FMRI Filippi
FMRI Filippi
fMRI
Techniques
and Protocols
Second Edition
NEUROMETHODS
Series Editor
Wolfgang Walz
University of Saskatchewan
Saskatoon, SK, Canada
Second Edition
Edited by
Massimo Filippi
Neuroimaging Research Unit, INSPE, Division of Neuroscience,
San Raffaele Scientific Institute and Vita-Salute San Raffaele
University, Milan, Italy
Editor
Massimo Filippi, MD, FEAN
Neuroimaging Research Unit
Institute of Experimental Neurology
Division of Neuroscience
San Raffaele Scientific Institute and
Vita-Salute San Raffaele University
Milan, Italy
Department of Neurology
Division of Neuroscience
San Raffaele Scientific Institute and
Vita-Salute San Raffaele University
Milan, Italy
Experimental life sciences have two basic foundations: concepts and tools. The Neuromethods
series focuses on the tools and techniques unique to the investigation of the nervous system
and excitable cells. It will not, however, shortchange the concept side of things as care has
been taken to integrate these tools within the context of the concepts and questions under
investigation. In this way, the series is unique in that it not only collects protocols but also
includes theoretical background information and critiques which led to the methods and
their development. Thus it gives the reader a better understanding of the origin of the
techniques and their potential future development. The Neuromethods publishing program
strikes a balance between recent and exciting developments like those concerning new ani-
mal models of disease, imaging, in vivo methods, and more established techniques, includ-
ing, for example, immunocytochemistry and electrophysiological technologies. New
trainees in neurosciences still need a sound footing in these older methods in order to apply
a critical approach to their results.
Under the guidance of its founders, Alan Boulton and Glen Baker, the Neuromethods
series has been a success since its first volume published through Humana Press in 1985. The
series continues to flourish through many changes over the years. It is now published under
the umbrella of Springer Protocols. While methods involving brain research have changed a
lot since the series started, the publishing environment and technology have changed even
more radically. Neuromethods has the distinct layout and style of the Springer Protocols
program, designed specifically for readability and ease of reference in a laboratory setting.
The careful application of methods is potentially the most important step in the process
of scientific inquiry. In the past, new methodologies led the way in developing new disci-
plines in the biological and medical sciences. For example, Physiology emerged out of
Anatomy in the nineteenth century by harnessing new methods based on the newly discov-
ered phenomenon of electricity. Nowadays, the relationships between disciplines and meth-
ods are more complex. Methods are now widely shared between disciplines and research
areas. New developments in electronic publishing make it possible for scientists that
encounter new methods to quickly find sources of information electronically. The design of
individual volumes and chapters in this series takes this new access technology into account.
Springer Protocols makes it possible to download single protocols separately. In addition,
Springer makes its print-on-demand technology available globally. A print copy can there-
fore be acquired quickly and for a competitive price anywhere in the world.
v
Preface
fMRI has gained a remarkable role as a tool for studying brain function due to its capability
to provide an invaluable insight into the mechanisms through which the human brain
works in healthy individuals and to explore the mechanisms associated with recovery of
function or clinical deterioration in patients with different neurological and psychiatric
conditions. The utility of this technique to monitor the effects of pharmacologic and reha-
bilitative treatments has also been recently demonstrated.
The second edition of this book aims at providing an up-to-date review of the main
methodological aspects of fMRI, as well as a state-of-the-art summary of the achievements
obtained by its application to the study of central nervous system functioning in the clinical
arena. Future evolutions of fMRI techniques are also discussed. The contributors of this
volume are all worldwide renowned scientists and physicians with a broad experience in the
technical development and clinical use of fMRI. Although the field is ample, based on a
series of very different disciplines and expanding at a dramatic pace every day, I believe that
this book provides an adequate background against which to plan and design new studies
to advance our knowledge on the physiology of the normal human brain and its change
following tissue injury.
Part I of the volume is aimed at providing the basic knowledge for the understanding
of the technical aspects of fMRI. It covers the basic principles of MRI and fMRI, the differ-
ent options that can be used to set up an fMRI experiment, and the steps of fMRI analysis,
from the preparation of data to the achievement of interpretable results. This part is there-
fore essential to introduce the readers to the “fMRI world” and make them able to inter-
pret with enough criticism the results of their own experiments. A chapter is devoted to the
advantages, caveats, and pitfalls of fMRI data acquired using high-field MR scanners. In
addition, although still in its infancy, the assessment of brain connectomics with functional
and structural imaging techniques is considered at length, given its potential for improving
the understanding of normal and pathological brain function.
Part II provides an overview of the main results derived from the application of fMRI
to the study of healthy individuals. Given its noninvasiveness, safety, and repeatability, fMRI
is rapidly replacing, whenever possible, other functional techniques, such as positron emis-
sion tomography, to image the function of the normal brain. In addition, due to its spatial
resolution, fMRI is commonly preferred to neurophysiological techniques to locate with
precision the areas activated during the performance of experimental tasks. What has been
achieved in the analysis of the main human functional systems with fMRI is illustrated,
including, among many other aspects, behavior, language, memory, emotion, sensation,
pain, vision, and hearing.
Part III is more clinically oriented and illustrates the main findings obtained by the
application of fMRI to assess the role of brain plasticity in the major neurological and psy-
chiatric conditions. The first chapter is devoted to fMRI studies of multiple sclerosis, since
there is a growing body of evidence that brain functional reorganization has an important
role, at least in same phases of the disease, in limiting the clinical consequences of MS-related
irreversible tissue damage. Therefore, MS can be viewed as a “model” to understand how
vii
viii Preface
pathology can affect the patterns of brain recruitment. The results obtained in other white
matter conditions, including isolated demyelinating myelitis and vasculitides, are then pre-
sented. The second chapter deals with stroke studies, which have shown consistently that
reorganization of surviving neuronal networks is one of the key factors underlying recovery
of function. The experimental caveats to be faced when studying patients with severe clini-
cal impairment are also reviewed. The following chapters cover psychiatric and neurode-
generative diseases, a field where fMRI is providing important pieces of information not
only for the understanding of the mechanisms underlying disease pathophysiology and
genesis of symptomatology, but also for planning and monitoring novel treatment strate-
gies. Then, two conditions, i.e., epilepsy and tumors, where fMRI is gaining an important
role in the presurgical evaluation of patients, are discussed. The last contribution of this
part describes the potential and some preliminary, but nevertheless promising, results on
the use of fMRI in the monitoring of pharmacological treatments and motor
rehabilitation.
Part IV is a glimpse into the future and presents novel approaches for the integration
of fMRI data with measures of damage assessed using structural MR techniques and the
application of fMRI to image spinal cord function. Finally, results derived from the applica-
tion of graph analysis to assess network abnormalities in patients with several neurological
and psychiatric disorders, including dementia, amyotrophic lateral sclerosis, multiple scle-
rosis, and schizophrenia, are presented.
