frequency of these TMS-evoked EEG oscillations
varies for different cortical areas (Ferrarelli et al., 2012; Rosanova et al., 2009): ~10 Hz
for visual cortex, ~20 Hz for parietal cortex, and ~30 Hz for prefrontal cortex. It has
been proposed that the frequency of these oscillations reflects the resonant frequency
of the stimulated cortical area (Rosanova et al., 2009), which is likely to be determined
by specific characteristics of the circuitry in that area. Therefore, TMS-evoked EEG
oscillations could be used to directly probe the integrity of cortical circuits in neuro-
psychiatric disorders.
To date, only a handful of studies have been conducted in this area. Ferrarelli and
colleagues (2008) examined the oscillations evoked by TMS to the frontal cortex in
healthy persons and schizophrenia patients and found that the oscillations were reduced
in total power and PLF in the patients. In a follow-up study, Ferrarelli and colleagues
(2012) obtained TMS-evoked oscillations to stimulation over prefrontal, premotor,
motor, and parietal cortex. They found that the frequency, total power, and PLF of the
TMS-evoked oscillation were lower in patients than controls for stimulation at the pre-
frontal and premotor sites. Canali and colleagues (2015) stimulated over premotor cortex
in patients with schizophrenia, bipolar disorder, and depression, and found a reduction
in the TMS-evoked oscillation frequency in all the patient groups compared to controls.
And in a study of first-episode psychosis patients, Ferrarelli and colleagues (2019) found
reduced beta/low gamma oscillations evoked by TMS of the motor cortex. The results of
these studies are consistent with the hypothesized gamma oscillation deficits in frontal
cortical areas in schizophrenia (see Section 18.3). TMS-evoked oscillations could thus
provide a powerful approach to investigate cortical circuits directly without needing to
use particular sensory stimuli or tasks to engage the brain region of interest.
scalp EEG, and functional neuroimaging, among others. The patterns of activity
observed with all of these methods can be classified into two categories: evoked
activity, which results directly from sensory or other stimulation; and intrinsic or
spontaneous activity, which is the ongoing activity does not directly result from
stimulation. Due to the highly recurrent architecture of connections in the brain,
spontaneous activity makes up most of neural activity, while evoked activity makes
up only a small portion (Raichle, 2010). Classical views of brain function have
considered spontaneous activity to be simply “noise”, with evoked activity simply
being added to the random ongoing activity. However, more recent evidence has
demonstrated that spontaneous activity is structured in space and time (Tsodyks
et al., 1999), and much of the variance in evoked responses can be accounted for
by spontaneous activity (Arieli et al., 1996). Spontaneous activity appears to reflect
the spatial and temporal connectivity patterns in local and large-scale networks,
and in part represents “predictions” of expected sensory input based on past ex-
perience, which interact with and modulate incoming stimulus evoked responses
(Ringach, 2009).
EEG is a powerful method with which to study spontaneous and evoked activity
in the brain as it is sensitive to neural activity across most of the neural frequency
spectrum. However, the neural generators and functional significance of many types
of spontaneous activity in the EEG are largely unknown. Infraslow (0.01–0.1 Hz) and
slow (0.1–1 Hz) oscillations in the EEG may correspond to fluctuations in the blood
oxygenation level dependent signal measured with functional magnetic resonance im-
aging, as these kinds of activity occur in the same frequency bands and may share the
same neural generators in large-scale networks of brain regions (Hiltunen et al., 2014).
At the opposite ends of the frequency and spatial spectra, broadband (as opposed
to narrowband) gamma activity (30–100 Hz) in the EEG may reflect asynchronous
neuronal spiking in local circuits (e.g., Burke et al., 2015; Yizhar et al., 2011). Recent
studies suggest that oscillatory activity across the frequency spectrum is hierarchically
organized (Buzsáki & Draguhn, 2004), with the phase of low-frequency oscillations
modulating the power of higher- frequency oscillations (reviewed in Canolty &
Knight, 2010).
EEG studies have repeatedly demonstrated that various types of sensory-and task-
evoked responses tend to be decreased in individuals with schizophrenia compared
to healthy control subjects. ERPs ranging from early sensory evoked components like
the P50, to purely cognitive components like the P300, generally show decreased, ra-
ther than increased amplitudes in schizophrenia (Javitt et al., 2008). And as reviewed
earlier, event-related oscillations typically show decreased evoked power and/or phase
locking. But while schizophrenia is generally associated with decreases in evoked
brain activity, spontaneous oscillatory activity during waking appears to be generally
increased in particular frequency bands. (In fact, findings of increased spontaneous
oscillation power in schizophrenia go back several decades; see Itil, 1977.) One major
question is whether evoked activity deficits are caused to some degree by increased
spontaneous activity.
446 KEVIN M. SPENCER
18.5.3 Sleep Oscillations
Investigations of spontaneous oscillatory activity during sleep in schizophrenia have
focused on two oscillations: slow (delta) waves and sleep spindles (7–15 Hz). Slow waves
448 KEVIN M. SPENCER
are generated in the cortex, whereas spindles originate in the thalamic reticular nu-
cleus (TRN) and are transmitted to the cortex (Steriade, 2006). Spindle generation in
the thalamus involves interactions among inhibitory interneurons and excitatory thal-
amocortical projection neurons in the TRN (e.g., Jacobsen et al., 2001). Corticothalamic
projections initiate spindles through glutamatergic inputs to TRN cells at NMDARs.
The oscillations tend to be linked through phase-amplitude coupling, such that sleep
spindles are more likely to occur during the up-state phase of the slow wave (e.g.,
Staresina et al., 2015). Both oscillations are hypothesized to support the consolidation of
long-term memories during sleep (Diekelmann & Born, 2010).
Compared to gamma oscillations, there have been relatively few studies of sleep
oscillations in schizophrenia, despite the detailed knowledge of the circuitry that
generates these oscillations. While spontaneous EEG activity during the resting state
or task performance tends to be increased in schizophrenia, spontaneous oscilla-
tory patterns during sleep tend to be reduced. The findings on slow waves are mixed,
with some reports of deficits (e.g., D’Agostino et al., 2018; Sarkar et al., 2010; Sekimoto
et al., 2011) but other studies not finding slow wave abnormalities (e.g., Ferrarelli et al.,
2007, 2010). In contrast, there is strong evidence that spindle generation is impaired in
schizophrenia (e.g., Ferrarelli et al., 2007, 2010; Manoach et al., 2014; Wamsley et al.,
2012). Spindle deficits have been reported in medication-naïve first-episode patients
(Manoach et al., 2014) and unaffected first-degree relatives of schizophrenia patients
(D’Agostino et al., 2018; Manoach et al., 2014), indicating that spindle impairment is a
trait marker of schizophrenia and is not due to antipsychotic medication. While spindle
impairment is associated with reduced thalamic volume (Buchmann et al., 2014), it has
also been correlated with abnormally increased thalamocortical functional connect-
ivity (Baran et al., 2019). With regards to symptomatology, spindle deficits have been
associated with psychotic symptom ratings (Ferrarelli et al., 2010; Wamsley et al., 2012).
Notably, there is substantial evidence that sleep spindle deficits are associated with
impairments in memory consolidation in schizophrenia (reviewed in Manoach et al.,
2016), pointing to spindle-generating circuitry as a potential target for treatment of both
psychotic symptoms and impaired cognition in schizophrenia.
18.6.1 NMDAR Hypofunction
The glutamate hypothesis of schizophrenia (Moghaddam & Javitt, 2012) offers clues to
the neural circuit abnormalities that may be responsible for both evoked gamma oscil-
lation deficits and increased spontaneous gamma activity in this disorder. In healthy
people, acute administration of NMDAR antagonists produces a constellation of posi-
tive and negative symptoms, cognitive deficits, and neurophysiological abnormalities
that resemble those found in schizophrenia (Krystal et al., 2003). In animals, the be-
havioral and neurophysiological abnormalities resulting from acute administra-
tion of NMDAR antagonists (reviewed in Amann et al., 2010) and genetic knockouts
of NMDARs (e.g., Belforte et al., 2010; Korotkova et al., 2010) resemble those found
in schizophrenia. In particular, acute NMDAR antagonism reduces the amplitude of
some ERPs (Amann et al., 2010) and gamma oscillations (e.g., Lazarewicz et al., 2010;
Leishman et al., 2015; Sivarao et al., 2016; Sullivan et al., 2015), in addition to increasing
broadband spontaneous gamma power (reviewed in Hunt & Kasicki, 2013). Genetic
ablation of NMDARs also increases spontaneous gamma power in adult animals (e.g.,
Carlén et al., 2012; Korotkova et al., 2010; Tatard-Leitman et al., 2015). Increased spon-
taneous gamma as a result of NMDAR hypofunction results from increased excitability
of pyramidal cells, either through disinhibition from reduced excitatory drive to fast-
spiking interneurons (Carlén et al., 2012; Homayoun & Moghaddam, 2007), and/or
by alterations in cell membrane properties of the pyramidal cells themselves (Tatard-
Leitman et al., 2015). With regard to increased spontaneous delta/theta power, in the
thalamus, NMDAR antagonism increases delta band activity, which is transmitted to
the hippocampus (Zhang et al., 2012) and the cortex (Hunt & Kasicki, 2013).
In healthy humans, besides the behavioral effects that resemble psychotic symptoms,
NMDAR antagonism via acute ketamine administration decreases the amplitude of
some ERPs (e.g., Gunduz-Bruce et al., 2012) and increases broadband gamma power
during rest (Muthukumaraswamy et al., 2015; Rivolta et al., 2015). This increase in spon-
taneous gamma power is consistent with increased cortical excitability, which has
450 KEVIN M. SPENCER
been shown with TMS to result from NMDAR antagonist administration (Di Lazzaro
et al., 2003). But in contrast to the animal model studies, ketamine in humans decreases
resting state cortical delta activity (Muthukumaraswamy et al., 2015).
Within the last two decades, research into neural oscillations has generated a body of
findings that has made a substantial impact on our conceptualization of schizophrenia,
understanding of its pathophysiology, and methods of testing new treatments. The
study of gamma oscillations has had the most success, due to the cross-fertilization of
clinical studies with neuropathological, basic, and computational research. As details
of the specific cellular elements that are disturbed in schizophrenia are revealed, it is
possible to construct hypotheses about the consequences of these circuit abnormalities
in animal and computational models and try to relate them to the human clinical data.
The field has matured in step with the broader field of oscillatory brain dynamics. In
most studies the potential sources of EEG artifacts are now understood and corrected
for, and the importance of having adequate numbers of epochs is appreciated. Studies
have better statistical power, and mass univariate statistics are being used with the ap-
propriate corrections for multiple tests. The influence of confounds like common refer-
ence and volume conduction is more widely appreciated for connectivity studies. Thus,
the potential for false positive findings seems to be less than before.
And yet, to this researcher, the promise of oscillation research for ultimately helping
develop new treatments for schizophrenia (and other neuropsychiatric disorders)
seems far from being fulfilled. The most complete body of research has been on the
gamma ASSR, which is being used as a biomarker in translational research, yet this
oscillation has limited relevance for cognition or symptoms in schizophrenia (except
possibly for auditory hallucinations). The areas of research into perceptual organ-
ization and working memory, while being the best motivated by psychological and
mechanistic theories, have not yet yielded reliable measures that can be used in transla-
tional research. So, much more work remains to be done. I still believe that as research
progresses into the neural mechanisms underlying different oscillations, and the nature
of neural circuit abnormalities in schizophrenia is understood in greater detail, it will
become possible to construct and test mechanistic hypotheses with greater precision,
both in vivo and in silico. Ultimately this research effort may come full circle, leading to
the use of neurostimulation methods to modulate and “correct” aberrant oscillations in
individuals with schizophrenia, and in this way, re-integrate the broken mind.
ACKNOWLEDGMENTS
This work was supported by grants I01 CX001443 from the US Department of Veterans
Affairs, R01 MH093450 from the National Institutes of Health, and an Independent
452 KEVIN M. SPENCER
Investigator Award from the Brain and Behavior Research Foundation. The author
reports no conflicts of interest.
REFERENCES
Amann, L. C., Gandal, M. J., Halene, T. B., Ehrlichman, R. S., White, S. L., McCarren, H. S.,
& Siegel, S. J. (2010). Mouse behavioral endophenotypes for schizophrenia. Brain Research
Bulletin, 83(3–4), 147–161.
Arieli, A., Sterkin, A., Grinvald, A., & Aertsen, A. (1996). Dynamics of ongoing ac-
tivity: Explanation of the large variability in evoked cortical responses. Science, 273,
1868–1871.
Baran, B., Karahanoğlu, F. I., Mylonas, D., Demanuele, C., Vangel, M., Stickgold, R., Anticevic,
A., & Manoach, D. S. (2019). Increased thalamocortical connectivity in schizophrenia
correlates with sleep spindle deficits: evidence for a common pathophysiology. Biological
Psychiatry: Cognitive Neuroscience and Neuroimaging, 4, 706–7 14.
Barch, D. M. (2005). The cognitive neuroscience of schizophrenia. Annual Review of Clinical
Psychology, 1, 321–353.
Belforte, J. E., Zsiros, V., Sklar, E. R., Jiang, Z., Yu, G., Li, Y., . . . Nakazawa, K. (2010). Postnatal
NMDA receptor ablation in corticolimbic interneurons confers schizophrenia- like
phenotypes. Nature Neuroscience, 13(1), 76–83.
Bleuler, E. (1911/ 1950). Dementia Praecox or the Group of Schizophrenias. International
Universities Press.
Boutros, N. N., Arfken, C., Galderisi, S., Warrick, J., Pratt, G., & Iacono, W. (2008). The status
of spectral EEG abnormality as a diagnostic test for schizophrenia. Schizophrenia Research,
99(1–3), 225–237.
Brenner, C. A., Kieffaber, P. D., Clementz, B. A., Johannesen, J. K., Shekhar, A., O’Donnell, B. F.,
& Hetrick, W. P. (2009). Event-related potential abnormalities in schizophrenia: A failure to
“gate in” salient information? Schizophrenia Research, 113(2–3), 332–338.
Brenner, C., Sporns, O., Lysaker, P. H., & O’Donnell, B. F. (2003). EEG synchronization to
modulated auditory tones in schizophrenia, schizoaffective disorder, and schizotypal per-
sonality disorder. American Journal of Psychiatry, 160(December), 2238–2240.
Buchmann, A., Dentico, D., Peterson, M. J., Riedner, B. A., Sarasso, S., Massimini, M., Tononi,
G., & Ferrarelli, F. (2014). Reduced mediodorsal thalamic volume and prefrontal cortical
spindle activity in schizophrenia. NeuroImage, 102, 540–547.
Burke, J. F., Ramayya, A. G., & Kahana, M. J. (2015). Human intracranial high-frequency ac-
tivity during memory processing: neural oscillations or stochastic volatility? Current
Opinion in Neurobiology, 31, 104–110.
Buzsáki, G. (2006). Rhythms of the brain. Oxford University Press.
Buzsáki, G. & Draguhn, A. (2004). Neuronal oscillations in cortical networks. Science, 304,
1926–1929.
Buzsáki, G. & Wang, X.-J. (2012). Mechanisms of gamma oscillations. Annual Review of
Neuroscience, 35, 203–225.
Canali, P., Sarasso, S., Rosanova, M., Casarotto, S., Sferrazza- Papa, G., Gosseries,
O., . . . Benedetti, F. (2015). Shared reduction of oscillatory natural frequencies in bipolar
disorder, major depressive disorder and schizophrenia. Journal of Affective Disorders, 184,
111–115.
A REVIEW OF OSCILLATORY BRAIN DYNAMICS IN SCHIZOPHRENIA 453
Canolty, R. T. & Knight, R. T. (2010). The functional role of cross-frequency coupling. Trends in
Cognitive Sciences, 14, 506–515.
Carlén, M., Meletis, K., Siegle, J. H., Cardin, J. A, Futai, K., Vierling-Claassen, D., . . . Tsai, L.-H.
(2012). A critical role for NMDA receptors in parvalbumin interneurons for gamma rhythm
induction and behavior. Molecular Psychiatry, 17(5), 537–548.
Cho, R. Y., Konecky, R. O., & Carter, C. S. (2006). Impairments in frontal cortical gamma
synchrony and cognitive control in schizophrenia. Proceedings of the National Academy of
Sciences of the United States of America, 103(52), 19878–19883.
Clementz, B. A., Blumenfeld, L. D., & Cobb, S. (1997). The gamma band response may account
for poor P50 suppression in schizophrenia. Neuroreport, 8(18), 3889–3893.
Clementz, B. A., Keil, A., & Kissler, J. (2004). Aberrant brain dynamics in schizo-
phrenia: Delayed buildup and prolonged decay of the visual steady-state response. Cognitive
Brain Research, 18(2), 121–129. doi:10.1016/j.cogbrainres.2003.09.007
Clementz, B., Sponheim, S., Iacono, W. G., & Beiser, M. (1994). Resting EEG in first-episode
schizophrenia patients, bipolar psychosis patients, and their first- degree relatives.
Psychophysiology, 31, 486–494.
D’Agostino, A., Castelnovo, A., Cavallotti, S., Casetta, C., Marcatili, M., Gambini, O., Canevini,
M., Tononi, G., Riedner, B., Ferrarelli, F., & Sarasso, S. (2018). Sleep endophenotypes of
schizophrenia: slow waves and sleep spindles in unaffected first-degree relatives. NPJ
Schizophrenia, 4(1), 2.
Davidson, M., Kapara, O., Goldberg, S., Yoffe, R., Noy, S., & Weiser, M. (2016). A nation-wide
study on the percentage of schizophrenia and bipolar disorder patients who earn minimum
wage or above. Schizophrenia Bulletin, 42(2), 443–447.
Diekelmann, S. & Born, J. (2010). The memory function of sleep. Nature Reviews Neuroscience,
11, 114–126.
Di Lazzaro, V., Oliviero, A., Profice, P., Pennisi, M. A., Pilato, F., Zito, G., . . . Tonali, P. A. (2003).
Ketamine increases human motor cortex excitability to transcranial magnetic stimulation.
The Journal of Physiology, 547(Pt 2), 485–496.
Ding, N. & Simon, J. Z. (2013). Power and phase properties of oscillatory neural responses in
the presence of background activity. Journal of Computational Neuroscience, 34(2), 337–343.
Draganova, R., Ross, B., Wollbrink, A., & Pantev, C. (2008). Cortical steady-state responses to
central and peripheral auditory beats. Cerebral Cortex, 18(5), 1193–1200.
Eckhorn, R., Bauer, R., Jordan, W., Brosch, M., Kruse, W., Munk, M., & Reitboeck, H. (1988).
Coherent oscillations: a mechanism of feature linking in the visual cortex? Multiple elec-
trode and correlation analyses in the cat. Biological Cybernetics, 60, 121–130.
Edgar, J. C., Chen, Y.-H., Lanza, M., Howell, B., Chow, V. Y., Heiken, K., . . . Cañive, J. M.
(2014). Cortical thickness as a contributor to abnormal oscillations in schizophrenia?
NeuroImage: Clinical, 4, 122–129.
Ethridge, L., Moratti, S., Gao, Y., Keil, A., & Clementz, B. A. (2011). Sustained versus tran-
sient brain responses in schizophrenia: The role of intrinsic neural activity. Schizophrenia
Research, 133(1–3), 106–111.
Farahani, E. D., Goossens, T., Wouters, J., & van Wieringen, A. (2017). Spatiotemporal recon-
struction of auditory steady-state responses to acoustic amplitude modulations: Potential
sources beyond the auditory pathway. NeuroImage, 148, 240–253.
Ferrarelli, F., Huber, R., Peterson, M. J., Massimini, M., Murphy, M., Riedner, B. A., . . . Tononi,
G. (2007). Reduced sleep spindle activity in schizophrenia patients. American Journal of
Psychiatry, 164(3), 483–492.
454 KEVIN M. SPENCER
Ferrarelli, F., Kaskie, R. E., Graziano, B., Reis, C. C., & Casali, A. G. (2019). Abnormalities
in the evoked frontal oscillatory activity of first-episode psychosis: A TMS/EEG study.
Schizophrenia Research, 206, 436–439.
Ferrarelli, F., Massimini, M., Peterson, M. J., Riedner, B. A., Lazar, M., Murphy, M. J., . . . Tononi,
G. (2008). Reduced evoked gamma oscillations in the frontal cortex in schizophrenia
patients: A TMS/EEG study. American Journal of Psychiatry, 165(8), 996–1005.
Ferrarelli, F., Peterson, M., Sarasso, S., Riedner, B. A., Murphy, M. J., Benca, R. M., . . . Tononi,
G. (2010). Thalamic dysfunction in schizophrenia suggested by whole-night deficits in slow
and fast spindles. American Journal of Psychiatry, 167(November), 1339–1348.
Ferrarelli, F., Sarasso, S., Guller, Y., Riedner, B. A, Peterson, M. J., Bellesi, M., ... Tononi, G.
(2012). Reduced natural oscillatory frequency of frontal thalamocortical circuits in schizo-
phrenia. Archives of General Psychiatry, 69, 766–774.
Frankle, W. G., Cho, R. Y., Prasad, K. M., Mason, N. S., Paris, J., Himes, M. L., . . . Narendran, R.
(2015). In vivo measurement of GABA transmission in healthy subjects and schizophrenia
patients. American Journal of Psychiatry, 172, 1148–1159.
Friston, K. J. & Frith, C. D. (1995). Schizophrenia: A disconnection syndrome? Clinical
Neuroscience, 3, 89–97.
Galambos, R., Makei, S., & Talmachoff, P. J. (1981). A 40-Hz auditory potential recorded from
the human scalp. Proceedings of the National Academy of Sciences of the United States of
America, 78(4), 2643–2647.
Gallinat, J., Winterer, G., Herrmann, C. S., & Senkowski, D. (2004). Reduced oscillatory
gamma-band responses in unmedicated schizophrenic patients indicate impaired frontal
network processing. Clinical Neurophysiology, 115(8), 1863–1874.
Ghorashi, S. & Spencer, K. M. (2015). Attentional load effects on beta oscillations in healthy and
schizophrenic individuals. Frontiers in Psychiatry, 6(October), 149.
Glantz, L. A. & Lewis, D. A. (2000). Decreased dendritic spine density on prefrontal cortical
pyramidal neurons in schizophrenia. Archives of General Psychiatry, 57(1), 65–73.
Gray, C. M., Konig, P., Engel, A. K., & Singer, W. (1989). Oscillatory responses in cat visual
cortex exhibit inter-columnar synchronization which reflects global stimulus properties.
Nature, 338, 334–337.
Gray, C. M. & Singer, W. (1989). Stimulus-specific neuronal oscillations in orientation columns
of cat visual cortex. Proceedings of the National Academy of Sciences of the United States of
America, 86, 1698–1702.
Grent-‘t-Jong, T., Gajwani, R., Gross, J., Gumley, A. I., Krishnadas, R., Lawrie, S. M., . . . Uhlhaas,
P. J. (2020). Association of magnetoencephalographically measured high- frequency
oscillations in visual cortex with circuit dysfunctions in local and large-scale networks
during emerging psychosis. JAMA Psychiatry, 77(8), 852–862.
Grent-’t-Jong, T., Gross, J., Goense, J., Wibral, M., Gajwani, R., Gumley, A. I., ... Uhlhaas, P.
J. (2018). Resting-state gamma-band power alterations in schizophrenia reveal E/I-balance
abnormalities across illness-stages. eLife, 7, e37799.
Grent-‘t-Jong, T., Rivolta, D., Sauer, A., Grube, M., Singer, W., Wibral, M., & Uhlhaas, P. J.
(2016). MEG-measured visually induced gamma-band oscillations in chronic schizo-
phrenia: Evidence for impaired generation of rhythmic activity in ventral stream regions.
Schizophrenia Research, 176(2–3), 177–185.
Grützner, C., Wibral, M., Sun, L., Rivolta, D., Singer, W., Maurer, K., & Uhlhaas, P. J. (2013).
Deficits in high-(>60 Hz) gamma-band oscillations during visual processing in schizo-
phrenia. Frontiers in Human Neuroscience, 7(March), 88.
A REVIEW OF OSCILLATORY BRAIN DYNAMICS IN SCHIZOPHRENIA 455
Guevara, R., Luis Pérez Velazquez, J., Nenadovic, V., Wennberg, R., Senjanovic, G., & Garcia
Dominguez, L. (2005). Phase synchronization measurements using electroencephalographic
recordings: What can we really say about neuronal synchrony? Neuroinformatics, 3, 301–313.
Gunduz-Bruce, H., Reinhart, R. M. G., Roach, B. J., Gueorguieva, R., Oliver, S., D’Souza, D.
C., . . . Mathalon, D. H. (2012). Glutamatergic modulation of auditory information pro-
cessing in the human brain. Biological Psychiatry, 71(11), 969–977.
Haenschel, C., Bittner, R. a, Waltz, J., Haertling, F., Wibral, M., Singer, W., . . . Rodriguez, E.
(2009). Cortical oscillatory activity is critical for working memory as revealed by deficits in
early-onset schizophrenia. Journal of Neuroscience, 29(30), 9481–9489.
Hajszan, T., Leranth, C., & Roth, R. H. (2006). Subchronic phencyclidine treatment decreases
the number of dendritic spine synapses in the rat prefrontal cortex. Biological Psychiatry,
60(6), 639–644.
Hall, M., Taylor, G., Salisbury, D. F., & Levy, D. L. (2011a). Sensory gating event-related
potentials and oscillations in schizophrenia patients and their unaffected relatives.
Schizophrenia Bulletin, 37(6), 1187–1199.
Hall, M., Taylor, G., Sham, P., Schulze, K., Rijsdijk, F., Picchioni, M., . . . Salisbury, D. F. (2011).
The early auditory gamma-band response is heritable and a putative endophenotype of
schizophrenia. Schizophrenia Bulletin, 37(4), 778–787.
Hall, S. D., Holliday, I. E., Hillebrand, A., Singh, K. D., Furlong, P. L., Hadjipapas, A., & Barnes,
G. R. (2005). The missing link: Analogous human and primate cortical gamma oscillations.
NeuroImage, 26, 13–17.
Hamm, J. P., Bobilev, A. M., Hayrynen, L. K., Hudgens-Haney, M. E., Oliver, W. T., Parker, D.
A., . . . Clementz, B. A. (2015). Stimulus train duration but not attention moderates γ-band
entrainment abnormalities in schizophrenia. Schizophrenia Research, 165(1), 97–102.
Hamm, J. P., Gilmore, C. S., Picchetti, N. A. M., Sponheim, S. R., & Clementz, B. A. (2011).
Abnormalities of neuronal oscillations and temporal integration to low-and high-frequency
auditory stimulation in schizophrenia. Biological Psychiatry, 69(10), 989–996.
Hebb, D. O. (1949). The organization of behavior. Wiley.
Herdman, A. T., Lins, O., Roon, P. Van, Stapells, D. R., Scherg, M., & Picton, T. W. (2002).
Intracerebral sources of human auditory steady-state responses. Brain Topography, 15, 69–86.
Herdman, A. T., Wollbrink, A., Chau, W., Ishii, R., Ross, B., & Pantev, C. (2003). Determination
of activation areas in the human auditory cortex by means of synthetic aperture magnetom-
etry. NeuroImage, 20(2), 995–1005.
Hiltunen, T., Kantola, J., Abou Elseoud, A., Lepola, P., Suominen, K., Starck, T., . . . Palva JM
(2014). Infra-slow EEG fluctuations are correlated with resting-state network dynamics in
fMRI. Journal of Neuroscience, 34, 356–362.
Hirano, Y., Oribe, N., Kanba, S., Onitsuka, T., Nestor, P. G., & Spencer, K. M. (2015).
Spontaneous gamma activity in schizophrenia. JAMA Psychiatry, 72(8), 813–821.
Hirano, Y., Oribe, N., Onitsuka, T., Kanba, S., Nestor, P. G., Hosokawa, T., ... Spencer, K. M.
(2020). Auditory cortex volume and gamma oscillation abnormalities in schizophrenia.
Clinical EEG and Neuroscience, 51, 244–251.
Hirvonen, J., Wibral, M., Palva, J. M., Singer, W., Uhlhaas, P., & Palva, S. (2017). Whole-brain
source-reconstructed MEG-data reveal reduced long-range synchronization in chronic
schizophrenia. eNeuro, 4(5), e0338–17.2017.
Hoftman, G. D., Dienel, S. J., Bazmi, H. H., Zhang, Y., Chen, K., & Lewis, D. A. (2018). Altered
gradients of glutamate and gamma-aminobutyric acid transcripts in the cortical visuo-
spatial working memory network in schizophrenia. Biological Psychiatry, 83(8), 670–679.
456 KEVIN M. SPENCER
Krishnan, G. P., Hetrick, W. P., Brenner, C. A., Shekhar, A., Steffen, A. N., & O’Donnell, B. F.
(2009). Steady state and induced auditory gamma deficits in schizophrenia. NeuroImage,
47(4), 1711–1719.
Krishnan, G. P., Vohs, J. L., Hetrick, W. P., Carroll, C. A., Shekhar, A., Bockbrader, M. A., &
O’Donnell, B. F. (2005). Steady state visual evoked potential abnormalities in schizophrenia.
Clinical Neurophysiology, 116(3), 614–624.
Krystal, J. H., D’Souza, D. C., Mathalon, D., Perry, E., Belger, A., & Hoffman, R. (2003). NMDA
receptor antagonist effects, cortical glutamatergic function, and schizophrenia: Toward a
paradigm shift in medication development. Psychopharmacology, 169(3–4), 215–233.
Kwon, J. S., O’Donnell, B. F., Wallenstein, G. V, Greene, R. W., Hirayasu, Y., Nestor, P.
G., . . . McCarley, R. W. (1999). Gamma frequency-range abnormalities to auditory stimula-
tion in schizophrenia. Archives of General Psychiatry, 56(11), 1001–1005.
Lachaux, J. P., Rodriguez, E., Martinerie, J., & Varela, F. J. (1999). Measuring phase synchrony in
brain signals. Human Brain Mapping, 8(4), 194–208.
Lazarewicz, M. T., Ehrlichman, R. S., Maxwell, C. R., Gandal, M. J., Finkel, L. H., & Siegel,
S. J. (2010). Ketamine modulates theta and gamma oscillations. Journal of Cognitive
Neuroscience, 22(7), 1452–1464.
Legget, K. T., Hild, A. K., Steinmetz, S. E., Simon, S. T., & Rojas, D. C. (2017). MEG and EEG
demonstrate similar test-retest reliability of the 40 Hz auditory steady-state response.
International Journal of Psychophysiology, 114, 16–23.
Leicht, G., Andreou, C., Polomac, N., Lanig, C., Schöttle, D., Lambert, M., & Mulert, C. (2015).
Reduced auditory evoked gamma band response and cognitive processing deficits in first
episode schizophrenia. The World Journal of Biological Psychiatry, 16(6), 387–397.
Leicht, G., Karch, S., Karamatskos, E., Giegling, I., Möller, H.-J., Hegerl, U., . . . Mulert, C. (2011).
Alterations of the early auditory evoked gamma-band response in first-degree relatives of
patients with schizophrenia: Hints to a new intermediate phenotype. Journal of Psychiatric
Research, 45(5), 699–705.
Leicht, G., Kirsch, V., Giegling, I., Karch, S., Hantschk, I., Möller, H.-J., . . . Mulert, C. (2010).
Reduced early auditory evoked gamma-band response in patients with schizophrenia.
Biological Psychiatry, 67(3), 224–231.
Leishman, E., O’Donnell, B. F., Millward, J. B., Vohs, J. L., Rass, O., Krishnan, G. P., . . . Morzorati,
S. L. (2015). Phencyclidine disrupts the auditory steady state response in rats. PLoS One,
10(8), e0134979.
Lemere, F. (1936). The significance of individual differences in the Berger rhythm. Brain, 59,
366–375.
Lenz, D., Fischer, S., Schadow, J., Bogerts, B., & Herrmann, C. S. (2011). Altered evoked γ-
band responses as a neurophysiological marker of schizophrenia? International Journal of
Psychophysiology, 79(1), 25–31.
Leung, M., Cheung, C., Yu, K., Yip, B., Sham, P., Li, Q., . . . McAlonan, G. (2011). Gray matter in
first-episode schizophrenia before and after antipsychotic drug treatment. Anatomical like-
lihood estimation meta-analyses with sample size weighting. Schizophrenia Bulletin, 37(1),
199–211.
Lewis, D. A., Curley, A. A., Glausier, J. R., & Volk, D. W. (2012). Cortical parvalbumin
interneurons and cognitive dysfunction in schizophrenia. Trends in Neurosciences,
35(1), 57–67.
Light, G. A., Hsu, J. L., Hsieh, M. H., Meyer-Gomes, K., Sprock, J., Swerdlow, N. R., & Braff, D.
L. (2006). Gamma band oscillations reveal neural network cortical coherence dysfunction
in schizophrenia patients. Biological Psychiatry, 60(11), 1231–1240.
458 KEVIN M. SPENCER
Manoach, D. S., Demanuele, C., Wamsley, E. J., Vangel, M., Montrose, D. M., Miewald,
J., . . . Keshavan, M. S. (2014). Sleep spindle deficits in antipsychotic-naive early course
schizophrenia and in non-psychotic first-degree relatives. Frontiers in Human Neuroscience,
8(October), 762.
Manoach, D. S., Pan, J. Q., Purcell, S. M., & Stickgold, R. (2016). Reduced sleep spindles in
schizophrenia: A treatable endophenotype that links risk genes to impaired cognition?
Biological Psychiatry, 80(8), 599–608.
McFadden, K. L., Steinmetz, S. E., Carroll, A. M., Simon, S. T., Wallace, A., & Rojas, D. C.
(2014). Test-retest reliability of the 40 Hz EEG auditory steady-state response. PLoS One,
9(1), e85748.
McGlashan, T. H. & Hoffman, R. E. (2000). Schizophrenia as a disorder of developmentally
reduced synaptic connectivity. Archives of General Psychiatry, 57(7), 637–648.
Minzenberg, M. J., Firl, A. J., Yoon, J. H., Gomes, G. C., Reinking, C., & Carter, C. S. (2010).
Gamma oscillatory power is impaired during cognitive control independent of medication
status in first-episode schizophrenia. Neuropsychopharmacology, 35(13), 2590–2599.
Minzenberg, M. J., Laird, A. R., Thelen, S., Carter, C. S., & Glahn, D. C. (2009). Meta-analysis
of 41 functional neuroimaging studies of executive function in schizophrenia. Archives of
General Psychiatry, 66(8), 811–822.
Minzenberg, M. J., Yoon, J. H., Cheng, Y., & Carter, C. S. (2015). Sustained modafinil
treatment effects on control- related gamma oscillatory power in schizophrenia.
Neuropsychopharmacology, 41(5), 1231–1240.
Moghaddam, B. & Javitt, D. (2012). From revolution to evolution: the glutamate hypothesis of
schizophrenia and its implication for treatment. Neuropsychopharmacology, 37(1), 4–15.
Mooney, C. M. & Ferguson, G. A. (1951). A new closure test. Canadian Journal of Psychology, 5,
129–133.
Mulert, C., Kirsch, V., Pascual-Marqui, R., McCarley, R. W., & Spencer, K. M. (2011). Long-
range synchrony of γ oscillations and auditory hallucination symptoms in schizophrenia.
International Journal of Psychophysiology, 79(1), 55–63.
Murray, R. M. & Lewis, S. W. (1987). Is schizophrenia a neurodevelopmental disorder? British
Medical Journal, 295(6600), 681–682.
Muthukumaraswamy, S. D., Shaw, A. D., Jackson, L. E., Hall, J., Moran, R., & Saxena, N. (2015).
Evidence that subanesthetic doses of ketamine cause sustained disruptions of NMDA and
AMPA-mediated frontoparietal connectivity in humans. Journal of Neuroscience, 35(33),
11694–11706.
Muthukumaraswamy, S. D. & Singh, K. D. (2011). A cautionary note on the interpretation
of phase-locking estimates with concurrent changes in power. Clinical Neurophysiology,
122(11), 2324–2325.
Muthukumaraswamy, S. D. & Singh, K. D. (2013). Visual gamma oscillations: The effects of
stimulus type, visual field coverage and stimulus motion on MEG and EEG recordings.
NeuroImage, 69, 223–230.
Muthukumaraswamy, S. D., Singh, K. D., Swettenham, J. B., & Jones, D. K. (2010). Visual
gamma oscillations and evoked responses: variability, repeatability and structural MRI
correlates. NeuroImage, 49(4), 3349–3357.
Narayanan, B., O’Neil, K., Berwise, C., Stevens, M. C., Calhoun, V. D., Clementz, B.
A., . . . Pearlson, G. D. (2014). Resting state electroencephalogram oscillatory abnormalities
in schizophrenia and psychotic bipolar patients and their relatives from the bipolar and
A REVIEW OF OSCILLATORY BRAIN DYNAMICS IN SCHIZOPHRENIA 459
Ranlund, S., Nottage, J., Shaikh, M., Dutt, A., Constante, M., Walshe, M., . . . Bramon, E.
(2014). Resting EEG in psychosis and at-risk populations -A possible endophenotype?
Schizophrenia Research, 153(1–3), 96–102.
Rapoport, J. L., Giedd, J. N., & Gogtay, N. (2012). Neurodevelopmental model of schizo-
phrenia: Update 2012. Molecular Psychiatry, 17(12), 1228–1238.
Rass, O., Forsyth, J. K., Krishnan, G. P., Hetrick, W. P., Klaunig, M. J., Breier, A., . . . Brenner,
C. A. (2012). Auditory steady state response in the schizophrenia, first-degree relatives, and
schizotypal personality disorder. Schizophrenia Research, 136(1–3), 143–149.
Riečanský, I., Kašpárek, T., Rehulová, J., Katina, S., & Přikryl, R. (2010). Aberrant EEG
responses to gamma-frequency visual stimulation in schizophrenia. Schizophrenia Research,
124(1–3), 101–109.
Ringach, D. L. (2009). Spontaneous and driven cortical activity: Implications for computation.
Current Opinion in Neurobiology, 19, 439–444.
Rivolta, D., Heidegger, T., Scheller, B., Sauer, A., Schaum, M., Birkner, K., . . . Uhlhaas,
P. J. (2015). Ketamine dysregulates the amplitude and connectivity of high-frequency
oscillations in cortical- subcortical networks in humans: evidence from resting- state
magnetoencephalography-recordings. Schizophrenia Bulletin, 41(5), 1105–1114.
Roach, B. J., Ford, J. M., Hoffman, R. E., & Mathalon, D. H. (2013). Converging evidence for
gamma synchrony deficits in schizophrenia. Supplements to Clinical Neurophysiology, 62,
163–180.
Roach, B. J. & Mathalon, D. H. (2008). Event-related EEG time-frequency analysis: An over-
view of measures and an analysis of early gamma band phase locking in schizophrenia.
Schizophrenia Bulletin, 34(5), 907–926.