The hope that has inspired this book is that it will be of help to clinicians and research-
ers in their daily life activity by providing a “user-friendly” summary of the field and the
necessary background against which to plan and carry out future and successful studies.
This is, indeed, an ever-growing and exciting field of research, where we have reached a lot
in the past few years, but where there is still a long journey ahead of us.
ix
x Contents
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 905
Contributors
xi
xii Contributors
EWALD MOSER • High Field MR Centre, Medical University of Vienna, Austria; Division
MR-Physics, Center for Medical Physics and Biomedical Engineering, Vienna, Austria
KEVIN MURPHY • School of Psychology, Cardiff University, Cardiff, Wales, UK
THOMAS E. NICHOLS • Oxford University Centre for Functional Magnetic Resonance
Imaging of the Brain (FMRIB), John Radcliffe Hospital, Oxford, UK
CHRISTOPHER NIMSKY • Neurochirurgische Klinik, Klinikum Stuttgart, Stuttgart,
Germany
GUY A. ORBAN • Dipartimento di Neuroscienze, Universita degli Studi di Parma, Parma,
Italy
ASPASIA ELENI PALTOGLOU • MRC Institute of Hearing Research, Nottingham, UK
SCOTT PELTIER • University of Michigan Functional MRI Laboratory and Biomedical
Engineering Department, University of Michigan, Ann Arbor, MI, USA
MARTIN PEPER • General and Biological Psychology Section, Faculty of Psychology,
Philipps-Universität Marburg, Marburg, Germany
ROBERT H. POWELL • Department of Clinical and Experimental Epilepsy, Institute of
Neurology, University College London, London, UK
PERRY F. RENSHAW • Brain Institute, University of Utah, Salt Lake City, UT, USA
SIMON ROBINSON • High Field MR Centre, Medical University of Vienna, Austria;
Department of Biomedical Imaging and Image-guided Therapy, Medical University
of Vienna, Austria
MARIA A. ROCCA • Neuroimaging Research Unit, Institute of Experimental Neurology,
Division of Neuroscience, San Raffaele Scientific Institute and Vita-Salute San Raffaele
University, Milan, Italy; Department of Neurology, Division of Neuroscience, San
Raffaele Scientific Institute and Vita-Salute San Raffaele University, Milan, Italy
STEPHEN M. SMITH • Oxford University Centre for Functional Magnetic Resonance
Imaging of the Brain (FMRIB), John Radcliffe Hospital, Oxford, UK
PATRICK STROMAN • Department of Diagnostic Radiology, c/o Center for Neuroscience
Studies, Queen's University, Kingston, ON, Canada; Department of Physics, c/o Center
for Neuroscience Studies, Queen's University, Kingston, ON, Canada
JING SUI • Brainnetome Center and National Laboratory of Pattern Recognition, Institute
of Automation, Chinese Academy of Sciences, Beijing, China; The Mind Research
Network and Lovelace Biomedical and Environmental Research Institute, Albuquerque,
NM, USA
RACHEL C. THORNTON • Department of Clinical and Experimental Epilepsy, Institute of
Neurology, University College London, London, UK
ARTHUR W. TOGA • USC Mark and Mary Stevens Neuroimaging and Informatics
Institute, Keck School of Medicine of USC, University of Southern California,
Los Angeles, CA, USA
INDRE V. VISKONTAS • Department of Neurology, Memory and Aging Center, University of
California, San Francisco, CA, USA
NICK S. WARD • Wellcome Trust Centre for Neuroimaging, Institute of Neurology, University
College London, London, UK
MARK W. WOOLRICH • Oxford University Centre for Functional Magnetic Resonance
Imaging of the Brain (FMRIB), John Radcliffe Hospital, Oxford, UK
Part I
Abstract
This chapter describes the basics of magnetic resonance imaging (MRI) and functional MRI (fMRI). It is
aimed at beginners in the field and does not require any previous knowledge. Complex technical issues are
made plausible by presenting plots and figures, rather than mathematical equations.
The part dealing with the basics of MRI covers spins, spin alignment in external magnetic fields, the
magnetic resonance effect, field gradients, frequency encoding, phase encoding, slice selection, k-space,
gradient echoes, and echo-planar imaging.
The part dealing with fMRI covers transverse relaxation times, the basics of the blood oxygen level-
dependent (BOLD) contrast, and the hemodynamic response.
Key words Spin, Field gradients, Frequency encoding, Phase encoding, Slice selection, k-Space,
Gradient echo, Echo-planar imaging, Transverse relaxation time, Blood oxygen level-dependent
1.1 Spins in an The first question arising is “What do we actually see in MRI?”
External Magnetic In general, we see protons. A proton is the nucleus of the hydro-
Field gen atom. Hydrogen is the most common element in tissue, so if we
are able to detect the presence of protons and display them with a
certain spatial resolution, it is fair to say that we can “see” tissue.
The detection of protons is based on a physical property
called the “spin.” A correct description of spins is only possible
with quantum mechanics and would be beyond the scope of this
book, so may it suffice to say that a spin is quite similar to a com-
pass needle. In particular, a compass needle carries a “magnetiza-
tion” which enables it to align in an external magnetic field and
produce a magnetic field itself (for example, a compass needle
can be used to attract small iron particles). A spin possesses an
elementary magnetization (albeit a tiny one), so it behaves in a
similar way.
Let us consider a simple object containing protons, for example a
container with water (Fig. 1). If there is no external magnetic field
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_1, © Springer Science+Business Media New York 2016
3
4 Ralf Deichmann
Fig. 1 (Left) Without an external magnetic field (labeled B) the spins point in dif-
ferent directions, the contributions of their respective magnetization vectors can-
cel out, so there is no net magnetization (labeled M). (Right) Inside an external
magnetic field B the spins align in parallel or anti-parallel direction, with a slight
majority of spins in parallel direction, giving rise to a net magnetization M
(usually labeled B), the spins will point in different directions, the con-
tributions of their respective magnetization vectors will cancel out, so
there is no net magnetization (usually labeled M). If, however, this
object is placed into an external magnetic field B (e.g., into the bore
of an MR scanner), the spins will align. According to the laws of quan-
tum mechanics, this alignment is either parallel or anti-parallel to the
external magnetic field, so once again one might assume that the sin-
gle magnetization vectors will cancel each other. However, a slight
majority of spins prefers the parallel direction. This results in a macro-
scopic net magnetization M which is parallel to B (Fig. 1, right). The
idea is now: any measurement of M would correspond to the detec-
tion of the presence of protons in the object. If M is measured with a
spatial resolution, we can display the result as an image. This is exactly
what is done in MR imaging. The measurement of M is based on a
physical effect which will be described in Subheading 1.2.