Rodriguez, E., George, N., Lachaux, J.-P., Martinerie, J., Renault, B., & Varela, F. J. (1999).
Perception’s shadow: Long-distance synchronization of human brain activity. Nature, 397,
430–433.
Rolls, E. T., Loh, M., Deco, G., & Winterer, G. (2008). Computational models of schizophrenia
and dopamine modulation in the prefrontal cortex. Nature Reviews Neuroscience, 9(9),
696–709.
Rosanova, M., Casali, A., Bellina, V., Resta, F., Mariotti, M., & Massimini, M. (2009). Natural
frequencies of human corticothalamic circuits. Journal of Neuroscience, 29(24), 7679–7685.
Sarkar, S., Katshu, M. Z. U. H., Nizamie, S. H., & Praharaj, S. K. (2010). Slow wave sleep deficits
as a trait marker in patients with schizophrenia. Schizophrenia Research, 124(1–3), 127–133.
https://doi.org/10.1016/j.schres.2010.08.013
Schulman, J. J., Cancro, R., Lowe, S., Lu, F., Walton, K. D., & Llinás, R. R. (2011). Imaging of thal-
amocortical dysrhythmia in neuropsychiatry. Frontiers in Human Neuroscience, 5(July), 69.
Selemon, L. D., & Goldman-Rakic, P. S. (1999). The reduced neuropil hypothesis: a circuit
based model of schizophrenia. Biological Psychiatry, 45, 17–25.
Sekimoto, M., Kato, M., Watanabe, T., Kajimura, N., & Takahashi, K. (2011). Cortical regional
differences of delta waves during all-night sleep in schizophrenia. Schizophrenia Research,
126(1–3), 284–290.
Siever, L. J. & Davis, K. L. (2004). The pathophysiology of schizophrenia disorders: perspectives
from the spectrum. American Journal of Psychiatry, 161(March), 398–413.
Singer, W. (1999). Neuronal synchrony: A versatile code for the definition of relations? Neuron,
24, 49–65.
Singer, W. & Gray, C. M. (1995). Visual feature integration and the temporal correlation hy-
pothesis. Annual Review of Neuroscience, 18, 555–586.
A REVIEW OF OSCILLATORY BRAIN DYNAMICS IN SCHIZOPHRENIA 461
Sivarao, D. V, Chen, P., Senapati, A., Yang, Y., Fernandes, A., Benitex, Y., . . . Ahlijanian, M. K.
(2016). 40 Hz auditory steady-state response is a pharmacodynamic biomarker for cortical
NMDA receptors. Neuropsychopharmacology, 41(9), 2232–2240.
Spencer, K. M. (2009). The functional consequences of cortical circuit abnormalities on
gamma oscillations in schizophrenia: insights from computational modeling. Frontiers in
Human Neuroscience, 3, 33.
Spencer, K. M. (2012). Baseline gamma power during auditory steady-state stimulation in
schizophrenia. Frontiers in Human Neuroscience, 5(January), 190.
Spencer, K. M., & Ghorashi, S. (2014). Oscillatory dynamics of Gestalt perception in schizo-
phrenia revisited. Frontiers in Psychology, 5(February), 68.
Spencer, K., Nestor, P., Niznikiewicz, M. A., Salisbury, D. F., Shenton, M. E., & McCarley, R.
W. (2003). Abnormal neural synchrony in schizophrenia. Journal of Neuroscience, 23(19),
7407–7411.
Spencer, K. M., Nestor, P. G., Perlmutter, R., Niznikiewicz, M. A, Klump, M. C., Frumin,
M., . . . McCarley, R. W. (2004). Neural synchrony indexes disordered perception and cogni-
tion in schizophrenia. Proceedings of the National Academy of Sciences of the United States of
America, 101(49), 17288–17293.
Spencer, K. M., Niznikiewicz, M. A, Nestor, P. G., Shenton, M. E., & McCarley, R. W. (2009).
Left auditory cortex gamma synchronization and auditory hallucination symptoms in
schizophrenia. BMC Neuroscience, 10, 85.
Spencer, K. M., Niznikiewicz, M. A., Shenton, M. E., & McCarley, R. W. (2008b). Sensory-
evoked gamma oscillations in chronic schizophrenia. Biological Psychiatry, 63(8),
744–747.
Spencer, K. M., Salisbury, D. F., Shenton, M. E., & McCarley, R. W. (2008a). Gamma-band
auditory steady-state responses are impaired in first episode psychosis. Biological Psychiatry,
64(5), 369–375.
Sponheim, S. R., Clementz, B. a, Iacono, W. G., & Beiser, M. (2000). Clinical and biological
concomitants of resting state EEG power abnormalities in schizophrenia. Biological
Psychiatry, 48(11), 1088–1097.
Staresina, B. P., Bergmann, T. O., Bonnefond, M., van der Meij, R., Jensen, O., Deuker, L., . . . Fell,
J. (2015). Hierarchical nesting of slow oscillations, spindles and ripples in the human hippo-
campus during sleep. Nature Neuroscience, 18(11), 1679–1686.
Steriade, M. (2006). Grouping of brain rhythms in corticothalamic systems. Neuroscience,
137(4), 1087–1106.
Sullivan, E. M., Timi, P., Hong, L. E., & O’Donnell, P. (2015). Effects of NMDA and GABA-
A receptor antagonism on auditory steady-state synchronization in awake behaving rats.
International Journal of Neuropsychopharmacology, 18(7), pyu118.
Sun, L., Castellanos, N., Grützner, C., Koethe, D., Rivolta, D., Wibral, M., . . . Uhlhaas, P. J.
(2013). Evidence for dysregulated high-frequency oscillations during sensory processing
in medication- naïve, first episode schizophrenia. Schizophrenia Research, 150(2–3),
519–525.
Sweet, R. A., Henteleff, R. A., Zhang, W., Sampson, A. R., & Lewis, D. A. (2009).
Reduced dendritic spine density in auditory cortex of subjects with schizophrenia.
Neuropsychopharmacology, 34(2), 374–389.
Tada, M., Nagai, T., Kirihara, K., Koike, S., Suga, M., Araki, T., . . . Kasai, K. (2016). Differential
alterations of auditory gamma oscillatory responses between pre-onset high-risk individuals
and first-episode schizophrenia. Cerebral Cortex, 26(3), 1027–1035.
462 KEVIN M. SPENCER
Tallon-Baudry, C., Bertrand, O., Delpuech, C., & Pernier, J. (1996). Stimulus specificity
of phase-locked and non-phase-locked 40 Hz visual responses in human. Journal of
Neuroscience, 16(13), 4240–4249.
Tan, H.-R. M., Gross, J., & Uhlhaas, P. J. (2015). MEG—measured auditory steady-state
oscillations show high test– retest reliability: A sensor and source- space analysis.
NeuroImage, 122, 417–426.
Tatard-Leitman, V. M., Jutzeler, C. R., Suh, J., Saunders, J. A., Billingslea, E. N., Morita,
S., . . . Siegel, S. J. (2015). Pyramidal cell selective ablation of N-methyl-D-aspartate receptor
1 causes increase in cellular and network excitability. Biological Psychiatry, 77(6), 556–568.
Taylor, G. W., McCarley, R. W., & Salisbury, D. F. (2013). Early auditory gamma band response
abnormalities in first hospitalized schizophrenia. Supplements to Clinical Neurophysiology,
62, 131–145.
Teale, P., Collins, D., Maharajh, K., Rojas, D. C., Kronberg, E., & Reite, M. (2008). Cortical
source estimates of gamma band amplitude and phase are different in schizophrenia.
NeuroImage, 42(4), 1481–1489.
Thuné, H., Recasens, M., & Uhlhaas, P. J. (2016). The 40-Hz auditory steady-state response in
patients with schizophrenia. JAMA Psychiatry, 73(11), 1145–1153.
Tsodyks, M., Kenet, T., Grinvald, A., & Arieli, A. (1999). Linking spontaneous activity of single
cortical neurons and the underlying functional architecture. Science, 286, 1943–1946.
Tsuchimoto, R., Kanba, S., Hirano, S., Oribe, N., Ueno, T., Hirano, Y., . . . Onitsuka, T. (2011).
Reduced high and low frequency gamma synchronization in patients with chronic schizo-
phrenia. Schizophrenia Research, 133(1–3), 99–105.
Uhlhaas, P. J., Linden, D. E. J., Singer, W., Haenschel, C., Lindner, M., Maurer, K., & Rodriguez,
E. (2006a). Dysfunctional long-range coordination of neural activity during Gestalt percep-
tion in schizophrenia. Journal of Neuroscience, 26(31), 8168–8175.
Uhlhaas, P. J., Phillips, W. a, Mitchell, G., & Silverstein, S. M. (2006b). Perceptual grouping in
disorganized schizophrenia. Psychiatry Research, 145(2–3), 105–117.
van Tricht, M. J., Ruhrmann, S., Arns, M., Müller, R., Bodatsch, M., Velthorst, E., . . . Nieman,
D. H. (2014). Can quantitative EEG measures predict clinical outcome in subjects at clin-
ical high risk for psychosis? A prospective multicenter study. Schizophrenia Research, 153(1–
3), 42–47.
Wamsley, E. J., Tucker, M. A., Shinn, A. K., Ono, K. E., McKinley, S. K., Ely, A. V, . . . Manoach,
D. S. (2012). Reduced sleep spindles and spindle coherence in schizophrenia: Mechanisms of
impaired memory consolidation? Biological Psychiatry, 71(2), 154–161.
Weinberger, D. R. (1987). Implications of normal brain development for the pathogenesis of
schizophrenia. Archives of General Psychiatry, 44(7), 660–669.
Wilson, T. W., Hernandez, O. O., Asherin, R. M., Teale, P. D., Reite, M. L., & Rojas, D. C. (2008).
Cortical gamma generators suggest abnormal auditory circuitry in early-onset psychosis.
Cerebral Cortex, 18(2), 371–378.
Winterer, G., Coppola, R., Goldberg, T. E., Egan, M. F., Jones, D. W., Sanchez, C. E., &
Weinberger, D. R. (2004). Prefrontal broadband noise, working memory, and genetic risk
for schizophrenia. American Journal of Psychiatry, 161(March), 490–500.
Winterer, G., Egan, M. F., Kolachana, B. S., Goldberg, T. E., Coppola, R., & Weinberger, D. R.
(2006). Prefrontal electrophysiologic “noise” and catechol-O-methyltransferase genotype
in schizophrenia. Biological Psychiatry, 60(6), 578–584.
A REVIEW OF OSCILLATORY BRAIN DYNAMICS IN SCHIZOPHRENIA 463
Winterer, G., Ziller, M., Dorn, H., Frick, K., Mulert, C., Wuebben, Y., . . . Coppola, R. (2000).
Schizophrenia: Reduced signal-to-noise ratio and impaired phase-locking during informa-
tion processing. Clinical Neurophysiology, 111(5), 837–849.
Wynn, J. K., Roach, B. J., Lee, J., Horan, W. P., Ford, J. M., Jimenez, A. M., & Green, M. F. (2015).
EEG findings of reduced neural synchronization during visual integration in schizophrenia.
PLoS One, 10(3), e0119849.
Yizhar, O., Fenno, L. E., Prigge, M., Schneider, F., Davidson, T. J., O’Shea, D. J., . . . Deisseroth, K.
(2011). Neocortical excitation/inhibition balance in information processing and social dys-
function. Nature, 477(7363), 171–178.
Zaehle, T., Lenz, D., Ohl, F. W., & Herrmann, C. S. (2010). Resonance phenomena in the human
auditory cortex: Individual resonance frequencies of the cerebral cortex determine electro-
physiological responses. Experimental Brain Research, 203(3), 629–635.
Zhang, Y., Yoshida, T., Katz, D. B., & Lisman, J. E. (2012). NMDAR antagonist action in thal-
amus imposes δ oscillations on the hippocampus. Journal of Neurophysiology, 107(11),
3181–3189.
Zhou, T.-H., Mueller, N. E., Spencer, K. M., Mallya, S. G., Lewandowski, K. E., Norris, L.
A., . . . Hall, M.-H. (2018). Auditory steady state response deficits are associated with
symptom severity and poor functioning in patients with psychotic disorder. Schizophrenia
Research, 201, 278–286.
CHAPTER 19
EEG FREQU E NC Y
T ECHNIQU ES FOR I MAG I NG
C ONTROL FU NC T I ONS I N
ANXIET Y
This chapter reviews and discusses the application of EEG frequency techniques to
imaging and understanding motivational/emotional and cognitive control functions
in anxiety. The first three sections present a historical account of how EEG frequency
techniques have been used to understand emotional, motivational, and cognitive
components of anxiety. The final section considers these different EEG frequency
metrics vis-à-vis theoretical models of anxiety and how they are being, or could be, used
to advance both science and treatment. Particularly, this chapter focuses on how anxiety
relates to motivational/emotional and cognitive control dynamics involved in balancing
bottom-up and top-down influences on behavior. As previous chapters in this volume
primarily review the measurement and nature of EEG frequency signals, here we in-
stead focus on the applications of these methods to the study of anxiety.
Anxiety is a common human experience that often involves a mix of cognitive,
physiological, and behavioral manifestations (Barlow, 2002). It lies on a continuum from
mild levels that can motivate individuals to increase vigilance and respond adaptively to
threat or challenge (Barlow, 2002), but at excessive levels represents the most common
mental health problem worldwide (U.S. Burden of Disease Collaborators, 2018). It is a
multi-dimensional construct that many have separated into anxious apprehension and
EEG FREQUENCY TECHNIQUES In Anxiety 465
anxious arousal subtypes, with the former dominated by verbal worries and ruminations
focused on future ambiguous threat and the latter by somatic tension and physiological
hyperactivity centered on clear immediate threats (Heller et al., 1997, Heller & Nitschke,
1998). Given its multifaceted nature, anxiety and its associated impairments have been
tackled from motivational/emotional, cognitive and neuropsychological perspectives
(Eysenck et al., 2007; Gray & McNaughton, 2000; Heller et al., 1997, Heller & Nitschke,
1998). In the review that follows, we examine anxiety across the spectrum of severity
and, thus, how EEG frequency metrics reflect the dynamic interplay of motivation,
emotion, and cognition in adaptive and maladaptive forms of anxiety.
Fox & Davidson, 1987) reported that in 10-month-old infants, distress, gaze aversion,
and crying in response to maternal separation was associated with greater relative right-
frontal brain activity. This was found to be both a trait and state marker of behavior,
as resting-state asymmetry was predictive of later anxious behaviors during the separ-
ation (Davidson & Fox, 1989; Fox et al., 1992), and asymmetry scores increased during
the separation (Fox & Davidson, 1987). Anxious behaviors, internalizing problems, and
diagnoses of anxiety disorders (based on direct observation and parent report) in pre-
school (Fox et al., 1995; Fox et al., 1996) and school-age children (Baving et al., 2002;
Schmidt et al., 1999) also demonstrated a relationship to relative right-frontal activa-
tion both at rest and during a stress induction. Similarly, child anxious temperament
(e.g., high reactivity and behavioral inhibition), which is predictive of the development
of clinically significant social anxiety (Hirshfeld-Becker et al., 2007), has been reported
to be associated with right anterior activity (McManis et al., 2002). The same patterns of
alpha frontal asymmetry can be observed in non-human primates (Kalin et al., 1998),
and can be decreased with the administration of anxiolytic benzodiazepines (Davidson
et al., 1992). In adults, self-reported trait-anxiety on the State-Trait-Anxiety Inventory
(STAI) (Spielberger, 1983) has been reported to correlate with relative right-frontal
asymmetry in healthy young adults (Adolph & Margraf, 2017; Blackhart et al., 2006).
In clinical populations, those with obsessive compulsive disorder (OCD) were found to
have greater relative right-frontal asymmetry (Ischebeck et al., 2014; Kuskowski et al.,
1993) and cognitive behavioral therapy has been found to shift asymmetry from relative
right to left activity in patients with social anxiety disorder (Moscovitch et al., 2011). In
summary, relative right-frontal alpha asymmetry seems to reflect tendencies to with-
draw from stress or novelty and a propensity towards anxious temperament.
The few EEG studies of alpha frontal asymmetry in patient populations (Buchsbaum
et al., 1985; Smith et al., 2016) have supported the hypothesis that anxious apprehen-
sion is predominately related to greater left-hemisphere activity, though findings in
non-patient populations have been mixed, possibly due to methodological differences.
For example, Heller and colleagues (1997) found that anxious apprehension, defined
as scores above the 90th percentile on the STAI, was associated with left-frontal asym-
metry, though as noted earlier, other studies have reported the opposite findings using
this measure (Adolph & Margraf, 2017; Blackhart et al., 2006) and one study found
no asymmetry differences (Smith et al., 2018). Studies that have used the Penn State
Worry Questionnaire (Meyer et al., 1990) as a measure of anxious apprehension have
also been mixed, with one study reporting greater left-frontal asymmetry (Smith et al.,
2016), and two reporting no asymmetry differences (Nitschke et al., 1999; Smith et al.,
2018). Hofmann et al. (2005) reported that a worry induction in healthy college students
resulted in increased left-frontal alpha asymmetry, whereas, Carter, Johnson, and
Borkovec (1986) reported no association between a worry induction and alpha asym-
metry, though greater left asymmetry was reported in the beta band. Recently, Nusslock
and colleagues (2018) reported that the typical pattern of relative right-frontal activity in
patients with depression was no longer present when considering those with comorbid
anxiety disorders (e.g., separation, GAD, and OCD). Together, these findings highlight
the importance of considering anxiety measurement and population in the interpret-
ation of alpha asymmetry findings.
With regards to anxious arousal and relative right- frontal asymmetry, all
investigations, with the exception of those in PTSD, have supported the hypothesis that
anxious arousal can be distinguished from anxious apprehension by relative right-frontal
activation. Self-reported anxious arousal (Mathersul et al., 2008; Nitschke et al., 1999;
Smith et al., 2016), anxiety inductions (Avram et al., 2010; Davidson et al., 2000; Isotani
et al., 2001) and clinical levels of anxious arousal (e.g., panic disorder) (Wiedemann
et al., 1999) have all been shown to correlate with relative right-frontal activation.
Reports in participants with PTSD have failed to find frontal asymmetry differences
in either direction (Kemp et al., 2010; Metzger et al., 2004; Rabe et al., 2006; Shankman
et al., 2008), though Rabe and colleagues (2006) did find right-frontal activity to be
associated with greater symptom severity when participants viewed trauma-related
pictures. Kemp and colleagues (2010) did not find differences between healthy controls
and participants with PTSD; however, they did find an association between PTSD se-
verity and right-frontal asymmetry within the PTSD group. These mixed findings in
PTSD could be due to the heterogeneous presentation of the disorder, age differences
in the populations studied, and the possibility that PTSD may be neurobiologically dis-
tinct from other anxiety disorders (Lobo et al., 2015; Shankman et al., 2008).
In addition to being associated with frontal alpha asymmetry, arousal has also been
related to alpha power in posterior regions (i.e., parietal and temporal) (Heller, 1993).
Indeed, PTSD patients demonstrate greater relative right-parietal activity (Kemp et al.,
2010; Metzger et al., 2004; Rabe, Beauducel, et al., 2006). Although some have reported
468 JASON S. MOSER et al.
similar findings in other anxious populations (Bruder et al., 1997; Heller et al., 1997;
Mathersul et al., 2008; Smith et al., 2016), there have been insufficient studies focusing
on posterior alpha activity to draw definitive conclusions.
attention network (Laufs et al., 2003; Mantini et al., 2007). Studies that have directly
examined correlates of alpha asymmetry have implicated the anterior cingulate cortex
(ACC; Gorka et al., 2015; Nash et al., 2012; Smith et al., 2018), the dorsal lateral prefrontal
cortex (Smith et al., 2018), the pre and postcentral gyri (Lubar et al., 2003) and the amyg-
dala (Zotev et al., 2016). Collectively, this suggests that alpha asymmetry could reflect
attention to motivationally relevant information and the execution of intended actions.
In sum, there has been great interest over the past 50 years in understanding how
alpha asymmetry relates to anxiety. Models of emotional valence, motivation, physio-
logical arousal, and verbal processing have been proposed to explain the reported
findings. However, because anxiety is a heterogeneous construct that may involve and/
or interact in complex ways with cognitive processes such as attention and motivation,
further research is needed to fully understand the functional significance of alpha asym-
metry in anxiety.
early attempts to understand relations between anxiety and FMT. Nonetheless, they
demonstrated that FMT was modulated in the context of anxiety.
A large number of studies has examined the role of medial frontal cortical activity
in cognitive control efforts, particularly when these efforts are aimed at reducing anx-
iety or the circumstances which contribute to anxiety. The cingulate, especially the an-
terior and midcingulate cortices (ACC and MCC, respectively), is a cortical area which
is highly interconnected with other brain systems, including emotional centers in the
limbic system, cognitive association areas in the prefrontal cortex, sensory and motor
processing areas in the parietal and occipital lobes, and deeper structures within the
brain (Van Den Heuvel et al., 2009). As such, the cingulate is well positioned to inte-
grate information from a variety of inputs and coordinate action-oriented outputs to
maximize behavioral adaptability (Laurienti et al., 2003). Furthermore, the MCC’s in-
volvement is critical whenever new or conflicting information must be evaluated for its
relevance to the current behavioral set (Botvinick et al., 2004). Indeed, researchers have
suggested the MCC has domain-general functionality that is engaged whenever there
is uncertainty about behavior and its outcomes (Cavanagh et al., 2012). Because uncer-
tainty is a central aspect to feelings of anxiety and to anxiety disorders, the MCC has
been a target for research efforts aimed at understanding the underpinnings of anxiety-
related brain dynamics (Cavanagh & Shackman, 2015; Shackman et al., 2011).
A growing body of work suggests a primary function of the MCC is to integrate cog-
nitive, affective, and sensory information to enable behavioral adaptation (Shackman
et al., 2011). MCC activity has been implicated when incoming information is novel,
conflicting, punishing, or difficult to assimilate into the current action plan. In par-
ticular, FMT oscillations in the MCC have been observed across a variety of tasks when-
ever cognitive control is needed to alter behavior (Cavanagh & Frank, 2014). Because of
its ability to facilitate long-range communication between disparate brain regions, theta
oscillations are a candidate mechanism for how the MCC helps organize and coordinate
neural processes related to action monitoring and response selection. Indeed, FMT has
been posited to represent an alarm signal that is integral for the recruitment of areas in
the prefrontal cortex that mediate cognitive control (Cavanagh & Frank, 2014). A wide
range of studies demonstrate that individuals high in anxiety generate larger FMT power
during cognitive tasks, suggesting anxiety and FMT hyperactivity are intimately related.
Moreover, larger FMT control signals are associated with more cautious or inhibited
responding, providing strong evidence that FMT is a biomarker that can be used to
study the intersection of anxiety and cognitive control (Cavanagh & Shackman, 2015).
event-related potential (ERP) paradigms. The wide array of tasks that elicit FMT is evi-
dence of the domain-general role that FMT plays in dealing with uncertainty through
the recruitment of cognitive control resources. Time-domain ERP components that
have been consistently related to FMT include the error-related negativity (ERN; Luu
et al., 2004; Trujillo & Allen, 2007), feedback negativity (FN; Foti et al., 2014; Watts
et al., 2018), and the N2 (Harper et al., 2014), all of which tend to be increased in anxious
individuals.
The ERN is a response-locked ERP component that is thought to reflect action-
monitoring that is involved in the recruitment of cognitive control during post-error
behavioral adjustment (Gehring et al., 2018). Studies investigating the time-frequency
dynamics have demonstrated that the ERN primarily reflects theta-band activity that
is generated in the MCC (Luu et al., 2004; Trujillo & Allen, 2007). ERN amplitude has
been positively associated with anxiety, worry, and behavioral inhibition in flanker
(Moser et al., 2012; Riesel et al., 2017), go/no-go (Moadab et al., 2010), and trial and
error learning (Judah et al., 2016) tasks (for reviews see (Cavanagh & Shackman, 2015;
Moser et al., 2013). The functional significance of the larger ERN in anxiety is hotly
debated, with some suggesting it reflects the degree to which errors are seen as aver-
sive (Weinberg et al., 2016), whereas others suggest it reflects compensatory deployment
of cognitive control to counteract the distracting effects of worry (Moser et al., 2013;
Schroder et al., 2017).
Another FMT-related ERP component thought to be generated by the MCC and
associated with anxiety is the feedback-locked FN (Foti et al., 2014; Watts et al., 2018).
Like other FMT-related ERP components, the FN can be understood as a signal from
the salience network indicating new information (e.g., feedback) is highly relevant for
adapting to changing circumstances. As such, the future-oriented emotional state of
anxiety can prime the salience network to be hypersensitive to performance-related
feedback, making individuals high in trait anxiety more likely to consistently produce
an elevated FN (Grupe & Nitschke, 2013). Congruent with this framework, FN ampli-
tude has shown positive associations with negative affect during a monetary incentive
delay task (Santesso et al., 2011) and behavioral inhibition in go/no-go (De Pascalis et al.,
2010) and spatial matching (Balconi & Crivelli, 2010) tasks.
Parallel to the ERN and FN, the N2 ERP component has been shown to be highly
related to FMT (Harper et al., 2014; Van Noordt et al., 2016) and is also sensitive to trait
and state anxiety differences. Although there is some debate as to which process(es) the
N2 most likely reflects, it is associated with performance monitoring, conflict resolution,
and response inhibition. Individuals high in anxiety have been shown to produce larger
N2s in a variety of contexts, including during go/no-go (Righi et al., 2009; Sehlmeyer
et al., 2010) and flanker (Schmid et al., 2015) tasks. Larger N2 amplitude to conflict trials
in anxious individuals might reflect the recruitment of compensatory effort to over-
come distraction by anxious thoughts or a regulation strategy aimed at reducing current
and future negative affect (Dreisbach & Fischer, 2012; Kool et al., 2010; Lindström et al.,
2013; Schouppe et al., 2012).
472 JASON S. MOSER et al.
Theta: 4–8 Hz
F8
Cz
Theta: 4–8 Hz
Time
Delta: 0.5–4 Hz
Cz Beta: >13 Hz
Figure 19.1 Simplistic graphic illustrating frontal midline theta and interchannel phase
synchrony cross frequency coupling, described in the text. Frontal midline theta metrics are
calculated at midline sites as theta power and phase synchrony between sites. Cross frequency
coupling metrics include correlations between amplitudes of slower (i.e., delta) and faster (i.e.,
beta) frequencies as well as correlations between phase and amplitude of these frequencies or
ratios of frequency power (i.e., slow wave/fast wave).
EEG FREQUENCY TECHNIQUES In Anxiety 473
of the association between anxiety and network-related FMT metrics (i.e., ICPS) that
will require further investigation to disentangle potentially confounding or moderating
effects.
1 Only one study has examined delta-gamma coupling (Knyazev et al., 2005), and the interpretation of
correlations over treatment during speech anticipation periods (Miskovic et al., 2011),
and that children of socially anxious parents showed increased delta-beta coupling
(Miskovic et al., 2011). However, negative associations between delta and beta have
also been reported in social anxiety (Harrewijn et al., 2016). Moreover, Harrewjin and
colleagues (2018b) recently found that this negative association between delta and beta
is heritable and speculated that the differences in the direction of coupling could be due
to different stress manipulations utilized by research groups (anticipatory stress vs. task-
induced stressor), or different ways of computing frequency power densities (i.e., abso-
lute vs. relative power). If the delta-beta correlation, however, is generally understood
as reflecting increased neural communication, then the mere presence of a relationship
between their frequencies could still indicate important cross-talk (regardless of direc-
tion), although the exact nature of this cross-talk remains vague. Further complicating
the relationship between anxiety and delta-beta coupling, one study showed that OCD
is characterized by delta-beta decoupling (Velikova et al., 2010). Clearly more research
will be needed to clarify how clinical levels of anxiety-related problems associate with
delta-beta coupling.
The previous section review indicates the exciting promise of EEG frequency techniques
to illuminate anxiety-related processes, including motivation and cognitive control.
In this final section, we first consider relevant cognitive, motivational, and emotional
theories that might help synthesize the reviewed literature. We then turn to potential
avenues for future work and finish with some brief concluding remarks.
19.5.1 Theoretical Considerations
Throughout this review, much of the literature seems to point to the presence of a dy-
namic interplay between cognition, motivation, and emotion in the links between EEG
frequency metrics and anxious states and traits. In particular, the interplay between
emotional processes involved in the generation and detection of internal or external
threats, motivational processes dedicated to approaching and/or avoiding these threats,
and cognitive processes involved in the monitoring and control of these forces seems to
be positioned front and center in anxiety-related situations and symptoms. Figure 19.2
shows a hypothetical model of the interplay between motivational conflicts and cog-
nitive control in anxiety, which is seen as intimately related to the dynamic process of
co-activation of approach and avoidance motivational systems that produce conflicts
Anxeity
Apprehension/
Arousal
Resolution/
Non-Resolution
Figure 19.2 A hypothetical model of the interplay between motivational conflicts and cogni-
tive control in anxiety.
478 JASON S. MOSER et al.
detected and weighed by the BIS during which ongoing behavior is inhibited and
controlled until there is a resolution of the conflict and a behavior is selected. Anxiety
is generally related to enhanced conflicts between motivational systems, inhibition, and
control. Alpha asymmetry and metrics and theta/beta ratio may index BIS-related in-
hibitory and related control functions. Apprehension and arousal subcomponents of
anxiety are proposed to have different relations to these dynamics. The model more
likely captures the role of apprehension whereas arousal is proposed to have more direct
links to fear and avoidance motivational systems that may be less related to control
systems.
There is much overlap between the concepts of emotion, motivation, and cogni-
tion, and, thus, we do not intend to communicate that they are completely separable
constructs (cf. Miller, 1996, 2010). Rather, we assert that consideration of these processes
is critical to better understanding what EEG frequency metrics might reflect with re-
spect to anxiety, as well as to evaluate the utility of EEG frequency metrics in detecting
risk, maintenance factors, and treatment targets for anxiety-related interference and
disorders. In the next section, we consider merging the theoretical contributions of Gray
and McNaughton (2000), Eysenck and colleagues (2007), and Heller and colleagues
(Heller et al., 1997, Heller & Nitschke, 1998) with findings from the cognitive neurosci-
ence of inhibitory control (over emotion) as offering a unique window into advancing
our knowledge of how EEG frequency metrics might reflect the dynamic interplay of
emotion, motivation, and cognition in anxiety.
Gray and McNaughton’s (2000) neuropsychological model of anxiety draws on
earlier work by Gray (1975, 1982) regarding motivation and personality (see also Corr
et al., 2013 for a more recent review). This work largely delineated the BAS and BIS as
two motivational systems relevant to personality and anxiety. The updated version of
this model, the Reinforcement Sensitivity Theory (RST), further decomposed the BIS
into the BIS and the fight-flight-freeze system (FFFS), which is involved in fear-related
behaviors aimed at defending an organism from active threat. The BIS, on the other
hand, is posited to be involved in the detection and resolution of approach-avoidance
conflicts—that is, conflicts driven by coactivation of the BAS and FFFS. The BIS there-
fore inhibits ongoing behavior to allow for risk assessment and scanning so as to resolve
the approach-avoid conflict by enacting the appropriate behavior in context.
The attentional control theory (ACT) of anxiety (Eysenck et al., 2007) more specif-
ically aims to describe the nature of the relationship between anxiety and cognition.
Eysenck and colleagues conclude that anxiety disrupts the balance of the bottom-up
salience system and the top-down control system such that the bottom-up system is
generally more active than the top-down control system. To resolve this imbalance in
the presence of task-related goals, anxious individuals are hypothesized to deploy com-
pensatory top-down control to meet the demands of the task at hand. We suggest that
the interplay of the bottom-up and top-down control systems in anxiety proposed by
Eysenck and colleagues bears striking resemblance to the approach-avoid conflicts
detected and resolved by the BIS in Gray’s RST. That is, anxiety often involves conflicts
between threat-driven, avoidance, or distraction processes and goal-driven, approach
EEG FREQUENCY TECHNIQUES In Anxiety 479
metrics to index these important dynamics given their millisecond precision and in-
herent ability to reflect activity of systems/networks.
19.5.2 Future Directions
There is a strong foundation of knowledge produced by the extant studies on
associations between anxiety and EEG frequency metrics, although a number of areas
are ripe for further investigation and theory development. One particularly curious
contradiction is the functional significance of theta activity. In the literature on FMT,
theta is hypothesized to signal the need for control whereby a larger theta signal leads
to greater control and adaptive behavioral adjustments. On the other hand, in the CFC
literature, theta—i.e., a slow wave—is hypothesized to reflect emotional arousal and
related subcortical processes such that its preponderance or dominance over faster wave
activity indicates lesser control. It will be important to reconcile these two literatures, as
currently they appear to be evolving in parallel. This reconciliation may be challenging
at present given the differences in paradigms and calculations of theta used in these two
literatures. Both the FMT and CFC literatures focus on frontal theta, and, therefore, it
follows that there should be some relation between these different metrics.
It also seems important to gain new insights into the exact relationship between anx-
iety constructs and alpha asymmetry. Indeed, it represents the historically largest litera-
ture on the association between anxiety and EEG frequency metrics, yet much is still
left unknown. The various competing models of alpha asymmetry likely occlude clarity
with respect to its association with anxiety, however, perhaps Gray and McNaughton’s
(2000) model of motivation and personality could shed some light. For example, given
the important role of the BIS in anxiety and other related EEG frequency metrics, it
could be difficult to isolate associations between alpha asymmetry and anxiety be-
cause of the likely co-activation of approach (left-sided activity) and avoidance (right-
sided activity) motivational systems. The work of Heller and colleagues (Sharp et al.,
2015) further demonstrates the importance of considering the heterogeneity of anxiety
constructs when examining associations between anxiety and alpha asymmetry, among
other neural measures of motivation and cognition. Thus, future work will need to be
more precise with setting up paradigms that test clearer predictions involving specific
motivational contexts and anxiety constructs.
In general, the literature on associations between anxiety and EEG frequency metrics
will have to grapple with different motivational contexts and anxiety constructs.
Enhanced precision in theory will drive more sophisticated methods for testing related
predictions concerning approach-avoidance conflicts and separable dimensions of anx-
iety. For example, whereas some have shown associations between “lumped” anxiety
constructs and FMT (Cavanagh & Schackman, 2015), others have shown that “splitting”
anxiety constructs reveals important distinctions (e.g., Moser et al., 2013).
Beyond the EEG frequency metrics and anxiety constructs themselves, this litera-
ture must also incorporate the roles of development and gender, both of which show
EEG FREQUENCY TECHNIQUES In Anxiety 481
important influences on anxiety and its associations with neural measures of mo-
tivation/emotion and cognitive control. For instance, there is growing evidence
demonstrating important changes in the association between anxiety and the ERN—a
theta-based ERP—across development (Moser et al., 2017). Gender also moderates this
association, such that females show a larger association (Moser et al., 2016) and a greater
change in association across development (Ip et al., 2019). Furthermore, racial, ethnic,
and cultural differences will have to be considered to fully understand the generaliz-
ability of these findings. For example, related research shows higher heart rate variability
(HRV) in Black individuals, which may reflect compensatory control mechanisms
deployed to manage the sustained and cumulative adverse effects of discrimination
(Kemp et al., 2016). Such a compensatory control response may also be reflected in the
EEG frequency metrics reviewed herein.
Additional methodological innovations will also further advance this area of clinical
neuroscience. For instance, the application of graph theoretic (Bolanos et al., 2013) and
directed network (Liu et al., 2014; Popov et al., 2018) methods to such EEG frequency
techniques is just beginning and will likely provide unique insights into associations
between anxiety and motivational and cognitive control systems. This work will also
further benefit from integrating data across the lab and the “real-world” through
combining EEG imaging techniques with experience-sampling methods and deploying
wearable technology (cf. Hur et al., 2019).
Finally, we would be remiss if we did not acknowledge the importance of considering
the psychological and philosophical machinations involved in attempting to make
meaning of associations between various psychological and biological measures. It is
important for future work in this area to avoid terminology such as “underlying” in
describing the involvement of neural oscillations in self-reported experiences of anx-
iety as they incorrectly insinuate a causal mechanism and further the fallacy that more
micro-level data (e.g., neural activity) are superior to more macro-level data (e.g., self-
report of anxiety; Sharp & Miller, 2019). Rather, such neural markers are likely involved
in the instantiation of experiences and resolution of motivational conflicts and anxiety
symptoms. The ways in which these neural, self-report, and behavioral measures are
combined and weighed should involve careful consideration of the goals to be achieved
(e.g., intervention development; See Taschereau-Dumouchel, Michel, Lau, Hofmann &
LeDoux, 2022 for a recent review of this approach).
19.5.3 Conclusions
The literature considering associations between anxiety and EEG frequency measures
has a long history and generally follows the historical trends present in other areas of
psychological science. More recently, early motivational and emotional models have
been overtaken by or integrated into cognitive models of anxiety and related constructs.
A long-view perspective on the literature suggests strong associations between anxiety-
related constructs and EEG frequency markers of motivational and cognitive control
482 JASON S. MOSER et al.
processes. A central theme in this literature is the involvement of the BIS that can be
seen reflected in a variety of EEG frequency metrics and generally reveals the dynamic
interplay between the co-activation of approach and avoidance motivations in anxious
individuals and states as the execution of adaptive behavioral responses is attempted.
This area of clinical neuroscience is ripe for future work and should involve advances in
theory, generalizability, and methodology.
REFERENCES
Adolph, D. & Margraf, J. (2017). The differential relationship between trait anxiety, depression,
and resting frontal α-asymmetry. Journal of Neural Transmission, 124(3), 379–386. https://
doi.org/10.1007/s00702-016-1664-9
Aftanas, L. I., Pavlov, S. V., Reva, N. V., & Varlamov, A. A. (2003). Trait anxiety impact on the
EEG theta band power changes during appraisal of threatening and pleasant visual stimuli.
International Journal of Psychophysiology, 50(3), 205–212.
Allen, J. J. B., Coan, J. A., & Nazarian, M. (2004). Issues and assumptions on the road from raw
signals to metrics of frontal EEG asymmetry in emotion. Biological Psychology, 67(1–2), 183–
218. https://doi.org/10.1016/j.biopsycho.2004.03.007
Allen, J. J. B. & Reznik, S. J. (2015). Frontal EEG asymmetry as a promising marker of depression
vulnerability: Summary and methodological considerations. Current Opinion in Psychology,
4, 93–97.https://doi.org/10.1016/j.copsyc.2014.12.017
Amodio, D. M. (2010). Coordinated roles of motivation and perception in the regulation of
intergroup responses: Frontal cortical asymmetry effects on the P2 event-related potential
and behavior. Journal of Cognitive Neuroscience, 22(11), 2609–2617. https://doi.org/10.1162/
jocn.2009.21395
Amodio, D. M., Devine, P. G., & Harmon- Jones, E. (2007). A dynamic model of
guilt: Implications for motivation and self-regulation in the context of prejudice: Research
article. Psychological Science, 8(6), 524–530. https://doi.org/10.1111/j.1467-9280.2007.01933.x
Angelidis, A., Hagenaars, M., van Son, D., van der Does, W., & Putman, P. (2018). Do not look
away! Spontaneous frontal EEG theta/beta ratio as a marker for cognitive control over
attention to mild and high threat. Biological Psychology, 135, 8–17.