1.2 The Larmor Let us assume that we disturb the realigned spins in a way that (at
Precession least for a short time) the magnetization vector M is not parallel to
B but tilted by a certain angle (how this can be achieved will be
discussed in Subheading 1.3). In this case, an interesting process
begins: the tilted magnetization vector rotates around the direction
of the external magnetic field. This movement, which resembles
closely the behavior of a spinning top, is called precession (Fig. 2).
The frequency is called Larmor frequency. It is important to note
that the Larmor frequency f is proportional to the external field B.
f = g B. (1)
In this formula, γ is the gyromagnetic ratio with a value of
42.58 MHz/T for hydrogen.
Principles of MRI and Functional MRI 5
Fig. 2 Behavior of the tilted magnetization: Precession with the Larmor frequency
1.3 The Magnetic The next physical effect is more or less the opposite of the one that
Resonance Effect has just been discussed. This time, the object is exposed to an exter-
nal RF signal which has exactly the Larmor frequency and is pro-
duced by a kind of “inbuilt FM broadcasting station” (Fig. 3, part 1).
This signal will tilt the magnetization which subsequently starts to
precede (Fig. 3, part 2). During precession, an RF signal is being sent
out which can be detected with a kind of “FM tuner” (Fig. 3, part 3).
It should be noted that this concept only works if the incoming
RF signal has exactly the spins’ Larmor frequency. Otherwise, the
magnetization will not be tilted and it is impossible to detect a signal.
Thus, we are dealing with a resonance effect, and this explains why
the imaging technique based on this effect is called magnetic reso-
nance imaging (MRI). The nuclear magnetic resonance effect was
first described independently by Bloch and Purcell in 1946 [1, 2].
We have now covered all the basic physical effects that are
required to understand how MRI works. To summarize it, all MR
experiments are based on the following principles:
● Put the object to be imaged into a strong external magnetic
field B. The spins will align and create a net magnetization M
which is parallel to B.
● Knowing B, calculate the Larmor frequency f = γB and send an
external RF pulse which has exactly this frequency. This will tilt
the magnetization. After a short time, the external RF has
served its duty and can be switched off.
6 Ralf Deichmann
Fig. 3 An external radiofrequency (RF) pulse that has exactly the Larmor fre-
quency tilts the magnetization (1). The tilted magnetization starts to precede (2)
and sends out an RF pulse with the Larmor frequency itself (3)
2 A One-Dimensional MR Experiment
Fig. 4 (a) Two glasses of water inside a closed box. (b) Magnetization inside both glasses after placing the box
inside an external magnetic field. (c) Reaction to an external radiofrequency (RF) pulse: the magnetization vec-
tors in both glasses start to precede and send out RF signals with the Larmor frequency. (d) Effect of an exter-
nal field gradient: the magnetic field strength at the position of both glasses is different, so radiofrequency
signals with different Larmor frequencies are being sent
Fig. 6 Frequency spectrum resulting from the Fourier transform. The frequency
corresponds to the position of the originating spin
Fig. 7 Schematic sketch of three MR experiments: in the absence of any gradients, a long signal can be
observed (1). In the presence of a gradient, the signal decays more rapidly (2). If the gradient is inverted, a
gradient echo occurs (3)
The third experiment (Fig. 7, part 3) starts off like the second
one, resulting in the same fast signal decay. However, after a while
the gradient is inverted (a negative gradient Gx means that the
magnetic field strength decreases in x-direction, rather than
increasing). This leads to a striking phenomenon: the signal which
seemed to have disappeared completely, suddenly comes back.
This effect is called gradient echo.
It is relatively simple to explain these effects. For illustration,
Fig. 8a shows the precession of four individual spins at different
positions in the first experiment after the initial RF pulse (the respec-
tive magnetization vectors are seen “from top”). Although the spins
are located at different positions, they are exposed to the same mag-
netic field because there are no gradients. Thus, they precede with
the same Larmor frequency; their magnetization vectors are always
parallel and add up to a relatively strong net magnetization.
In theory, this should go on forever. In practice, the signal decays
due to transverse relaxation effects which will be discussed later.
Figure 8b shows the respective sketch for the second experi-
ment. Due to the field gradient, the spins are exposed to different
field strengths and precede with different Larmor frequencies. After
a relatively short time, they are completely dephased, i.e., the mag-
netization vectors cancel each other and there is no net magnetiza-
tion. The result is a fast signal decay, as depicted in Fig.7 (part 2) .
Fig. 8 (continued) dephase, reducing the duration of the signal. (c) Explanation of the third experiment: the
gradient inversion leads to a change from anti-clockwise to clockwise rotation, so spins rephase. A strong
signal, the so-called gradient echo can be observed when the magnetization vectors are parallel again
Fig. 8 (a) Explanation of the first experiment: spins at different positions still have the same Larmor frequencies
as the magnetic field is homogeneous. As a consequence, their magnetization vectors remain parallel and sum
up to a strong net magnetization over a relatively long time. (b) Explanation of the second experiment: spins
at different positions have different Larmor frequencies due to the field gradient. Their magnetization vectors
12 Ralf Deichmann
Fig. 9 Signal losses due to the finite gradient rise time: by the time the gradient
has reached its full amplitude, the signal has decayed considerably
Fig. 10 Solution of the problem imposed by the finite gradient rise times: the
initial negative gradient creates a gradient echo and thus a strong signal at a
time when the read gradient has reached its full amplitude. The maximum of the
echo occurs when the shaded areas are identical
Fig. 11 If the gradient axes are chosen as shown on the left-hand side, the result of the one-dimensional
experiment with a read gradient in x-direction is a profile in anterior/posterior direction
5 The k-Space
Fig. 12 Definition of a data point’s k-value as the area under the gradient before
the data point is sampled
Fig. 13 Gradient echo experiment: the data points constituting the gradient echo
have increasing k-values, ranging from a negative to a positive value. The k-value
is zero for the central data point where the shaded areas are identical and the
echo has maximum amplitude
Fig. 14 Definition of a data point’s kx- and ky-values as the areas under the
respective gradients before the data point is sampled
6 Two-Dimensional Acquisition
Fig. 16 The basis of two-dimensional imaging: several data points with different
combinations of kx- and ky-values have to be sampled, filling the two-
dimensional k-space
Principles of MRI and Functional MRI 17
7 Slice-Selective Excitation
Fig. 19 If the gradient axes are chosen as shown on the left-hand side, the result
of the two-dimensional experiment with a read gradient in x-direction and a
phase gradient in y-direction is an image in the axial plane, showing an overlay
of all axial slices
Fig. 20 The basis of slice selective excitation: due to the slice gradient Gz spins
have spatially dependent Larmor frequencies. A radiofrequency pulse with fre-
quency f0 can only excite spins whose Larmor frequency corresponds to f0. These
spins are located in a plane perpendicular to the gradient direction
Fig. 22 If the gradient axes are chosen as shown on the left-hand side, the result
of the three-dimensional experiment with a read gradient in x-direction, a phase
gradient in y-direction, and a slice gradient in z-direction is an image in the axial
plane with a finite slice thickness
8 Echo-Planar Imaging
Fig. 23 Acquisition of multiple gradient echoes by successive read gradient inversion. All echoes cover the
same k-values, but the order is reversed when comparing even and odd echoes
Fig. 24 The basis of echo planar imaging: due to intermediate “blips” (shaded gradient pulses) the echoes have
increasing ky-values, thus covering different lines in k-space
so-called blip, shaded in Fig. 24). The ky-value of the second echo is
determined by the sum of the areas under the initial phase gradient
and this blip, so ky is increased and the second echo covers another
horizontal, slightly “higher” line in k-space in reverse direction. This
concept of intermediate blips is maintained throughout the remain-
ing acquisition, resulting in a meander-like journey through k-space.
22 Ralf Deichmann
Fig. 25 A complete echo planar imaging experiment. The central echo which
covers the center of k-space has the highest amplitude
Fig. 27 Signal decay in compartments with a long T2* value (solid line) and a short
T2* value (dashed line). If a short echo time (TE) is chosen (1), there is a high
signal amplitude, but hardly any contrast between both compartments. For an
intermediate TE (2), there is a good contrast and still a sufficient signal ampli-
tude. For a long TE (3), the signal in both compartments has decayed
Fig. 28 T2* weighted brain images acquired with echo time (TE) values of 10, 50, and 200 ms. At 10 ms, con-
trasts are relatively low. At 50 ms, a nice T2* contrast can be observed (arrow). At 200 ms, signal has mostly
decayed and only cerebrospinal fluid is visible
Fig. 29 Change of physiological parameters, the concentration of deoxyhemoglobin, and the T2* weighted
signal amplitude in response to neuronal activation
Fig. 31 Definition of echo time (TE) for an echo-planar imaging (EPI) sequence
a wasted waiting time (as one might have expected from Fig. 26),
but can be used for the acquisition of the first half of the echo train,
being another reason why EPI allows for a high temporal resolu-
tion in fMRI experiments.
References
1. Bloch F, Hansen WW, Packard M (1946) 5. Ogawa S, Lee TM, Nayak AS, Glynn P (1990)
Nuclear induction. Phys Rev 69:127 Oxygenation-sensitive contrast in magnetic res-
2. Purcell EM, Torrey HC, Pound RV (1946) onance image of rodent brain at high magnetic
Resonance absorption by nuclear magnetic fields. Magn Reson Med 14:68–78
moments in a solid. Phys Rev 69:37–38 6. Kwong KK, Belliveau JW, Chesler DA et al
3. Lauterbur PC (1973) Image formation by (1992) Dynamic magnetic resonance imaging
induced local interactions: examples employing of human brain activity during primary sensory
nuclear magnetic resonance. Nature stimulation. Proc Natl Acad Sci USA 89:
242:190–191 5675–5679
4. Mansfield P (1977) Multiplanar image forma- 7. Buxton RB, Uludag K, Dubowitz DJ, Liu TT
tion using NMR spin echoes. J Phys C (2004) Modeling the hemodynamic response
10:L55–L58 to brain activation. NeuroImage 23:S220–S233
Chapter 2
Abstract
The chapter gives an overview of peripheral devices commonly used in fMRI experiments, and it addresses
the principles, performance aspects, and specifications of fMRI hardware. The general guidelines for
MR-compatible hardware are also discussed. The target audience is quite broad and mathematical descrip-
tions are kept to a minimum and qualitative descriptions are favored whenever possible.
Key words Functional MRI, Hardware, Peripheral devices, MRI, Multimodal acquisition,
Neuroimaging
1 Introduction
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_2, © Springer Science+Business Media New York 2016
29
30 Luis Hernandez-Garcia et al.
Fig. 1 These photos were taken during the installation of an MRI scanner at Resurgens Orthopaedics in Atlanta,
GA on the 19th floor of their Crawford facility. Courtesy of Resurgens Orthopaedics, Atlanta, GA
one must find a site with sufficient space for the scanner. Access is
also important, as the main magnet is typically delivered and
installed in one piece. Furthermore, MRI magnets are filled with
cryogens, which are delivered in large dewars. Thus there must be
a wide path to the loading dock that avoids stairs and is clear of
obstacles. Typically, MRI scanners are housed in basements or
ground floors near the loading docks of hospitals, although excep-
tions exist, as the one shown in Fig. 1.
Next one must consider how the magnetic field will affect its sur-
roundings. Before the magnet is installed, one must consider how far
the magnetic field will extend. Most people who work with MRI are
familiar with the enormous forces that the main magnetic field can
exert on objects in its proximity (i.e. the magnet room). Modern MRI
scanners are actively shielded and contain the magnetic field fairly well
within the magnet room. Even with shielding, sometimes a subtle but
significant magnetic field can extend beyond the walls of the magnet
room. Thus it is important to keep in mind two questions: how well
the magnetic field is contained, and what sort of equipment is in the
rooms adjacent to the magnet room. The first question is usually
answered in terms of the location of the “5 Gauss line.”
The United States FDA regulates that the general public (any-
one not working with an MRI scanner or being scanned) not be
exposed to static magnetic fields over 5 G (5 × 10−4 T), and thus the
5 G boundary must be contained inside the magnet room. MRI
scanner vendors will typically provide contour plots of the m agnetic
field superimposed on the blueprints of the room and provide con-
sultation on the location of the scanner. One must realize, however,
that smaller magnetic fields will extend beyond the walls of the
32 Luis Hernandez-Garcia et al.
Fig. 2 A passively shielded 4.0 T magnet encased in a hexagonal iron cage to partially contain the main mag-
netic field
Introduction to Functional MRI Hardware 33
Active Shielding
15
Original B field
B field after shielding
10
Magnetic field (T esla)
-5 Operating region
(inside of bore)
-10
-15
-300 -200 -100 0 100 200 300
Distance from the center of the bore
Fig. 3 A plot of the magnetic field strength along the radial direction. The thick vertical lines denote the location
of the coil’s windings. The thinner line denotes the operating region where the sample is placed. Ideally, this
region should have a flat magnetic field, and additional fields (shims) are necessary
34 Luis Hernandez-Garcia et al.
Fig. 4 The streaks in the image are caused by the presence of electromagnetic
frequency noise at a single frequency. This is typically introduced by the pres-
ence of badly shielded electronic equipment in the room. The AC power supply
running at 60 Hz to the device is in this case the culprit of the artifact
sufficient to let RF noise into the room that ruins images, so test-
ing of the room shielding must be stressed. Figure 4 shows an
example of the effects of an RF noise source on an MR image
Good shielding of the room is not enough. MR-related
electronic equipment produces RF noise and can act as an
antenna that passively carries RF noise from the outside into the
room. There are also a number of peripherals that are needed
for functional MRI in order to provide stimulation and record
data from the subject (see Sect. 4). It is preferable to keep all
electronics out of the magnet room, but if the electronic equip-
ment must be inside, it must be tested thoroughly for RF noise.