Avram, J., Balteş, F. R., Miclea, M., & Miu, A. C. (2010). Frontal EEG activation asymmetry
reflects cognitive biases in anxiety: Evidence from an emotional face Stroop task. Applied
Psychophysiology Biofeedback, 35, 285–292. https://doi.org/10.1007/s10484-010-9138-6
Balconi, M. & Crivelli, D. (2010). FRN and P300 ERP effect modulation in response to feed-
back sensitivity: The contribution of punishment-reward system (BIS/BAS) and behaviour
identification of action. Neuroscience Research, 66(2), 162–172.
Barlow, D. H. (1991). Disorders of emotion. Psychological Inquiry, 2(1), 58–7 1. https://doi.org/
10.1207/s15327965pli0201_15
Barlow, D. H. (2002). Anxiety and its disorders (2nd ed.). Guilford Press.
Bartholomew, M. E., Yee, C. M., Heller, W., Miller, G. A., & Spielberg, J. M. (2019).
Reconfiguration of brain networks supporting inhibition of emotional challenge.
NeuroImage, 186, 350–357.
Baving, L., Laucht, M., & Schmidt, M. H. (2002). Frontal brain activation in anxious school
children. Journal of Child Psychology and Psychiatry and Allied Disciplines, 43(2), 265–274.
https://doi.org/10.1111/1469-7610.00019
EEG FREQUENCY TECHNIQUES In Anxiety 483
Blackhart, G. C., Minnix, J. A., & Kline, J. P. (2006). Can EEG asymmetry patterns predict fu-
ture development of anxiety and depression? A preliminary study. Biological Psychology,
72(1), 46–50. https://doi.org/10.1016/j.biopsycho.2005.06.010
Bolanos, M., Bernat, E.M., He, B., & Aviyente, S. (2013). A weighted small world network
measure for assessing functional con-nectivity. Journal of Neuroscience Methods, 212, 133–42.
Botvinick, M. M., Cohen, J. D., & Carter, C. S. (2004). Conflict monitoring and anterior cingu-
late cortex: an update. Trends in Cognitive Sciences, 8(12), 539–546.
Brookshire, G., & Casasanto, D. (2012). Motivation and motor control: Hemispheric special-
ization for approach motivation reverses with handedness. PLoS One, 7(4), e36036. https://
doi.org/10.1371/journal.pone.0036036
Bruder, G. E., Fong, R., Tenke, C. E., Leite, P., Towey, J. P., Stewart, J. E., . . . Quitkin, F. M. (1997).
Regional brain asymmetries in major depression with or without an anxiety disorder: A
quantitative electroencephalographic study. Biological Psychiatry, 41, 939–948. https://doi.
org/10.1016/S0006-3223(96)00260-0
Buchsbaum, M. S., Hazlett, E., Sicotte, N., Stein, M., Wu, J., & Zetin, M. (1985). Topographic
EEG changes with benzodiazepine administration in generalized anxiety disorder. Biological
Psychiatry, 20(8), P837–842. https://doi.org/10.1016/0006-3223(85)90208-2
Buzzell, G. A., Barker, T. V., Troller-Renfree, S. V., Bernat, E. M., Bowers, M. E., Morales,
S., . . . Fox, N. A. (2018). Adolescent cognitive control, theta oscillations, and social motiv-
ation. bioRxiv [online], 366831.
Cavanagh, J. F., Bismark, A. W., Frank, M. J., & Allen, J. J. (2019). Multiple dissociations be-
tween comorbid depression and anxiety on reward and punishment processing: Evidence
from computationally informed EEG. Computational Psychiatry, 3, 1–17.
Cavanagh, J. F. & Frank, M. J. (2014). Frontal theta as a mechanism for cognitive control. Trends
in Cognitive Sciences, 18(8), 414–421.
Cavanagh, J. F., Meyer, A., & Hajcak, G. (2017). Error-specific cognitive control alterations
in generalized anxiety disorder. Biological Psychiatry: Cognitive Neuroscience and
Neuroimaging, 2(5), 413–420.
Cavanagh, J. F. & Shackman, A. J. (2015). Frontal midline theta reflects anxiety and cognitive
control: Meta-analytic evidence. Journal of Physiology-Paris, 109(1-3), 3–15.
Cavanagh, J. F., Zambrano‐Vazquez, L., & Allen, J. J. (2012). Theta lingua franca: A common
mid‐frontal substrate for action monitoring processes. Psychophysiology, 49(2), 220–238.
Carter, W. R., Johnson, M. C., & Borkovec, T. D. (1986). Worry: An electrocortical analysis.
Advances in Behaviour Research and Therapy, 8(4), 193–204. https://doi.org/10.1016/
0146-6402(86)90004-4
Cohen, M. X. & Donner, T. H. (2013). Midfrontal conflict-related theta-band power reflects
neural oscillations that predict behavior. Journal of Neurophysiology, 110(12), 2752–2763.
Corr, P. J., DeYoung, C. G., & McNaughton, N. (2013). Motivation and personality: A neuro-
psychological perspective. Social and Personality Psychology Compass, 7(3), 158–175.
Davidson, R. J. (1979). Frontal versus perietal EEG asymmetry during positive and negative
affect. Psychophysiology, 16(2), 202–203.
Davidson, R. J. (1984). Affect, Cognition, and Hemispheric Specialization. In C. E. Izard, J.
Kagan, & R. B. Zajonc (Eds.), Emotions, cognition, and behavior (pp. 320–365). Cambridge
University Press.
Davidson, R. J. (1988). EEG measures of cerebral asymmetry: Conceptual and methodo-
logical issues. International Journal of Neuroscience, 39(1–2), 71–89. https://doi.org/10.3109/
00207458808985694
484 JASON S. MOSER et al.
Davidson, R. J. & Fox, N. A. (1982). Asymmetrical brain activity discriminates between positive
and negative affective stimuli in human infants. Science, 218(4578), 1235–1237. https://doi.
org/10.1126/science.7146906
Davidson, R. J. & Fox, N. A. (1989). Frontal brain asymmetry predicts infants’ response to ma-
ternal separation. Journal of Abnormal Psychology, 98(2), 127–131. https://doi.org/10.1037/
0021-843X.98.2.127
Davidson, R. J., Kalin, N. H., & Shelton, S. E. (1992). Lateralized effects of diazepam on frontal
brain electrical asymmetries in rhesus monkeys. Biological Psychiatry, 32(5), 438–451.
https://doi.org/10.1016/0006-3223(92)90131-I
Davidson, R. J., Marshall, J. R., Tomarken, A. J., & Henriques, J. B. (2000). While a phobic
waits: Regional brain electrical and autonomic activity in social phobics during antici-
pation of public speaking. Biological Psychiatry, 47(2), 85–95. https://doi.org/10.1016/
S0006-3223(99)00222-X
De Pascalis, V., Varriale, V., & D’Antuono, L. (2010). Event-related components of the punish-
ment and reward sensitivity. Clinical Neurophysiology, 121(1), 60–76.
Dreisbach, G. & Fischer, R. (2012). Conflicts as aversive signals. Brain and Cognition,
78(2), 94–98.
Ekman, P. & Friesen, W. V. (1971). Constants across cultures in the face and emotion. Journal of
Personality and Social Psychology, 17(2), 124–129. https://doi.org/10.1037/h0030377
Ellis, J. S., Watts, A. T., Schmidt, N., & Bernat, E. M. (2018). Anxiety and feedback processing
in a gambling task: Contributions of time-frequency theta and delta. Biological Psychology,
136, 1–12.
Engel, A. K. & Fries, P. (2010). Beta-band oscillations—signaling the status quo? Current
Opinion in Neurobiology, 20(2), 156–165.
Eysenck, M. W., Derakshan, N., Santos, R., & Calvo, M. G. (2007). Anxiety and cognitive per-
formance: Attentional control theory. Emotion, 7(2), 336.
Field, T., Martinez, A., Nawrocki, T., Pickens, J., Fox, N. A., & Schanberg, S. (1998). Music shifts
frontal EEG in depressed adolescents. Adolescence, 33(129), 109–116.
Foti, D., Weinberg, A., Bernat, E. M., & Proudfit, G. H. (2014). Anterior cingulate activity to
monetary loss and basal ganglia activity to monetary gain uniquely contribute to the feed-
back negativity. Clinical Neurophysiology, 126(7), 1338–1347.
Fox, N. A. (1991). If it’s not left, it’s right. American Psychologist, 46(8), 863–872. https://doi.org/
10.1037/0003-066X.46.8.863
Fox, N. A., Bell, M. A., & Jones, N. A. (1992). Individual differences in response to stress and
cerebral asymmetry. Developmental Neuropsychology, 8(2–3), 161–184. https://doi.org/
10.1080/87565649209540523
Fox, N. A., & Davidson, R. J. (1987). Electroencephalogram asymmetry in response to the
approach of a stranger and maternal separation in 10-month-old infants. Developmental
Psychology, 23(2), 233–240. https://doi.org/10.1037/0012-1649.23.2.233
Fox, N. A., Rubin, K. H., Calkins, S. D., Marshall, T. R., Coplan, R. J., Porges, S. W., . . . Stewart,
S. (1995). Frontal activation asymmetry and social competence at four years of age. Child
Development, 66(6), 1770-1784. https://doi.org/10.1111/j.1467-8624.1995.tb00964.x
Fox, N. A., Schmidt, L. A., Calkins, S. D., Rubin, K. H., & Coplan, R. J. (1996). The role of frontal
activation in the regulation and dysregulation of social behavior during the preschool
years. Development and Psychopathology, 8(1), 89–102. https://doi.org/10.1017/S095457940
0006982
EEG FREQUENCY TECHNIQUES In Anxiety 485
Gable, P., & Harmon- Jones, E. (2008). Relative left frontal activation to appetitive
stimuli: Considering the role of individual differences. Psychophysiology, 45(2), 275–278.
https://doi.org/10.1111/j.1469-8986.2007.00627.x
Gainotti, G. (1972). Emotional Behavior and Hemispheric Side of the Lesion. Cortex, 8(1), 41–
55. https://doi.org/10.1016/S0010-9452(72)80026-1
Gehring, W. J., Goss, B., Coles, M. G. H., Meyer, D. E., & Donchin, E. (2018). The error-related
negativity. Perspectives in Psychological Science, 13(2), 200–204. doi:10.1177/1745691617715310
Gorka, S. M., Phan, K. L., & Shankman, S. A. (2015). Convergence of EEG and fMRI measures
of reward anticipation. Biological Psychology, 112, 12–19. https://doi.org/10.1016/j.biopsy
cho.2015.09.007
Gray, J. A. (1975). Elements of a two-process theory of learning. Academic Press.
Gray, J. A. (1982). The neuropsychology of anxiety: An enquiry into the functions of the septo-
hippocampal system. Oxford University Press.
Gray, J. A. & McNaughton, N. (2000). The neuropsychology of anxiety: An enquiry into the
functions of the septo-hippocampal system (2nd ed.). Oxford University Press.
Grupe, D. W. & Nitschke, J. B. (2013). Uncertainty and anticipation in anxiety: an integrated
neurobiological and psychological perspective. Nature reviews Neuroscience, 14(7), 488.
Hall, E. E., Ekkekakis, P., & Petruzzello, S. J. (2010). Predicting affective responses to exercise
using resting EEG frontal asymmetry: Does intensity matter? Biological Psychology, 83(3),
201–206. https://doi.org/10.1016/j.biopsycho.2010.01.001
Harmon-Jones, E. (2007). Trait anger predicts relative left frontal cortical activation to anger-
inducing stimuli. International Journal of Psychophysiology, 66(2), 154–160. https://doi.org/
10.1016/j.ijpsycho.2007.03.020
Harper, J., Malone, S. M., & Bernat, E. M. (2014). Theta and delta band activity explain N2 and
P3 ERP component activity in a go/no-go task. Clinical Neurophysiology, 125(1), 124–132.
Harrewijn, A., Van der Molen, M. J. W., & Westenberg, P. M. (2016). Putative EEG measures of
social anxiety: Comparing frontal alpha asymmetry and delta–beta cross-frequency correl-
ation. Cognitive, Affective, & Behavioral Neuroscience, 16(6), 1086–1098.
Harrewijn, A., van der Molen, M. J., van Vliet, I. M., Houwing-Duistermaat, J. J., & Westenberg,
P. M. (2018a). Delta-beta correlation as a candidate endophenotype of social anxiety: A two-
generation family study. Journal of Affective Disorders, 227, 398–405.
Harrewijn, A., van der Molen, M. J., van Vliet, I. M., Tissier, R. L., & Westenberg, P. M. (2018b).
Behavioral and EEG responses to social evaluation: A two-generation family study on social
anxiety. NeuroImage: Clinical, 17, 549–562.
Heller, W. (1993). Neuropsychological mechanisms of individual differences in emotion,
personality, and arousal. Neuropsychology, 7(4), 476–489. https://doi.org/10.1037/
0894-4105.7.4.476
Heller, W. & Nitschke, J. B. (1998). The puzzle of regional brain activity in and anxiety: The im-
portance of subtypes and comorbidity. Cognition & Emotion, 12(3), 421–447.
Heller, W., Nitschke, J. B., Etienne, M. A., & Miller, G. A. (1997). Patterns of regional brain ac-
tivity differentiate types of anxiety. Journal of Abnormal Psychology, 106(3), 376–385. https://
doi.org/10.1037/0021-843X.106.3.376
Hirshfeld-Becker, D. R., Biederman, J., Henin, A., Faraone, S. V., Davis, S., Harrington, K., &
Rosenbaum, J. F. (2007). Behavioral inhibition in preschool children at risk is a specific pre-
dictor of middle childhood social anxiety: A five-year follow-up. Journal of Developmental and
Behavioral Pediatrics, 28(3), 225–233. https://doi.org/10.1097/01.DBP.0000268559.34463.d0
486 JASON S. MOSER et al.
Hofmann, S. G., Moscovitch, D. A., Litz, B. T., Kim, H. J., Davis, L. L., & Pizzagalli, D. A. (2005).
The worried mind: Autonomic and prefrontal activation during worrying. Emotion, 5(4),
464–475. https://doi.org/10.1037/1528-3542.5.4.464
Hur, J., Stockbridge, M. D., Fox, A. S. & Shackman, A. J. (2019). Dispositional negativity, cogni-
tion, and anxiety disorders: An integrative translational neuroscience framework. Progress
in Brain Research, 247, 375–436.
Ip, K. I., Liu, Y., Moser, J., Mannella, K., Hruschak, J., Bilek, E., . . . Fitzgerald, K. (2019).
Moderation of the relationship between the error-related negativity and anxiety by age
and gender in young children: A preliminary investigation. Developmental Cognitive
Neuroscience, 39, 100702.
Ischebeck, M., Endrass, T., Simon, D., & Kathmann, N. (2014). Altered frontal EEG asymmetry
in obsessive-compulsive disorder. Psychophysiology, 51(7), 596–601. https://doi.org/10.1111/
psyp.12214
Isotani, T., Tanaka, H., Lehmann, D., Pascual-Marqui, R. D., Kochi, K., Saito, N., . . . Sasada,
K. (2001). Source localization of EEG activity during hypnotically induced anxiety and re-
laxation. International Journal of Psychophysiology, 41(2), 143–153. https://doi.org/10.1016/
S0167-8760(0 0)00197-5
Jenkinson, N. & Brown, P. (2011). New insights into the relationship between dopamine, beta
oscillations and motor function. Trends in Neurosciences, 34(12), 611–618.
Jensen, O., Kaiser, J., & Lachaux, J. P. (2007). Human gamma-frequency oscillations associated
with attention and memory. Trends in Neurosciences, 30(7), 317–324.
Judah, M. R., Grant, D. M., Frosio, K. E., White, E. J., Taylor, D. L., & Mills, A. C. (2016).
Electrocortical evidence of enhanced performance monitoring in social anxiety. Behavior
Therapy, 47(2), 274–285.
Kalin, N. H., Larson, C., Shelton, S. E., & Davidson, R. J. (1998). Asymmetric frontal brain
activity, cortisol, and behavior associated with fearful temperament in rhesus monkeys.
Behavioral Neuroscience, 112(2), 286–292. https://doi.org/10.1037/0735-7044.112.2.286
Kemp, A. H., Griffiths, K., Felmingham, K. L., Shankman, S. A., Drinkenburg, W., Arns,
M., . . . Bryant, R. A. (2010). Disorder specificity despite comorbidity: Resting EEG alpha
asymmetry in major depressive disorder and post-traumatic stress disorder. Biological
Psychology, 85(2), 350–354. https://doi.org/10.1016/j.biopsycho.2010.08.001
Kemp, A. H., Koenig, J., Thayer, J. F., Bittencourt, M. S., Pereira, A. C., Santos, I. S., ... &
Benseñor, I. M. (2016). Race and resting-state heart rate variability in Brazilian civil servants
and the mediating effects of discrimination: An ELSA-Brasil cohort study. Psychosomatic
Medicine, 78(8), 950–958.
Keune, P. M., Schönenberg, M., Wyckoff, S., Mayer, K., Riemann, S., Hautzinger, M., & Strehl,
U. (2011). Frontal alpha-asymmetry in adults with attention deficit hyperactivity dis-
order: Replication and specification. Biological Psychology, 87(2), 306–310. https://doi.org/
10.1016/j.biopsycho.2011.02.023
Knyazev, G. G., Savostyanov, A. N., & Levin, E. A. (2005). Uncertainty, anxiety and brain
oscillations. Neuroscience Letters, 387, 121–125.
Knyazev, G. G. & Slobodskaya, H. R. (2003). Personality trait of behavioral inhibition is
associated with oscillatory systems reciprocal relationships. International Journal of
Psychophysiology, 48(3), 247–261.
Knyazev, G. G., Schutter, D. J., & van Honk, J. (2006). Anxious apprehension increases coup-
ling of delta and beta oscillations. International Journal of Psychophysiology, 61(2), 283–287
EEG FREQUENCY TECHNIQUES In Anxiety 487
Knyazev, G. G. (2007). Motivation, emotion, and their inhibitory control mirrored in brain
oscillations. Neuroscience & Biobehavioral Reviews, 31(3), 377–395.
Knyazev, G. G. (2011). Cross-frequency coupling of brain oscillations: an impact of state anx-
iety. International Journal of Psychophysiology, 80(3), 236–245.
Kool, W., McGuire, J. T., Rosen, Z. B., & Botvinick, M. M. (2010). Decision making and the
avoidance of cognitive demand. Journal of Experimental Psychology, 139(4), 665.
Kuskowski, M. A., Malone, S. M., Kim, S. W., Dysken, M. W., Okaya, A. J., & Christensen, K. J.
(1993). Quantitative EEG in obsessive-compulsive disorder. Biological Psychiatry, 33(6), 423–
430. https://doi.org/10.1016/0006-3223(93)90170-I
Laufs, H., Kleinschmidt, A., Beyerle, A., Eger, E., Salek-Haddadi, A., Preibisch, C., & Krakow,
K. (2003). EEG-correlated fMRI of human alpha activity. NeuroImage, 19(4), 1463-1476.
https://doi.org/10.1016/S1053-8119(03)00286-6
Laurienti, P. J., Wallace, M. T., Maldjian, J. A., Susi, C. M., Stein, B. E., & Burdette, J. H. (2003).
Cross‐modal sensory processing in the anterior cingulate and medial prefrontal cortices.
Human Brain Mapping, 19(4), 213–223.
Ledoux, J. (1978). The neurology of emotion. Casa Editrice.
Liu, Y., Moser, J. S., Aviyente, S. (2014). Network community structure detection for directional
neural networks inferred from multichannel multisubject EEG data. IEEE Transactions on
Biomedical Engineering, 61, 1919–1930.
Lindström, B. R., Mattsson-Mårn, I. B., Golkar, A., & Olsson, A. (2013). In your face: Risk
of punishment enhances cognitive control and error-related activity in the corrugator
supercilii muscle. PloS One, 8(6), e65692.
Lobo, I., Portugal, L. C., Figueira, I., Volchan, E., David, I., Garcia Pereira, M., & De Oliveira, L.
(2015). EEG correlates of the severity of posttraumatic stress symptoms: A systematic review
of the dimensional PTSD literature. Journal of Affective Disorders, 183, 210–220. https://doi.
org/10.1016/j.jad.2015.05.015
LoBue, V., Coan, J. A., Thrasher, C., & DeLoache, J. S. (2011). Prefrontal asymmetry and Parent-Rated
temperament in infants. PLoS One, 6(7): e22694. https://doi.org/10.1371/journal.pone.0022694
Lubar, J. F., Congedo, M., & Askew, J. H. (2003). Low-resolution electromagnetic tomog-
raphy (LORETA) of cerebral activity in chronic depressive disorder. International Journal of
Psychophysiology, 49(3), 175–185. https://doi.org/10.1016/S0167-8760(03)00115-6
Luu, P., Tucker, D. M., & Makeig, S. (2004). Frontal midline theta and the error-related nega-
tivity: neurophysiological mechanisms of action regulation. Clinical Neurophysiology, 115(8),
1821–1835.
Mantini, D., Perrucci, M. G., Del Gratta, C., Romani, G. L., & Corbetta, M. (2007).
Electrophysiological signatures of resting state networks in the human brain. Proceedings
of the National Academy of Sciences of the United States of America, 104(32), 13170–13175.
doi:10.1073/pnas.0700668104
Mathersul, D., Williams, L. M., Hopkinson, P. J., & Kemp, A. H. (2008). Investigating models of
affect: Relationships among EEG alpha asymmetry, depression, and anxiety. Emotion, 8(4),
560–572. https://doi.org/10.1037/a0012811
McManis, M. H., Kagan, J., Snidman, N. C., & Woodward, S. A. (2002). EEG asymmetry,
power, and temperament in children. Developmental Psychobiology, 41(2), 169–177. https://
doi.org/10.1002/dev.10053
Metzger, L. J., Carson, M. A., Paulus, L. A., Paige, S. R., Lasko, N. B., Pitman, R. K., & Orr,
S. P. (2004). PTSD arousal and depression symptoms associated with increased right-sided
488 JASON S. MOSER et al.
Moser, J. S., Moran, T. P., Schroder, H. S., Donnellan, M. B., & Yeung, N. (2013). On the relation-
ship between anxiety and error monitoring: A meta-analysis and conceptual framework.
Frontiers in Human Neuroscience, 7, 466. doi:10.3389/fnhum.2013.00466
Moser, J. S., Moran, T. P., Kneip, C., Schroder, H. S., & Larson, M. J. (2016). Sex moderates the
association between symptoms of anxiety, but not obsessive compulsive disorder, and error-
monitoring brain activity: A meta-analytic review. Psychophysiology, 53, 21–29.
Moser, J.S. (2017). The nature of the relationship between anxiety and the error-related nega-
tivity across development. Current Behavior Neuroscience Reports, 4(4), 309–321.
Nash, K., Inzlicht, M., & McGregor, I. (2012). Approach-related left prefrontal EEG asymmetry
predicts muted error-related negativity. Biological Psychology, 91(1), 96–102. https://doi.org/
10.1016/j.biopsycho.2012.05.005
Nitschke, J. B., Heller, W., Palmieri, P. A., & Miller, G. A. (1999). Contrasting patterns of brain
activity in anxious apprehension and anxious arousal. Psychophysiology, 36(5), 628–637.
https://doi.org/10.1017/S0048577299972013
Nusslock, R., Shackman, A. J., McMenamin, B. W., Greischar, L. L., Davidson, R. J., & Kovacs,
M. (2018). Comorbid anxiety moderates the relationship between depression history
and prefrontal EEG asymmetry. Psychophysiology, 55(1), e12953. https://doi.org/10.1111/
psyp.12953
Nusslock, R., Walden, K., & Harmon-Jones, E. (2015). Asymmetrical frontal cortical activity
associated with differential risk for mood and anxiety disorder symptoms: An RDoC per-
spective. International Journal of Psychophysiology, 98(2, Pt.2), 249–261. https://doi.org/
10.1016/j.ijpsycho.2015.06.004
Peterson, C. K. & Harmon-Jones, E. (2009). Circadian and seasonal variability of resting
frontal EEG asymmetry. Biological Psychology, 80(3), 315–320. https://doi.org/10.1016/
j.biopsycho.2008.11.002
Poppelaars, E. S., Harrewijn, A., Westenberg, P. M., & van der Molen, M. J. W. (2018). Frontal
delta-beta cross-frequency coupling in high and low social anxiety: An index of stress regu-
lation? Cognitive, Affective & Behavioral Neuroscience, 18(4), 764–777.
Popov, T., Westner, B. U., Silton, R. L., Sass, S. M., Spielberg, J. M., Rockstroh, B., Heller, W., &
Miller, G. A. (2018). Time course of brain network reconfiguration supporting inhibitory
control. Journal of Neuroscience, 38, 4348–4356.
Putman, P., Verkuil, B., Arias-Garcia, E., Pantazi, I., & van Schie, C. (2014). EEG theta/beta
ratio as a potential biomarker for attentional control and resilience against deleterious
effects of stress on attention. Cognitive, Affective, & Behavioral Neuroscience, 14(2), 782–791.
Putman, P., Arias-Garcia, E., Pantazi, I., & van Schie, C. (2012). Emotional Stroop interference
for threatening words is related to reduced EEG delta–beta coupling and low attentional
control. International Journal of Psychophysiology, 84, 194–200.
Putman, P., van Peer, J., Maimari, I., & van der Werff, S. (2010). EEG theta/beta ratio in relation
to fear-modulated response-inhibition, attentional control, and affective traits. Biological
Psychology, 83(2), 73–78.
Rabe, S., Beauducel, A., Zöllner, T., Maercker, A., & Karl, A. (2006). Regional brain electrical
activity in posttraumatic stress disorder after motor vehicle accident. Journal of Abnormal
Psychology, 115(4), 687–698. https://doi.org/10.1037/0021-843X.115.4.687
Rabe, S., Zöllner, T., Maercker, A., & Karl, A. (2006). Neural correlates of posttraumatic growth
after severe motor vehicle accidents. Journal of Consulting and Clinical Psychology, 74(5),
880–886. https://doi.org/10.1037/0022-006X.74.5.880
490 JASON S. MOSER et al.
Reznik, S. J., & Allen, J. J. B. (2018). Frontal asymmetry as a mediator and moderator of
emotion: An updated review. Psychophysiology, 55(1), e12965. https://doi.org/10.1111/
psyp.12965
Riesel, A., Goldhahn, S., & Kathmann, N. (2017). Hyperactive performance monitoring as a
transdiagnostic marker: Results from health anxiety in comparison to obsessive–compulsive
disorder. Neuropsychologia, 96, 1–8.
Righi, S., Mecacci, L., & Viggiano, M. P. (2009). Anxiety, cognitive self-evaluation and per-
formance: ERP correlates. Journal of Anxiety Disorders, 23(8), 1132–1138.
Sackeim, H. A., Greenberg, M. S., Weiman, A. L., Gur, R. C., Hungerbuhler, J. P., & Geschwind,
N. (1982). Hemispheric asymmetry in the expression of positive and negative emotions.
Neurologic evidence. Archives of Neurology, 39(4), 210–218.
Santesso, D. L., Bogdan, R., Birk, J. L., Goetz, E. L., Holmes, A. J., & Pizzagalli, D. A. (2011).
Neural responses to negative feedback are related to negative emotionality in healthy adults.
Social Cognitive and Affective Neuroscience, 7(7), 794–803.
Scheeringa, R., Petersson, K. M., Kleinschmidt, A., Jensen, O., & Bastiaansen, M. C. M. (2012).
EEG alpha power modulation of fMRI resting-state connectivity. Brain Connectivity, 2(5),
254–264. https://doi.org/10.1089/brain.2012.0088
Schmid, P. C., Kleiman, T., & Amodio, D. M. (2015). Neural mechanisms of proactive and re-
active cognitive control in social anxiety. Cortex, 70, 137–145.
Schmidt, L. A., Fox, N. A., Schulkin, J., & Gold, P. W. (1999). Behavioral and psycho-
physiological correlates of self- presentation in temperamentally shy children.
Developmental Psychobiology, 35(2), 119–135. https://doi.org/10.1002/(SICI)1098-2302(199
909)35:2<119::AID-DEV5>3.0.CO;2-G
Schmidt, B., Kanis, H., Holroyd, C. B., Miltner, W. H., & Hewig, J. (2018). Anxious
gambling: Anxiety is associated with higher frontal midline theta predicting less risky
decisions. Psychophysiology, 55(10), e13210.
Schouppe, N., De Houwer, J., Richard Ridderinkhof, K., & Notebaert, W. (2012).
Conflict: Run! Reduced Stroop interference with avoidance responses. The Quarterly
Journal of Experimental Psychology, 65(6), 1052–1058.
Schroder, H. S., Glazer, J. E., Bennett, K. P., Moran, T. P., & Moser, J. S. (2017). Suppression of
error-preceding brain activity explains exaggerated error monitoring in females with worry.
Biological Psychology, 122, 33–41. doi:10.1016/j.biopsycho.2016.03.013
Schutter, D. J. & Honk, J. V. (2004). Decoupling of midfrontal delta–beta oscillations after tes-
tosterone administration. International Journal of Psychophysiology, 53, 71–73.
Schutter, D. J. & Van Honk, E. J. (2005). Salivary cortisol levels and the coupling of midfrontal
delta-beta oscillations. International Journal of Psychophysiology, 55, 127–129.
Sehlmeyer, C., Konrad, C., Zwitserlood, P., Arolt, V., Falkenstein, M., & Beste, C. (2010).
ERP indices for response inhibition are related to anxiety- related personality traits.
Neuropsychologia, 48(9), 2488–2495.
Shackman, A. J., Salomons, T. V., Slagter, H. A., Fox, A. S., Winter, J. J., & Davidson, R. J. (2011).
The integration of negative affect, pain and cognitive control in the cingulate cortex. Nature
Reviews Neuroscience, 12(3), 154–167. doi:10.1038/nrn2994
Shankman, S. A., Silverstein, S. M., Williams, L. M., Hopkinson, P. J., Kemp, A. H.,
Felmingham, K. L., . . . Clark, C. R. (2008). Resting electroencephalogram asymmetry and
posttraumatic stress disorder. Journal of Traumatic Stress, 21(2), 190–198. https://doi.org/
10.1002/jts.20319
EEG FREQUENCY TECHNIQUES In Anxiety 491
Sharp, P. B., Miller, G. A., & Heller, W. (2015). Transdiagnostic dimensions of anx-
iety: neural mechanisms, executive functions, and new directions. International Journal of
Psychophysiology, 98(2), 365–377.
Sharp, P. B. & Miller, G. A. (2019). Reduction and autonomy in psychology and neuroscience: A
call for pragmatism. Journal of Theoretical and Philosophical Psychology, 39(1), 18.
Shi, R., Sharpe, L., & Abbott, M. (2019). A meta-analysis of the relationship between anxiety
and attentional control. Clinical Psychology Review, 72, 101754.
Silberman, E. K. & Weingartner, H. (1986). Hemispheric lateralization of functions related to
emotion. Brain and Cognition, 5(3), 322–353. https://doi.org/10.1016/0278-2626(86)90035-7
Smith, E. E., Cavanagh, J. F., & Allen, J. J. B. (2018). Intracranial source activity (eLORETA)
related to scalp-level asymmetry scores and depression status. Psychophysiology, 55(1),
e13019. https://doi.org/10.1111/psyp.13019
Smith, E. E., Zambrano-Vazquez, L., & Allen, J. J. B. (2016). Patterns of alpha asymmetry in
those with elevated worry, trait anxiety, and obsessive-compulsive symptoms: A test of the
worry and avoidance models of alpha asymmetry. Neuropsychologia, 85, 118–126. https://doi.
org/10.1016/j.neuropsychologia.2016.03.010
Spielberg, J. M., Miller, G. A., Heller, W., & Banich, M. T. (2015). Flexible brain network recon-
figuration supporting inhibitory control. Proceedings of the National Academy of Sciences of
the United States of America, 112(32), 10020–10025.
Spielberger, C. D. (1983). State-Trait Anxiety Inventory (STAI). Mind Garden, 94061(650), 261–
3500. https://doi.org/10.1002/9780470479216.corpsy0943
Steiner, A. R. W., & Coan, J. A. (2011). Prefrontal asymmetry predicts affect, but not beliefs
about affect. Biological Psychology, 88, 65–7 1. https://doi.org/10.1016/j.biopsycho.2011.06.010
Suetsugi, M., Mizuki, Y., Ushijima, I., Kobayashi, T., Tsuchiya, K., Aoki, T., & Watanabe, Y.
(2000). Appearance of frontal midline theta activity in patients with generalized anxiety dis-
order. Neuropsychobiology, 41(2), 108–112.
Taschereau-Dumouchel, V., Michel, M., Lau, H., Hofmann, S. G., & LeDoux, J. E. (2022).
Putting the “mental” back in “mental disorders”: a perspective from research on fear and
anxiety. Molecular Psychiatry. https://doi.org/10.1038/s41380-021-01395-5
Trujillo, L. T. & Allen, J. J. (2007). Theta EEG dynamics of the error-related negativity. Clinical
Neurophysiology, 118(3), 645–668.
U.S. Burden of Disease Collaborators. (2018). The state of US health, 1990-2016. Burden
of diseases, injuries, and risk factors among US states. Journal of the American Medical
Association, 319, 1444–1472.
Van Den Heuvel, M. P., Mandl, R. C., Kahn, R. S., & Hulshoff Pol, H. E. (2009). Functionally
linked resting‐state networks reflect the underlying structural connectivity architecture of
the human brain. Human Brain Mapping, 30(10), 3127–3141.
van der Molen, M. J., Harrewijn, A., & Westenberg, P. M. (2018). Will they like me? Neural and
behavioral responses to social-evaluative peer feedback in socially and non-socially anxious
females. Biological Psychology, 135, 18–28.
Van Noordt, S. J., Campopiano, A., & Segalowitz, S. J. (2016). A functional classifica-
tion of medial frontal negativity ERPs: Theta oscillations and single subject effects.
Psychophysiology, 53(9), 1317–1334.
Van Peer, J. M., Roelofs, K., & Spinhoven, P. (2008). Cortisol administration enhances the
coupling of midfrontal delta and beta oscillations. International Journal of Psychophysiology,
67, 144–150.
492 JASON S. MOSER et al.
van Son, D., De Blasio, F. M., Fogarty, J. S., Angelidis, A., Barry, R. J., & Putman, P. (2019).
Frontal EEG theta/beta ratio during mind wandering episodes. Biological Psychology,
140, 19–27.
Velikova, S., Locatelli, M., Insacco, C., Smeraldi, E., Comi, G., & Leocani, L. (2010).
Dysfunctional brain circuitry in obsessive–compulsive disorder: source and coherence ana-
lysis of EEG rhythms. NeuroImage, 49(1), 977–983.
Velo, J. R., Stewart, J. L., Hasler, B. P., Towers, D. N., & Allen, J. J. B. (2012). Should it matter
when we record? Time of year and time of day as factors influencing frontal EEG asym-
metry. Biological Psychology, 91(2), 283–291. https://doi.org/10.1016/j.biopsycho.2012.06.010
Wacker, J., Mueller, E. M., Pizzagalli, D. A., Hennig, J., & Stemmler, G. (2013). Dopamine-D2-
receptor blockade reverses the association between trait approach motivation and frontal
asymmetry in an approach-motivation context. Psychological Science, 24(4), 489–497.
https://doi.org/10.1177/0956797612458935
Watts, A. T., Tootell, A. V., Fix, S. T., Aviyente, S., & Bernat, E. M. (2018). Utilizing time-
frequency amplitude and phase synchrony measure to assess feedback processing in a
gambling task. International Journal of Psychophysiology, 132, 203–212.
Weinberg, A., Meyer, A., Hale-Rude, E., Perlman, G., Kotov, R., Klein, D. N., & Hajcak, G.
(2016). Error-related negativity (ERN) and sustained threat: Conceptual framework and
empirical evaluation in an adolescent sample. Psychophysiology, 53(3), 372–385. doi:10.1111/
psyp.12538
Wiedemann, G., Pauli, P., Dengler, W., Lutzenberger, W., Birbaumer, N., & Buchkremer, G.
(1999). Frontal brain asymmetry as a biological substrate of emotions in patients with
panic disorders. Archives of General Psychiatry, 56(1), 78–84. https://doi.org/10.1001/archp
syc.56.1.78
Zhang, D. & Gu, R. (2018). Behavioral preference in sequential decision‐making and its associ-
ation with anxiety. Human Brain Mapping, 39(6), 2482–2499.
Zotev, V., Yuan, H., Misaki, M., Phillips, R., Young, K. D., Feldner, M. T., & Bodurka, J. (2016).
Correlation between amygdala BOLD activity and frontal EEG asymmetry during real-time
fMRI neurofeedback training in patients with depression. NeuroImage: Clinical, 11, 224–238.
https://doi.org/10.1016/j.nicl.2016.02.003
Pa rt I V
CHAPTER 20
20.1 Introduction
dissect these methods both by the information they yield about the underlying physio-
logical interactions as well as by limitations posed by the typical confounders in MEG
and EEG data.
Phase and amplitude time series of neuronal activity in electrophysiological signals can
be extracted by filtering. When estimated for two or more such time series, bivariate
correlation metrics for pairs of such phase and amplitude estimates can be applied to
uncover the patterns of correlations in these data, that is, to quantify the phase and amp-
litude FC.
Phase is a quantitative description of the state of a given process/component and its
temporal evolution within a period (or a cycle) of the oscillation. From micro-electrode
recordings to macroscopic EEG signals, observations of genuine phase correlations
imply correlations of neuronal excitability fluctuations at the cellular level and the po-
tential for consistent spike-timing relationships.
The physiological and mechanistic implications of amplitude, on the other hand, are
always dependent on the experimental setting because the amplitude measurement is
dependent on biological and physical signal generation and conduction mechanisms.
The population-level summation of electric fields leading to the measured signal amp-
litude is largely dependent on the coherence of the underlying synaptic inputs. This is
because cancellation of electric currents effectively suppresses signals from incoherent
sources. Thus, the amplitude of oscillations is primarily an approximate index of local
synchronization and only secondarily dependent on the net magnitude of source
currents.
Phase and amplitude of neuronal oscillations thus reflect partially distinct physio-
logical phenomena and, correspondingly, inter-areal phase and amplitude correlations
tap into physiologically and functionally distinct forms of large-scale coordination of
neuronal activities.