If it is indeed noisy, care must be taken to shield the equipment
to prevent image artifacts (Copper mesh is very useful for build-
ing RF shields).
Consider the case of a button response box. Typically, one
keeps the bulk of the electronic components out of the magnet
room, but the buttons themselves must be in the scanner and they
need to communicate with the response recording electronics.
Even if fiber-optic technology is used to carry signals into the
room, RF noise can enter through the same opening as the fiber
optic cabling. The solution is to build a “penetration panel” into
the shield. This is a panel on the wall that is outfitted with wave-
guides and filtered connectors. The role of a waveguide is to block
any electromagnetic radiation that is not parallel with the direction
of the waveguide, and at the same time, to guide the EM waves
Introduction to Functional MRI Hardware 35
Fig. 5 On the left is the penetration panel that connects the MRI scanner electronics to the magnet hardware.
The image on the right is a second penetration panel used for all the additional stimulus/response equipment
needed for fMRI
degraded and thus they require its own cooling system to keep
them stable. This system is usually a cooling loop involving a water
chiller that can remove about 14,000 BTU per hour.1
It is also noteworthy that the scanner’s performance can be
affected by other unexpected environmental factors. Outside mag-
netic fields and vibrations can be an issue. For example, nearby
construction can produce vibrations that will affect the MR signal’s
stability if the floor is not adequately mechanically damped. It is
thus important to carry out vibration tests of the site before instal-
lation proceeds. Large moving objects, such as nearby trains can
also generate magnetic fields that affect the scanner’s stability [4]
One other issue that can cause a great deal of grief to investiga-
tors is the production of small electromagnetic spikes inside the
magnet room. These occur when metal objects in the room vibrate
(typically because of the gradients) and bang against each other or
when arcing occurs across badly-soldered connections in home-
made equipment. The RF receiver hardware is sensitive enough to
pick up these spikes. Spikes in the k-space data translate into stripe
patterns in the images (Fig. 6). It is thus very important to make
sure that all metal equipment is well secured.
A useful option to consider for a functional MRI lab is a mock
MRI scanner. This is advantageous since it allows subjects to get
1
These numbers are based on the specification of a 3 T scanner by General
Electric (MR750).
Introduction to Functional MRI Hardware 37
M1
M2
M1 - Bore Inner diameter
M2 – Head coil Inner diameter
M3 distance from isocenter to edge of bore
M3
3.1 The Magnet We will begin by considering the magnet. Typical MRI magnets,
like the one shown in Fig. 9, are large solenoid coils made of super-
conducting metal (niobium alloys, typically). They are kept cooled
at approximately 4 K by liquid Helium in order to achieve and
maintain superconductivity.2 The magnet is “ramped up” to field
by introducing a current through a pair of leads that produces the
desired magnetic field. Once the specified current has been built
up, the circuit is closed such that the current “re-circulates”
through the coil constantly and there is no need to supply more
power to it. It is crucial to maintain the low temperature to prevent
the coil from resisting the current flow.
A new, exciting development is the “dry magnet.” The term
refers to super-conducting magnets that are cooled without liquid
2
For more information on superconductivity, see [5]
Introduction to Functional MRI Hardware 39
sensors that sound an alarm when the oxygen level falls below safe
levels. It must be stressed that all personnel be trained in emer-
gency quench procedures in case the emergency systems fail.
Having considered what can happen when the magnet fails, let us
now return to the more cheerful subject of what the magnet can do.
The key parameter in the magnet is its field strength (B0), as it
determines many properties of the images. Primarily, field strength
determines the amount of spins that align with and against the
field. The higher the field strength, the larger the population of
aligned spins that can contribute to the MR signal. More specifi-
cally, the population of spins aligned with the magnetic field (n+)
and against it (n−) is described by the Boltzmann equation
æ n+ ö æ g hB0 ö
ç ÷ = exp ç ÷ (1)
è n- ø è kT ø
v 0 = g B0 (2)
1 ì tc 4t c ü
aí + 2 2 ý
(3)
T1 î1 + w0 t c 1 + 4w0 t c þ
2 2
and
1 ì 5t c 2t c ü
a í3t c + + 2 2 ý
(4)
T2 î 1 + w0 t c 1 + 4w0 t c þ
2 2
1 1 1 1
= + = + gDB0 (5)
T2 * T2 T2 ¢ T2
B0¢ = B0 (1 - c ) (6)
Fig. 10 Illustration of the dielectric effect on a high-resolution, T1 weighted, SPGR image. The center of the
image appears brighter, because of the formation of standing waves in the RF pattern. As a result, the center
of the field of view receives a higher flip angle than the periphery of the image. The image on the right has
been corrected by removing the low frequency spatial oscillation with a 2D FIR filter
based on subtle signal changes over time and therefore, drifts act as
significant confounds, especially in slow paradigms. Statistical and
signal processing tools do exist to reduce these drifts effects, but it
is much more desirable that they be reduced during acquisition.
Unfortunately, there are many sources of drift in the MRI hard-
ware, so it is important that the magnet undergo extensive stability
testing before it becomes operational and that quality control tests
including stability measurements be performed regularly. The
scanner’s stability can be measured on a phantom over a small
region of interest. Figure 11 illustrates a typical stability test.
The physical configuration and shape of the magnet also plays
an important role in many of these parameters. While “open” MRI
systems exist and are used for large or claustrophobic subjects,
their field strength is typically not sufficient for functional MRI
applications and their use is limited to clinical applications that do
not demand high-quality imaging. Among the closed bore sys-
tems, one can choose between short and long bore systems. Short
bore systems are intended for head-only applications and can
sometimes offer improved performance over smaller regions. Long
Fig. 11 A typical stability test showing the time course and its frequency content in a phantom (above)
44 Luis Hernandez-Garcia et al.
m0 IdL ´ r
B (r ) = ò (7)
4p r
3
3.2 Magnetic Field In order to produce images, an MRI scanner needs a spatially varying
Gradients magnetic field (see image reconstruction chapter) under tight control
by the user. This is accomplished by using an additional set of coils
that add extra magnetic fields to the main field. A set of such coils is
shown in Fig. 12. By supplying customized current waveforms to
these coils, the user can change the distribution of the magnetic
field’s shape at will. In broad terms, by varying gradient’s strength
can be varied over time during the pulse sequence, one can obtain
MR signals whose phase distribution is a function of the spatial distri-
bution of the sample.
The ideal gradient set is capable of quickly changing the mag-
netic field as a linear function of spatial location along each of the
Cartesian axes. Typical gradients in clinical and functional MRI are
between 10 to 40 mT/m, but specialized gradient inserts exist that
can produce larger gradients (in the range of 100 mT/m). Small
bore animal systems can be outfitted with more powerful gradients
(up to approximately 400 mT/m). The main challenges in MRI
gradient design and construction usually consist of producing lin-
ear gradients in space and time, and the production of eddy cur-
rents. Motivated by the need to achieve finer spatial resolution and
better axon fiber tracking through diffusion tensor images (DTI) ,
Massachusetts General Hospital has developed a high-performance
gradient insert that can achieve up to 300 mT/m in a human sys-
tem, whereas standard clinical gradients rarely exceed 50 mT/m.
This system is utilized at present primarily for experiments con-
cerning the “Human Connectome Project” [12] (www.human-
connectomeproject.org).
The spatial linearity of the gradients must be maintained over
the volume of the sample to be imaged, or the images will appear
warped (although these distortions can be corrected during recon-
struction if the true shape of the gradient is known). The spatial
Fig. 12 Gradient coils from Doty Scientific (reproduced with permission of Doty Scientific)
46 Luis Hernandez-Garcia et al.
Input Current
Uncompensated
Compensated
Output Gradient
Uncompensated
Compensated
Fig. 15 (a) Transmit/receive birdcage head coil. (b)Receive-only quadrature occipital coil (c) 8-channel phased-
array coil
Fig. 16 Comparison of fMRI images obtained using conventional (a) and parallel imaging (b) in the inferior
brain. In this example, parallel imaging with an 8-channel phased-array coil was used to reduce data readout
time by a factor of 2. This reduced signal loss and image distortions due to susceptibility, particularly in the
region indicated by the arrows. (Courtesy of Yoon Chung Kim, University of Michigan Functional MRI Laboratory)
3.4 Computing Functional MRI experiments generate large datasets. The raw data
Resources alone for a single subject, scanned for one hour, weighs in between
100 and 200 MB. Combine this with the space needed for image
reconstructions and analysis, and one should budget a storage and
scratch space of around 2 GB for every hour of scanning. In this
section, we will provide some guidelines and suggestions on how
to set up a computing environment to handle all this data. The two
major factors influencing the design of your environment will be
(1) money, and (2) the expertise available, in terms of computer
systems and MRI data processing.
The data stream in a typical laboratory consists of four main
stages. The first stage is the MRI scanner, which produces either
raw, unreconstructed data, or reconstructed images that are ready
for post-processing and analysis. If a lab has an MR physicist at its
disposal, then the former case is often true, since an MR physicist
may develop image reconstruction codes that provide improved
image quality over vendor-provided software, and that can add in
improved reconstruction techniques as they are developed.
Assuming images have been reconstructed, the second stage con-
sists of slice timing correction, realignment, coregistration, warp-
ing, and smoothing, which are all operations that prepare the
dataset for statistical modeling and analysis. The third stage is sta-
tistical modeling and analysis. The fourth stage is data backup,
though it is advisable to make backups of data at more than one
point in the stream.
The majority of fMRI labs maintain one powerful workstation
or computer cluster to do the bulk of their processing. Users can
log into this computer from their own machines to initiate process-
ing or view and download the preprocessed data for local analysis.
The central advantage to this model is that the large and compli-
cated software packages used to process fMRI need only be main-
tained on one machine, which greatly simplifies maintenance. A
secondary advantage is that this model allows users greater flexibil-
ity in choosing the operating system of the computers they use; the
52 Luis Hernandez-Garcia et al.
5 Subject Monitoring
5.1 Scanner In order to match the recorded external signals with the fMRI data
Synchronization being recorded, synchronization with the start of the scan must be
achieved. This can be done using a TTL pulse to/from the scanner
from/to the external device or recording media. For instance, a
logic pulse from the MR scanner to the computer recording physi-
ological noise can be set to trigger the recording sequence.
Commercial MR scanners from the main vendors (GE, Siemens,
Philips) all have the capability to send or receive TTL sync pulses.
5.3 Cardiac Monitoring the cardiac waveform can be achieved in several ways.
Monitoring The most common solutions are pulse oximeters or ECG patches.
The primary cardiac harmonic frequency lies in the 0.5–2.0 Hz
range, with both the first and secondary harmonics shown to affect
the fMRI signal [31].
Pulse oximetry refers to indirectly monitoring the oxygen lev-
els in the extremities to monitor the cardiac waveform. This is most
often accomplished in fMRI labs by using an LED and photodiode
that clips to the subject’s finger, connected to a data acquisition
board (see Fig. 17). Several MR scanner vendors offer this as part
of the MR system (GE, Siemens), and stand-alone monitoring
units from commercial vendors are also available (Invivo, Biopac).
Normal setup with compliant subjects allows adequate sampling of
cardiac rhythm, as seen in Fig. 18. Drawbacks include the fact that
2000
1500
1000
500
-500
0 5 10 15 20 25 30
Fig. 18 Cardiac waveform acquired during an fMRI scan. (Data acquired on a 3.0 T GE scanner, using a pulse
oximeter with a sampling rate of 40 Hz)
Introduction to Functional MRI Hardware 57
QRS
Complex
ST
PR Segment
T
P Segment
PR Interval
Q
S
QT Interval
Fig. 19 Example of ECG patch (Courtesy of Invivo, www.invivocorp.com). With a schematic of a typical QRS
waveform
subject motion may corrupt the signal, and motor tasks may be
impeded with the oximeter placed on the finger (alternative place-
ment on the ear or toe is possible).
ECG patches located over the heart allow high-fidelity monitor-
ing of cardiac electrical activity. This allows identification of features
beyond the simple cardiac peaks, such the QRS complex during the
depolarization of the ventricles (Fig. 19). Disadvantages of ECG
recording include increased setup complexity, and patient comfort.
5.4 Respiratory The respiratory rhythm has a normal frequency range of 0.1–0.5 Hz.
Monitoring Motion of the chest during respiration combined with the changes
in oxygen saturation in the lungs lead to a modulation of the local
magnetic field that can affect the phase of the MR signal at the posi-
tion of the head during scanning (Fig. 20), which can lead to modu-
lation of the recorded MR signal intensity. Mitigation of respiratory
effects on the MR signal can include scanning during breath-holds,
modified pulse sequences to sample and account for the modulation
in magnetic field [32] and recording of the respiratory signal to use
as nuisance covariates in post-processing analysis [33].
Monitoring respiratory rhythm can be accomplished by using
a plethysmograph (pressure belt) around the waist of a subject, like
the one shown in Fig. 21, or by using a nasal cannula to monitor
expired CO2 concentration. A sample respiratory waveform is
shown in Fig. 22.