(a)
(b)
Single trails Averaged data
Power
Arrhythmic signals/
‘1/f noise’
Frequency Frequency
(c)
(d)
Power
Quasiperiodic
oscillations
Power
(Periodic) oscillations
Frequency Frequency
(e)
Hit
Miss 100
μV
50 s
0.01 0.1 1 10 50
Frequency (Hz)
Figure 20.1 Conceptual illustration of aperiodic, quasiperiodic, and periodic signals. (A) The
time series of 1/f noise yield both (B) in a single realization (left) and in averaged data (right)
1/f power spectra. (C) Signals composed of mixtures of transient periodicities from multiple
sources or quasiperiodic oscillators with a large variability in oscillation frequencies give rise
to power spectra (D) that may exhibit peaks at the corresponding frequencies (green and blue
bars) in single trials and yet the averaged spectra may not exhibit peaks. Only oscillations with
a stable period retain peaks in averaged power spectra. (E) Human behavioral performance is
characterized by power-law scale-free dynamics where the detected (Hits) and undetected
(Misses) stimuli in a threshold-stimulus detection task are clustered with power-law and long-
range temporal correlations. Adapted from Palva and Palva 2018.
BIVARIATE FUNCTIONAL CONNECTIVITY MEASURES 499
This chapter discusses the principal approaches for measuring phase and amplitude
correlations from neuronal data. Further, it aims to illustrate the adaptation of phase-
locking methods to the assessment of frequency-resolved amplitude correlations, as
well as to quantification of cross-frequency correlations, such as phase-amplitude and
cross-frequency phase correlations, among signals in different frequency bands.
All these interaction metrics are based on estimates of phase and amplitude. To ex-
plicitly represent frequency-band-limited phase and amplitude dynamics, most real-
valued continuous signals can be transformed by filtering to a complex form. A popular
choice is to convolve the signal with a complex wavelet, such as the Gabor or Morlet
wavelet (Sinkkonen et al., 1995; Tallon-Baudry et al., 1996). Such Gaussian wavelet
filtering yields optimal time-frequency-resolution compromise and is superior to other
filtering approaches in most cases (Sinkkonen et al., 1995). The Morlet wavelet w is
defined by:
( )
W (t , f 0 ) = A exp −t 2 / 2σt2 exp (2iπ f 0t ) (1)
where t denotes time, i the imaginary unit, f0 the center frequency of the wavelet, and σt
the standard deviation of the wavelet’s Gaussian envelope in time domain. A is a scaling
parameter and for conventional filtering that aims to yield a filtered signal with the same
amplitude as the original signal, A should be set to A =2/∑ (w ) where∑ (w ) denotes
the integral of wavelet’s amplitude envelope (simply the sum of wavelet values for a dis-
crete realization). Notably, this yields a wavelet spectrum that is flat for 1/f noise. The
wavelet family should be designed with a constant ratio m
m = f 0 / σ f (2)
where σf the standard deviation of the wavelet’s gaussian envelope in frequency do-
main and given by σf = 1/2 πσt . This ratio, or “the m parameter” thus defines the time-
frequency compromise so that small values (such as m =5) favors time resolution
whereas larger values (such as m =7.5) yields improved frequency resolution at the ex-
pense of time resolution.
In applications where other forms of filtering are used, and the filter yields real valued
output, the complex form can be obtained by finding the imaginary part with the Hilbert
transform. As an important non-linear filtering approach, empirical mode decompos-
ition (Huang & Wu, 2008) is an adaptive and data-driven method that identifies major
500 J. MATIAS PALVA and SATU PALVA
signal components iteratively from the fastest to slowest frequencies and yields time
series that may follow frequency fluctuations better than conventional fixed-frequency
filtering.
For segments of stationary signals, the phase estimation can be achieved with spec-
tral measures as well, such as with the Fourier transform. These three approaches are
formally equivalent (Bruns, 2004). However, spectral methods yield constant-sized
frequency bins while wavelets can and should be scaled with frequency so that they
have constant time-frequency resolution. This is often overlooked but is as a funda-
mental disadvantage for the spectral methods. For any given fixed time-window size,
different frequencies will have different amounts of cycles in the window where they
are presumed stationary. Because neuronal oscillations are “stationary” for just a few
cycles, spectral measures will be optimal for a very limited frequency range, and for
most other frequencies will include too few or too many cycles of the oscillations in the
time window.
In the filtering approach, let us denote the filtering operation with a center frequency
f generically with the operator Tf so that the filtered signal x(t,f) is obtained from the ori-
ginal signal xo(t) by
where i is the imaginary unit, ax(t,f) denotes the amplitude and θ x (t , f ) the phase of
x(t,f), and Tf,complex a complex filter such as the Morlet wavelet transform. If the filtering
is performed with a real-valued filter, Tf,real, the imaginary signal is obtained by the
Hilbert transform, so that
{ }
x(t , f ) = Tf,real [x 0 (t)] + i H Tf,real [x 0 (t)] (3)
The time series of phase and amplitude enable the estimation of bivariate phase
and amplitude correlations between any pair of signals. In addition to phase
BIVARIATE FUNCTIONAL CONNECTIVITY MEASURES 501
synchronization per se, a range of phase and amplitude interactions can be quantified
with the phase- locking estimation formalism and will be described here first.
Phase locking is statistically indicated by a non-random phase or phase-difference
distribution.
The generic n:m-phase difference ωnm of signals x and y (Tass et al., 1998) is given by
ωn:m = nθ x − mθ y (4)
where the small integers n and m determine the frequency ratio nfx = mfy of the center
frequencies of filters Tfx and Tfy. For assessing the classical within-frequency phase syn-
chrony, n = m =1. This phase difference is conveniently obtained in complex form from
the filtered x and y simply by
where the asterisk* denotes the complex conjugate. Hence, for frequencies fx and fy, the
ns × ns-sized pair-wise complex phase difference matrix Φ n:m of a single sample with
each signal against each other signal is given by the complex outer product, ⊗, so that
This expression thus provides the phase difference with which one may quantify
classical within-frequency phase synchronization but also cross-frequency phase syn-
chronization (CFS). CFS is a cross-frequency coupling (CFC) mechanism that indicates
a non-random phase difference between different low-(LF) and high-frequency (HF)
frequency pairs at small-integer ratios n:m (LF:HF). CFS enables temporally precise co-
ordination of neuronal processing by establishing the possibility for systematic spike-
timing relationships among oscillatory assemblies at different frequencies and can thus
serve functional integration and coordination across within-frequency synchronized
large-scale networks (Fell & Axmacher, 2011; Palva et al., 2005; Palva & Palva, 2017;
Siebenhühner et al., 2016; Siebenhühner et al., 2020).
The phase-difference approach can also be used to represent more complex bivariate
relationships such as phase-amplitude coupling (PAC), or “nested oscillations”, that are
characterized by the phase-locking of the amplitude envelope of a faster oscillation with
the phase of a slower oscillation. PAC has been suggested to reflect the regulation of sen-
sory information processing in faster frequencies by excitability fluctuations imposed
by slower oscillations (Canolty & Knight, 2010; Fell & Axmacher, 2011; Jensen & Colgin,
2007; Jensen & Lisman, 2013, Schroeder & Lakatos, 2009). The complex phase difference
matrix, Φ PAC, is given by filtering the amplitude envelopes Ax of the faster oscillation at
fx with the filter Tfy that was used to obtain the slower oscillation at fy < fx (Vanhatalo
et al., 2004):
502 J. MATIAS PALVA and SATU PALVA
*
ΦAA = Tfz ( Ax )/ | Tfz ( Ax ) | ⊗ Tfz ( Ay )/ | Tfz ( Ay ) | (9)
With the phase or phase difference data, the presence of statistically significant phase
locking can be assessed with a number of circular statistics. The most common of them
is the phase locking value/factor (PLV, PLF) that is given by an average of complex
phases (Sinkkonen et al., 1995). Let us first define the complex PLV (cPLV) as simply the
average across the complex phase differences
BIVARIATE FUNCTIONAL CONNECTIVITY MEASURES 503
arg(xf1)
arg(yf1)
(b) n:m phase synchrony
Re(xf1)
Re(yf 2)
arg(xf1)
(f2/f1)arg (xf1)
arg(yf 2)
(c) Phase-amplitude coupling
Amplitude
Re(xf1)
|yf 2|
Re(yf 2)
arg(xf1)
arg(Ff 1(|yf 2|))
Figure 20.2 Schematic illustration of within-frequency and cross-frequency phase and amp-
litude interactions. (A) Within-frequency synchronization is a case of generic n:m-phase syn-
chrony with n = m =1. Re(.) denotes the real part of a complex filtered signal and arg(.) its phase.
(B) A simulated example of 1:4 phase synchronization. (C) An example of phase-amplitude coup-
ling where the phase of the slow and the amplitude of the fast oscillations are correlated.
Adapted from Palva & Palva, 2012.
where nt is the number of complex phase difference Φ samples pooled across trials and/
or time. PLV is the absolute value of cPLV (PLV =|cPLV|) and is 1 for perfect coup-
ling (delta-function phase distribution) and approaches 0 for a uniform phase distribu-
tion when nt → ∞. If the phase-difference samples Φ are independent and the marginal
phase distributions are uniform, the no-interaction null hypothesis is characterized by
uniformly distributed Φ and can be tested with the Rayleigh test.
When Φ are pooled across time, that is, when the independence condition is not
met because of redundancy, and/or when the underlying process is not sinusoidal (see
Nikulin et al., 2007), surrogate data are needed for statistical testing. For spontaneous
data, the surrogates may be constructed by random rotation of the time series before
estimation of the phase differences in order to preserve endogenous and filter-induced
autocorrelations and avoid the underestimation of surrogate distribution (Palva
et al, 2005; Siebenhühner et al., 2020). For event-related data, the surrogates may be
constructed by trial shuffling. If the interaction estimates involve components that are
time-locked to the events, such as evoked responses, the artificial interactions that they
cause must be accounted for with forward-and-inversed-modeled surrogate data in trial
shuffling (Hirvonen et al., 2018).
504 J. MATIAS PALVA and SATU PALVA
One-to-one-phase and amplitude correlations, but not CFS or PAC, estimated with PLV
are directly influenced by linear mixing such as volume conduction in electrophysio-
logical recordings, which leads to inflated PLV values and false positive phase correl-
ation observations in the presence of no true correlations. For amplitude correlations,
BIVARIATE FUNCTIONAL CONNECTIVITY MEASURES 505
the same applies also to correlation coefficients. This problem is pervasive across all
forms of electrophysiological recordings: from micro-electrodes to intra-cranial stereo-
EEG and electrocorticography, and to scalp EEG, volume conduction causes linear
mixing of electric potentials. While volume conduction does little to influence MEG,
signal mixing between cortical sources and scalp sensors leads to comparable linear
mixing (Figure 20.3). Finally, demixing the sensor signals by source modeling leaves re-
sidual linear mixing, known as source leakage (Palva et al., 2018).
This problem can be partially solved by noting that linear mixing always has a zero-
time and zero-phase lag influence, which is fully captured by the real part of PLV or that
of coherence. On the other hand, as true neuronal interactions supposedly often involve
conduction-delay-related phase lags, excluding the zero-lag contribution becomes an
(a) (b)
1 1
0.8 0.8
m=0
0.6 m = 0.1 0.6
m = 0.2
iPLV
PLV
m = 0.3
0.4 m = 0.4 0.4
m = 0.6
0.2 0.2
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
(c) co (d) co
1 1
0.8 0.8
0.6 0.6
iPLV
PLV
0.4 0.4
0.2 0.2
0 0
–1 –0.75 –0.5 –0.25 0 0.25 0.5 0.75 1 –1 –0.75 –0.5 –0.25 0 0.25 0.5 0.75 1
Figure 20.3. Phase coupling measures PLV and iPLV are influenced not only by the phase
coupling strength cΘ, but also by the phase difference nϕxy and linear mixing m between the
studied signals. Here, phase coupling and linear signal mixing (m =0 (blue), 0.1 (green), 0.2 (red),
0.3 (violet), 0.4 (cyan), and 0.6 (orange)) were simulated between two signals for a range of coup-
ling and phase lag values. (A) PLV between the signals as a function of cΘ and m for nϕxy =−0.3.
Open circles at cΘ =0.4 indicate the coupling strength used in C and D. (B) iPLV between the
signals as a function of cΘ and m. (C) PLV as a function of nϕxy, when cΘ was set to 0.4. Notably,
PLV is greatly influenced by the phase difference when signal mixing is strong even though
without mixing, the magnitude of PLV is independent of the phase lag. Open circles at nϕxy =−0.3
indicate the nϕxy used in A and B. (D) The strength of iPLV depends on the phase difference and is
biased towards large phase difference so that iPLV is abolished when nϕxy =0 or nϕxy =±π.
Adapted from Palva et al., 2018.
506 J. MATIAS PALVA and SATU PALVA
approach for obtaining a measure that focuses on neuronal phase coupling and is in-
sensitive to linear mixing. Measures such as imaginary coherence (iCoh, Nolte et al.,
2004), the imaginary part of PLV (iPLV) (Palva & Palva, 2012), and weighted phase-lag
index (wPLI, Vinck et al., 2011) ignore the real part of the complex phase differences.
These measures are hence insensitive to linear mixing while revealing the true phase-
lagged interactions. Of these, wPLI is superior to iPLV and iCoh in not being influenced
by the magnitude of linear mixing. The obvious shortcoming of each these approaches
is, nonetheless, their insensitivity to true near-zero-phase lag interactions and the
dependence of the correlation value on the phase difference per se in addition to the
strength of the interaction.
The equivalent for iPLV and wPLI for amplitude correlations is orthogonalized cross
correlation (oCC) where the two signals are orthogonalized prior to the estimation of
the correlation coefficient (Brookes et al., 2012, Hipp et al., 2012). It is important to note
that the orthogonalized amplitude correlation coefficients are not independent of con-
current phase coupling. In fact, they are non-trivially affected by the presence of true
phase coupling and linear mixing in a phase-difference dependent manner and may
yield both false positive and negative findings (Palva et al., 2018b).
The exclusion of artificial interactions with coupling measures that are not inflated by
linear mixing, such as iPLV, wPLI, and oCC, has been often claimed to categorically
“account” for the problem of linear mixing causing false positive detections. Although
these measures are de facto immune to artificial false positives and can be overly conser-
vative by missing true near-zero-phase interactions, they still yield abundant “spurious”
or “ghost” false positive interactions due to signal spread. This is because field spread
in the vicinity of a true nonzero phase interaction mirrors the true interactions into
spurious ghost interactions, that appear as false positives with any bivariate interaction
measure in a manner that can never be accounted for in bivariate fashion (Figure 20.4).
There are two principal approaches to addressing the problem of spurious “ghost”
interactions. One is to perform symmetric multivariate orthogonalization where the
cortical-parcel-source signals are orthogonalized in a multivariate fashion (rather than
bivariate as for oCC) (Colclough et al., 2015). Symmetric orthogonalization overcomes
the problem of spurious ghost interactions by simultaneously removing zero phase-lag
components from all source time series using Löwdin orthogonalization for gradient
descent. All-to-all amplitude correlations are estimated with partial correlation of amp-
litude envelopes to keep direct and remove indirect interactions. Because the partial
correlation matrix is expected to be sparse, a graphical lasso regularization of the inverse
BIVARIATE FUNCTIONAL CONNECTIVITY MEASURES 507
(a)
PLV (b)
iPLV
0.14 min
0.32
PLV
0.51
0.69 max
c=0
Artificial
iPLV
0.28 0.12
PLV
0.49 0.17
0.69 max 0.21 max
True
True
c = 0.4
Artificial
Spurious
Spurious
Figure 20.4 (A) Linear mixing leads to observations of false positive artificial PLV
interactions, within the mixing region even in the absence of true correlations. Activity of 169
uncoupled (c =0) sources (black dots) in a 13 × 13 grid was simulated and the 20 strongest PLV
edges of the two sources-of-interest (centers of the cyan and red regions) are visualized. The cyan
and red color gradients indicate the Gaussian mixing strength. (B) iPLV analysis of the same
data as in A shows that iPLV does not lead to observations of artificial interactions in the ab-
sence of true interactions. (C) True phase correlations are mirrored into false positive spurious
correlations, between the two mixing regions in the condition where there is a true interaction
(c =0.9) between two sources-of-interest (here the centers of the mixing regions). (D) Even
though iPLV does not yield artificial false positives, it detects the spurious “ghost” interactions
similarly to PLV. Spurious correlations arise because any two sources in separate mixing regions
partially retain the nonzero phase difference of their center sources.
covariance matrix is applied to penalize near-zero elements, which reduces noise in the
partial correlation graph. This leads symmetric orthogonalization to attenuate both
signal-leakage caused and true indirect couplings. The main limitations of this method
are that it is principally applicable only to the estimation of amplitude correlations and
508 J. MATIAS PALVA and SATU PALVA
that it is limited by the rank of the data due to its dependence on singular value decom-
position. For MEG/EEG data that are preprocessed with signal space separation (SSS)
and temporal SSS methods, the rank of the data (~degrees of freedom) is often limited
to 60–70 (Haumann et al., 2016). Thus, symmetric orthogonalization is applicable only
to cortical networks with less than ~60 parcels in total, such as the 19 regions per hemi-
sphere used in Colclough et al. (2015).
Another approach to controlling spurious connections is “hyperedge bundling”,
which uses any bivariate connectivity mapping as the basis and bundles observed
connections on the basis of their proximity in source geometry. Hyperedge bundling
reduces the false positive rate by a factor of 10-100 with moderate to little decrease on
the true positive rate. The main advantages of this approach are that it does not involve
complex mathematical transformations prior to interaction estimation and may be used
with any bivariate metric in a manner that retains the metric’s sensitivity. Moreover,
hyperedge bundling enables high-resolution analyses with hundreds of cortical parcels,
that is, redundant oversampling of the source space, which safeguards the analysis out-
come from the possibility that the actual neuronal source constellations or degrees of
freedom in the data are different from those of the used parcellation scheme. In the
worst-case scenario, coarse parcellations can misrepresent or miss source areas that fall
in between the parcels or are much smaller than the parcels.
Finally, the overall effects of signal mixing on estimates of inter-areal phase-and
amplitude coupling can be mitigated by optimizing the MEG/EEG inverse modelling
procedure. The inverse transformed source dipole time series can be collapsed into
parcel time series in a manner that maximizes the source reconstruction accuracy and
minimizes the effects of source leakage (Korhonen et al., 2014). The fidelity-weighted
inverse modeling has higher reconstruction accuracy than a regular inverse operator for
the given parcellation, because it gives greater weight to sources with better reconstruc-
tion accuracy for the signals from the parcels they belong into.
Initial MEG and EEG source connectivity studies used region- of-
interest-
based
approaches for phase and amplitude correlation analyses but this approach has been all
but replaced by data-driven all-to-all connectivity mapping, or “connectomics”, studies.
Data-driven analyses approaches reveal the most robust effects in data regardless of the
original hypothesis and as such provide more rigorous hypothesis testing compared to
hypothesis-driven analyses that may be confounded by several forms of biases and ex-
plicit errors such as circularity in analysis targeting (Kriegeskorte, 2009). Data-driven
analysis of both local and especially large-scale oscillatory interactions provide a com-
prehensive view on brain dynamics that is not biased by selection criteria. In addition,
BIVARIATE FUNCTIONAL CONNECTIVITY MEASURES 509
several approaches for network analysis as described below can be applied to compen-
sate for the presence of artefactual and spurious connections.
Graph theory and complex network theory enable the analysis and prediction of the
behaviors of a wide variety of complex systems. In human brain networks, these have
included graph theoretical network analysis (Bullmore & Sporns, 2009; Rubinov &
Sporns, 2010) that can also be used to characterize the properties of oscillatory networks
or connectomes that have been obtained with data-driven analysis approaches (Palva
et al., 2010a; Palva et al., 2010b; Siebenhühner et al., 2016; Siebenhühner et al., 2020).
This approach describes oscillatory networks at three levels: at the entire graph level
(i.e., the connectome), level of edges (i.e., connections), and vertices (i.e., brain regions
or parcels). Vertex centrality estimates can be used to identify those vertices (brain
regions) that are hubs in networks of interareal oscillatory interactions and likely to
play an important role in neuronal communication. Several metrics can, in addition, be
used to characterize specific topological attributes such as clustering, path lengths, and
modularity of oscillatory networks (Palva et al., 2010b; Zhigalov et al., 2017).
In addition to source mixing described, estimates of CFC may be inflated by false posi-
tive couplings arising from non-sinusoidal and nonzero mean signals. False positives
are caused by the artificial higher- frequency components produced when non-
sinusoidal signals are filtered into narrow bands and by artificial lower-frequency
components arising from filtering of nonzero-mean waveforms. Because these are
ubiquitous in electrophysiological signals, their filter artefacts constitute a signifi-
cant confounder to CFS and PAC estimation. Approaches based on waveform analysis
(Kramer et al., 2008; Aru et al., 2015, Van Driel et al., 2015) have been proposed to reduce
the artefactual connections arising from non-sinusoidal signals. Nevertheless, filter-
artefact-caused spurious CFC, in particular CFS, is difficult to dissociate from genuine
CFC by inspection of the waveform shape of any single signal in isolation. Local CFC
estimates are thus prone to ambiguous results. CFC is necessarily genuine when there
is evidence for two distinct coupled processes while spurious CFC arises from a single
process with signal components distributed to distinct frequency bands because of filter
artefacts (Figure 20.5). These two distinct processes can be identified by connection-by-
connection testing of whether CFC can unambiguously be attributed to two separable
processes using graph-theoretical approaches (Figure 20.5, Siebenhühner et al., 2020).
If CFC is spurious between areas A and B, these regions are necessarily also connected
by 1:1 phase synchrony and local CFC that together may lead to a spurious observation
of inter-areal CFC. The absence of either PS or local CFC, in contrast, indicates that the
observed inter-areal CFC cannot be attributable to a single source and is thus genuine.
510 J. MATIAS PALVA and SATU PALVA
Frequency
two sinusoidal processes high
+ low
OR:
Frequency Frequency
one non-sinusoidal process high
low
(b) two separable processes
B high
A low
A B A B
Area Area
1:1 coupled slow osciallations
B
Frequency
B high
A low
A
A B A B
Area Area
1:1 coupled fast oscillations
B B Frequency high
A low
A A B A B
Area Area
(c)
Genuine inter-areal CFC, unambiguous Genuine inter-areal CFC, ambiguous
high high
Frequency
low low
high high
low low
A B A B A B A B A B A B
Figure 20.5 Schemata for dissociating the genuine from putatively spurious CFC. (A) Spurious
observations of local are seen when a non-sinusoidal signal is filtered. (B) Inter-areal CFC can be
proven to be genuine if it unambiguously originates from separable neuronal signals. Spurious
inter-areal CFC is always accompanied by spurious local CFC at one or both locations that are
also inter-areally coupled via 1:1 phase synchrony either at low frequency or high frequency
so that a “triangle motif ” with the observed (spurious) inter-areal CFC is formed. (C) Graph-
theory based approach can identify all triangle motifs that might contain spurious inter-areal
CFC. This approach only identifies those inter-areal CFC observations as genuine, which are not
part of a full triangle motif, whereas all others are discarded. Since this may include also genuine
connections, this approach provides a lower bound for the number of genuine connections.
Modified from Siebenhühner et al., 2020.
BIVARIATE FUNCTIONAL CONNECTIVITY MEASURES 511
REFERENCES
Abbas, A., Belloy, M., Kashyap, A., Billings, J., Nezafati, M., Schumacher, E., & Keilholz,
S. (2019). Quasi-periodic patterns contribute to functional connectivity in the brain.
NeuroImage, 191, 193–204. doi:10.1016/j.neuroimage.2019.01.07
Aru, J., Aru, J., Priesemann, V., Wibral, M., Lana, L., Pipa, G., Singer, W., & Vicente, R. (2015).
Untangling cross-frequency coupling in neuroscience. Current Opinion in Neurobiology, 31,
51–61. doi:10.1016/j.conb.2014.08.002
Aydore, S., Pantazis, D., & Leahy, R. M. (2013). A note on the phase locking value and its
properties. NeuroImage, 74, 231–244.
Belloy, M. E., Naeyaert, M., Abbas, A., Shah, D., Vanreusel, V., van Audekerke, J., ... Verhoye M
(2018). Dynamic resting state fMRI analysis in mice reveals a set of quasi-periodic patterns
and illustrates their relationship with the global signal NeuroImage, 180(Pt B), 463–484.
doi:10.1016/j.neuroimage.2018.01.075.
Brookes, M. J., Woolrich, M. W., & Barnes, G. R. (2012). Measuring functional connectivity
in MEG: A multivariate approach insensitive to linear source leakage. NeuroImage, 63(2),
910–920.
Bruns, A. (2004). Fourier-, Hilbert-and wavelet-based signal analysis: Are they really different
approaches? Journal of Neuroscience Methods, 137, 321–332.
Bullmore, E. & Sporns, O. (2009). Complex brain networks: Graph theoretical analysis of
structural and functional systems. Nature Reviews Neuroscience, 10, 186–198.
Canolty, R. T. & Knight, R. T. (2010). The functional role of cross-frequency coupling. Trends in
Cognitive Science, 14(11), 506–515. doi:10.1016/j.tics.2010.09.001
Colclough, G. L., Brookes, M. J., Smith, S. M., & Woolrich, M.W. (2015). A symmetric multi-
variate leakage correction for MEG connectomes. NeuroImage, 117, 439–448. doi:10.1016/
j.neuroimage.2015.03.071.
Fell, J. & Axmacher, N. (2011). The role of phase synchronization in memory processes. Nature
Reviews Neuroscience, 12(2), 105–118. doi:10.1038/nrn2979
Hipp, J. F., Hawellek, D. J., Corbetta, M., Siegel, M., & Engel, A. K. (2012). BOLD fMRI correl-
ation reflects frequency-specific neuronal correlation. Nature Neuroscience, 15, 884–889.
Hirvonen, J., Monto, S., Wang, S. H., Palva, J. M., & Palva, S. (2018). Dynamic large-scale net-
work synchronization from perception to action. Network Neuroscience, 2(4), 442–463.
https://doi.org/10.1162/netn_a_0003
Huang, N. E. & Wu, Z. (2008). A review on Hilbert‐Huang transform: Method and its
applications to geophysical studies. Reviews of Geophysics [online], 46, 1–23.
Kramer, M. A., Tort, A. B. L., & Kopell, N. J. (2008). Sharp edge artifacts and spurious coupling
in EEG frequency comodulation measures. Journal of Neuroscience Methods, 170(2), 352–357.
doi:10.1016/j.jneumeth.2008.01.020
Kriegeskorte, N., Simmons, W. K., Bellgowan, P. S., & Baker, C. I. (2009). Circular analysis in
systems neuroscience: The dangers of double dipping. Nature Neuroscience, 12(5), 535–40.
doi:10.1038/nn.2303
Jensen, O. & Colgin, L. L. (2007). Cross-frequency coupling between neuronal oscillations.
Trends in Cognitive Sciences, 11, 267–269. doi:10.1016/j.tics.2007.05.003
Lisman, J. E. & Jensen, O. (2013). The theta-gamma neural code. Neuron, 77, 1002–1016.
doi:10.1016/j.neuron.2013.03.007
512 J. MATIAS PALVA and SATU PALVA
Tass, P., Rosenblum, M. G., Weule, J., Kurths, J., Pikovsky, A., Volkmann, J., Schnitzler, A.,
& Freund, H.-J. (1998). Detection of n:m phase locking from noisy data: Application to
magnetoencephalography. Physics Review Letters, 81, 3291.
Siebenhühner, F., Wang, S. H., Arnulfo, G., Lampinen, A., Nobili, L., Palva, J. M., & Palva,
S. (2020). Resting-state cross-frequency coupling networks in human electrophysiological
recordings. PLoS Biology, 18(5), e3000685. https://doi.org/10.1371/journal.pbio.3000685.
Siebenhühner, F., Wang, S. H., Palva, J. M., & Palva, S. (2016). Cross-frequency synchroniza-
tion connects networks of fast and slow oscillations during visual working memory main-
tenance. Elife, 5, e13451. doi:10.7554/eLife.13451
van Driel, J., Cox, R., & Cohen, M. X. (2015). Phase-clustering bias in phase–amplitude
cross-frequency coupling and its removal. Journal of Neuroscience Methods, 254, 60–72.
doi:10.1016/j.jneumeth.2015.07.014
Vanhatalo, S., Palva, J. M., Holmes, M. D., Miller, J. W., Voipio, J., & Kaila, K. (2004). Infraslow
oscillations modulate excitability and interictal epileptic activity in the human cortex during
sleep. Proceedings of the National Academy of Science of the United States of America, 101,
5053–5057.
Vinck, M., Oostenveld, R., van Wingerden, M., Battaglia, F., & Pennartz, C. M. (2011). An
improved index of phase-synchronization for electrophysiological data in the presence of
volume-conduction, noise and sample-size bias. NeuroImage, 55, 1548–1565.
Vinck, M., van Wingerden, M., Womelsdorf, T., Fries, P., & Pennartz, C. M. (2010). The pair-
wise phase consistency: A bias- free measure of rhythmic neuronal synchronization.
NeuroImage, 51, 112–122.
Zhang, X., Pan, W. J., & Keilholz, S. D. (2020). The relationship between BOLD and neural
activity arises from temporally sparse events. NeuroImage, 207, 116390. doi:10.1016/
j.neuroimage.2019
Zhigalov, A., Arnulfo, G., Nobili, L., Palva, S., & Palva, J. M. (2017). Modular co-organization of
functional connectivity and scale-free dynamics in the human brain. Network Neuroscience,
1(2), 143–165.
CHAPTER 21
MULTIVARIAT E MET H OD S
FOR FU NC T I ONA L
C ONNECTIVIT Y A NA LYSI S
SELIN AVIYENTE
21.1 Introduction
& Yener Mutlu, 2011; Dimitriadis et al., 2013). As a measure for functional connect-
ivity, PLV was introduced by Lachaux and colleagues (1999) and it estimates the syn-
chrony between two signals by looking at the circular variance of their phase difference
across trials. In comparison to other linear and nonlinear methods, PLV is more sen-
sitive to nonlinear effects (Jalili et al., 2013). In addition, this metric contributes to the
assessment of brain rhythms and their related cognitive processes, for example, alpha,
beta, delta, and theta in the low-frequency bands and gamma bands in the higher
frequencies (Pereda et al., 2005; Lachaux et al., 1999; Aydore et al., 2013; Aviyente
et al., 2011).
Although PLV is a promising measure for quantifying functional connectivity, it
is still limited due to its bivariate nature. Specifically, it does not provide information
regarding the integration across multiple regions in the brain. In addition, functional
connectivity results from bivariate measures are difficult to interpret and computation-
ally expensive for systems with large number of regions. In order to overcome these
drawbacks, researchers propose multivariate phase synchrony measures (Stam & Van
Dijk, 2002; Carmeli et al., 2005; Mutlu & Aviyente, 2012; Al-Khassaweneh et al., 2016).
The two main approaches to quantifying multivariate synchronization are spectral and
graph theoretic methods.
The application of bivariate measures to multivariate data sets with N time-series
results in an N × N matrix of bivariate indices, which leads to a large amount of mostly
redundant information. Therefore, it is necessary to reduce the complexity of the data
set in such a way to reveal the relevant underlying structures using multivariate ana-
lysis methods. The basic approach used for multivariate phase synchronization is to
trace the observed pairwise correspondences back to a smaller set of direct interactions
using approaches such as partial coherence adapted to phase synchronization (Schelter
et al., 2006). Another complementary way to achieve such a reduction is cluster
analysis—a separation of the parts of the system into different groups such that the
signal interdependencies within each group tend to be stronger than in between groups
(Newman, 2006a, 2006b). Allefeld and colleagues have proposed two complementary
approaches to identify synchronization clusters and applied their methods to EEG data
(Allefeld & Kurths, 2004; Allefeld et al., 2007; Allefeld et al., 2005; Allefeld & Kurths,
2003). Allefeld and Kurths (2004) present a mean-field approach that assumes the ex-
istence of a single synchronization cluster that all oscillators contribute to a different ex-
tent. The authors define the to-cluster synchronization strength of individual oscillators
to identify multivariate synchronization. This method has the disadvantage of assuming
a single cluster and thus cannot identify the underlying clustering structure. Allefeld
and colleagues (2007) introduce an approach that addresses the limitation of the single
cluster approach using methods from random matrix theory. This method is based on
the eigenvalue decomposition of the pairwise bivariate synchronization matrix and
appears to allow identification of multiple clusters. Each eigenvalue greater than 1 is
associated with a synchronization cluster and quantifies its strength within the data set.
The internal structure of each cluster is described by the corresponding eigenvector.
516 SELIN AVIYENTE
Combining the eigenvalues and the eigenvectors, one can define a participation index
for each oscillator and its contribution to different clusters. This method assumes that
the synchrony between systems belonging to different clusters (i.e. between-cluster
synchronization) is equal to zero and requires an adjustment for proper computation
of the participation indices in the case that there is between-cluster synchronization.
Despite the usefulness of eigenvalue decomposition for the purposes of cluster iden-
tification, Allefeld and Bialonski (2007) demonstrate that there are important special
cases—clusters of similar strength that are slightly synchronized to each other—where
the assumed one-to-one correspondence of eigenvectors and clusters is completely lost.
Other alternative measures that quantify multivariate relationships include the directed
transfer function and Granger causality defined for an arbitrary number of channels
(Granger, 1969; Baccala & Sameshima, 2001). Both of these methods have been applied
to study interdependencies and causal relationships, however, are limited to stationary
processes and linear dependencies.
On the other hand, graph theory provides the means for characterizing the
functional connections in the brain using a complex network model (Bullmore &
Sporns, 2009). Functional connectivity networks are constructed by considering
the different brain regions or electrodes/sensors as nodes and the relationships be-
tween different nodes, quantified by bivariate functional connectivity measures such
as PLV and correlation, as edges. In this manner, functional connectivity networks
can take advantage of the widely available set of techniques for characterizing com-
plex networks. In terms of brain networks, these measures have been grouped as
measures of functional segregation and functional integration (Rubinov & Sporns,
2010). Measures of functional segregation include the clustering coefficient, tran-
sitivity, and modularity (Rubinov & Sporns, 2010; de Vico Fallani et al., 2014). On
the other hand, measures of functional integration include the characteristic path
length and the global efficiency. By computing measures that characterize network
structure, such as the small-world measure and the degree distribution, it has been
shown that functional connectivity networks exhibit features of complex networks,
including the small-world network (Bullmore & Sporns, 2009; Bassett & Bullmore,
2017; Bassett & Bullmore, 2006), and both small-world and scale-free networks
(van den Heuvel et al., 2008). Although graph theoretic measures have contributed
greatly to the advancements in the study of functional connectivity networks, these
measures present some drawbacks. Measures employed in the characterization of
network structure such as the mean clustering coefficient, the characteristic path
length and the global efficiency may be affected by certain characteristics of the net-
work. Examples include how nodes with low degree affect the clustering coefficient,
and the dependence of the characteristic path length and the global efficiency in the
shortest path between nodes, when networks may rely on other mechanisms than
the shortest path for communication.
This chapter reviews spectral, direct multivariate, and graph theoretic measures for
quantifying multivariate synchrony within a group of oscillators.
MULTIVARIATE METHODS FOR FUNCTIONAL CONNECTIVITY ANALYSIS 517
21.2.1 S-Estimator
S-estimator is one of the most commonly used multivariate synchronization metrics. It
quantifies the amount of synchronization within a group of oscillators using the eigen-
value spectrum of the correlation, covariance or functional connectivity matrix:
λ log ( λ i )
N
S = 1+
∑ i =1 i
, (21.1)
log ( N )
where λ i s are the N normalized eigenvalues. This measure is complementary to the en-
tropy of the normalized eigenvalues of the correlation matrix. The more dispersed the
eigenspectrum is the higher the entropy would be. If all of the oscillations in a group are
completely synchronized, that is, the entries of the pairwise functional connectivity ma-
trix are all equal to 1, then all of the eigenvalues except one will be equal to zero, and the
value of S will be equal to 1 indicating perfect multivariate synchrony. This measure can
quantify the amount of synchronization within a group of signals and thus is useful as a
global complexity measure.
21.2.2 Omega Complexity
A variation of S-estimator is Ω-complexity. Using the eigenvalues of the correlation ma-
trix as λ i , the omega complexity can be computed as:
N λ λ
Ω = exp − ∑ i log i . (21.2)
i =1 N N
between a value close to 0 (for minimum synchrony) and 1 (for maximum synchrony),
one can compute Omega as 1/ Ω.
λ1 ( f ) − λ 2 ( f )
GFS ( f ) = . (21.3)
λ1 ( f ) + λ 2 ( f )
If the sine-cosine clouds lie on a straight line, one of the eigenvalues equals to 0 and
the covariance is completely explained by a single principal component; the GFS takes
a value of 1 for such cases. This corresponds to complete phase synchrony at a given fre-
quency. In a non phase-synchronized case, the two eigenvalues are close to each other,
leading to GFS values close to 0. In order to obtain GFS values in a certain frequency
range, similar to cross-coherence, one can get the average of GFS values over that fre-
quency range.
Omega complexity and S-estimator are all based on the computation of the covari-
ance or bivariate connectivity matrix. Therefore, the metrics are not direct measures
quantifying multivariate synchrony and will be influenced by the bias affecting bivariate
measures. A solution to this problem is to use multivariate phase synchrony (MPS)
measure, which is indeed an extension to PLV. Having extracted the instantaneous
phases from the individual time-series (using the Hilbert transform, for instance), the
MPS is computed as
L N
1
MPS = ∑⋅
LN t =1
∑e
jΦi (t )
(21.4)
i =1
MULTIVARIATE METHODS FOR FUNCTIONAL CONNECTIVITY ANALYSIS 519
MPS measures the mean phase coherence between the time-series, averaged over the
observation samples. It ranges from 0 for completely unsynchronized systems to 1 for
completely synchronized systems. However, this measure does not directly quantify
the relationship of phases across oscillators as it relies on the absolute phase rather than
the phase differences. For this reason, in this section we discuss a more recent method
to quantify multivariate phase synchrony directly from the phase differences of the
observed oscillators (Al-Khassaweneh et al., 2016; Mutlu & Aviyente, 2012).
γ k
M −1 (t ) = sin (θ (t )) ×…× sin (θkM −2 (t )) × cos (θkM −1 (t )) ,
k
1
For example, γ 1k (t ) is the x coordinate of a vector on the unit circle at angular pos-
ition θ1k (t ) , while γ 2k (t ) is the x coordinate of a vector on a circle with radius sin θ1k (t ) ( )
at angular position θ (t ) . Similar analysis applies to the remaining θ (t ) s. Thus, every
k
2
k
i
i −1
γ ik (t ) is just the x coordinate of a vector on a circle with radius rik (t ) = ∏ sin θkj (t ) ,
j =1
( )
520 SELIN AVIYENTE
for i = 2, 3,..., M and with a phase θik (t ) . The equation for ri (t ) shows that as i increases,
γ ik (t ) will have less impact on the overall synchrony. This means that the choice of the
first phase difference, θ1k (t ) , will have a high impact on the measured synchrony.