5.5 Galvanic Skin Galvanic skin response (GSR) is a measure of the electrical resistance
Response (GSR) of the skin, a physical property that has been shown to increase in
response to subject arousal, mental effort, or stress. It is monitored
58 Luis Hernandez-Garcia et al.
Fig. 20 Phase difference between inspiration and expiration for a coronal slice
3500
3000
2500
2000
1500
0 10 20 30 40 50 60
Fig. 22 Respiratory waveform acquired during an fMRI scan. (Data acquired on a 3.0 T GE scanner, using a
pulse oximeter with a sampling rate of 40 Hz)
Introduction to Functional MRI Hardware 59
5.6 Head Motion Severe head motion during an fMRI scan can severely corrupt the
Tracking data. Besides minimizing patient motion using cushioning and
restraints, a measure of head motion may also be collected to help
correct data in post-processing. This may be done using modified
pulse sequences (PACE), or by using external motion tracking
[33–35]. Again, patient comfort and visual path should be taken
into consideration.
5.7 Eye Tracking Patient gaze and fixation time is important for several types of
fMRI paradigms and pathological types. In addition, eye motion
can be a source of variance in fMRI scans [36]. Thus, tracking eye
position can be desirable. The common method is to monitor the
position of infrared (IR) light that is reflected off the eye of the
subject. This involves transmitting IR light to the subject’s eye,
and then recording it, using MR-compatible equipment. Several
vendors provide hardware solutions including both long-range and
short-range cameras. These systems typically have on the order of
0.1–1° spatial resolution and accuracy, with working ranges of
10–25° horizontally and vertically, and 60–120 Hz sampling rate.
In acquiring a physical eye-tracking system for an fMRI lab,
consideration should be given to the optical path for the eye-
tracking system, taking into account the MR bore, head coil, and
visual stimulus presentation system; signal integrity of the MR
data; ease of setup for the MR techs and scanners. This will also
involve peripheral equipment located in the scanner room or con-
trol room: usually a camera, power supply, video monitor for real-
time display of the subject’s eye, and a PC.
6 Multimodal fMRI
Fig. 23 EEG recorded during fMRI acquisition, before (top) and after (bottom) MR artifact correction
Fig. 25 MRI compatible TMS coil and holder apparatus (image courtesy of Dr.
A. Thielscher of the Max Plank Institute for Biological Cybernetics http://www.
kyb.mpg.de/)
64 Luis Hernandez-Garcia et al.
7 Conclusions
References
1. Shellock FG (2002) Reference manual for mag- 10. Tropp J (2004) Image brightening in samples
netic resonance safety, implants, and devices. of high dielectric constant. J Magn Reson
Saunders, Oxford, UK 167:12–24
2. Shellock FG, Crues JV 3rd (2002) MR safety 11. Schneider E, Glover G (1991) Rapid
and the American College of Radiology white in vivo proton shimming. Magn Reson Med
paper. AJR Am J Roentgenol 178:1349–1352 18:335–347
3. Train JJ (2003) Magnetic resonance compat- 12. Dylan Tisdall M, Witzel T, Tountcheva V,
ible equipment. Anaesthesia 58:387, Author McNab JA, Adad JC, Kimmlingen R, Hoecht
reply 387 P, Eberlein E, Heberlein K, Schmitt F, Thein
4. Durand E, van de Moortele PF, Pachot-Clouard H, Wedeen Van J, Rosen BR, Wald LL (2012)
M, Le Bihan D (2001) Artifact due to B0 fluc- Improving SNR in high b-value diffusion
tuations in fMRI: correction using the k-space imaging using Gmax = 300 mT/m human gra-
central line. Magn Reson Med 46:198–201 dients, Proc ISMRM 2012
5. Tinkham M (2004) Introduction to supercon- 13. Gach HM, Lowe IJ, Madio DP et al (1998)
ductivity, 2nd edn, Dover Books on Physics. A programmable pre-emphasis system. Magn
Dover Publications, Mineola, NY Reson Med 40:427–431
6. Radebaugh R (2009) Cryocoolers: the state 14. Wysong RE, Madio DP, Lowe IJ (1994) A
of the art and recent developments. J Phys novel eddy current compensation scheme for
Condens Matter 21:164219 pulsed gradient systems. Magn Reson Med
7. Williams DS, Detre JA, Leigh JS, Koretsky AP 31:572–575
(1992) Magnetic resonance imaging of perfu- 15. Mansfield P, Chapman B (1986) Active mag-
sion using spin inversion of arterial water. Proc netic screening of coils for static and time-
Natl Acad Sci U S A 89:212–216 dependent magnetic field generation in NMR
8. Yang QX, Wang J, Zhang X et al (2002) Analysis imaging. J Phys E Sci Instrum 19:540–545
of wave behavior in lossy dielectric samples at 16. Edelstein WA, Kidane TK, Taracila V et al
high field. Magn Reson Med 47:982–989 (2005) Active-passive gradient shielding for
9. Collins CM, Liu W, Schreiber W, Yang QX, MRI acoustic noise reduction. Magn Reson
Smith MB (2005) Central brightening due to Med 53:1013–1017
constructive interference with, without, and 17. Pruessmann KP et al (1999) SENSE: sensitivity
despite dielectric resonance. J Magn Reson encoding for fast MRI. Magn Reson Med
Imaging 21:192–196 42(5):952–962
66 Luis Hernandez-Garcia et al.
47. Rothwell JC (1999) Paired-pulse investiga- SP, Silva MT, Paulus W, Pascual-Leone A
tions of short-latency intracortical facilitation (2005) Anodal transcranial direct current stim-
using TMS in humans. Electroencephalogr ulation of prefrontal cortex enhances working
Clin Neurophysiol Suppl 51:113–119 memory. Exp Brain Res 166(1):23–30
48. Ilmoniemi RJ, Ruohonen J, Karhu J (1999) 55. Dieckhöfer A, Waberski TD, Nitsche
Transcranial magnetic stimulation–a new tool M, Paulus W, Buchner H, Gobbelé R
for functional imaging of the brain. Crit Rev (2006) Transcranial direct current stimu-
Biomed Eng 27:241–284 lation applied over the somatosensory
49. Bastings EP et al (1998) Co-registration of corti- cortex – differential effect on low and
cal magnetic stimulation and functional magnetic high frequency SEPs. Clin Neurophysiol
resonance imaging. Neuroreport 9:1941–1946 117(10):2221–2227
50. Bohning DE et al (1998) Echoplanar BOLD 56. Wagner T, Valero-Cabre A, Pascual-Leone A
fMRI of brain activation induced by concur- (2007) Noninvasive human brain stimulation.