Equation (21.5) may also be interpreted as follows. Every γ ik (t ) is the x projection of
the y coordinate of the previous γ ik−1 (t ) on the x-axis with a phase θik (t ) , i.e. define x
and y coordinates of the rotating vector for each trial k as
( )
γ kx1 (t ) = cos θ1k (t ) ,
γ (t ) = sin (θ (t )) ,
k
y1
k
1
γ k
xM (t ) = sin (θ (t )) ×…× sin (θkM −1 (t )) × cos (θkM (t )) ,
k
1
( )
coordinates by dik (t ) = rik (t ) = ∏ sin θkj (t ) for i = 2, 3,..., M . This will result in unit
j =1
radius for all i. Therefore, the multivariate phase synchrony measure, hyper-torus phase
synchrony (HTS), is given by
N
1
HTS (t ) = ∑ D (t ) k
(21.7)
N× M k =1 2
γ kx (t ) γ ky (t ) γ kx (t ) γ ky M (t )
where D k (t ) = k1 , k1 , …, kM , .
d1 (t ) d1 (t ) d M (t ) d Mk (t )
This HTS metric can also be simplified as
MULTIVARIATE METHODS FOR FUNCTIONAL CONNECTIVITY ANALYSIS 521
1
HTS (t ) = PLV12 (t ) + …+ PLVM2 (t )
M
1 M
= ∑PLVi2 (t ),
M i =1
(21.8)
where PLVi quantifies the synchronization of each oscillator with respect to a common
reference angle with θik (t, ω ) as described above and PLVi is given by
N
1
PLVi (t ) =
N
∑ exp ( jθ (t ))
k =1
k
i
In the last decade, connectivity- based methods have had a prominent role in
characterizing normal brain organization as well as alterations due to various brain
disorders. Functional connectivity networks (FCNs) are obtained by recording physio-
logical signals from different brain regions and then computing the pairwise similarity
(Stam & Reijneveld, 2007; Bullmore & Bassett, 2011). In this context, the nodes of the
network correspond to the brain regions and the edges correspond to their functional
connectivity. In the context of synchronization, the N × N bivariate connectivity ma-
trix, A can be treated as the adjacency matrix of the graph.
As with other real-world connected systems and relational data, studying the top-
ology of interactions in the brain has profound implications in the comprehension of
complex phenomena, such as the emergence of coherent behaviour and cognition or
the capability to functionally reorganize after brain lesions (i.e. brain plasticity) (Sporns,
2018). In practice, graph metrics (or indices) such as clustering coefficient, path length
522 SELIN AVIYENTE
and efficiency measures are often used to characterize the “small-world” properties of
brain networks. Centrality metrics such as degree, betweenness, closeness, and eigen-
vector centrality are used to identify the crucial areas within the network. Community
structure analysis, which detects the groups of regions more densely connected between
themselves than expected by chance, is also essential for understanding brain network
organization and topology.
The definition of the nodes (or vertices) for brain graphs is modality specific. In
sensor-based modalities, such as EEG, brain nodes are commonly assigned directly
to sensors or to electrodes. However, volume conduction in EEG causes the signal at
each sensor to be a mixture of blurred activity from different inner cortical sources.
This effect can either be ignored, in which case brain nodes will suffer from a biased
non-neural dependence, or it can be addressed in several ways, such as using spatial
filters (Kayser & Tenke, 2006), choosing functional connectivity metrics that attenuate
volume conduction (Stam et al., 2007), or using cortical source reconstruction. After
defining brain nodes, assigning links between them is the subsequent crucial modelling
step. In functional neuroimaging, the links of a brain graph are given by evaluating the
similarity between two brain signals, through functional connectivity (FC) measures
such as phase synchrony, Pearson’s correlation or coherence. FC methods fall into
two broad categories: those measuring symmetric mutual interaction (undirected
weighted links) and those measuring asymmetric information propagation (directed
weighted links). FC across N recording sites can be described by an N × N adjacency
matrix A containing all the pairwise FC measures aij corresponding to the weighted
links of the brain graph. Since most of the graph theoretic metrics are defined on
binary (unweighted) graphs, initial efforts on the analysis of brain network topologies
were implemented via the binarization of the weights using some arbitrarily chosen
thresholds with some good success. A simple way of building a graph from a weighted
FC matrix is to apply a threshold τ to each element of the matrix, such that if aij ≥ τ ,
then an edge is drawn between the corresponding nodes, but if aij < τ , no edge is drawn
(Achard et al., 2006). This thresholding operation thus binarizes the weight matrix and
converts the continuously variable edge weights to either 1 or 0. By varying the threshold
τ used to construct a binary graph from a continuous weight matrix, the connection
density of the network is made denser or less dense. If the threshold is low and many
weak weights are added to the graph as edges then the connection density will increase;
if the threshold is high and only the strongest weights are represented as edges, then
the connection density will decrease. Since the choice of τ may be arbitrary, a lot of the
binarization techniques rely on choosing a threshold to achieve a certain graph density.
Moreover, in recent work, instead of choosing a single τ , a range of τ values have been
chosen to generate multiple binary graphs from a single FC network. In this manner, the
topology of the connectivity network can be analyzed across different scales. However,
thresholding poses the problem of over-simplifying FCNs and, more importantly, there
is no generally accepted criterion to select the threshold (Bassett & Bullmore, 2017; Lee
MULTIVARIATE METHODS FOR FUNCTIONAL CONNECTIVITY ANALYSIS 523
et al., 2012). Moreover, the size and density of the thresholded network vary based on
the chosen threshold value (Rubinov & Sporns, 2010). Recent studies show that the sig-
nificance of the difference between groups is strongly dependent on the threshold par-
ameter, that is, the power of the statistical analysis varies with the threshold (Langer
et al., 2013). Recently, extensions of graph theoretic measures have been proposed
for weighted networks to address some of these issues (Bolanos et al., 2013; Bassett
et al., 2011). It has also been shown that graph theoretic measures, such as the clustering
measure and the small-world parameter, are very sensitive to the size of the network,
that is, the number of nodes, and the density of the connections. Thus, comparing two
networks with different edge density may lead to wrong conclusions, making it diffi-
cult to disentangle experimental effects from those introduced by differences in the
average degree (Bassett & Bullmore, 2017; Muldoon et al., 2016). For these reasons, in
this chapter we focus on metrics suitable for undirected and weighted links.
Graphs can be investigated at different levels of scale, and specific measures cap-
ture graph attributes at local (nodal) and global (network-wide) scales (Kruschwitz
et al., 2015; de Vico Fallani et al., 2014). Nodal measures include simple statistics such
as node degree or strength, while global measures express networkwide attributes
such as the path length or the efficiency. Intermediate scales can be accessed via hier-
archical neighborhoods around single graph elements, or by considering subgraphs
or motifs. Motifs are defined as subsets of network nodes and their mutual edges
whose patterns of connectivity can be classified into distinct motif categories. In em-
pirical networks, these categories often occur in characteristic frequencies that can be
compared with distributions from appropriate (random) null models. Even though a
multitude of graph theoretics have been defined to analyze complex networks, here we
focus on metrics that are particularly useful for characterizing functional connectivity
networks.
21.4.1 Small-World
Watts and Strogatz (1998) showed that graphs with many local connections and a
few random long distance connections are characterized by a high cluster coefficient
(like ordered networks) and a short path length (like random networks). Such near-
optimal networks, which are intermediate between ordered and random networks,
are designated as small-world networks. Many neuroimaging studies have shown that
both structural and functional brain networks shared similar small-world properties of
short path length and high clustering (Bassett & Bullmore, 2006; Bassett & Bullmore,
2017). Small-world networks are simultaneously strongly clustered and integrated.
This phenomenon of small-worldness is captured by the small-world parameter, which
is the ratio of the normalized clustering coefficient to the normalized path length. For
a weighted network, the small-world parameter is given as (Rubinov & Sporns, 2010;
Humphries & Gurney, 2008):
524 SELIN AVIYENTE
C w / Crand
w
σw = (21.10)
Lw / Lwrand
where C and Crand are the clustering coefficients of the network and a random network
with the same degree distribution, respectively, and L and Lrand are the characteristic
path lengths of the network and a random network with the same degree distribution,
respectively. In this definition, the clustering coefficient is a measure of segregation and
reflects mainly the fraction of clustered connectivity available around individual nodes.
The clustering coefficient for a weighted network is defined as (Onnela et al., 2005):
1 2tiw
Cw = ∑
N i N ki ( ki − 1)
, (21.11)
where tiw is the weighted geometric mean of the triangles around a node i.
Similarly, the characteristic path length of the network is the average shortest path
length between all pairs of nodes in the network. For a weighted network it is calculated
as (Rubinov & Sporns, 2010):
1 ∑ j N , j ≠i dij (21.12)
w
L = ∑ i N
w
,
N (n − 1)
where dijw is the shortest weighted path length between node i and j.
Even though the small-world parameter has been widely used to characterize func-
tional brain networks, it has some shortcomings. First, it is not clear in most cases how
the small-worldness of the brain network relates to biological properties. Second, a
single parameter that summarizes the network topology is not sufficient and does not
tell the whole story about the network structure. Third, the small-world scalar σ can be
greater than 1 even in cases when the normalized path length is much greater than one;
because it is defined as a ratio, if C >> 1 and L > 1 , the scalar σ > 1 . This means that a
small-world network will always have σ > 1 , but not all networks with σ > 1 will be
small-world (some of them may have greater path length than random graphs). Finally,
the measure is strongly driven by the density of the graph, and denser networks will
naturally have smaller values of σ even if they are in fact generated from an identical
small-world model.
21.4.2 Modularity
Among the most widely encountered and biologically meaningful aspects of brain
networks is their organization into distinct network communities or modules (Meunier,
MULTIVARIATE METHODS FOR FUNCTIONAL CONNECTIVITY ANALYSIS 525
Lambiotte et al., 2009; Laumann et al., 2015; Chavez et al., 2010; Fair et al., 2009; Ferrarini
et al., 2009; Meunier, Archard et al., 2009; Power et al., 2011). Modules are useful to par-
tition larger networks into basic “building blocks,” that is, internally densely connected
clusters that are more weakly interconnected among each other. Modular partitions have
neurobiological significance as their boundaries separate functionally related neural
elements, define critical bridges and hubs that join communities, channel and restrict the
flow of neural signals and information, and limit the uncontrolled spread of perturbations.
There are numerous computational techniques for extracting communities and
modules from complex networks. One of the most widely used approaches in network
neuroscience is modularity maximization, which aims to divide a given network into a set of
nonoverlapping communities by maximizing a global objective function, the modularity
metric (Newman, 2006a). Originally, this metric was formulated to detect communities
whose internal density of connections is maximal, relative to a degree-preserving null
model. A good partitioning of a network is expected to have high modularity Q with
( ) (
Q = fraction of edges within communities − expected fraction of such edges , where )
the expected fraction of edges is evaluated for a random graph.
While modularity characterizes the brain network at a finer scale compared to the small-
world parameter, it still has some shortcomings. First, modularity maximization has some
shortcomings such as coming up with degenerate solutions, that is, numerous partitions
may result in the same value of the modularity metric. Second, modularity optimization
cannot identify modules below a certain size. One way to address these issues and repre-
sent the full multi-scale structure of brain networks is to perform consensus clustering
across multiple spatial resolutions—an approach that combines sampling the entire range
of possible spatial resolutions with a hierarchical consensus clustering procedure.
21.4.3 Centrality
Numerous measures quantify the potential of individual nodes and edges to influ-
ence the global state of the network. Many of them allow the identification of network
hubs generally defined as highly central parts of the network (Sporns et al., 2007). The
number of connections maintained by a node (its degree) or the combined weight
of these connections (its strength) often provides a strong indicator of influence or
centrality. Other measures of centrality take advantage of the layout of the shortest
paths within the network and record the number of such paths that pass through a
given node or edge—a measure called the betweenness. Another way to approach
centrality is by referencing the relation of nodes and edges to a network’s commu-
nity structure. The participation coefficient quantifies the diversity of a given node’s
connections across multiple modules—high participation indicates that many of
these connections are made across modules, thus linking structurally and function-
ally the distinct communities (Guimera & Nunes Amaral, 2005). This measure is par-
ticularly useful in brain networks as it can be applied to both structural and functional
network data.
526 SELIN AVIYENTE
This section illustrates the differences between the various multivariate metrics
introduced in this chapter. Moran and colleagues (2015) used an EEG dataset from a
previously published cognitive control-related error processing study. The study was
designed following the experimental protocol approved by the Institutional Review
Board (IRB) of the Michigan State University. The data collection was performed in
accordance with the guidelines and regulation established by this protocol. Written and
informed consent was collected from each participant before data collection.
The experiment consisted of a speeded-reaction Flanker task (Eriksen & Eriksen,
1974), in which subjects identified the middle letter on a five-letter string, being con-
gruent (e.g. MMMMM) or incongruent (e.g. MMNMM) with respect to the Flanker
letters. Flanker letters (e.g. MM MM) were shown during the first 35 ms of each trial, and
during the following 100 ms the Flanker and target letters were shown on the screen.
This was followed by an inter-trial interval of variable duration ranging from 1200 ms
to 1700 ms. A total of six blocks consisting of 80 trials composed the experiment, and
letters were changed between blocks. EEG responses were recorded by the 64 elec-
trode ActiveTwo system (BioSemi, Amsterdam, The Netherlands). The sampling fre-
quency was 512 Hz. Trials containing artifacts were rejected and volume conduction was
reduced through the Current Source Density (CSD) Toolbox (Kayser & Tenke, 2006).
A total of 18 subjects and 58 channels were considered for the analysis, for which the
total number of error trials ranged from 20 to 61. The same number of correct responses
was chosen randomly. In this example, we explore the effectiveness of the different
multivariate measures by applying them to the N × N bivariate functional connectivity
matrices corresponding to error-related negativity (ERN) and the correct-related nega-
tivity (CRN). Previous studies have shown that the ERN is associated with increased
synchronization in the theta band (4–8 Hz) between electrodes in the central and lat-
eral frontal regions (Aviyente et al., 2011). For this reason, an FCN was constructed for
each subject by averaging the PLV over the time window 25–75 ms and the frequency
bins corresponding to the theta band per subject and response type. This results in two
FCNs of size 58 × 58 per subject, one corresponding to error responses and the other to
correct responses.
Table 21.1 summarizes the results of applying four of the metrics discussed in this
chapter. For the S-estimator and Omega complexity, in both cases the results indicate
that the FCNs for the correct response have higher levels of synchronization compared
to the error response. However, these measures do not necessarily quantify how this
increased synchronization is distributed across the network and does not reflect the
organization of the network. The graph theoretic metrics, on the other hand, address
this issue. We note that the small-world measure is slightly higher for error networks
compared to correct ones. In particular, the FCNs constructed from error responses
are small-world while the ones from correct response are not. This difference between
MULTIVARIATE METHODS FOR FUNCTIONAL CONNECTIVITY ANALYSIS 527
other hand, lateral-medial synchrony is significantly different than that within the oc-
cipital regions.
21.6 Conclusions
This chapter reviewed different metrics that have been introduced to quantify multi-
variate synchronization in the brain. The first group of methods focus on spectral
properties of the bivariate phase synchrony matrix and as such quantify the spread of
the eigenvalues of this N × N matrix. Even though these measures are successful at
capturing the global synchronization within the brain, they do not pinpoint to the ac-
tual mechanisms underlying the observed global synchrony values. Graph theoretic
metrics, on the other hand, can characterize the network topology across multiple scales
ranging from the micro to the macro-scale. Thus, they can provide explanations to the
observed synchrony values in terms of different network models such as small-world
and modular structures. In order to better characterize brain networks, these analyses
need to be performed at the subnetwork level. The recently proposed hyper-torus phase
synchrony measure addresses this issue by computing the synchronization within a
group of electrodes or a subnetwork. This type of analysis offers the best trade-off be-
tween bivariate metrics that focus on pairs of electrodes and multivariate metrics that
focus on the whole brain.
REFERENCES
Achard, S., Salvador, R., Whitcher, B., Suckling, J., & Bullmore, E. T. (2006). A resilient, low-
frequency, small-world human brain functional network with highly connected association
cortical hubs. Journal of Neuroscience, 26(1), 63–72.
Al-Khassaweneh, M., Villafane-Delgado, M., Yener Mutlu, A., & Aviyente, S. (2016). A measure
of multivariate phase synchrony using hyperdimensional geometry. IEEE Transactions on
Signal Processing, 64(11), 2774–2787.
Allefeld, C. & Bialonski, S. (2007). Detecting synchronization clusters in multivariate time
series via coarse-graining of Markov chains. Physical Review E, 76(6), 66207–66215.
Allefeld, C., Frisch, S., & Schlesewsky, M. (2005). Detection of early cognitive processing by
event-related phase synchronization analysis. Neuroreport, 16(1), 13–16.
Allefeld, C. & Kurths, J. (2003). Multivariate phase synchronization analysis of EEG data. IEICE
Transactions on Fundamentals of Electronics, Communications and Computer Sciences,
86(9), 2218–2221.
Allefeld, C. & Kurths, J. (2004). An approach to multivariate phase synchronization analysis
and its application to event-related potentials. International Journal of Bifurcation and
Chaos, 14(2), 417–426.
Allefeld C., Müller, M., & Kurths, J. (2007). Eigenvalue decomposition as a generalized syn-
chronization cluster analysis. International Journal of Bifurcation Chaos, 17, 3493–3497.
MULTIVARIATE METHODS FOR FUNCTIONAL CONNECTIVITY ANALYSIS 529
Aviyente, S., Bernat, E. M., Evans, W. S., & Sponheim, S. R. (2011). A phase synchrony measure
for quantifying dynamic functional integration in the brain. Human Brain Mapping,
32(1), 80–93.
Aviyente, S. & Yener Mutlu, A. (2011). A time-frequency-based approach to phase and phase
synchrony estimation. IEEE Transactions on Signal Processing, 59(7), 3086–3098.
Aydore, S., Pantazis, D., & Leahy, R. M. (2013). A note on the phase locking value and its
properties. Neuroimage, 74, 231–244.
Baccala, L. A. & Sameshima, K. (2001). Partial directed coherence: A new concept in neural
structure determination. Biological Cybernetics, 84(6), 463–474.
Bassett D. S. & Bullmore, E. (2006). Small-world brain networks. The Neuroscientist, 12(6),
512–523.
Bassett D. S. & Bullmore, E. T. (2017). Small-world brain networks revisited. The Neuroscientist,
23(5), 499–516.
Bassett, D. S., Wymbs, N. F., Porter, M. A., Mucha, P. J., Carlson, J. M., & Grafton, S. T. (2011).
Dynamic reconfiguration of human brain networks during learning. Proceedings of the
National Academy of Sciences, 108(18) 7641–7646.
Bastos, A. M. & Schoffelen, J. M. (2016). A tutorial review of functional connectivity analysis
methods and their interpretational pitfalls. Frontiers in Systems Neuroscience [online],
9, 175.
Bolanos, M., Bernat, E. M., He, B., & Aviyente, S. (2013). A weighted small world network
measure for assessing functional connectivity. Journal of Neuroscience Methods, 212(1),
133–142.
Bullmore E. T. & Bassett, D. S. (2011). Brain graphs: Graphical models of the human brain
connectome. Annual Review of Clinical Psychology, 7, 113–140.
Bullmore, E. & Sporns, O. (2009). Complex brain networks: Graph theoretical analysis of
structural and functional systems. Nature Reviews Neuroscience, 10(3), 186–198.
Buzsáki, G. & Draguhn, A. (2004). Neuronal oscillations in cortical networks. Science,
304(5679), 1926–1929.
Carmeli, C., Knyazeva, M. G., Innocenti, G. M., & De Feo, O. (2005). Assessment of EEG syn-
chronization based on state-space analysis. Neuroimage, 25(2), 339–354.
Chavez, M., Valencia, M., Navarro, V., Latora, V., & Martinerie, J. (2010). Functional modu-
larity of background activities in normal and epileptic brain networks. Physical Review
Letters, 104(11), 118701.
Dimitriadis, S. I., Laskaris, N. A., & Tzelepi, A. (2013). On the quantization of time-varying
phase synchrony patterns into distinct functional connectivity microstates (fcμstates) in a
multi-trial visual ERP paradigm. Brain Topography, 26(3), 397–409.
de Vico Fallani, F., Richiardi, J., Chavez, M., & Achard, S. (2014). Graph analysis of functional
brain networks: Practical issues in translational neuroscience. Philosophical Transactions of
the Royal Society B: Biological Sciences, 369(1653), 20130521.
Eriksen, B. A. & Eriksen, C. W. (1974). Effects of noise letters upon the identification of a target
letter in a nonsearch task. Perception & Psychophysics, 16(1), 143–149.
Fair, D. A., Cohen, A. L., Power, J. D., Dosenbach, N. U. F., Church, J. A., Miezin, F. M.,
Schlaggar, B. L., & Petersen, S. E. (2009). Functional brain networks develop from a “local to
distributed” organization. PLoS Computational Biology, 5(5), e1000381.
Ferrarini, L., Veer, I. M., Baerends, E., van Tol, M-J., Renken, R. J., van der Wee, N.
J. A., . . . Milles, J. (2009). Hierarchical functional modularity in the resting-state human
brain. Human Brain Mapping, 30(7), 2220–2231.
530 SELIN AVIYENTE
Power, J. D., Cohen, A. L., Nelson, S. M., Wig, G. S., Barnes, K. A., Church, J. A., . . . Petersen, S.
E. (2011). Functional network organization of the human brain. Neuron, 72(4), 665–678.
Quiroga, R. Q., Kraskov, A., Kreuz, T., & Grassberger, P. (2002). Performance of different
synchronization measures in real data: A case study on electroencephalographic signals.
Physical Review E, 65(4), 041903.
Rubinov, M. & Sporns, O. (2010). Complex network measures of brain connectivity: Uses and
interpretations. NeuroImage, 52(3), 1059–1069.
Schelter, B., Winterhalder, M., Dahlhaus, R., Kurths, J., & Timmer, J. (2006). Partial phase
synchronization for multivariate synchronizing systems. Physical Review Letters, 96(20),
208103.
Sporns, O. (2018). Graph theory methods: Applications in brain networks. Dialogues in Clinical
Neuroscience, 20(2), 111.
Sporns, O., Honey, C. J., & Kötter, R. (2007). Identification and classification of hubs in brain
networks. PloS One, 2(10), e1049.
Stam, C. J., Nolte, G., & Daffertshofer, A. (2007). Phase lag index: Assessment of functional
connectivity from multi-channel EEG and MEG with diminished bias from common
sources. Human Brain Mapping, 28(11), 1178–1193.
Stam C. J. & Reijneveld, J. C. (2007). Graph theoretical analysis of complex networks in the
brain. Nonlinear Biomedical Physics, 1(1), 1–19.
Stam, C. J. & Van Dijk, B. W. (2002). Synchronization likelihood: An unbiased measure of
generalized synchronization in multivariate data sets. Physica D: Nonlinear Phenomena,
163(3–4), 236–251.
Tononi, G., Sporns, O., & Edelman, G. M. (1994). A measure for brain complexity: relating
functional segregation and integration in the nervous system. Proceedings of the National
Academy of Sciences of the United States of America, 91(11), 5033–5037.
Uhlhaas, P. J. & Singer, W. (2006). Neural synchrony in brain disorders: Relevance for cognitive
dysfunctions and pathophysiology. Neuron, 52(1), 155–168.
Uhlhaas P. J. & Singer, W. (2010). Abnormal neural oscillations and synchrony in schizo-
phrenia. Nature Reviews Neuroscience, 11(2), 100–113.
van den Heuvel, M. P., Stam, C. J., Boersma, M., & Hulshoff Pol, H. E. (2008). Small-world and
scale-free organization of voxel-based resting-state functional connectivity in the human
brain. NeuroImage, 43(3), 528–539.
Watts, D. J. & Strogatz, S. H. (1998). Collective dynamics of “small-world” networks. Nature,
393(6684), 440–442.
Yener Mutlu, A. & Aviyente, S. (2012). Hyperspherical phase synchrony for quantifying multi-
variate phase synchronization. 2012 IEEE Statistical Signal Processing Workshop (SSP),
888–891.
CHAPTER 22
BRAIN STIM U L AT I ON
APPROAC H E S TO
INVESTIGAT E E E G
OSCILL AT I ONS
22.1 Introduction
Several of the previous chapters in this book laid out the numerous associations be-
tween oscillatory brain activity and various domains of cognitive functioning that have
been discovered over the course of the last century. At the same time, altered patterns of
this oscillatory brain activity have been observed in many neurological and psychiatric
diseases (Herrmann & Demiralp, 2005; Uhlhaas & Singer, 2006, 2012).
The recording of brain signals using electro-or magnetoencephalography
(EEG/ MEG) or by means of invasive techniques (e.g., intracortical EEG or
electrocorticography), has strongly contributed to our knowledge in this area, and
will continue to do so. However, as these methods can only provide observational
data, inference about the functional role of brain oscillations for cognitive functions
remains correlational. That is, oscillatory activity is observed as the dependent vari-
able while participants are engaged in different cognitive tasks or task conditions,
which are experimentally manipulated (independent variable, Figure 22.1A). Whether
the observed oscillatory activity has a direct causal influence on the investigated
function, or whether it reflects byproducts of the underlying neural processing,
cannot be resolved by this type of experimental design. In order to demonstrate such
causal relationships, one needs to revert the design and experimentally manipulate
the oscillatory activity in question (independent variable) and measure the resulting
BRAIN STIMULATION APPROACHES TO INVESTIGATE EEG OSCILLATIONS 533
Performance [%]
5 80
0
5 10 15 20 60
Frequency [Hz]
20 40
15
B 10 20
5
0 5 0
10 15 20
Frequency [Hz]
Figure 22.1 Experimental designs to investigate brain oscillations. (A) In traditional EEG
experiments participants are exposed to different experimental conditions (e.g. a task free fix-
ation interval and a stimulus). An EEG is recorded and the frequency content of the signal is
compared between task conditions and correlated with task performance. (B) In order to estab-
lish causal relationships between the brain oscillation and the cognitive process, one needs to
revert the design and manipulate a feature of the oscillation of interest (here power around 10 Hz)
by an intervention. If behavioral changes are observed in response to the intervention, this can be
taken as evidence for a causal involvement of the oscillations in the investigated process.
behavioral changes as the dependent variable (Bergmann et al., 2016; Herrmann et al.,
2016b; Figure 22.1B).
In principle, a variety of methods allow researchers to modulate brain oscillations.
However, many of these approaches, such as pharmacological interventions (e.g.,
Kopp et al., 2004), optogenetics (Sohal, 2012), or intracranial electrical stimulation
(Alagapan et al., 2016; Fröhlich & McCormick, 2010), are highly invasive and their ap-
plication mostly restricted to animal models and small groups of patients. For example,
light-driven activation of fast-spiking interneurons at 40 Hz has been shown to cause
gamma band increase of local field potentials in mice (Cardin et al., 2009) and low-
frequency direct cortical stimulation in the alpha range has been shown induce state-
dependent modulation of oscillatory brain activity in the alpha and theta band in
epilepsy patients (Alagapan et al., 2016). While invasive stimulation offers potential for
more reliable effects due to stronger and more focal perturbation of brain activity, they
require opening of the scalp and skull or even penetration of brain tissue making them
unsuitable for application in healthy human subjects. Other methods like rhythmic
534 FLORIAN H. KASTEN and CHRISTOPH S. HERRMANN
stimulation with sensory stimuli (e.g., rhythmic visual stimulation with flickering
light causes steady state visual evoked responses) or neurofeedback training have been
shown to modulate oscillatory activity in the brain and can be applied in healthy human
subjects. However, steady-state responses exhibit their effects on brain oscillations
indirectly via the sensory systems and it is debated whether the observed oscillatory
activity reflects a modulation of brain oscillations or merely a superposition of event-
related potentials (but see Notbohm et al., 2016). Neurofeedback training can be used
in many different contexts, including studying higher cognitive processes. However, the
intervention is rather time consuming as the training usually requires several sessions
on separate days before its effects can be measured.
During the past decades, a branch of techniques has evolved to noninvasively
modulate brain activity by either applying magnetic pulses via transcranial magnetic
stimulation (Barker et al., 1985; Hallett, 2000), or weak electric currents by means of
transcranial electrical stimulation (tES). The latter is an umbrella term covering sev-
eral electrical stimulation methods including transcranial direct current stimulation
(tDCS), transcranial alternating current stimulation (tACS), and transcranial random
noise stimulation (Woods et al., 2016). Among these methods, the rhythmic, repetitive
application of TMS (rTMS) and tACS are considered particularly promising to study
the functional role of brain oscillations in cognition. The latter works via the application
of an alternating current usually of sinusoidal shape. Although, depending on the spe-
cific limitations of the hardware used for stimulation, almost any type of waveform can
be created (Figure 22.2). For example, recent work applied tACS using the envelope of
tDCS tACS
Figure 22.2 Stimulation waveforms used in tES and TMS. Top: Waveforms conventionally
used for tDCS and tACS. Middle: Alternative waveforms used to stimulate oscillatory activity
in the brain. For otDCS, a sinusoidal tACS waveform is combined with a DC offset. TACS using
sawtooth waves cause large power in the spectrum at harmonic frequencies, which makes re-
sidual artifacts easier to detect. In amplitude modulated-tACS a high-frequency carrier wave-
form is modulated in amplitude by a low-frequency waveform at the frequency of the target brain
oscillation. For envelope-tACS the stimulation waveform is extracted from the envelope of a
speech signal. Bottom: Waveform shapes of mono-and biphasic TMS pulses.
BRAIN STIMULATION APPROACHES TO INVESTIGATE EEG OSCILLATIONS 535
human speech signals to improve speech intelligibility (Riecke et al., 2018; Wilsch et al.,
2018). Others administered tACS with sawtooth waves or amplitude modulated sine
waves that are supposed to allow easier removal of stimulation artifacts from concur-
rent electrophysiological recordings (Dowsett & Herrmann, 2016; Kasten et al., 2018b;
Witkowski et al., 2016). Some researchers also applied alternating currents together with
a DC offset, which has been referred to as oscillating transcranial direct current stimu-
lation (otDCS; Marshall et al., 2006; Neuling et al., 2012a). RTMS and tACS are believed
to synchronize/entrain endogenous brain oscillations to the externally applied driving
force, thus specifically targeting the brain oscillation of interest and probing its causal
role during a particular task (Herrmann et al., 2013; Reato et al., 2013; Thut et al., 2011b,
2011a). In addition, effects outlasting the duration of stimulation by several minutes or
even hours have been observed in many rTMS and tACS studies (Veniero et al., 2015).
These findings give rise to the hope that in the future these methods may offer new
treatment options for psychiatric and neurological conditions by restoring dysfunc-
tional oscillatory activity.
TMS exploits the principles of electromagnetic induction of electric fields in the brain
(Barker et al., 1985; Wagner et al., 2009). To this end, strong, transient currents are fed
through a coil of wire placed above the scalp in the proximity of the targeted brain
area. The high-intensity current creates a magnetic field with magnetic flux passing
perpendicularly to the plane of the coil. The rapidly changing magnetic field in turn
passes through the skull and induces current in the brain tissues underneath the scalp,
flowing in loops parallel to the coil plane (Figure 22.3A; Hallett, 2000; Wagner et al.,
2009). The resulting electric fields cause changes in membrane polarization sufficiently
strong to modulate neural excitability by changing membrane polarization and even
trigger the firing of action potentials (Wagner et al., 2009). Different coil shapes, like
circular, double cone, or figure of eight, are available that vary with respect to their in-
tensity, focality, and the depth of brain areas that can be reached with the stimulation
(Hallett, 2000; Klomjai et al., 2015). TMS can be administered using different stimula-
tion protocols. The application of a single TMS pulse can already elicit effects and im-
pair visual perception (Amassian et al., 1989). Traditionally, low-frequency rTMS (<1
Hz) is considered to suppress cortical excitability while high-frequency rTMS (>5Hz)
is considered facilitatory (Klomjai et al., 2015). Variants of rTMS include theta burst
stimulation, where short bursts of high-frequency stimulation (40 Hz) are repeated at
theta frequency (5 Hz). Besides the overall excitatory and inhibitory effects of rTMS,
the method can also be used to modulate and investigate brain oscillations (Hanslmayr
et al., 2014; Thut et al., 2011b).
536 FLORIAN H. KASTEN and CHRISTOPH S. HERRMANN
Intensity [a.u.]
oscillation adapts fully synchronized
EEG
tACS
0
∆Frequency [Hz]
(g)
Synchronization
area
Intensity [a.u.]
Figure 22.3 Modes and mechanisms of stimulation. (A) Figure-eight coil TMS and induced
magnetic/electric field. (B) Simple tACS montage with electrodes placed above Cz and Oz of the
international 10–20 system to target occipital-parietal regions. (C & D) Electrode montages to
modulate inter-hemispheric synchronization. Stimulation between electrode pairs is applied ei-
ther in-phase (C) or anti-phase (D). (E) Synchronization of endogenous brain oscillations to ex-
ternal stimulation via rTMS (top) and tACS (bottom). The external driving force causes the brain
oscillation to adapt in frequency and phase to the stimulation. Due to the increased synchronous
neuronal activity, an amplitude increase is expected. Please note that real EEG recordings are
strongly corrupted by artifacts arising from stimulation. (F) The “Arnold tongue” illustrates the
relationship between the intensity of a driving force coupled to an oscillator and their frequency
differences on the resulting synchronization. The farther apart the frequency of the driving
force and the oscillator, the higher the intensity needed to cause the oscillator to synchronize.
If the frequencies of the oscillator and driving force match, very small intensities are needed.
(G) Arnold tongues at different n:m (n perturbations within m oscillatory cycles) frequency
ratios.
BRAIN STIMULATION APPROACHES TO INVESTIGATE EEG OSCILLATIONS 537
In contrast to TMS, tES methods like tDCS and tACS have much more subtle effects
on neuronal activity. Here, a direct or alternating current is passed through the scalp
by two or more stimulation electrodes (Figure 22.3B–D). Commonly, saline soaked
sponge electrodes, fixed to the scalp via rubber bands, or electrically conductive rubber
electrodes attached using a conductive, adhesive paste are used for this purpose.
Some systems also make use of Ag/AgCl electrodes similar to those used in standard
EEG systems. Results from simulations of the current flow through the head, as well
as findings from invasive recordings in animals and human cadavers suggest that a
large proportion of the current is directly shunted through the skin (Miranda et al.,
2006; Neuling et al., 2012b; Opitz et al., 2016; Vöröslakos et al., 2018). However, at the
same time, small amounts of the injected current still reach the target regions in the
brain. In case of tDCS, these currents shift the resting potentials of the stimulated
neurons, resulting in slight depolarization of the cell membrane under the anode and
hyperpolarization under the cathode. While these changes in membrane potentials are
too small to directly elicit neural firing, early work on animal models in vivo and in
vitro suggests that they can alter neuronal excitability by shifting the membrane poten-
tial such that more or less incoming excitatory postsynaptic potentials are required to
reach the cells’ firing threshold (Bikson et al., 2004; Bindman et al., 1964; Creutzfeldt
et al., 1962; Jefferys, 1981).
When an alternating current is applied (e.g., via tACS), the electric field rhythmic-
ally alternates between hyperpolarizing and depolarizing the cell membrane during
the negative and positive half cycles of the sinusoidal stimulation waveform. Computer
simulations as well as in vivo and in vitro recordings demonstrated that these phasic
modulations of membrane polarization temporally align neural firing to the frequency
of the externally applied stimulation (Deans et al., 2007; Fröhlich & McCormick, 2010;
Krause et al., 2019; Ozen et al., 2010; Reato et al., 2010). The electric field strengths neces-
sary to achieve this effect are comparable to the electric fields reaching cortical pyram-
idal cells during tES (Antal & Herrmann, 2016; Opitz et al., 2016).
Although the mechanisms of tACS and rTMS differ in terms of their effects on the
cellular level, both techniques are assumed to cause synchronization/entrainment of the
endogenous oscillatory brain activity to the external driving force (Thut et al., 2011a).
Synchronization phenomena can be observed everywhere in nature, from chirping of
crickets to the beating of the heart, as well as in human-made technical systems like pen-
dulum clocks or electric circuits. The basic principles of these phenomena are universal
and can be described using the framework of synchronization theory (Pikovsky et al.,
2003). Synchronization requires the presence of a so-called self-sustained oscillator.
Such oscillators are characterized by an internal source of energy, allowing the oscillator
to exhibit some sort of rhythmic activity until the energy source is consumed (Pikovsky
et al., 2003). If two or more oscillators with similar eigenfrequencies (i.e., the frequency
at which a system tends to oscillate in the absence of any external driving force) are
weakly coupled, their rhythmic activity synchronizes; that is, they align their rhythmic
activity in frequency and phase (Pikovsky et al., 2003). Similarly, if coupled to an ex-
ternal driving force, the oscillator synchronizes in frequency and phase to the external
538 FLORIAN H. KASTEN and CHRISTOPH S. HERRMANN
signal. Importantly, the further apart the frequency of the external force is to the eigen-
frequency of the oscillator, the more energy is needed to synchronize it (Pikovsky et al.,
2003). Driving forces at frequencies too far away from the oscillators’ eigenfrequency
and with too little strength cannot achieve synchronization. In contrast, driving forces
that exactly match the frequency of the oscillator can achieve synchronization even at
very small intensities. When the relationship between frequency difference and driving
force strength on the synchronization strength is visualized in a diagram with synchron-
ization strength coded in color, a triangular shape can be observed, which has been
referred to as the Arnold Tongue (Figure 3G; Pikovsky et al., 2003). Besides 1:1 frequency
matching between oscillator and driving force, synchronization can also occur if the
driving force perturbs the oscillator at harmonic or subharmonic frequencies. Such har-
monic entrainment can occur for any n:m (n perturbations within m oscillatory cycles)
frequency relationship (where n, m ∈ ). For example, an oscillator with a frequency of
10 Hz synchronizes to driving forces with frequencies around 10 Hz, but also to those
with 2.5 Hz, 3.33 Hz, 6.66 Hz, 15 Hz, 20 Hz, and 30 Hz, corresponding to the 1:3, 1:2, 2:3,
3:2, 2:1, and 3:1 Arnold tongues (Figure 22.3G). The strength of synchronization decays
the higher the harmonic frequency of the external driving force is (Pikovsky et al., 2003).
In the human brain, synchronization of oscillatory activity to external driving
forces has been observed in response to flickering light stimulation (Herrmann, 2001;
Notbohm et al., 2016) and rTMS (Herrmann et al., 2016a; Thut et al., 2011b). Further,
the effect of sinusoidal alternating currents on local field potentials and multi-unit ac-
tivity follows the rules of synchronization theory in computer simulations as well as
in in vivo and in vitro recordings (Fröhlich & McCormick, 2010; Krause et al., 2019;
Negahbani et al., 2018; Ozen et al., 2010; Reato et al., 2010). However, as recordings of M/
EEG activity during stimulation are contaminated by a strong, electromagnetic artifact,
direct evidence for these mechanisms of actions for tACS effects in humans is largely
missing so far (an overview of approaches for tACS artifact removal in M/EEG and their
associated problems is reviewed in Kasten & Herrmann, 2019).
These mechanisms describe the general framework currently assumed to underlie
effects of rTMS and tACS during stimulation (also called online effects). A common ob-
servation in brain stimulation experiments is that behavioral and physiological effects
persist after the stimulation is switched off (reviewed in Veniero et al., 2015). These
aftereffects (or offline effects) can last for several minutes or even hours after both rTMS
(Schindler et al., 2008; Schutter et al., 2001) and tACS (Kasten et al., 2016; Neuling et al.,
2013; Wischnewski et al., 2018). After tACS, these outlasting effects tend to be rather fre-
quency specific. The application of rTMS, in contrast, seems to elicit more broadband
aftereffects in all frequency bands irrespective of the stimulation frequency (Veniero
et al., 2015). While the pattern of effects observed during this sustained period may
appear similar to those observed during stimulation, it is important to emphasize that
the underlying mechanisms responsible for those effects may be different. This phe-
nomenon has been well documented for effects of tDCS on motor evoked potentials
(MEPs). Here, the selective, pharmacological blockage of NMDA receptors abolished
the induction of long-lasting aftereffects, while not affecting online effects of the
BRAIN STIMULATION APPROACHES TO INVESTIGATE EEG OSCILLATIONS 539
stimulation (Nitsche et al., 2003). In contrast, blockage of calcium and sodium channels
abolished or reduced the enhancing effect of anodal tDCS on MEP size during and after
stimulation (Nitsche et al., 2003). These results indicate that the online effect of tDCS
depends on membrane polarization, modulating the conductance of sodium and cal-
cium channels. Offline effects appear to be caused by NMDA receptor mediated plas-
ticity, although online effects seem to be necessary to induce the offline effect (Nitsche
et al., 2003).
While the mechanism underlying online effects of rTMS and tACS is assumed to be
entrainment, offline effects have been suggested to be caused by entrainment echoes, a
state of sustained synchronization after stimulation is switched off (Hanslmayr et al.,
2014), or processes of spike-timing dependent plasticity (STDP; Vossen et al., 2015;
Zaehle et al., 2010). STDP relies on the temporal relation of pre-and postsynaptic
potentials. If a presynaptic potential repeatedly precedes a postsynaptic potential, the
synaptic connection is strengthened, which is referred to as long-term potentiation
(LTP). If the presynaptic potential repeatedly follows the postsynaptic potential, long-
term depression (LTD) occurs and the synaptic connection is weakened (Markram
et al., 1997). In a neural circuit that generates an oscillation, pathways exist by which the
activity of a neuron is fed back to its own synaptic input connections via other neurons.
Depending on the length and speed of these connections, each spike requires a certain
time to travel through these circuits, which determines its intrinsic frequency (Zaehle
et al., 2010). Following the principles of STDP, repetitive stimulation (e.g., with rTMS or
tACS) with frequencies slightly below the circuits’ intrinsic frequency should lead to a
strengthening of synaptic connections because stimulation can systematically precede
the circulating action potentials. The application of other frequencies in turn, cannot
cause such effects, as the stimulation does not match the circuits’ intrinsic frequency,
and membrane polarization does not coincide with incoming neural spiking (Vossen
et al., 2015; Zaehle et al., 2010).
Although entrainment echoes may exist during the first few seconds after stimula-
tion, the STDP model seems more suited to explain the long-lasting changes elicited
by tACS and rTMS, which have been observed for up to an hour (Kasten et al., 2016;
Veniero et al., 2015; Vossen et al., 2015). Wischnewski and colleagues (2018) provided
direct evidence for such involvement of synaptic plasticity in the generation of tACS
aftereffects. In that study, administration of an NMDA receptor antagonist was observed
to abolish aftereffects induced by tACS in the beta range over the primary motor cortex.
This suggests a role of NMDA receptor-mediated plasticity in the generation of tACS
aftereffects similar to those observed after tDCS (Nitsche et al., 2003).
During the design of a brain stimulation experiment a couple of choices have to be made
that can severely impact both the stimulation success and the interpretability of results.
540 FLORIAN H. KASTEN and CHRISTOPH S. HERRMANN
This section provides an overview of these choices and examples of how different
features of brain oscillations are targeted to investigate their causal role for different
domains of cognitive functioning.
Usually, the design process starts with identifying a brain oscillation that correlates
with a cognitive function, or ideally an output measure of a concrete task (e.g., reac-
tion time or performance). Depending on the kind of association, brain stimulation can
be used to target different features of the oscillation, such as its amplitude, frequency,
or phase. Some relationships may also depend on cross-frequency interactions (Palva
et al., 2005) or synchronization between distinct brain areas (Siegel et al., 2012). Another
important aspect is related to the stimulation montage. One needs to decide about the
positioning and orientation of the stimulation electrodes or the TMS coil. Usually, the
observed correlations between a cognitive function and a brain oscillation are limited to
specific areas of the brain. Oscillations in the same frequency range may serve different
functions within distinct areas of the brain. Based on neuroimaging results, for ex-
ample, from source localization of oscillatory activity, a target region in the brain is
derived and a montage is chosen that maximizes the electric field strength in this area.
A couple of software tools are available that simulate the expected current flow caused
by transcranial magnetic and electrical stimulation (e.g., ROAST: Huang et al., 2017a;
Simnibs: Thielscher et al., 2015). Recent validation studies using invasive electrophysi-
ology in animals and humans indicate that these models are able to predict the spa-
tial extent of the electric field in the brain quite accurately but tend to overestimate the
strength of the induced electric fields (Huang et al., 2017b; Opitz et al., 2016). Even more
advanced tools are able to automatically optimize electrode montages to target specific
brain areas with specific directions of current flow (Baltus et al., 2018b; Huang et al.,
2018; Saturnino et al., 2019; Wagner et al., 2016).
As soon as a feature of an oscillation and a brain region are selected for stimulation,
appropriate control conditions for the experiment must be established. The choice
of adequate control conditions is not a trivial problem. Stimulation using TMS or
tACS can induce visual and somatosensory perceptions that may confound results if
compared to a stimulation-free control condition (Schutter, 2016; Turi et al., 2013, 2014).
In addition, one would ideally compare the selected stimulation condition with stimu-
lation of all other possible stimulation frequencies at all possible brain regions in order
to demonstrate frequency and region-specific effects. However, due to a huge param-
eter space this is practically impossible. Let us consider a very simple case of tACS with
only two electrodes that can be placed on any of the 32 locations of the international
10–20 system for EEG electrodes. In addition, stimulation frequencies are limited to
only one frequency within each of the five major EEG frequency bands. This param-
eter space already amounts to almost 5,000 possible control conditions. The inclusion
of more fine-grained electrode locations and stimulation frequencies or stimulating at
different intensities easily increases the number of possible control conditions to sev-
eral million. In practice, stimulation of the target oscillations is thus usually compared
to one or two control frequencies applied to the same region, or the same stimulation
frequency is applied to one or two control regions. Other researchers also compare
BRAIN STIMULATION APPROACHES TO INVESTIGATE EEG OSCILLATIONS 541
If the frequency of a brain oscillation is of interest, one would try to speed up or slow
the target oscillation by applying stimulation at frequencies slightly above or below its
intrinsic frequency. According to synchronization theory, such stimulation should shift
the intrinsic frequency in the brain towards the frequency of the external driving force
(Pikovsky et al., 2003). This approach has been used, for example, to investigate the causal
role of occipital alpha oscillations in the generation of the sound-induced double-flash
illusion (Cecere et al., 2015). The illusion occurs when two sound beeps are presented
within a time window of ~100 ms together with a single visual flash. Participants per-
ceive a second illusory flash in this kind of experiment (Shams et al., 2002). The authors
were able to alter the length of the temporal window during which the illusion occurs by
stimulating participants ± 2 Hz above or below their individual alpha frequency (Cecere
et al., 2015). In another experiment, the perceived frequency of an illusory jitter was
modulated in a similar manner by applying amplitude modulated tACS ± 1 Hz above or
below participants’ individual alpha frequency (Minami & Amano, 2017). In the audi-
tory domain, tACS has been used to test whether the frequency of gamma oscillations
determines the temporal resolution of the auditory system. In that study, participants
were stimulated above or below their individual gamma frequency (Baltus et al., 2018a,
2018b). By accelerating the individual gamma frequency, participants were able to de-
tect smaller gaps in continuous sound streams of noise. However, when the individual
gamma frequency was decelerated, no modulation of gap detection performance was
observed (Baltus et al., 2018a, 2018b). Beyond investigations into basic sensory pro-
cessing, the frequency of brain oscillations had also been modulated in the context of
working memory performance. It is well known that humans can uphold 7 ± 2 items in
their working/short term memory (Miller, 1956). Later, this capacity has been associated
with individual gamma and theta frequencies. Specifically, it is argued that the number
of gamma cycles that fit into the positive half wave of a theta oscillation determines the
capacity (Lisman & Idiart, 1995; Lisman & Jensen, 2013). Vosskuhl and colleagues (2015)
tested this relationship and demonstrated that slowing down participants’ theta fre-
quency with tACS such that more gamma cycles fit into one theta oscillation increased
working memory capacity in a digit span task.
Investigations into the role of oscillatory phase or phase relationships for cogni-
tion are comparatively challenging as they require precise stimulus timing or complex
stimulation protocols. Neuling and colleagues (2012a) presented brief auditory targets
at specific phases of a 10 Hz otDCS waveform applied to the auditory cortex. In line
with previous correlational evidence (Mathewson et al., 2009), participants’ detection
performance systematically varied depending on the phase of the 10-Hz stimulation
at which the target was presented. Similar modulations of detection performance with
tACS phase was found for 4-Hz stimulation in an auditory detection task (Riecke et al.,
2015) and for stimulation of somatosensory cortex at mu frequency in a somatosensory
detection task (Gundlach et al., 2016). In a different type of experiment, authors applied
tACS to two distinct brain regions with the same stimulation waveform applied either
in or out of phase (with 0-or 180-degree phase shift, Figure 22.3C,D). This way, stimula-
tion is thought to either increase (in phase) or decrease (out of phase) the synchrony of
BRAIN STIMULATION APPROACHES TO INVESTIGATE EEG OSCILLATIONS 543
the targeted brain regions within a specific frequency band. Such modulation of inter-
regional synchrony, when applied in the gamma frequency range, modulates percep-
tion of ambiguous movements (Helfrich et al., 2014a; Strüber et al., 2014). Increasing/
decreasing fronto-parietal synchronization in the theta frequency range in the left
hemisphere with in-phase vs. anti-phase tACS has been shown to increase/decrease
reaction times in a visual memory-matching task (Polanía et al., 2012). Other authors
extended this concept to even target interpersonal synchrony (Novembre et al., 2017;
Szymanski et al., 2017). In one study, tACS in the beta range was simultaneously applied
to the motor cortex of two participants (Novembre et al., 2017). Here, in-phase stimula-
tion in the beta range increased synchrony in a joint finger tapping task as compared to
anti-phase stimulation or sham. No such effect was found for stimulation in other fre-
quency bands. In another study, simultaneous tACS with same-phase same-frequency
vs. different-phase different-frequency in the theta range over fronto-parietal sites did
not show effects on a synchronous drumming task (Szymanski et al., 2017). An im-
portant concern about in-phase vs. anti-phase stimulation protocols has been raised
recently. In order to directly infer a phase dependent relationship of a behavioral
measure and a brain oscillation, it is crucial that the only difference between stimu-
lation conditions is in the phase difference between the stimulated areas. Saturnino
and colleagues (2017) simulated the electric fields created by different stimulation
montages targeting inter-regional synchrony with in-phase and anti-phase stimulation.
Their results demonstrated that the electric field patterns differed during in-and anti-
phase stimulation for many of the employed montages, giving rise to the concern that
effects of the stimulation could also originate from differences in field strength or even
differences in the brain regions that have been (co-)stimulated during the in-and anti-
phase stimulation (Saturnino et al., 2017). However, the authors also suggest stimulation
montages that can avoid such confounds. Especially, montages using a stimulation elec-
trode in the center, surrounded by several return electrodes or a large ring electrode, are
recommended to avoid such confounds (Saturnino et al., 2017).
Besides direct targeting of a brain oscillation with stimulation of the same frequency
band, sometimes cross-frequency interactions are of interest or used to indirectly
modulate an oscillation via stimulation of a different frequency band. For example,
oscillations in the alpha and gamma range are well known to show an antagonistic re-
lationship. When alpha oscillations increase, gamma oscillations are suppressed, and
vice versa (Jensen & Mazaheri, 2010). Some authors therefore applied stimulation in
the gamma frequency range to suppress alpha oscillations (Boyle & Frohlich, 2013). As
increased activity in the alpha band is negatively correlated with vigilance. This antag-
onistic stimulation approach has recently been used to counteract increasing reaction
times in a sustained attention task (Loffler et al., 2018). Along the same lines, rTMS in the
alpha range has been observed to alter alpha-gamma cross-frequency coupling during
a visual working memory task (Hamidi et al., 2009). An exceptionally sophisticated
tACS cross-frequency protocol has been used to study theta-gamma coupling in a
working memory task. As discussed earlier, short-term or working memory capacity is
thought to depend on the number of gamma oscillations that fit into the positive half of
544 FLORIAN H. KASTEN and CHRISTOPH S. HERRMANN
a theta cycle (Lisman and Idiart, 1995; Lisman and Jensen, 2013), and can be increased
by slowing the frequency of fronto-parietal theta oscillations (Vosskuhl et al., 2015).
To further study the relationship of theta and gamma oscillations in working memory,
Alekseichuk and colleagues (2016) applied tACS using short trains of a high-frequency
stimulation at different gamma frequencies superimposed on either the positive or the
negative half wave of a slow stimulation waveform in the theta range. In that study, only
gamma stimulation applied during the positive but not during the negative half of the
theta cycle fostered participants’ working memory performance.
The aforementioned studies exemplify how brain stimulation can be used to study
the role of various features of oscillatory brain activity for cognition. Especially tACS
can be used very flexibly. As it allows high control over timing and waveform shapes, it
can be used to study complex phase or cross-frequency relationships. In contrast, rTMS
protocols allow to stimulate during specific, transient time periods of a cognitive pro-
cess as it induces stronger/faster effects.
to obtain meaningful results from concurrent TMS-EEG, one needs to account for these
artifacts—ideally by avoiding/reducing them when performing the experiment, for ex-
ample, by carefully arranging the EEG cables relative to the coil or by slowing down the
recharging times of the TMS device (Veniero et al., 2009). The remaining distortions in
the signal have to be removed by advanced data processing procedures.
The data segment containing the massive, sharp artifact of the TMS pulse (lasting 10–
25 ms after the pulse) is commonly removed from the data and subsequently replaced
by a linear or cubic interpolation (Bergmann et al., 2012; Thut et al., 2011b). The same
approach can be used to get rid of the recharger artifact following the pulse (Thut
et al., 2011b). However, here the latency of the artifact has to be carefully identified as
the artifact strongly depends on the TMS intensity (Veniero et al., 2009). To remove
physiological artifacts from eye movements or muscle activity, blind-source separation
methods such as principle component analysis (PCA), independent component ana-
lysis (ICA), or signal space projection are widely used (Ilmoniemi et al., 2015; Rogasch
et al., 2017).
Transcranial alternating current stimulation can be measured using both EEG and
MEG. Again, the stimulation contaminates the recorded signals with a strong elec-
tromagnetic artifact and recording hardware with sufficient dynamic range is needed
to avoid saturation of the recordings. When combined with EEG, direct connection
of stimulation and recording electrodes via bridges of gel or saline solution should be
avoided. When used in the MEG, stimulation parameters should be carefully tested
on a phantom head to rule out that the electromagnetic artifact imposes harm to the
sensor array. In contrast to the transient stimulation pulses used in TMS, tACS applies
a continuous alternating current. Consequently, recordings will not contain artifact-
free segments as long as the stimulation is switched on, precluding the use of interpol-
ation methods as with TMS. During the last couple of years different strategies have
been employed aiming to suppress the tACS artifact from M/EEG recordings (Kasten &
Herrmann, 2019).
In concurrent tACS-EEG, some authors created a template of the artifact waveform
and subtracted it from the EEG recording (Dowsett & Herrmann, 2016; Helfrich et al.,
2014b; Kohli & Casson, 2015; Voss et al., 2014). The method assumes that the stimulation
artifact has a stable size and shape over time, while signals originating from the brain
fluctuate randomly. To create the template, multiple EEG segments, time-locked to the
same phase of the artifact waveform (e.g., the zero-crossing of the signal), are averaged.
This way, brain activity represented in the segments averages out, while the shape and
size of the artifact is retained. The subtraction of the template should in turn remove the
artifact from the EEG signal, while leaving the superimposed brain activity intact. As
the approach achieved non-optimal results, some authors subsequently applied PCA to
remove residual artifacts from the data (Helfrich et al., 2014b).
In the MEG, spatial filtering approaches, such as synthetic aperture magnetom-
etry (Soekadar et al., 2013) and linearly constrained minimum variance (LCMV)
beamforming (Neuling et al., 2015), have been suggested to suppress artifacts from
tACS and tDCS. Beamformers have been designed to separate signals originating from
BRAIN STIMULATION APPROACHES TO INVESTIGATE EEG OSCILLATIONS 547
different directions and have widespread applications for example in radar and sonar
technologies (Van Veen & Buckley, 1988). In neuroscience, beamformers are used to lo-
calize sources of brain activity seen in M/EEG signals and can work in both the time (Van
Veen et al., 1997) and the frequency domains (Gross et al., 2001). The filters are designed
to pass signals from a specific location in the brain, while attenuating signals from all
other sources. To obtain a spatial map of brain activation, multiple filters with different
spatial pass-bands are constructed on a predefined grid of possible source locations
(Van Veen et al., 1997). The crucial feature of the LCMV beamformer in the context of
concurrent tACS-MEG is its insensitivity to highly correlating sources (Neuling et al.,
2015). The spatial filters of the LCMV beamformer are designed to minimize the vari-
ance of the output signal at each source location (hence the name). If two or more spa-
tially distinct, highly correlating sources are present in the data, the common variance
of the signals cancels to minimize the filter output (Van Veen et al., 1997). Such high
correlations between distinct sources are unlikely to naturally occur in the brain and
the LCMV beamformer is relatively robust to moderate correlations between sources
(Van Veen et al., 1997). During concurrent tACS-MEG, the massive stimulation arti-
fact propagates to virtually all sensors with high consistency. Thus, the beamformer can
cancel out large proportions of the artifact waveform (Neuling et al., 2015). As a conse-
quence, however, the artifact suppression capabilities of the beamformer are naturally
limited by the degree to which the artifact signals are correlated (or uncorrelated) over
the sensor array (Mäkelä et al., 2017).
Both methods, the template subtraction and the beamformer approach, assume a sta-
tionary, invariant artifact waveform, but this assumption has recently been challenged.
Physiological processes such as heartbeat and respiration can lead to small changes
in body impedance and elicit small head movements that can modify the size of the
recorded artifact waveform compromising the artifact suppression capabilities of these
methods (Noury et al., 2016; Noury & Siegel, 2017). These systematic changes in artifact
size manifest in an amplitude modulation of the tACS waveform that causes side-bands
around the main stimulation frequency, which survive artifact suppression attempts
(Noury et al., 2016). Additionally, these processes may also affect artifact suppression
performance directly at the stimulation frequency. Variations in artifact strength cannot
be incorporated when constructing a template of the artifact to be subtracted from EEG
data. As a consequence, any deviation of the artifact strength from the averaged tem-
plate remains in the signal as a residual artifact. In a similar manner, artifact suppression
capabilities of LCMV beamforming might be corrupted if the systematic changes re-
duce the signal correlation of the artifact waveform over sensors (Mäkelä et al., 2017).
While the presence of residual tACS artifacts hinders the proper analysis of stimu-
lation effects on spontaneous brain oscillations, some authors argue that the methods
attenuate the stimulation artifact sufficiently to analyze tACS effects on event-related
oscillations (Kasten et al., 2018a; Neuling et al., 2017; Noury & Siegel, 2018), because the
residual artifact cancels out if two intervals (e.g., a pre-and a post-stimulus interval) that
contain a similar residual artifact are contrasted. Importantly, this approach can only
work if the absolute difference of the conditions is computed. Relative measures that
548 FLORIAN H. KASTEN and CHRISTOPH S. HERRMANN
involve a division of one condition by the other, as done for example to compute event-
related (de-) synchronization (Pfurtscheller & Lopes Da Silva, 1999), are vulnerable to
systematic bias by the residual tACS artifact (Kasten et al., 2018a; Kasten & Herrmann,
2019). Further, the subtraction requires that the artifact size is uncorrelated with the
task. Such systematic modulations of the artifact size may arise if a task elicits head
movements or changes in body impedance (e.g., emotional stimuli) and should be ruled
out by appropriate control analyses (Kasten et al., 2018a; Kasten & Herrmann, 2019).
In order to improve the performance of artifact cleaning approaches or to directly
avoid artifact contamination within the frequency range of interest, the use of alterna-
tive stimulation waveforms such as sawtooth waves (Dowsett & Herrmann, 2016) or
stimulation with an amplitude modulated, high frequency waveform (Witkowski et al.,
2016) has been proposed (Figure 22.2 shows examples of both waveforms). Sawtooth
waves contain strong harmonic peaks in the frequency spectrum that allow for easier
detection and rejection of data segments contaminated by residual artifacts (Dowsett
& Herrmann, 2016). Amplitude modulated waveforms consist of a high-frequency car-
rier signal, which is modulated in amplitude by the lower-frequency stimulation wave-
form. In principle, such a signal only contains power at the carrier frequency and two
sidebands, but not at the frequency of the modulating waveform itself, thus shifting
the stimulation artifact into higher frequencies and avoiding the spectral overlap be-
tween the brain signal of interest and the artifact (Witkowski et al., 2016). Recently, a
computer simulation demonstrated that a cortical network, oscillating in the alpha fre-
quency range, can be entrained to the modulation frequency of such a signal, although
the effect was substantially weaker as compared to stimulation with a pure sine wave
(Negahbani et al., 2018). Further, there are challenges to the assumption that M/EEG
recordings of AM-tACS are completely artifact free in the range of the modulation fre-
quency as nonlinear behavior of the involved hardware has been shown to reintroduce
such artifacts (Kasten et al., 2018a; Minami & Amano, 2017).
In general, the application of rTMS and tACS is considered safe if applied within
normal dosage ranges (intensities and durations) and by trained personnel. Expert
congresses have yielded publications that detail guidelines for TMS and tES safety
(Antal et al., 2017;Rossi et al. 2020). This section provides a brief introduction to safety
aspects associated with rTMS and tACS. For details, the reader is referred to the afore-
mentioned publications.
The most severe adverse effect associated with rTMS is the potential to induce
seizures. Such events are reported extremely rarely and in most cases stimulation
parameters were chosen outside the recommended dosage ranges. However, in some
patient groups or under specific medication the threshold for seizures may be lowered
(Rossi et al., 2020). Dosage limits for rTMS depend on the number of applied pulses,
BRAIN STIMULATION APPROACHES TO INVESTIGATE EEG OSCILLATIONS 549
stimulation intensity, frequency, and inter-train intervals. Contraindication for the use
of TMS is the presence of hardware containing metallic parts in the proximity of the
TMS coil (e.g., cochlear implants, pulse generators, or medical pumps). The strong mag-
netic force may cause malfunction of the devices. When combined with EEG, rTMS
can cause heating of the electrodes and an increase in the risk of skin burns. The use of
smaller electrodes (pallet or annulus shaped electrodes), can greatly reduce the amount
of heating (Ilmoniemi & Kičić, 2010).
Transcranial electrical stimulation is generally considered safe within the range of
0–4 mA stimulation intensity for <60 minutes over a single session (Antal et al., 2017).
For tACS, stimulation durations of up to 45 minutes at 1.5 mA have been used without
severe adverse effects (Laczó et al., 2012). The application of electric currents via scalp
electrodes bears the risk of skin burns, if electrodes are incorrectly applied (e.g., drying
of sponge electrodes or use of tap water instead of saline solution). So far, there are no
confirmed incidence of seizures reported in the context of tES (Antal et al., 2017). In
addition to adverse effects that impose serious safety issues, it should be acknowledged
that both methods can cause discomfort for participants. The application of electric
currents can lead to the sensation of tingling and itching or heating under the electrodes.
Further, parts of the current may polarize cells of the retina and induce a flickering sen-
sation (so-called phosphenes). The amount of visual and somatosensory sensation
depends on stimulation intensities, frequencies, and electrode montages (Turi et al.,
2013, 2014). Comparatively high intensities may even be painful for some participants.
TMS can induce phosphenes and contractions of scalp muscles in the proximity of the
stimulated brain area. In order to increase participant safety and comfort, standardized
questionnaires are available to screen participants for risk factors (diseases, medica-
tion, etc.) and exclusion criteria (implants) and to evaluate commonly reported adverse
effects (Antal et al., 2017; Brunoni et al., 2011; Rossi et al., 2009).
measures (e.g., Fekete et al., 2018; Stecher & Herrmann, 2018; Veniero et al., 2017). There
is a wide range of potential factors that may affect stimulation success or even the dir-
ection of effects. For example, there is accumulating evidence that brain stimulation
effects are state dependent (Feurra et al., 2013; Neuling et al., 2013; Ruhnau et al., 2016;
Silvanto et al., 2008), with some brain states being more susceptible to stimulation than
others. For example, when participants close their eyes, alpha oscillations tend to in-
crease strongly. During such a state of strong activity, brain stimulation failed to further
increase power in the alpha band (Neuling et al., 2013; Ruhnau et al., 2016). Nevertheless,
some involvement of the stimulated oscillation in the a given task or stimulation setting
seems necessary to see effects (Feurra et al., 2013). Such results indicate that the overall
context of the stimulation may play a crucial role. Subtle differences in the context,
such as the lightning conditions of the room, may modulate or mask stimulation effects
(Stecher et al., 2017).
Apart from contextual influences, differences between individuals could po-
tentially explain large proportions of effect variability. Of increasing interest are
differences in individual anatomy. Humans differ with respect to skull thickness and
folding of the cortex. Applying stimulation at standard locations (e.g., according to
EEG electrode positions) can result in substantially different electric field patterns in
the brain (Laakso et al., 2015). Recent work has shown that these e-field differences
can account for a large amount of variability of brain stimulation effects (Antonenko
et al., 2019; Kasten et al., 2019). Utilizing improved modelling software to individu-
alize stimulation montages may increase the reliability of stimulation effects (Huang
et al., 2018; Saturnino et al., 2019; Wagner et al., 2016). In the context of aftereffects,
a long list of factors is known to have the potential to modulate the induction of cor-
tical plasticity by means of non-invasive brain stimulation. Those factors include
participants’ sex, age, genetics, and use of medication or psychoactive substances
(Ridding & Ziemann, 2010). Unfortunately, most of the work has been done in the
context of tDCS and TMS protocols not targeting brain oscillations. Thus far, direct
investigations into the role of these factors for modulatory effects of rTMS and tACS
on brain oscillations is largely missing and requires more investigation. Individual
factors may not only alter the induction of plasticity dependent aftereffects, but also
influence the brain oscillation of interest. For example, nicotine has been shown to
diminish or abolish the induction of aftereffects in the context of tDCS and TMS
protocols (Grundey et al., 2012; Thirugnanasambandam et al., 2011). At the same
time, nicotine can alter the frequency and amplitudes of brain oscillations in the
alpha and theta range (Domino et al., 2009; Herning et al., 1983). Tobacco smoking
initially reduces power of alpha and theta oscillations, while tobacco withdrawal is
associated with a power increase (Herning et al., 1983). In the context of stimula-
tion protocols targeting alpha or theta oscillations, such effects could severely con-
found experimental results. Strengthening the understanding of determinants of
brain stimulation effects has the potential to reduce the variability of stimulation
outcomes, for example by identifying participant (or patient) groups not responding
to brain stimulation.
BRAIN STIMULATION APPROACHES TO INVESTIGATE EEG OSCILLATIONS 551
Dysfunctional patterns of oscillatory activity in the brain have been associated with
a wide range of psychiatric and neurological disease (Herrmann & Demiralp, 2005;
Uhlhaas & Singer, 2006, 2012). TACS is seen as an especially promising technique to
restore such dysfunctional oscillations and reduce related symptoms as devices are
small, portable, comparably cheap and relatively easy to apply after short training
(Antal et al., 2017; Bikson et al., 2018). Further, the devices allow high levels of control
over stimulation waveforms, enabling the design of stimulation protocols for very spe-
cific applications. For example, recently two studies applied tACS using the envelope
of human speech to improve speech intelligibility, offering potential to combine brain
stimulation with hearing aids (Riecke et al., 2018; Wilsch et al., 2018). There are also
recent results from a first clinical trial, applying tACS to restore reduced power in the
alpha band in order to reduce auditory hallucinations in patients with schizophrenia
(Ahn et al., 2019; Mellin et al., 2018). While the authors find some indication that tACS
may be capable of inducing long-term changes to dysfunctional oscillatory activity and
lead to symptom improvement, a lot more research is needed to demonstrate its clin-
ical value.
A disadvantage of rTMS and tACS is that both methods can only target superfi-
cial regions in the brain, but cannot reach areas in the depth of the brain without co-
stimulating the overlaying cortex. Recently, stimulation with interfering electric fields
(so-called temporal interference stimulation, TIS) has been proposed as a possible
solution for this (Grossmann et al., 2017). TIS is based on the idea that alternating
currents with slightly different high frequencies are injected to the brain via two pairs
of electrodes. While superficial areas of the brain are stimulated at high frequencies out-
side of the physiological range, stimulation gives rise to an amplitude modulation or
beat-frequency in regions where the two electric fields overlap (von Conta et al. 2021).
Such amplitude-modulated stimulation waveforms can in principle interact with neural
oscillations, however, substantially higher stimulation intensities may be required in
comparison to conventional tACS (Negahbani et al. 2018, Esmaipour et al. 2020). While
there is first evidence from computational modelling and from animal models, the
effects of TIS in humans have so far only been demonstrated in one study applying TIS
to superficial areas of the motor cortex (Ma et al. 2022). Overall, TIS has great poten-
tial to offer new applications and stimulation targets for non-invasive brain stimulation.
However, more research is needed to establish the efficacy of TIS in humans, especially
in targeting deep brain regions, and its underlying mechanisms.
22.7 Conclusions
oscillatory activity and human cognition. Further, they may offer new pathways to treat
neurological and psychiatric disease. However, more research is needed to explore the
underlying mechanisms of the stimulation and understand the processes and factors
determining stimulation success and the direction of effects. In particular, the improve-
ment of analysis methods for concurrent recordings of brain activity during stimulation
can strongly contribute to our knowledge in this area.
REFERENCES
Ahn, S., Mellin, J. M., Alagapan, S., Alexander, M. L., Gilmore, J. H., Jarskog, L. F., &
Fröhlich, F. (2019). Targeting reduced neural oscillations in patients with schizophrenia
by transcranial alternating current stimulation. NeuroImage, 186, 126–136. doi:10.1016/
j.neuroimage.2018.10.056
Alagapan, S., Schmidt, S. L., Lefebvre, J., Hadar, E., Shin, H. W., & Frӧhlich, F. (2016).
Modulation of cortical oscillations by low-frequency direct cortical stimulation is state-
dependent. PLOS Biology, 14, e1002424. doi:10.1371/journal.pbio.1002424
Alekseichuk, I., Turi, Z., Amador de Lara, G., Antal, A., & Paulus, W. (2016). Spatial working
memory in humans depends on theta and high gamma synchronization in the prefrontal
cortex. Current Biology, 26, 1513–1521. doi:10.1016/j.cub.2016.04.035
Amassian, V. E., Cracco, R. Q., Maccabee, P. J., Cracco, J. B., Rudell, A., & Eberle, L. (1989).
Suppression of visual perception by magnetic coil stimulation of human occipital
cortex. Electroencephalography and Clinical Neurophysiology, 74, 458–462. doi:10.1016/
0168-5597(89)90036-1
Antal, A., Alekseichuk, I., Bikson, M., Brockmöller, J., Brunoni, A. R., Chen, R., . . . Paulus,
W. (2017). Low intensity transcranial electric stimulation: Safety, ethical, legal regula-
tory and application guidelines. Clinical Neurophysiology, 128(9), 1774–1718. doi:10.1016/
j.clinph.2017.06.001
Antal, A., Bikson, M., Datta, A., Lafon, B., Dechent, P., Parra, L. C., & Paulus, W. (2014).
Imaging artifacts induced by electrical stimulation during conventional fMRI of the brain.
NeuroImage, 85, 1040–1047. doi:10.1016/j.neuroimage.2012.10.026
Antal, A. & Herrmann, C. S. (2016). Transcranial alternating current and random noise stimu-
lation: Possible mechanisms. Neural Plasticity, 2016, 3616807. doi:10.1155/2016/3616807
Antonenko, D., Thielscher, A., Saturnino, G. B., Aydin, S., Ittermann, B., Grittner, U.,
et al. (2019). Towards precise brain stimulation: Is electric field simulation related to
neuromodulation? Brain Stimulation, 12(5), 1159–1168. doi:10.1016/j.brs.2019.03.072
Baltus, A., Vosskuhl, J., Boetzel, C., & Herrmann, C. S. (2018a). Transcranial alternating current
stimulation modulates auditory temporal resolution in elderly people. European Journal of
Neuroscience, 51(5), 1328–1338. doi:10.1111/ejn.13940
Baltus, A., Wagner, S., Wolters, C. H., & Herrmann, C. S. (2018b). Optimized auditory
transcranial alternating current stimulation improves individual auditory temporal reso-
lution. Brain Stimulation, 11, 118–124. doi:10.1016/j.brs.2017.10.008
Barker, A. T., Jalinous, R., & Freeston, I. L. (1985). Non-invasive magnetic stimulation of human
motor cortex. The Lancet, 1(8437), 1106–1107. doi:10.1016/s0140-6736(85)92413-4
Bergmann, T. O., Karabanov, A., Hartwigsen, G., Thielscher, A., & Siebner, H. R. (2016).
Combining non- invasive transcranial brain stimulation with neuroimaging and
BRAIN STIMULATION APPROACHES TO INVESTIGATE EEG OSCILLATIONS 553
Esmaeilpour, Zeinab, Greg Kronberg, Davide Reato, Lucas C. Parra, and Marom Bikson.
“Temporal Interference Stimulation Targets Deep Brain Regions by Modulating Neural
Oscillations.” Brain Stimulation, 14, no. 1 (January 1, 2021): 55–65. https://doi.org/10.1016/
j.brs.2020.11.007.
Fekete, T., Nikolaev, A. R., Knijf, F. De, Zharikova, A., & van Leeuwen, C. (2018). Multi-
electrode alpha tACS during varying background tasks fails to modulate subsequent alpha
power. Frontiers in Neuroscience, 12, 428. doi:10.3389/fnins.2018.00428
Feurra, M., Pasqualetti, P., Bianco, G., Santarnecchi, E., Rossi, A., & Rossi, S. (2013). State-
dependent effects of transcranial oscillatory currents on the motor system: What you think
matters. Journal of Neuroscience, 33, 17483–9. doi:10.1523/JNEUROSCI.1414-13.2013
Fröhlich, F. & McCormick, D. A. (2010). Endogenous electric fields may guide neocortical net-
work activity. Neuron, 67, 129–143. doi:10.1016/j.neuron.2010.06.005
Goldman, R. I., Stern, J. M., Engel, J., & Cohen, M. S. (2002). Simultaneous EEG and fMRI of
the alpha rhythm. Neuroreport, 13, 2487–2492. doi:10.1097/00001756-200212200-00022
Gross, J., Kujala, J., Hamalainen, M., Timmermann, L., Schnitzler, A., & Salmelin, R. (2001).
Dynamic imaging of coherent sources: Studying neural interactions in the human brain.
Proceedings of the National Academy of Sciences of the United States of America, 98, 694–699.
doi:10.1073/pnas.98.2.694
Grossman, Nir, David Bono, Nina Dedic, Suhasa B. Kodandaramaiah, Andrii Rudenko, Ho-
Jun Suk, Antonino M. Cassara, et al. “Noninvasive Deep Brain Stimulation via Temporally
Interfering Electric Fields.” Cell, 169, 6 (2017): 1029– 1041.e16. https://
doi.org/
10.1016/
j.cell.2017.05.024.
Grundey, J., Thirugnanasambandam, N., Kaminsky, K., Drees, A., Skwirba, A. C., Lang,
N., . . . Nitsche, M. A. (2012). Rapid effect of nicotine intake on neuroplasticity in non-
smoking humans. Frontiers in Pharmacology, 3, 1–9. doi:10.3389/fphar.2012.00186
Gundlach, C., Müller, M. M., Nierhaus, T., Villringer, A., & Sehm, B. (2016). Phasic modulation
of human somatosensory perception by transcranially applied oscillating currents. Brain
Stimulation, 9, 712–7 19. doi:10.1016/j.brs.2016.04.014
Hallett, M. (2000). Transcranial magnetic stimulation and the human brain. Nature, 406, 147–
150. doi:10.1038/35018000
Hamidi, M., Slagter, H. A., Tononi, G., & Postle, B. R. (2009). Repetitive transcranial mag-
netic stimulation affects behavior by biasing endogenous cortical oscillations. Frontiers in
Integrative Neuroscience, 3, 14. doi:10.3389/neuro.07.014.2009
Hanslmayr, S., Matuschek, J., & Fellner, M. C. (2014). Entrainment of prefrontal beta
oscillations induces an endogenous echo and impairs memory formation. Current Biology,
24, 904–909. doi:10.1016/j.cub.2014.03.007
Helfrich, R. F., Knepper, H., Nolte, G., Strüber, D., Rach, S., Herrmann, C. S., et al. (2014a).
Selective modulation of interhemispheric functional connectivity by HD-tACS shapes per-
ception. PLoS Biology, 12(12), e1002031. doi:10.1371/journal.pbio.1002031
Helfrich, R. F., Schneider, T. R., Rach, S., Trautmann-Lengsfeld, S. A., Engel, A. K., &
Herrmann, C. S. (2014b). Entrainment of brain oscillations by transcranial alternating
current stimulation. Current Biology, 24, 333–339. doi:10.1016/j.cub.2013.12.041
Herning, R. I., Jones, R. T., & Bachman, J. (1983). EEG changes during tobacco withdrawal.
Psychophysiology, 20, 507–512. doi:10.1111/j.1469-8986.1983.tb03004.x
Herrmann, C. S. (2001). Human EEG responses to 1–100 Hz flicker: Resonance phenomena in
visual cortex and their potential correlation to cognitive phenomena. Experimental Brain
Research, 137, 346–53. doi:10.1007/s002210100682
BRAIN STIMULATION APPROACHES TO INVESTIGATE EEG OSCILLATIONS 555
Endogenous but Not Exogenous Visual-Spatial Attention.” Scientific Reports, 10, no. 1
(December 2020): 12270. https://doi.org/10.1038/s41598-020-68992-2.
Kasten, F. H., Maess, B., & Herrmann, C. S. (2018a). Facilitated event- related power
modulations during transcranial alternating current stimulation (tACS) revealed by con-
current tACS-MEG. eNeuro [online] 5(3). doi:10.1523/ENEURO.0069-18.2018
Kasten, F. H., Negahbani, E., Fröhlich, F., & Herrmann, C. S. (2018b). Non-linear transfer
characteristics of stimulation and recording hardware account for spurious low-frequency
artifacts during amplitude modulated transcranial alternating current stimulation (AM-
tACS). NeuroImage, 179, 134–143. doi:10.1016/j.neuroimage.2018.05.068
Klimesch, W., Sauseng, P., & Gerloff, C. (2003). Enhancing cognitive performance with re-
petitive transcranial magnetic stimulation at human individual alpha frequency. European
Journal of Neuroscience, 17, 1129–1133. doi:10.1046/j.1460-9568.2003.02517.x
Klink, Katharina, Sven Paßmann, Florian H. Kasten, and Jessica Peter. “The Modulation of
Cognitive Performance with Transcranial Alternating Current Stimulation: A Systematic
Review of Frequency-Specific Effects.” Brain Sciences, 10, no. 12 (December 2020): 932.
https://doi.org/10.3390/brainsci10120932.
Klomjai, W., Katz, R., & Lackmy-Vallée, A. (2015). Basic principles of transcranial magnetic
stimulation (TMS) and repetitive TMS (rTMS). Annals of Physical and Rehabilitation
Medicine, 58, 208–213. doi:10.1016/j.rehab.2015.05.005
Kohli, S. & Casson, A. J. (2015). Removal of transcranial a.c. current stimulation artifact from
simultaneous EEG recordings by superposition of moving averages. 2015 37th Annual
International Conference of the IEEE Engineering in Medicine and Biology Society (EMBC),
2015, 3436–3439. doi:10.1109/EMBC.2015.7319131
Kopp, C., Rudolph, U., Low, K., & Tobler, I. (2004). Modulation of rhythmic brain activity by di-
azepam: GABAAreceptor subtype and state specificity. Proceedings of the National Academy
of Sciences of the United States of America, 101, 3674–3679. doi:10.1073/pnas.0306975101
Krause, M. R., Vieira, P. G., Csorba, B. A., Pilly, P. K., & Pack, C. C. (2019). Transcranial
alternating current stimulation entrains single- neuron activity in the primate brain.
Proceedings of the National Academy of Sciences of the United States of America, 116, 5747–
5755. doi:10.1073/pnas.1815958116
Laakso, I., Tanaka, S., Koyama, S., De Santis, V., & Hirata, A. (2015). Inter-subject vari-
ability in electric fields of motor cortical tDCS. Brain Stimulation, 8, 906–913. doi:10.1016/
j.brs.2015.05.002
Laczó, B., Antal, A., Niebergall, R., Treue, S., & Paulus, W. (2012). Transcranial alternating
stimulation in a high gamma frequency range applied over V1 improves contrast percep-
tion but does not modulate spatial attention. Brain Stimulation, 5, 484–491. doi:10.1016/
j.brs.2011.08.008
Laufs, H., Kleinschmidt, A., Beyerle, A., Eger, E., Salek-Haddadi, A., Preibisch, C., & Krakow,
K. (2003). EEG-correlated fMRI of human alpha activity. NeuroImage, 19, 1463–1476.
doi:10.1016/S1053-8119(03)00286-6
Lisman, J. E. & Idiart, M. A. (1995). Storage of 7+/-2 short-term memories in oscillatory
subcycles. Science, 267(5203), 1512–1515. doi:10.1126/science.7878473
Lisman, J. E. & Jensen, O. (2013). The theta-gamma neural code. Neuron, 77, 1002–1016.
doi:10.1016/j.neuron.2013.03.007
Loffler, B. S., Stecher, H. I., Fudickar, S., de Sordi, D., Otto-Sobotka, F., Hein, A., & Herrmann,
C. S. (2018). Counteracting the slowdown of reaction times in a vigilance experiment with
BRAIN STIMULATION APPROACHES TO INVESTIGATE EEG OSCILLATIONS 557
Neuling, T., Wagner, S., Wolters, C. H., Zaehle, T., & Herrmann, C. S. (2012b). Finite-element
model predicts current density distribution for clinical applications of tDCS and tACS.
Frontiers in Psychiatry, 3, 1–10. doi:10.3389/fpsyt.2012.00083
Nitsche, M. A, Fricke, K., Henschke, U., Schlitterlau, A., Liebetanz, D., Lang, N., et al.
(2003). Pharmacological modulation of cortical excitability shifts induced by transcranial
direct current stimulation in humans. Journal of Physiology, 553, 293–301. doi:10.1113/
jphysiol.2003.049916
Notbohm, A., Kurths, J., & Herrmann, C. S. (2016). Modification of brain oscillations via
rhythmic light stimulation provides evidence for entrainment but not for superpos-
ition of event-related responses. Frontiers in Human Neuroscience, 10, 10. doi:10.3389/
fnhum.2016.00010
Noury, N., Hipp, J. F., & Siegel, M. (2016). Physiological processes non-linearly affect electro-
physiological recordings during transcranial electric stimulation. NeuroImage, 140, 99–109.
doi:10.1016/j.neuroimage.2016.03.065
Noury, N. & Siegel, M. (2017). Phase properties of transcranial electrical stimulation
artifacts in electrophysiological recordings. NeuroImage, 158, 406– 416. doi:10.1016/
j.neuroimage.2017.07.010
Noury, N. & Siegel, M. (2018). Analyzing EEG and MEG signals recorded during tES, a reply.
NeuroImage, 167, 53–61. doi:10.1016/j.neuroimage.2017.11.023
Novembre, G., Knoblich, G., Dunne, L., & Keller, P. E. (2017). Interpersonal synchrony
enhanced through 20 Hz phase-coupled dual brain stimulation. Social Cognitive and
Affective Neuroscience, 12(4), 662–670. doi:10.1093/scan/nsw172.
Opitz, A., Falchier, A., Yan, C. G., Yeagle, E. M., Linn, G. S., Megevand, P., et al. (2016).
Spatiotemporal structure of intracranial electric fields induced by transcranial electric
stimulation in humans and nonhuman primates. Scientific Reports, 6, 1–11. doi:10.1038/
srep31236
Ozen, S., Sirota, A., Belluscio, M. A., Anastassiou, C. A., Stark, E., Koch, C., et al. (2010).
Transcranial electric stimulation entrains cortical neuronal populations in rats. Journal of
Neuroscience, 30, 11476–11485. doi:10.1523/JNEUROSCI.5252-09.2010
Palva, J. M., Palva, S., & Kaila, K. (2005). Phase synchrony among neuronal
oscillations in the human cortex. Journal of Neuroscience, 25, 3962–72. doi:10.1523/
JNEUROSCI.4250-04.2005
Pfurtscheller, G., & Lopes Da Silva, F. H. (1999). Event-related EEG/MEG synchronization and
desynchronization: Basic principles. Clinical Neurophysiology, 110, 1842–1857. doi:10.1016/
S1388-2457(99)00141-8
Pikovsky, A., Rosenblum, M., & Kurths, J. (2003). Synchronization: A Universal concept in
nonlinear sciences. Cambridge University Press. doi:10.1063/1.1554136
Polanía, R., Nitsche, M. A., Korman, C., Batsikadze, G., & Paulus, W. (2012). The importance
of timing in segregated theta phase-coupling for cognitive performance. Current Biology, 22,
1314–1318. doi:10.1016/j.cub.2012.05.021
Reato, D., Rahman, A., Bikson, M., & Parra, L. C. (2010). Low-intensity electrical stimula-
tion affects network dynamics by modulating population rate and spike timing. Journal of
Neuroscience, 30, 15067–15079. doi:10.1523/JNEUROSCI.2059-10.2010.
Reato, D., Rahman, A., Bikson, M., & Parra, L. C. (2013). Effects of weak transcranial alternating
current stimulation on brain activity-a review of known mechanisms from animal studies.
Frontiers in Human Neuroscience, 7, 687. doi:10.3389/fnhum.2013.00687
BRAIN STIMULATION APPROACHES TO INVESTIGATE EEG OSCILLATIONS 559
Shams, L., Kamitani, Y., & Shimojo, S. (2002). Visual illusion induced by sound. Cognitive
Brain Research, 14, 147–152. doi:10.1016/S0926-6410(02)00069-1
Siegel, M., Donner, T. H., & Engel, A. K. (2012). Spectral fingerprints of large-scale neuronal
interactions. Nature Reviews Neuroscience, 13, 121–134. doi:10.1038/nrn3137
Silvanto, J., Muggleton, N., & Walsh, V. (2008). State-dependency in brain stimulation
studies of perception and cognition. Trends in Cognitive Sciences, 12, 447–454. doi:10.1016/
j.tics.2008.09.004
Soekadar, S. R., Witkowski, M., Cossio, E. G., Birbaumer, N., Robinson, S. E., & Cohen, L. G.
(2013). In vivo assessment of human brain oscillations during application of transcranial
electric currents. Nature Communications, 4, 2032. doi:10.1038/ncomms3032
Sohal, V. S. (2012). Insights into cortical oscillations arising from optogenetic studies. Biol.
Psychiatry 71, 1039–1045. doi:10.1016/j.biopsych.2012.01.024
Stecher, H. I. & Herrmann, C. S. (2018). Absence of alpha-tACS aftereffects in darkness reveals
importance of taking derivations of stimulation frequency and individual alpha variability
into account. Frontiers in Psychology, 9, 1–9. doi:10.3389/fpsyg.2018.00984
Stecher, H. I., Pollok, T. M., Strüber, D., Sobotka, F., & Christoph, S. (2017). Ten minutes of α-
tACS and ambient illumination independently modulate EEG α-power. Frontiers in Human
Neuroscience, 11. doi:10.3389/fnhum.2017.00257
Stonkus, R., Braun, V., Kerlin, J. R., Volberg, G., & Hanslmayr, S. (2016). Probing the causal
role of prestimulus interregional synchrony for perceptual integration via tACS. Scientific
Reports, 6, 32065. doi:10.1038/srep32065
Strüber, D., Rach, S., Neuling, T., Herrmann, C. S., & Kar, K. (2015). On the possible role of
stimulation duration for after- effects of transcranial alternating current stimulation.
Frontiers in Cellular Neuroscience, 9, 311. doi:10.3389/fnsys.2015.00148
Strüber, D., Rach, S., Trautmann-Lengsfeld, S. A., Engel, A. K., & Herrmann, C. S. (2014).
Antiphasic 40 Hz oscillatory current stimulation affects bistable motion perception. Brain
Topography, 27, 158–171. doi:10.1007/s10548-013-0294-x
Szymanski, C., Müller, V., Brick, T. R., von Oertzen, T., & Lindenberger, U. (2017). Hyper-
transcranial alternating current stimulation: Experimental manipulation of inter-brain syn-
chrony. Frontiers in Human Neuroscience, 11, 1–15. doi:10.3389/fnhum.2017.00539
Thielscher, A., Antunes, A., & Saturnino, G. B. (2015). Field modeling for transcranial magnetic
stimulation: A useful tool to understand the physiological effects of TMS? 2015 37th Annual
International Conference of the IEEE Engineering in Medicine and Biology Society (EMBC),
2015, 222–225. doi:10.1109/EMBC.2015.7318340
Thirugnanasambandam, N., Grundey, J., Adam, K., Drees, A., Skwirba, A. C.,
Lang, N., . .
. Nitsche, M. A. (2011). Nicotinergic impact on focal and non- focal
neuroplasticity induced by non- invasive brain stimulation in non- smoking humans.
Neuropsychopharmacology, 36, 879–86. doi:10.1038/npp.2010.227
Thut, G., Schyns, P. G., & Gross, J. (2011a). Entrainment of perceptually relevant brain
oscillations by non- invasive rhythmic stimulation of the human brain. Frontiers in
Psychology, 2, 1–10. doi:10.3389/fpsyg.2011.00170
Thut, G., Veniero, D., Romei, V., Miniussi, C., Schyns, P., & Gross, J. (2011b). Rhythmic TMS
causes local entrainment of natural oscillatory signatures. Current Biology, 21, 1176–1185.
doi:10.1016/j.cub.2011.05.049
Turi, Z., Ambrus, G. G., Ho, K.-A., Sengupta, T., Paulus, W., & Antal, A. (2014). When size
matters: Large electrodes induce greater stimulation-related cutaneous discomfort than
smaller electrodes at equivalent current density. Brain Stimulation, 7, 460–467. doi:10.1016/
j.brs.2014.01.059
BRAIN STIMULATION APPROACHES TO INVESTIGATE EEG OSCILLATIONS 561
Turi, Z., Ambrus, G. G., Janacsek, K., Emmert, K., Hahn, L., Paulus, W., et al. (2013). Both
the cutaneous sensation and phosphene perception are modulated in a frequency-specific
manner during transcranial alternating current stimulation. Restorative Neurology and
Neuroscience, 31, 275–85. doi:10.3233/RNN-120297
Uhlhaas, P. J. & Singer, W. (2006). Neural synchrony in brain disorders: Relevance for cognitive
dysfunctions and pathophysiology. Neuron, 52, 155–168. doi:10.1016/j.neuron.2006.09.020
Uhlhaas, P. J. & Singer, W. (2012). Neuronal dynamics and neuropsychiatric disorders: Toward
a translational paradigm for dysfunctional large-scale networks. Neuron, 75, 963–980.
doi:10.1016/j.neuron.2012.09.004
Van Veen, B. D. & Buckley, K. M. (1988). Beamforming: A versatile approach to spatial filtering.
IEEE ASSP Magazine, 5, 4–24. doi:10.1109/53.665
Van Veen, B. D., Van Drongelen, W., Yuchtman, M., & Suzuki, A. (1997). Localization of
brain electrical activity via linearly constrained minimum variance spatial filtering. IEEE
Transactions in Biomedical Engineering, 44, 867–880. doi:10.1109/10.623056
Veniero, D., Benwell, C. S. Y., Ahrens, M. M., & Thut, G. (2017). Inconsistent effects of parietal
α-tACS on pseudoneglect across two experiments: A failed internal replication. Frontiers in
Psychology, 8, 1–14. doi:10.3389/fpsyg.2017.00952
Veniero, D., Bortoletto, M., & Miniussi, C. (2009). TMS-EEG co-registration: On TMS-
induced artifact. Clinical Neurophysiology, 120, 1392–1399. doi:10.1016/j.clinph.2009.04.023
Veniero, D., Vossen, A., Gross, J., & Thut, G. (2015). Lasting EEG/MEG aftereffects of rhythmic
transcranial brain stimulation: Level of control over oscillatory network activity. Frontiers in
Cellular Neuroscience, 9, 477. doi:10.3389/fncel.2015.00477
Violante, I. R., Li, L. M., Carmichael, D. W., Lorenz, R., Leech, R., Hampshire, A., Rothwell,
J. C., & Sharp, D. J. (2017). Externally induced frontoparietal synchronization modulates
network dynamics and enhances working memory performance. Elife, 6. doi:10.7554/
eLife.22001
Conta, Jill von, Florian H. Kasten, Branislava Ćurčić-Blake, André Aleman, Axel Thielscher,
and Christoph S. Herrmann. “Interindividual Variability of Electric Fields during
Transcranial Temporal Interference Stimulation (TTIS).” Scientific Reports, 11, no. 1 (October
13, 2021): 20357. https://doi.org/10.1038/s41598-021-99749-0.
Vöröslakos, M., Takeuchi, Y., Brinyiczki, K., Zombori, T., Oliva, A., Fernández- Ruiz,
A., . . . Berényi, A. (2018). Direct effects of transcranial electric stimulation on brain circuits
in rats and humans. Nature Communications, 9, 483. doi:10.1038/s41467-018-02928-3
Voss, U., Holzmann, R., Hobson, A., Paulus, W., Koppehele-Gossel, J., Klimke, A., & Nitsche,
M. A. (2014). Induction of self-awareness in dreams through frontal low current stimulation
of gamma activity. Nature Neuroscience, 17, 810–812. doi:10.1038/nn.3719
Vossen, A., Gross, J., & Thut, G. (2015). Alpha power increase after transcranial alternating
current stimulation at alpha frequency (α-tACS) reflects plastic changes rather than entrain-
ment. Brain Stimulation, 8, 499–508. doi:10.1016/j.brs.2014.12.004
Vosskuhl, J., Huster, R. J., & Herrmann, C. S. (2015). Increase in short-term memory capacity
induced by down-regulating individual theta frequency via transcranial alternating current
stimulation. Frontiers in Human Neuroscience, 9, 257. doi:10.3389/fnhum.2015.00257
Vosskuhl, J., Huster, R. J., & Herrmann, C. S. (2016). BOLD signal effects of transcranial
alternating current stimulation (tACS) in the alpha range: A concurrent tACS–fMRI study.
NeuroImage, 140, 118–125. doi:10.1016/j.neuroimage.2015.10.003
Wagner, S., Burger, M., & Wolters, C. H. (2016). An optimization approach for well-targeted
transcranial direct current stimulation. SIAM Journal of Applied Mathematics, 76, 2154–2174.
doi:10.1137/15M1026481
562 FLORIAN H. KASTEN and CHRISTOPH S. HERRMANN
Wagner, T., Rushmore, J., Eden, U., & Valero-Cabre, A. (2009). Biophysical foundations
underlying TMS: Setting the stage for an effective use of neurostimulation in the cognitive
neurosciences. Cortex, 45, 1025–1034. doi:10.1016/j.cortex.2008.10.002
Weinrich, C. A., Brittain, J. S., Nowak, M., Salimi-Khorshidi, R., Brown, P., & Stagg, C. J. (2017).
Modulation of long- range connectivity patterns via frequency- specific stimulation of
human cortex. Current Biology, 27, 3061-3068.e3. doi:10.1016/j.cub.2017.08.075
Wilsch, A., Neuling, T., Obleser, J., & Herrmann, C. S. (2018). Transcranial alternating current
stimulation with speech envelopes modulates speech comprehension. NeuroImage, 172,
766–774. doi:10.1016/j.neuroimage.2018.01.038
Wischnewski, M., Engelhardt, M., Salehinejad, M. A., Schutter, D. J. L. G., Kuo, M.-F., &
Nitsche, M. A. (2018). NMDA receptor-mediated motor cortex plasticity after 20 Hz
transcranial alternating current stimulation. Cerebral Cortex, 29(7), 2924–2931. doi:10.1093/
cercor/bhy160
Wischnewski, M. & Schutter, D. J. L. G. (2017). After-effects of transcranial alternating current
stimulation on evoked delta and theta power. Clinical Neurophysiology, 128, 2227–2232.
doi:10.1016/j.clinph.2017.08.029
Witkowski, M., Garcia-Cossio, E., Chander, B. S., Braun, C., Birbaumer, N., Robinson, S. E., &
Soekadar, S. R. (2016). Mapping entrained brain oscillations during transcranial alternating
current stimulation (tACS). NeuroImage, 140, 89–98. doi:10.1016/j.neuroimage.2015.10.024
Woods, A. J., Antal, A., Bikson, M., Boggio, P. S., Brunoni, A. R., Celnik, P., . . . Nitsche, M. A..
(2016). A technical guide to tDCS, and related non-invasive brain stimulation tools. Clinical
Neurophysiology. 127, 1031–1048. doi:10.1016/j.clinph.2015.11.012
Wöstmann, M., Vosskuhl, J., Obleser, J., & Herrmann, C. S. (2018). Opposite effects of
lateralised transcranial alpha versus gamma stimulation on auditory spatial attention. Brain
Stimulation, 11, 752–758. doi:10.1016/j.brs.2018.04.006
Zaehle, T., Rach, S., & Herrmann, C. S. (2010). Transcranial alternating current stimulation
enhances individual alpha activity in human EEG. PLoS One 5, 13766. doi:10.1371/journal.
pone.0013766
CHAPTER 23
PAR AM ETERIZI NG NE U RA L
F IELD P OTENTIA L DATA
BRADLEY VOYTEK
23.1 Introduction
There is a growing literature focusing on the fact that oscillations are nonsinusoidal
(Sherman et al., 2016; Cole & Voytek, 2017, 2018, 2019; Cole et al., 2017; Jackson et al.,
2019); specific time-domain characteristics of their nonsinusoidality potentially capture
dynamics of the underlying neural circuit (Sherman et al., 2016; Cole & Voytek, 2018).
Most analyses of oscillations are conducted on canonically defined frequency bands.
This is often done without consideration of the aperiodic component, which has a 1/f-
like characteristic in the power spectrum that likely arises from the double-exponential
time domain shape of postsynaptic currents (Freeman & Zhai, 2009; Gao et al., 2017).
Because power at any given frequency is a mixture of both periodic and aperiodic ac-
tivity, measuring band-limited power with the presumption that the resulting numer-
ical power value captures only oscillatory power is problematic (Haller et al., 2020).
First and foremost, this chapter highlights problems regarding the traditional inter-
pretation of field potential data, and then presents recent approaches for minimizing the
damage that our presumptions cause (as related to physiological interpretations of field
potential analyses). For example, because field potentials include a mixture of features—
aperiodic, periodic, and transients—one cannot be certain that band-limited power
values extracted from traditional analyses includes an oscillation at all. Therefore, if
we wish to talk about oscillations and their various functional correlates, we must first
verify—to the best of our ability—that an oscillation is present above and beyond the
aperiodic signal.
As an additional caveat, we must also ensure that spectral power above the aperi-
odic signal reflects an oscillation at that frequency, rather than a harmonic of a slower,
nonsinusoidal rhythm (Cole & Voytek, 2017). Finally, because the aperiodic signal of
neural power spectra has received less attention than oscillations, we discuss the origins
of this signal, its possible physiological interpretations, and emerging research regarding
its cognitive and behavioral relevance. I argue that these three features—oscillations,
nonsinusoidal waveforms, and the aperiodic signal—simultaneously exist in the same
field potential data, but likely have different physiological interpretations. Without
careful parameterization of all of these features simultaneously, it is easy to misinterpret
their physiological relevance. The goal of this chapter is to help researchers mitigate the
propagation of potential physiological (mis)interpretations of their oscillation results
by encouraging explicit parameterization of field potential data.
Oscillations are one of the most widely used analytic approaches in human neuro-
science. Part of the excitement about neural oscillations is their ubiquity: they are
conserved across species (Bullock, 1981), correlate with numerous cognitive and be-
havioral processes (Engel et al., 2001; Varela et al., 2001), are disrupted in a variety of
neurological and psychiatric disorders (Voytek & Knight, 2015), and have been causally
linked to neural information routing and the shaping of spiking networks (Varela et al.,
PARAMETERIZING NEURAL FIELD POTENTIAL DATA 565
2001, 2001; Fries, 2005). Mechanistically, oscillatory networks are said to coordinate
distributed neural ensembles via synchronizing spike-timing via phase coordination
of spiking (Fries, 2005); that is, oscillations have been argued to be the mechanism by
which brain regions form functional, dynamic communication networks (Fries, 2005;
Voytek et al., 2015a; Helfrich & Knight, 2016).
The vast majority of studies examining neural oscillations assume classic, canonical
bands of interest. These are approximately defined as: delta (1–4 Hz), theta (4–8 Hz),
alpha (8–12 Hz), beta (16–30 Hz), and low gamma (30–60 Hz). However, these fre-
quency bands are only loosely related to the underlying physiology; there exists a great
deal of variability across species (Bullock, 1981), age (Obrist, 1954), and cognitive/be-
havioral state (Klimesch, 1999; Haegens et al., 2014; Samaha et al., 2015). One concern
with such a priori, band-limited analyses is that they are often performed without first
examining the power spectrum to account for this intra-and interindividual variability
in, for example, center frequency. This means that predetermined frequency bands may
include nonoscillatory activity from outside the true physiological oscillatory band—
whose center frequency and bandwidth may not fall exactly within a canonical band—
thus masking crucial behaviorally and physiologically relevant information.
While there are several methods for identifying individual differences in oscillations,
most are restricted to identifying the frequency at which the power spectrum peaks
within a specific sub-band (Haegens et al., 2014); additionally, these methods are limited
to finding only one oscillation within a specified band while ignoring other potentially
physiologically relevant oscillations and their inter-relationships (Donoghue et al.,
2020). Importantly, all of these methods measure total band power, not band power rela-
tive to the aperiodic signal, thus conflating the two processes (Donoghue et al., 2020).
To further complicate matters, oscillations are often not tonic and sustained, but ra-
ther appear in brief, intermittent bursts (Feingold et al., 2015; Jones, 2016; Peterson &
Voytek, 2017). The functional significance of bursting is supported by a rapidly growing
literature, primarily based on cortical LFP recordings from rodents and nonhuman
primates (Feingold et al., 2015; Lundqvist et al., 2016; Sherman et al., 2016). Example
works have shown that oscillatory burst probability increases during a variety of cog-
nitive and behavioral tasks such as working memory (Lundqvist et al., 2016) or motor
control and action completion (Feingold et al., 2015; Sherman et al., 2016). Historically,
the separation between the functional role of bursting and sustained oscillations can
be highlighted by research into the visual alpha rhythm. The earliest EEG studies of
alpha treated it as an idling rhythm (Jasper & Penfield, 1949; Klimesch, 2012), one that is
sustained while at rest and becomes intermittent when awake and attentive. The idling
perspective was complicated by studies revealing that alpha power in posterior (visual)
regions increases with working memory load across many paradigms (Jensen, 2002).
The fact that alpha power increases during active working memory conflicts with higher
alpha power during eyes closed restful states. Thus, the idling hypothesis was updated
to an account where alpha acts to rhythmically inhibit neural gain (Jensen & Mazaheri,
2010). In this updated view, regions not receiving sensory information, or otherwise
engaging in active behavioral responses, were said to be suppressed to avoid introducing
566 BRADLEY VOYTEK
noise into the larger network. When stimulated, task-relevant regions are disinhibited,
restoring cortical processing. (For an in-depth perspective on the alpha rhythm, see
Chapter 10.)
Recent computational work from our lab has challenged this classic view of alpha as
a suppression rhythm (Peterson & Voytek, 2017). When alpha oscillations are modeled
specifically as balanced gain modulation, alpha can suppress visual detection. When
suppression acts over long periods (e.g., >1 sec) this strongly reduces the overall gain,
per the standard view. That is, this low-gain state reflects an “idle” state of activity,
serving to inhibit firing output for any given input. Surprisingly, however, the model
shows that when alpha is in a short, bursting mode (<0.5 sec, or 5 cycles), alpha actu-
ally enhances neuronal firing. That is, modeling and physiological work uncovered an
enticing result: that oscillatory bursts may play a different functional role compared to
sustained oscillations of the same frequency. This result—that oscillatory bursts may
have a different physiological effect than sustained rhythms—highlights the import-
ance of parameterizing not just the power of an oscillation, but its short-time temporal
dynamics.
Neural oscillations are not smoothly varying sinusoids. Rather, they manifest many
different nonsinusoidal characteristics. Across brain regions, species, and frequencies,
there do exist a variety of stereotyped nonsinusoidal shapes such as human motor
cortical beta oscillations (Cole et al., 2017; Jackson et al., 2019), sleep slow oscillations
(Steriade et al., 1993; Amzica & Steriade, 1998), the sensorimotor mu rhythm (Kuhlman,
1978; Arroyo et al., 1993), and especially the rodent hippocampal theta rhythm (Belluscio
et al., 2012; Cole & Voytek, 2018). The nonsinusoidal shape of oscillation waveforms is
an exciting, re-emerging candidate for potentially indexing circuit-level physiology,
such as synaptic input synchrony (Sherman et al., 2016). It is becoming increasingly
apparent that nonsinusoidal oscillatory waveform shape carries physiological informa-
tion (Sherman et al., 2016; Cole & Voytek, 2017, 2018). Because there does not seem to
be a theoretical reason why brain oscillations should be sinusoidal, it may be that the
diverse set of neuronal activation and synaptic current dynamics present in different
oscillations determine the waveform shape in the nearby field potential.
Several nonsinusoidal features of an oscillation can be parameterized, such as a
waveform’s symmetry (the ratio between the rise period and the decay period), the
sharpness of each oscillation peak, the rise and decay times for each cycle’s voltage, and so
on. To quantify these features, an oscillation can explicitly be broken up into individual
cycles by identifying the locations of peaks and troughs and then computing features
on the raw voltage time series (Figure 23.1) (Cole & Voytek, 2019). We recently reviewed
studies that have related the waveform shape of brain oscillations to neurophysiology
(Cole & Voytek, 2017). One of the most well-supported of these relationships shows that
PARAMETERIZING NEURAL FIELD POTENTIAL DATA 567
–200 51 19 29
26 26
17 22 20 33 23
–400 18 21 24
19 21 26
–600 20
–800 15
–1000
3.0 3.2 3.4 3.6 3.8 4.0
Time (S)
amplitude period time_rise time_trough time_decay time_peak volt_rise volt_trough volt_decay volt_peak
0 27.442521 46 26 18 20 22 77.542938 –18.474257 127.051091 108.576834
1 34.980215 48 37 19 11 21 123.623268 –84.545266 143.613947 59.068681
2 19.290058 40 31 24 9 29 161.573095 –87.419348 126.497350 39.078002
3 18.898692 53 7 31 46 30 113.252293 15.490505 58.663243 74.153747
4 20.070945 29 16 10 13 13 149.175399 31.155748 97.587049 128.742797
(b)
Sharp-peak oscillation
Voltage
Power
0.0 0.5 1.0 0 10 20 30 40 50
Sinusoid-like oscillation
Voltage
Power
Until recently, the aperiodic signal has been largely referred to as electrophysiological
“noise” or the “background.” There is mounting evidence, however, that this signal
carries physiological information (Gao et al., 2017) and is dynamically altered by cog-
nitive and perceptual states (Waschke et al., 2017; Dave et al., 2018), as well as in aging
(Voytek et al., 2015b; Waschke et al., 2017; Dave et al., 2018) and disease (Peterson et al.,
2018; Robertson et al., 2019; Veerakumar et al., 2019).
This aperiodic signal has previously been referred to as the 1/f background and/or
neural noise, because power at any given frequency is inversely related to the frequency
PARAMETERIZING NEURAL FIELD POTENTIAL DATA 569
itself (1/f ) and looks like colored (pink or brown) “noise” seen in digital signal pro-
cessing. However, this is likely not just “noise”, because, as noted, field potentials are
dominated by postsynaptic currents across relatively large populations (Buzsáki et al.,
2012; Pesaran et al., 2018). These currents are driven by input from neurons that are
integrated across a large number of neurons and thus can appear noise-like. However,
we have shown that the 1/f-like nature of the signal can arise simply as a consequence
of the nonlinear nature of post-synaptic currents, and that changes in the exponent—
or degree—of the 1/f drop-off may reflect the relative balance of excitatory and inhibi-
tory (EI) currents coming in to the region (Gao et al., 2017). For this reason, we refer
to this feature as the aperiodic signal, rather than referring to it as “noise” (Donoghue
et al., 2020).
The prospect that the aperiodic signal may index EI balance is exciting, but much
more work is required to assess the strength and accuracy of this relationship.
Unfortunately, because field potential spectra are a combination of periodic and aperi-
odic components, and because the large-power periodic oscillatory bumps can easily
lead to mismeasurements of the aperiodic component, a method for carefully and ac-
curately separating and parametrizing those components is critical. This is especially
important given the mounting computational and animal single-unit evidence that
suggests that EI balance may be the physiological basis for top-down gain modulation
(Chance et al., 2002) and persistent delay period activity required for working memory
maintenance (Lim & Goldman, 2013).
In order for researchers to state that oscillatory power in frequency X has altered as a
function of task, condition, behavior, or group, they must first demonstrate that such
an oscillation exists in their data, and that the apparent change in oscillatory power is
due to a true reduction in that oscillation as opposed to a change in its center frequency
moving it outside the analyzed band, a shift in its bandwidth, or from a shift in the aperi-
odic signal. Additionally, these oscillations need to be demonstrated to be true rhythms,
and not harmonics caused by nonsinusoidal features of a rhythm at another frequency.
Problematically, standard analytic approaches conflate periodic parameters
(center frequency, power, bandwidth) with aperiodic ones (offset, exponent). This
compromises physiological interpretations. Additionally, separating bursting from
sustained oscillations can be difficult, if not impossible, if just looking at the spectral
representation of the data (Jones, 2016). To further compound the problem, a single
nonsinusoidal, rhythmic process that has “sharp” extrema will manifest, in the power
spectrum, as an oscillatory bump at the frequency of the rhythm as well as exhibiting
smaller bumps at integer harmonics of the rhythm’s primary frequency. This means that
570 BRADLEY VOYTEK
a sharp 10-Hz oscillation will also appear as a bump at 20 Hz, and possibly at 30 Hz and
beyond. In such a scenario—one that is common due to the ubiquity of nonsinusoidal
brain rhythms—a narrowband analysis at, say, 20 Hz, would give the researcher the im-
pression that a 20-Hz beta oscillation exists in their data in conjunction with the 10-Hz
alpha, when in fact all that is present is a nonsinusoidal 10-Hz rhythm. Further, should
this 10-Hz sharpness change in relation to behavioral state, such as a task-related reduc-
tion in sharpness resulting in a “smoother” 10-Hz shape, power in the 20-Hz harmonic
would concomitantly be reduced. This could easily lead to the results being interpreted
as a task-related reduction in 20-Hz beta power, when in fact no such beta oscillation
truly existed in the first place.
There are currently several algorithms for identifying oscillations in specific ways that
have attempted to address some of these concerns individually, but never conjointly. In
particular, an approach called BOSC (Better OSCillation Detector) (Hughes et al., 2012),
begins by fitting a linear regression to the log-log PSD to estimate the aperiodic signal.
This is used to determine a power threshold, which is then used in combination with a
duration threshold to define oscillations in wavelet-based decompositions of the time
series data. However, a significant limitation of this and other similar methods is that a
simple linear fit of the background spectrum can be significantly skewed by the presence
of oscillations—especially large oscillations—and therefore mischaracterizes the aperi-
odic signal.
Another similar approach is the irregular-resampling auto-spectral analysis (IRASA)
method, which seeks to explicitly separate the periodic and aperiodic components
through a resampling procedure (Wen & Liu, 2016). Though conceptually similar, this
resampling method is computationally much more expensive, and may have trouble
separating large amplitude oscillations from the aperiodic signal. Other methods, such
as principle component variants fail to separate periodic and aperiodic features, and re-
quire manual component selection (Miller et al., 2012). Consistent with previous work,
we show that the aperiodic signal is of significant physiological and behavioral interest,
although all of these methods treat it as a nuisance variable, such as correcting for it via
spectral whitening rather than a feature to be explicitly modeled.
To overcome the limitations of traditional narrowband analyses, to reduce the errors
caused by conflating periodic and aperiodic features, and to address the nonsinusoidal
nature of field potential signals, we recently developed several open-source algorithms
for parameterizing field potential data. The first is a semi-automated parameterization
of neural power spectra (Donoghue et al., 2020). Spectra are parameterized as a linear
combination of the aperiodic component and putative periodic oscillations (Figure
23.2). This method thus operates upon frequency representations of time-series field
potential data. The algorithm considers the PSD as the linear sum of an aperiodic signal
upon which there oscillatory “bumps,” or frequency regions of power over and above
this aperiodic signal, can exist. These bumps are considered to be putative oscillations
and modeled individually as Gaussians. Each Gaussian then characterizes the fre-
quency definition of the oscillation, whereby the mean, amplitude, and standard devi-
ation can be interpreted as the center frequency, power, and bandwidth, respectively,
PARAMETERIZING NEURAL FIELD POTENTIAL DATA 571
fit
iterate
(e) remove Gaussians from
original PSD
10 20 30 40
frequency (Hz)
peak Gaussian fit
(b) 2std (h)
remove aperiodic signal assess goodness of fit
final fit
halt itting at noise floor (f) re-fit aperiodic
fit
peak
2std
Figure 23.2 Parametrization algorithm applied to field potential data. (A) The power spectral
density (PSD) is first fit with an estimated aperiodic signal (blue), defined by two parameters: a
slope and an offset. (B) The estimated aperiodic portion of the signal is subtracted from the raw
PSD, the residuals of which are assumed to be a mix of periodic oscillatory peaks and noise.
(C) The maximum (peak) of the residuals is found. If this peak is above the noise floor (2std; red
dashed line) then a Gaussian (green) is fit around this peak based on the peak’s frequency, ampli-
tude, and estimated bandwidth. The fitted Gaussian is then subtracted, and the process is iterated
until the noise floor is reached (bottom). These values are used as seeds for the multi-Gaussian
fitting in D. (D) Having identified the number of putative oscillations, based on the number of
peaks above the noise floor, multi-Gaussian fitting is then performed on the aperiodic-adjusted
signal from B to account for the joint power contributed by all the putative oscillations, together.
(E) This multi-Gaussian model is then subtracted from the original PSD from A. (F) This is done
to give a better estimate of the aperiodic signal—one that is less corrupted by the large oscillations
present in the original PSD. (G) This re-fit aperiodic signal is combined with the multi-Gaussian
model to give the final fit. (H) The final fit—here parameterized as a line (aperiodic signal) and
two Gaussians (putative oscillations)—captures >99% of the variance of the original PSD. In
this example, the extracted parameters for the aperiodic signal are: broadband offset =-21.4 au;
slope =-1.12 au/Hz. Two Gaussians were found, with the parameters: (1) frequency =10.0 Hz
amplitude =0.69 au, bandwidth =3.18 Hz; (2) frequency =16.3 Hz, amplitude =0.14 au, band-
width =7.03 Hz.
This figure appears in https://doi.org/10.1038/s41593-020-00744-x
of the oscillation. The final outputs of the algorithm are the definition of the fit aperi-
odic signal, and the definitions for N Gaussians, where N is the number of oscillations
found in the PSD. Notably, this algorithm extracts all these parameters together in a
manner that accounts for potentially overlapping oscillations; it also minimizes the
degree to which they are confounded and requires no specification of canonical oscil-
lation bands.
572 BRADLEY VOYTEK
As for waveform shape methods, multiple approaches have been developed. In par-
ticular, instantaneous phase estimation methods for hippocampal theta have been
modified to consider the peak and trough locations in a broadband signal to account
for nonsinusoidal features. Another approach, empirical mode decomposition, has
the theoretical capability of extracting a nonsinusoidal oscillation in a single compo-
nent (Sweeney-Reed et al., 2018), although in practice on neural signals the rhythms of
interest spread across multiple components, meaning the decomposition does not reli-
ably separate components of interest.
Though there are still many open questions regarding approaches to analyzing wave-
form shape, we have developed a set of available methods that cover critical basics
(Figure 23.1). Our current toolbox for parameterizing cycle-by-cycle waveform features
(bycycle) breaks an oscillation up into individual cycles by identifying the locations of
peaks and troughs and then computing features on the raw voltage time-series. After the
signal is segmented into cycles, each cycle is characterized by a set of parameters. The
amplitude of the cycle is computed as the average voltage difference between the trough
and the two adjacent peaks. The period is defined as the time between the two peaks.
Rise-decay symmetry is the fraction of the period that was composed of the rise time.
Peak-trough symmetry is the fraction of the period, encompassing the previous peak
and current trough, that was composed of the peak. The distributions of these features
can be computed across all cycles in a signal in order to compare oscillation properties
in different neural signals.
However, in order to analyze cycle-by-cycle features, we must first identify, in the time
domain, whether a signal has an oscillatory cycle or not. After computing features for
each cycle, following the waveform shape algorithm given, an additional step is done to
determine whether each cycle is part of an oscillatory burst. Traditional burst-detection
methods are relatively simple: bursts are detected by comparing instantaneous, band-
limited power at each time point to the median power (usually three times median)
in the channel of interest (Feingold et al., 2015). When the amplitude rises above this
threshold, the signal is said to have entered a bursting state; when it falls below another
threshold (e.g., 1.5 times median) it is said to exit the burst state. While this is accept-
able in some cases, for very stable oscillations and very intermittent oscillations, this cri-
terion becomes unreliable because it will often fail to separate a stable oscillation from
brief nonoscillatory segments. Further, this median-threshold procedure has never
been explicitly tested against ground-truth simulated neural data. Oscillatory bursts
are identified as time periods in which consecutive cycles in the time-series had similar
amplitudes, similar periods, and rise and decay flanks that are predominantly mono-
tonic (Cole & Voytek, 2019).
Unlike our approaches, these other methods are not generalizable across the nu-
merous different forms of field potential data, nor are they designed and optimized to
be parallelizable or to run in a cloud environment. Additionally, none are designed to
examine all of the features in concert, which we show is critical for teasing apart the
physiological interpretations of changes to oscillations, the aperiodic signal, and wave-
form shape.
PARAMETERIZING NEURAL FIELD POTENTIAL DATA 573
23.6 Discussion
The fact that both periodic oscillations and the aperiodic signal are seen in field poten-
tial data hints at the promise for those signals to bridge across those spatial scales. In
studying neural oscillations, the analyses we apply are often very complicated, math-
ematically intensive, and full of assumptions, both explicit and implicit. Therefore,
careful consideration of the methods applied to our data is paramount, as seemingly
arbitrary choices in the hyperparameters of our analysis, such as the minimum number
of spikes required for including a neuron in an analysis or the length and precise cutoff
frequencies of a filter, can have large impacts on the results and ultimate conclusions.
Often, in-depth knowledge of the techniques is required in order to appropriately
choose hyperparameters and assure the validity of our conclusions. Because consid-
erable effort is required to obtain this knowledge, this means that we will often make
honest mistakes in our analysis and as peer reviewers, because we often miss the statis-
tical confounds that may underlie highly impactful results.
Therefore, we first advocate paying deliberate and careful attention to the raw data to
help gain a maximal understanding of broad features; complicated methods can often
transform our data in ways we do not expect. If we do not understand how we see the
ultimate effect by looking at the raw data, there is reason for concern, or at least further
investigation. It can similarly be useful to apply multiple methods to our data. When it
comes to analyzing neural oscillations, there are several analytic options to choose from
(e.g., in the frequency or time-domains), and so this choice should be made consciously.
The choice for the analytic method applied can strongly impact the ultimate conclu-
sion. As we showed, phase-amplitude coupling analysis and sharpness analysis were
capturing the same phenomena in the data. However, the former favored a conclusion
of coupled oscillations, whereas we favored the latter conclusion concerning synchrony
of transmembrane currents. The choice of developing a time-domain approach to
analyzing neural oscillations was very deliberate: as physical processes, including neural
dynamics, happen over time (as opposed to being generated in the frequency domain),
there are advantages in analyzing signals in this natural domain and directly measuring
and accounting for non-stationarities in the dynamics.
While it is certainly a biased perspective, we believe that neural oscillations research
would be better positioned if analysis of these rhythms combined time-domain and
frequency-domain approaches, using careful parameterization. Time-domain cycle-by-
cycle parameterization offers several advantages over frequency-domain approaches.
First, it directly quantifies waveform asymmetry, which is only indirectly and ambigu-
ously captured in spectral analysis (i.e., similar harmonic patterns can be generated by
different oscillations that produce diverse waveforms). Second, cycle-by-cycle param-
eterization inherently runs an oscillation detection algorithm, so oscillatory features
are only analyzed on appropriate portions of the signal (i.e., where the oscillation is
observable). Third, cycle-by-cycle parameterization offers time-resolved estimates of
574 BRADLEY VOYTEK
oscillatory features with an appropriate degree of temporal resolution: not only is sym-
metry measured for a single cycle, but the estimates of amplitude and frequency are
intuitively measured in terms of peak-to-trough voltage and trough-to-trough time,
respectively. However, “instantaneous” estimates of amplitude and frequency that are
comparatively used offer amplitude and frequency estimates at every point in time.
We have previously demonstrated that the cycle-by-cycle approach can provide more
sensitive measures of amplitude and frequency, compared to more traditional filter-
and-Hilbert transform approaches, by better differentiating simulated experimental
conditions (Cole & Voytek, 2019).
One downside of the time-domain approach outlined is that it can miss relatively
weaker oscillations, and that fine-tuning of oscillation-detection hyperparameters can
influence results. For this reason, we advocate including a frequency-domain param-
eterization approach to help identify which frequencies have power above the aperiodic
signal, during which times. In this way, both parameterization approaches can be used
synergistically.
To date, tens of thousands of studies have been published regarding neural
oscillations, their function, and their behavioral, cognitive, and disease correlates.
Most of these studies have been conducted using traditional approaches that assume
canonical frequency bands—delta, theta, alpha/mu, beta, and gamma. Often these os-
cillation bands are described as having roles themselves, such as theta being equated
with memory, and alpha with attention and wakefulness. This approach arises due to
letting predefined bands guide analyses, rather than allowing the data to guide the
analyses. That is, canonical band analyses commit us to tacit acceptance of predefined
oscillatory bands having a functional role, rather than considering the underlying
physiological mechanisms that may generate different spectral features and addressing
inter-individual differences. With proper parameterization, it is possible to broaden
our perspective, allowing us to take full advantage of the rich variance present in oscil-
latory data. This increases analytical power and can potentially provide greater insight
into physiological mechanisms underlying oscillations and the role that oscillatory
variability may play in explaining individual differences in cognitive functioning in
health, aging, and disease.
ACKNOWLEDGEMENTS
I sincerely thank Scott Cole, Thomas Donoghue, Richard Gao, Matar Haller, Erik Peterson,
Natalie Schaworonkow, Avgusta Shestyuk, Tammy Tran, and Roemer van der Meij for all the
years of hard work, careful thought, and patience with these ideas. My support for this work
came from a Sloan Research Fellowship (FG-2015-66057), the Whitehall Foundation (2017-12-
73), the National Science Foundation under grant BCS-1736028, and the National Institute of
General Medical Sciences grant 1R01GM134363-01. The author declares no competing finan-
cial interests. Portions of this chapter are based on/informed by Donoghue et al., 2020; Cole &
Voytek, 2018; Peterson & Voytek, 2018.
PARAMETERIZING NEURAL FIELD POTENTIAL DATA 575
References
Amzica, F. & Steriade, M. (1998). Electrophysiological correlates of sleep delta waves.
Electroencephalography and Clinical Neurophysiology, 107, 69–83.
Arroyo, S., Lesser, R. P., Gordon, B. , Uematsu, S., Jackson, D., & Webber, R. (1993). Functional
significance of the mu rhythm of human cortex: An electrophysiologic study with subdural
electrodes. Electroencephalography and Clinical Neurophysiology, 87, 76–87.
Belluscio, M. A., Mizuseki, K., Schmidt, R., Kempter, R., & Buzsáki, G. (2012). Cross-frequency
phase-phase coupling between theta and gamma oscillations in the hippocampus. Journal of
Neuroscience, 32, 423–435.
Bruns, A., Eckhorn, R., Jokeit, H., & Ebner, A. (2000). Amplitude envelope correlation detects
coupling among incoherent brain signals: Neuroreport, 11, 1509–1514.
Bullock, T. H. (1981). Neuroethology deserves more study of evoked responses. Neuroscience,
6(7), 1203–1215.
Buzsáki, G. (2004). Large-scale recording of neuronal ensembles. Nature Neuroscience, 7,
446–451.
Buzsáki, G., Anastassiou, C. A., & Koch, C. (2012). The origin of extracellular fields and
currents—EEG, ECoG, LFP and spikes. Nature Reviews Neuroscience, 13, 407–420.
Buzsáki, G. & Draguhn, A. (2004). Neuronal oscillations in cortical networks. Science, 304,
1926–1929.
Canolty, R. T., Edwards, E., Dalal, S. S., Soltani, M., Nagarajan, S. S., Kirsch, H. E., . . . Knight,
R. T. (2006). High gamma power is phase-locked to theta oscillations in human neocortex.
Science, 313, 1626–1628.
Canolty, R. T. & Knight, R. T. (2010). The functional role of cross-frequency coupling. Trends in
Cognitive Sciences, 14, 506–515.
Chance, F. S., Abbott, L. F., & Reyes, A. D. (2002). Gain modulation from background synaptic
input. Neuron, 35, 773–782.
Cole, S. & Voytek, B. (2019). Cycle- by- cycle analysis of neural oscillations. Journal of
Neurophysiology, 122(2), 849–861.
Cole, S. R., van der Meij, R., Peterson, E. J., de Hemptinne, C., Starr, P. A., & Voytek, B. (2017).
Nonsinusoidal beta oscillations reflect cortical pathophysiology in Parkinson’s disease.
Journal of Neuroscience, 37, 4830–4840.
Cole, S. R. & Voytek, B. (2017). Brain oscillations and the importance of waveform shape.
Trends in Cognitive Sciences, 21, 137–149.
Cole, S. R. & Voytek, B. (2018). Hippocampal theta bursting and waveform shape reflect CA1
spiking patterns. bioRxiv [online]. http://biorxiv.org/lookup/doi/10.1101/452987
Dave, S., Brothers, T. A., & Swaab, T. Y. (2018). 1/f neural noise and electrophysiological indices
of contextual prediction in aging. Brain Research, 1691, 34–43.
de Hemptinne, C., Swann, N. C., Ostrem, J. L., Ryapolova-Webb, E. S., San Luciano, M.,
Galifianakis, N. B., & Starr, P. A. (2015). Therapeutic deep brain stimulation reduces cortical
phase-amplitude coupling in Parkinson’s disease. Nature Neuroscience, 18, 779–786.
Engel, A. K., Fries, P., & Singer, W. (2001). Dynamic predictions: Oscillations and synchrony in
top–down processing. Nature Reviews Neuroscience, 2, 704–7 16.
Feingold, J., Gibson, D. J., DePasquale, B., & Graybiel, A. M. (2015). Bursts of beta oscilla-
tion differentiate postperformance activity in the striatum and motor cortex of monkeys
performing movement tasks. Proceedings of the National Academy of Sciences of the United
States of America, 112, 13687–13692.
576 BRADLEY VOYTEK
Freeman, W. J. & Zhai, J. (2009). Simulated power spectral density (PSD) of background elec-
trocorticogram (ECoG). Cognitive Neurodynamics, 3, 97–103.
Fries, P. (2005). A mechanism for cognitive dynamics: neuronal communication through
neuronal coherence. Trends in Cognitive Sciences, 9, 474–480.
Gao, R., Peterson, E. J., & Voytek, B. (2017) Inferring synaptic excitation/inhibition balance
from field potentials. NeuroImage, 158, 70–78.
Haegens, S., Cousijn, H., Wallis, G., Harrison, P. J., & Nobre, A. C. (2014). Inter-and intra-
individual variability in alpha peak frequency. NeuroImage, 92, 46–55.
Donoghue, T., Haller, M., Peterson, E., Varma, P., Sebastian, P., Gao, R., . . . Voytek, B. (2020).
Parameterizing neural power spectra into periodic and aperiodic components. Nature
Neuroscience, 23,1655–1665.
Helfrich, R. F. & Knight, R. T. (2016). Oscillatory dynamics of prefrontal cognitive control.
Trends in Cognitive Sciences, 20, 916–930.
Hughes, A. M., Whitten, T. A., Caplan, J. B., & Dickson C. T. (2012). BOSC: A better oscilla-
tion detection method, extracts both sustained and transient rhythms from rat hippocampal
recordings. Hippocampus, 22, 1417–1428.
Jackson, N., Cole, S. R., Voytek, B., & Swann, N. C. (2019). Characteristics of waveform
shape in Parkinson’s disease detected with scalp electroencephalography. eNeuro, 6(3),
ENEURO.0151-19.2019.
Jasper, H. & Penfield, W. (1949). Electrocorticograms in man: Effect of voluntary movement
upon the electrical activity of the precentral gyrus. Archiv für Psychiatrie & Nervenkrankh,
183, 163–174.
Jensen, O. (2002). Oscillations in the alpha band (9-12 Hz) increase with memory load during
retention in a short-term memory task. Cerebral Cortex, 12, 877–882.
Jensen, O. & Mazaheri, A. (2010). Shaping functional architecture by oscillatory alpha ac-
tivity: Gating by inhibition. Frontiers in Human Neuroscience, 4, 186. doi:10.3389/
fnhum.2010.00186
Jones, S. R. (2016). When brain rhythms aren’t ‘rhythmic’: Implication for their mechanisms
and meaning. Current Opinion in Neurobiology, 40, 72–80.
Klimesch, W. (1999). EEG alpha and theta oscillations reflect cognitive and memory perform-
ance: A review and analysis. Brain Research Reviews, 29, 169–195.
Klimesch, W. (2012). Alpha-band oscillations, attention, and controlled access to stored infor-
mation. Trends in Cognitive Sciences, 16, 606–617.
Kopell, N. J., Gritton, H. J., Whittington, M. A., & Kramer, M. A. (2014). Beyond the
connectome: The dynome. Neuron, 83, 1319–1328.
Kuhlman, W. N. (1978). Functional topography of the human mu rhythm.
Electroencephalography and Clinical Neurophysiology, 44, 83–93.
Lim, S. & Goldman, M. S. (2013). Balanced cortical microcircuitry for maintaining informa-
tion in working memory. Nature Neuroscience, 16,1306–1314.
Lundqvist, M., Rose, J., Herman, P., Brincat, S. L., Buschman, T. J., & Miller, E. K. (2016).
Gamma and beta bursts underlie working memory. Neuron, 90, 152–164.
Miller, K. J., Hermes, D., Honey, C. J., Hebb, A. O., Ramsey, N. F., Knight, R. T., Ojemann,
J. G., & Fetz, E. E. (2012). Human motor cortical activity is selectively phase-entrained on
underlying rhythms. PLoS Computational Biology, 8, e1002655.
Obrist, W. D. (1954). The electroencephalogram of normal aged adults. Electroencephalography
and Clinical Neurophysiology, 6, 235–244.
PARAMETERIZING NEURAL FIELD POTENTIAL DATA 577
Pesaran, B., Vinck, M., Einevoll, G. T., Sirota, A., Fries, P., Siegel,
M., . . . Srinivasan, R. (2018). Investigating large-s cale brain dynamics using field po-
tential recordings: analysis and interpretation. Nature Neuroscience, http://w ww.nat
ure.com/articles/s41593-018-0171-8
Peterson, E. J., Rosen, B. Q., Campbell, A. M., Belger, A., & Voytek, B. (2018). 1/f neural noise is
a better predictor of schizophrenia than neural oscillations. bioRxiv [online]. http://biorxiv.
org/lookup/doi/10.1101/113449
Peterson, E. J. & Voytek, B. (2017). Alpha oscillations control cortical gain by modulating
excitatory-inhibitory background activity. bioRxiv [online]. http://biorxiv.org/lookup/doi/
10.1101/185074
Robertson, M. M., Furlong, S., Voytek, B., Donoghue, T., Boettiger, C. A., & Sheridan, M. A.
(2019). EEG power spectral slope differs by ADHD status and stimulant medication ex-
posure in early childhood. Journal of Neurophysiology, 122(6), 2427–2437. doi:10.1152/
jn.00388.2019.
Samaha, J., Bauer, P., Cimaroli, S., & Postle, B. R. (2015). Top-down control of the phase of
alpha-band oscillations as a mechanism for temporal prediction. Proceedings of the National
Academy of Sciences of the United States of America, 112, 8439–8444.
Sherman, M. A., Lee, S., Law, R., Haegens, S., Thorn, C. A., Hämäläinen, M. S., Moore, C. I., &
Jones, S. R. (2016). Neural mechanisms of transient neocortical beta rhythms: Converging
evidence from humans, computational modeling, monkeys, and mice. Proceedings of the
National Academy of Sciences of the United States of America, 113, E4885–E4894.
Steriade, M., Contreras, D., & Curr, F. (1993). The slow (4 Hz) oscillation in reticular thalamic
and thalamocortical neurons: Scenario of sleep rhythm generation in interacting thalamic
and neocortical networks. Journal of Neuroscience, 13(8), 3284–3299.
Sweeney-Reed, C. M., Nasuto, S. J., Vieira, M. F., & Andrade, A. O. (2018). Empirical mode de-
composition and its extensions applied to EEG analysis: A review. Advances in Data Science
and Adaptive Analysis, 10(2), 1840001.
Tort, A. B. L., Fontanini, A., Kramer, M. A., Jones-Lush, L. M., Kopell, N. J., & Katz, D. B. (2010).
Cortical networks produce three distinct 7-12 Hz rhythms during single sensory responses
in the awake rat. Journal of Neuroscience, 30, 4315–4324.
Varela, F., Lachaux, J-P., Rodriguez, E., & Martinerie, J. (2001). The brainweb: Phase synchron-
ization and large-scale integration. Nature Reviews Neuroscience, 2, 229–239.
Vaz, A. P., Yaffe, R. B., Wittig, J. H., Inati, S. K., & Zaghloul, K. A. (2017). Dual origins of
measured phase-amplitude coupling reveal distinct neural mechanisms underlying epi-
sodic memory in the human cortex. NeuroImage, 148, 148–159.
Veerakumar, A., Tiruvadi, V., Howell, B., Waters, A. C., Crowell, A. L., Voytek, B., . . . Mayberg,
H. S. (2019). Field potential 1/f activity in the subcallosal cingulate region as a candidate
signal for monitoring deep brain stimulation for treatment-resistant depression. Journal of
Neurophysiology, 122(3), 1023–1035.
Voytek, B., D’Esposito, M., Crone, N., & Knight, R. T. (2013). A method for event-related phase/
amplitude coupling. NeuroImage, 64, 416–424.
Voytek, B., Kayser, A. S., Badre, D., Fegen, D., Chang, E. F., Crone, N. E., . . . D’Esposito, M.
(2015a). Oscillatory dynamics coordinating human frontal networks in support of goal
maintenance. Nature Neuroscience, 18, 1318–1324.
Voytek, B. & Knight, R. T. (2015). Dynamic network communication as a unifying neural basis
for cognition, development, aging, and disease. Biological Psychiatry, 77, 1089–1097.
578 BRADLEY VOYTEK
Voytek, B., Kramer, M. A., Case, J., Lepage, K. Q., Tempesta, Z. R., Knight, R. T., & Gazzaley, A.
(2015b). Age-related changes in 1/f neural electrophysiological noise. Journal of Neuroscience,
35, 13257–13265.
Waschke, L., Wöstmann, M., & Obleser, J. (2017). States and traits of neural irregularity in
the age-varying human brain. Scientific Reports, 7, art no. 17381. https://doi.org/10.1038/s41
598-017-17766-4
Wen, H. & Liu, Z. (2016). Separating fractal and oscillatory components in the power spectrum
of neurophysiological signal. Brain Topography, 29, 13–26.
Index
For the benefit of digital users, indexed terms that span two pages (e.g., 52–53) may, on occasion,
appear on only one of those pages.
Tables and figures are indicated by t and f following the page number
A affective reactions
ACC see anterior cingulate cortex (ACC) see also affective processes, asymmetric
action potentials 23–24, 46–47 frontal cortical activity
active inhibition 206–8 and manipulations of asymmetric frontal
active sensing 51 cortical activity 222–24
ActiveTwo system 526 neurofeedback 223
adaptive control hypothesis 182 situational 224
adenosine triphosphate (ATP) 410–11 unilateral hand contractions 223–24
adolescence and early adulthood negative affect 224
age as discrete or continuous 334–35 positive affect 224, 226–27
specific age-related associations 330–31 affective valence
time-frequency activity in 324–47 and asymmetric frontal cortical
advantages of time-frequency activity 221–22
activity 325–27 defining 221
characterizing the results 339–42, 340f and motivational direction 221, 227–28
covariation in patterns of change 342–45 trait 222
and development 328–31, 335 age factors
empirical exploration 335–46 age-related associations, developmental
extending of findings 338 research 330–31
methodological issues 333–34 altered oscillations 270
pubertal stage 334 aperiodic signal 568
research design issues 331–33 cognitive decline in aging 92–93, 401–2
statistical analysis 334–35 and development 270
task difficulty vs. task demands 333–34 research design 331–32
Adrian, E. D. 145–46 and sleep 402–3, 408, 413
affective processes Alayrangues, J. 270–7 1
see also affective reactions algorithms
asymmetric frontal cortical activity artefact correction 34
parenting and child frontal EEG cognitive tasks 44–45
asymmetry 309–10 complementary 157
socioemotional outcome 309 data-driven/adaptive time frequency
temperament 308–9 representations 71–72
differential functions of 6–9 Hz distributed source modeling 380
activity 310–11 eye-tracking data 157
580 Index
bivariate parallel process models 339–42, 341t combining with neuroimaging 544–48
bivariate phase synchrony 519 computer simulations 537
blood-oxygen-level dependent (BOLD) invasive 533–34
signals 359 safety aspects 548–49
brain stimulation 545 targets for 539–44
fMRI 388–89, 410–11 transcranial alternating current
sleep 410–11 stimulation (tACS) 165, 263, 363, 422,
source localization, cortical 388–89, 391–92 534–35, 540–41, 542–44, 547, 551
Bonnefond, M. 207–8 transcranial direct current stimulation
Borkovec, T. D. 467 (tDCS) 534–35
Born–Jordan RID 70–7 1 transcranial electrical stimulation
boundary element method (BEM) 383, 384, (tES) 534–39, 536f
387 transcranial magnetic stimulation
Bowers, M. E. 100 (TMS) 444
bradykinesia 272–74 rhythmic repetitive application
brain (rTMS) 534–39, 536f, 541, 548
see also amygdala; anterior cingulate cortex transcranial random noise
(ACC); brain stimulation; cerebral stimulation 534–35
cortex; cerebrospinal fluid (CSF); broadband activity 496–97
frequency analysis; frequency bands; Brodbeck, V. 388
frontal cortex; gray matter (GM); burst-detection methods 572
hippocampus; hypothalamus; motor Button, K. S. 10–11
cortex; neuroimaging; prefrontal cortex;
regions of interest (ROIs); thalamus; C
ventral tegmental area (VTA); visual Canali, P. 444
cortex; white matter (WM) capacitance 18, 23, 37–38
classical views of function 444–45 capacitors 21–22
default mode network, regions Carter, W. R. 467
within 392–93 Carver, C. S. 225
distribution of tissues 384 cat neocortex 146
electrochemical machine 15 cat olfactory bulb 145–46
experimental human research 359 cat visual cortex 146, 147
functional connectivity (FC), applying EEG causality, in monkeys
source analysis 388–92 fronto-parietal network 139
hemispheres see hemispheres, brain sensorimotor systems 134–35
oscillations see oscillations, neural visual neocortex 136–38
oscillatory dynamics 434–51 Cavanagh, J. F. 97, 469–70
rhythms of 4 CDA see contralateral delay activity
skull-to-brain conductivity ratio, (CDA)
uncertainty 384–85 cell membranes 23
subcortical regions 474 central executive and attentional control
whole-brain functional connectivity function 364–66
(FC) 394 cerebral cortex 293–94, 353
whole-brain networks 435–36 cerebrospinal fluid (CSF) 383, 384
brain computer interfaces 275 CFC see cross-frequency coupling (CFC)
brain stimulation 532–52 chaos theory 50
challenges/future directions 549–51 Chen, R. 263
Index 587
left and right regions 221, 236, 363 graph theoretic metrics 521–25
medial 96 illustrative example 526–28
structural development 358 spectral methods for multivariate
theta-band activity 387 synchronization 517–18
TMS 444 resting-state 468–69
VTA projections 359 small-world parameter 523–24
frontal eye fields (FEF) 330–31 whole-brain 394
frontal midline theta (FMT) functional connectivity networks (FCNs) 521,
activity during Stage 1 and REM sleep 415–16 526–27
anterior cingulate cortex (ACC) 179–80 functional magnetic resonance imaging
and anxiety 469–74 (fMRI) 212
broader impact/future directions 188 and anxiety 468–69
characteristics 179–83 BOLD response 388–89, 410–11
clinical applications 187–88 brain stimulation 545
conflict-related 182–83 combined with electrocorticography 359
and cross-frequency coupling 479–80 feedback negativity (FN)/reward positivity
enhanced 182 (RewP) 100
ERP components 179–80 resting-state research 391–92
ERP findings linked to FMT source localization 378–79, 388–89, 394
response 180–82 EEG inverse solutions 382, 383
error-related 472–73 Funder, D. C. 244
feedback-related 473 fusiform face area (FFA) 378–79
functional aspects 179–80
history 178–79 G
model specimen of cortical theta 178–88 GABA see gamma-aminobutyric acid
primer on as an index of medial frontal (GABA)
cortex function 469–70 Gable, P. A. 232, 233–34, 235
processes 179 Galambos, R. 44, 146
response-conflict 182–83 gamma activity
single-trial 472–73 see also prokinetic gamma
translational potential 183–87 and ADHD 162–63
whether ERP theta and FMT reflect the and ASD 162
same process 182–83 clinical relevance 160–64
frontal-related negativity (FRN) 360 early studies 146
fronto-basal ganglia circuits 275 evoked 150–51, 161
fronto-parietal network, in monkeys 139 frequencies, in traditional EEG
Fukuda, K. 127 research 149–50
full-band EEG 410–11 high and low gamma 149–50
functional connectivity (FC) induced (total) 147–49, 151–52
bivariate measures 495–509 and microsaccades 158–60
clustering coefficient 524 methodological aspects 155–56
coherence 265–69 modality-specific effects 161
EEG source analysis, applying 388–92, 394 neurofeedback 164–65
executive control network (ECN) 366 neurophysiological mechanisms 153–55
multivariate methods 514–28 oscillatory synchrony 153
direct measures of multivariate phase-locked early evoked 162
synchronization 518–21 pyramidal cells 153, 161–62
594 Index
periodic stimulation, generating Parkinson’s disease (PD) 163, 183–84, 265, 267, 272
oscillations by 54–56, 56f and dementia 422
phase and amplitude correlations 497 Parrino, L. 411
pre-event oscillatory activity 123–26 partial correlation matrix 506–8
in prefrontal cortex, of monkeys 138–39 partial path diagram, bivariate growth
properties 204 model 343–45, 344f
quantifying and categorizing 42f, 42–44 parvalbumin-expressing (PV+) inhibitory
quasi-periodic 496, 497, 498f neurons 448–49
resting state 446 Pascual-Marqui, R. D. 380–81
ripples/HFO oscillations 424 PCA see principal component analysis (PCA)
sensorimotor function see sensorimotor Pearson correlation coefficients 342–43
system pedunculopontine nucleus (PPN) 267–68
sensory and perceptual gamma, in Penn State Worry Questionnaire 467
schizophrenia 435–42 perception
sinusoidal oscillators 184–85 alpha-band changes during 52–53
slow delta and alpha 367–68 action-perception processes 302–4
spatio-temporal patterns and travelling Gestalt perception 439–41
waves 50–51 oscillations as substrate of representations
spontaneous, in schizophrenia 435, 444–48 in 59
and evoked activity 444–45 sensory and perceptual gamma
as “noise” 444–45, 447 oscillations 435–42
sleep 447–48 Perez, V. B. 438
during tasks 447 periodicity, sleep spindles 421
subthreshold 47 periodic stimulation, generating oscillations
synchrony and oscillatory by 54–56, 56f
communication 51–52 Peterson, E. J. 245
TMS-evoked 435 PFC see prefrontal cortex (PFC)
traditional demarcation of oscillatory Pfurtscheller, G. 206–7
events 45–46 phase-amplitude coupling (PAC) 501, 568
working memory task, changes during 43f, 43 phase and amplitude correlations of
oscillatory communication 51–52 oscillations 497
Ozer, D. J. 244 phase-and amplitude-time series
estimation 499–500
P phase coherency 44–45
P2-N2-P3-Slow Wave complex 95 phase coupling and causality, in monkeys
P2-RewP-P3-Slow Wave complex 102 fronto-parietal network 139
P3-related processes sensorimotor systems 134–35
components 94 visual neocortex 136–38
ERP methodologies 92, 94–96 Wiener–Granger (WG) causality 134–35,
oddball task 94–96 137–38, 139
P3a subcomponent 94 phase-frequency regression 266–67
P3b subcomponent 94 phase-locking
reduced time-domain P3 95–96 see also phase-locking value (PLV)
pallidal activity 273 gamma activity 162
Papenberg, G. 330–31 indices of 44–45
paradoxical sleep 404 inter-trial 55–56, 56f
parametrization algorithm 570–7 1, 571f quantification of 42–43, 502–3
602 Index
phase locking factor (PLF) 436–37, 438, 440, potassium ion channels 23–24
447, 450 potential energy 18, 37–38
phase-locking value (PLV) power
and alternative measures 502–3 see also alpha/beta power and oscillations;
functional connectivity (FC) 514–15 alpha power; beta power; gamma
imaginary part (iPLV) 505–6 power; theta-beta power ratio
linear mixing inflating 504–6 baseline-corrected 336–38
measurement of phase synchrony and definitions of EEG frequency research 3–4
CFC 504 enhanced beta/gamma 423
single trial phase-locking value (S-PLV) 83 oscillations 91
phase-phase coupling 568 peak power, in sleep 409
phase spectrum 28, 29f, 37–38 a priori calculations, conducting 10–11
phase synchronization index 82–83 semi-automated parameterization 570–7 1
phase synchrony spectral see spectral power
see also synchrony time-frequency 336–38
inter-regional 391 power spectrum 28, 37–38, 65–66
inter-trial phase synchrony (ITPS) 83 predictive coding 53, 128
measures 82–83 prefrontal cortex (PFC) 140, 359
phosphenes 549 and amygdala 310
photic stimulation 146 and attention 359
physiological artefacts 156–58 and cognitive control 470
pictorial dot-probe paradigm 364–65 dorsolateral 330–31, 468–69
Pivik, R. T. 299–300 lateral 473
plasticity macaque monkey 131–32, 138–39
Hebbian 53–54 and medial frontal cortex 470
oscillations as drivers of 57–58 oscillations in 138–39
spike-timing dependent plasticity and posterior parietal cortex 139, 140
(STDP) 539 right lateral 234
synaptic 153–54, 270 schizophrenia, impairments in 311,
PLV see phase-locking value (PLV) 442–44
polarity 88–90 and sleep 412–13
population coding 264 and TMS-evoked oscillations 444
positive affect and visual working memory 140
and approach motivation 224–25, 226–27 and VTA 359
and asymmetric frontal cortical Prehn-Kristensen, A. 163
activity 224, 226–27 “preparing to overcome prepotency” (POP)
positive occipital sharp transients task, 443
(POSTs) 418 pre-register specific hypotheses 9–10
positron emission tomography (PET) 378–79, pre-stimulus beta-frequency coherence 134,
382 135f, 136–37, 137f
raclopride 359 primary motor cortex 261f, 261
Possel, P. 240–41 principal component analysis (PCA) 34, 35,
posterior alpha band, ontogenesis 297 91, 342–43, 546
post-synaptic potentials 4–5, 47, 205–6 developmental research, adolescence 336,
post-synaptic receptor molecules 132–33 337f, 339, 340f
post-traumatic stress disorder (PTSD) 241, process H (sleep homeostasis) 402–3, 404
467–68 prokinetic gamma 264–65
Index 603
realistic head models, source analysis 384 sleep spindles and cognition 422
recordings 419–20 as a “splitting of the mind” 435–36
in schizophrenia 444–45 spontaneous oscillations 435, 444–48
scalp-Laplacian 35–36 resting state oscillations 446
scalp location 35 sleep oscillations 447–48
scalp recordings 153, 159 during tasks 447
in ADHD 354, 355f transcranial magnetic stimulation
alpha activity 52–53 (TMS) 444
EEG and MEG 450 unmedicated 161, 437, 438, 443, 446
field dynamics 65 visual evoked and induced gamma 441–42
frontal midline theta (FMT) 179 working memory 434, 442–44
gamma activity 147, 155–56 schizotypal personality disorder 434, 437
potentials 365, 436 Schmeichel, B. J. 236
resting-state frontocentral theta Schöne, B. 226
activity 353 Schutter, D. J. 475
scalp topography 88–90, 102 selective attention, alpha-band changes
theta 296 during 52–53
Schaffer, C. E. 239–40 sensorimotor system
schizophrenia cognitive aspects 274–75
see also antipsychotic medication coherence
auditory gamma oscillations 436–39 cortico-cortical 268–69
auditory steady-state response cortico-spinal 266–67
(ASSR) 436–38, 439 cortico-subcortical 267–68
early auditory-evoked gamma band as a measure of functional connectivity
response (EAGBR) 436, 438–39 (FC) 265–69
auditory hallucinations 434 frequency bands see frequency bands
chronic 161, 437, 443, 446 long-term changes in oscillatory
circuit mechanisms underlying oscillation activity 269–73
abnormalities 448–50 development and aging 270
NMDAR (N-methyl-D-aspartate learning 270–7 1
receptor) hypofunction 438, 449–50 movement disorders 271–73
reduced synaptic connectivity 450 movement-related time-frequency spectra
delusions 434 for primary motor cortex 261f, 261
early-onset 434 oscillatory activity 259–76
first-episode 437, 438, 443 phase coupling and causality, in
future directions 451 monkeys 134–35
and gamma activity 160–61 understanding 259–61
hallucinations 437 sensory and perceptual gamma oscillations
oscillatory brain dynamics 434–51 auditory 436–39
PFC impairments 311, 442–44 auditory steady-state response
psychotic symptoms 434, 437 (ASSR) 436–38, 439
scalp EEG 444–45 early auditory-evoked gamma band
sensory and perceptual gamma response (EAGBR) 436, 438–39
oscillations 435–42 visual 439–42
auditory 436–39 Gestalt perception 439–41
visual 439–42 visual evoked and induced
Singer and Eckhorn labs 435 gamma 441–42
606 Index
reward-related delta (RewP-delta) 101, 102, ssVEPs see steady-state visual evoked
103–4 potentials (ssVEPs)
Society of Psychophysiological Research 301 standardized low-resolution brain
sodium ion channels 23–24 electromagnetic tomography
source analysis studies 475 (sLORETA) 233, 380–81, 387–88, 392
source localization, cortical 377–94 State-Trait Anger Expression
see also inverse model Questionnaire 231, 234–35
application of, in frequency steady-state evoked potentials 146
analysis 386–94 steady-state responses 533–34
clinical applications 388 steady-state visual evoked potentials
forward and inverse model 377 (ssVEPs) 54–56, 149f
frequency analysis 386–88 Steriade, M. 412–13
frequency bands 262–63 Sternberg paradigm 443
functional brain connectivity, applying EEG Stewart, J. L. 240
source analysis 388–92 STFT see short-time Fourier transform
pediatric populations 385–86, 392–94 (STFT)
procedures 377 stimulus evoked responses 444–45
realistic models and structural Stroganova, T. A. 297, 311–12
MRI 383–86 Strogatz, S. H. 523
techniques/approaches 378–83 Stroop-like tasks 306
applying a priori information as structural equation model (SEM) 339
constraints to inverse solution 382–83 structural MRI 383
distributed source modeling 378, 379–82 and realistic models see realistic head
equivalent current dipole (ECD) 378–79 models, source analysis
“inverse problem” with EEG source substance use disorder (SUD) 97
analysis 381–82 subthalamic deep brain stimulation 568
source space 379 subthalamic nucleus (STN) 260–61, 261f, 267,
spatial filtering 546–47 272f, 272, 275–76
spatio-temporal patterns 50–51 Summerfield, C. 152
spectral plots 301, 302f Sun, L. 441–42
spectral power 106, 564 superior temporal gyrus (STG) 450
averaging 273 suprachiasmatic nucleus (SCN) 402
band-limited relative 410 supragranular labeled neurons (SLN) 138
changes in 271 supragranular layers 138
distributions 354, 355f surface negative deflection 412–13
and oscillations 43, 44–45, 54–55 Sutton, S. K. 94
overall 157 SWA see slow wave activity (SWA)
sensorimotor system 261f, 261, 271–72, 272f SWS see slow wave sleep (SWS)
sleep 410 synchronization 294, 310–11
task-related 157 see also desynchronization; event-related
spectrogram 104 synchronization (ERS); multivariate
spectrum of spectra approach 411–12 synchronization
Spencer, K. M. 439, 440, 441–42, 447 global field synchronization (GFS) 518
spike timing 47, 134–35, 153–54, 497 phase synchronization index 82–83
coordination of 185, 360, 564–65 synchrony
spike-timing dependent plasticity (STDP) 539 see also phase synchrony; time-frequency-
spike trains 45, 46–47, 48 based phase synchrony analysis
608 Index
V W
Van Dijk, B. W. 126 waking EEG 297
Van Honk, J. 475 Watts, D. J. 523
Van Peer, J. M. 475 wavelet analysis
van Tricht, M. J. 446 continuous wavelet transform (CWT) 68–
Van Wijk, M. 266 69, 80, 84
ventral lateral thalamus 267, 275–76 Morlet wavelet 103, 120f, 499
ventral tegmental area (VTA) 359, 360 time-frequency decomposition (TFDs)
ventrolateral and dorsolateral prefrontal methods 68–69, 76–77, 80
cortex of the right hemisphere quantifying and categorizing brain
(VLPFC and DLPFC) 330–31 oscillations 43
vertex centrality 509 wavelet-based phase estimation 80
vertex sharp transients 418 wavelet time-frequency resolution trade-off
vertical electro-oculogram (VEOG) 158 31f, 31
vigilance, effective and affective facets of 179 wavelet transforms 91–92
visual cortex Webb, C. 101
see also macaque monkey weighted phase-lag index (wPLI) 389–91,
alpha oscillations 206–7 505–6
attentional influences on 154–55 White, T. L. 225
Index 611