rent transcranial magnetic stimulation. Invest Annu Rev Biomed Eng 9:527–565
Radiol 33:336–340 57. Radman T, Ramos RL, Brumberg JC, Bikson
51. Bohning DE et al (1999) A combined TMS/ M (2009) Role of cortical cell type and mor-
fMRI study of intensity-dependent TMS over phology in subthreshold and suprathreshold
motor cortex. Biol Psychiatry 45:385–394 uniform electric field stimulation in vitro. Brain
52. Bohning DE et al (2000) BOLD-f MRI response Stimul 2:215–228
to single-pulse transcranial magnetic stimulation 58. Antal A et al (2011) Transcranial direct current
(TMS). J Magn Reson Imaging 11:569–574 stimulation over the primary motor cortex dur-
53. Nitsche MA, Paulus W (2000) Excitability ing fMRI. Neuroimage 55(2):590–596
changes induced in the human motor cortex 59. Weber MJ et al (2014) Prefrontal transcranial
by weak transcranial direct current stimulation. direct current stimulation alters activation and
J Physiol 527(Pt 3):633–639 connectivity in cortical and subcortical reward
54. Fregni F, Boggio PS, Nitsche M, Bermpohl F, systems: A tDCS‐fMRI study. Hum Brain
Antal A, Feredoes E, Marcolin MA, Rigonatti Mapp 35(8):3673–3686
Chapter 3
Abstract
In this chapter, we discuss technical considerations regarding pulse sequence selection and sequence
parameter selection that can affect fMRI studies. The major focus is on optimizing MRI data acquisitions
for blood oxygen level-dependent signal detection. Specific recommendations are made for generic 1.5, 3,
and 7 T MRI scanners.
Key words MRI, fMRI, Pulse sequences, Blood oxygen level-dependent (BOLD), Echoplanar
imaging, Spiral imaging, Multiband imaging, Motion
1 Introduction
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_3, © Springer Science+Business Media New York 2016
69
70 Mark J. Lowe and Erik B. Beall
1 1 1
*
= + ¢ (1)
T2 T2 T2
Fig 1 Steady-state MR signal as a function of repetition time for tissue with three
different T1 relaxation times
2.1.1 The Spin Echo The time evolution of MR signal immediately after excitation is
referred to as free induction decay (FID). During the FID, the
processes governing the loss of signal coherence are a combination
of T1, T2, and T2′-related processes. In tissue, T2′ will have a large
effect on the loss of signal. T2′ effects are what is referred to as
reversible processes. The loss of phase coherence from these effects
can be reversed by applying a refocusing RF pulse. Figure 2 illus-
trates the sequence timing, using a pulse sequence timing diagram.
Application of a refocusing RF pulse at a time t after the initial
excitation pulse, will result in a complete refocusing of the revers-
ible dephasing effects at a time 2 t. This is referred to as a Spin
Echo. The time 2 t is usually called TE. The MR signal from a
given pulse sequence can be derived from the Bloch equations. For
a SE acquisition, the MR signal will be given by:
ì æ TE ö ü
æ -TE ö ïï ç TR - 2 ÷ æ -TR ö ïï
SSE µ exp ç ÷ í1 - 2 exp ç - ÷ + exp ç ÷ý (3)
è T2 ø ï ç T1 ÷ è T1 ø ï
ïî è ø ïþ
Fig 2 Pulse sequence diagram of a spin echo acquisition. The envelope of the
signal indicates the FID, while the peak of the echo is modulated by T2 according
to Eq. (3)
Fig 3 Steady-state MR signal as a function of echo time for tissue with three
different T2 relaxation times
2.1.2 The Gradient It is possible to perform the spatial encoding for MRI during the
Recalled Echo FID. An echo can be created by increasing the dephasing through
application of a brief field gradient along a particular direction and
then reversing it while acquiring the signal data. This is called a
gradient recalled echo (GRE), or a field echo. The sequence
74 Mark J. Lowe and Erik B. Beall
Table 1
Approximate relaxation times for gray and white matter at 1.5 T and 3 T
1.5 T 3T
Fig 5 Changes in MR brain signal intensity during the first-pass transit of intrave-
nously administered paramagnetic contrast agent. Triangle (Δ) symbols represent
the time course of signal during photic stimulation and circle (o) symbols represent
the time course during rest (darkness). Reproduced with permission from Ref. [1]
MRI. Gadolinium chelates will not cross the blood brain barrier.
Thus, a bolus injection of such a paramagnetic material will cause
a transient change in the T2 relaxation near arterial blood vessels
that are perfusing brain tissue. If this is done while rapidly acquir-
ing T2*-weighted MR images of the brain region that is active, one
will observe a decrease in the measured signal intensity that is
monotonically related to the volume of gadolinium passing
through. One can infer directly the volume of blood perfusing this
region from the signal decrease.
Functional contrast can thus be obtained in MRI by comparing
the regional perfusion, measured with bolus contrast injection, while
performing a task to that measured while at rest. Figure 5 is an
example of the difference in the MRI signal evolution from the same
brain region in visual cortex while undergoing photic stimulation
and in darkness. One can see that the area under the curve for photic
stimulation is larger than that for rest, indicating that the volume of
blood perfusing the tissue was increased during stimulation.
76 Mark J. Lowe and Erik B. Beall
2.3 Endogenous There are two principal mechanisms for generating contrast in MR
Functional Contrast images using endogenous features related to neuronal activation.
Both of these mechanisms are related to the hemodynamic response
to an increase in neuronal activation. One of these is a regional increase
in blood flow and the other is a concomitant increase in the oxygen-
ation content of the blood perfusing tissue near activated neurons.
Far and away the most commonly employed fMRI acquisitions
utilize the fact that regional brain activation results in a local
increase in blood oxygenation. This is called BOLD contrast. The
contrast in BOLD stems from the fact that oxygenated hemoglo-
bin is a weakly diamagnetic molecule, while deoxygenated hemo-
globin is a strongly paramagnetic molecule. The relative increase in
the concentration of oxygenated hemoglobin in the vessels perfus-
ing activated tissue results in an increase in the T2 and T2* relax-
ation times in the affected brain regions. Thus, methods utilizing
BOLD contrast for fMRI employ acquisition techniques that are
sensitive to changes in T2 and T2*. Because of the flexibility of T2
and T2* acquisition methods, this has become the contrast of
choice for the vast majority of fMRI experiments. For this reason,
the remainder of this chapter will focus on acquisition strategies to
acquire BOLD-weighted MRI data and we will discuss methods to
optimize these depending on experimental needs.
3.1 The Pulse As stated above, the pulse sequence refers to the specific acquisi-
Sequence tion strategy in which spatial encoding and magnetization read-out
is performed, providing the basic structure of the RF pulses and
field gradients used. Because conventional MRI is based on Fourier
spatial encoding, it has become convention within MRI to discuss
pulse sequences in the context of k-space, another name for the
Fourier conjugate of coordinate space. K-space is essentially the
image in the spatial frequency domain, and most pulse sequences
acquire image data in this domain. There are a variety of advan-
tages to this; most importantly that a coordinate space image can
be produced simply by performing a 2-dimensional Fourier trans-
form (typically computed using the Fast Fourier Transform, or
FFT) on sequentially acquired MRI data.
A pulse sequence for reading one arbitrary line of k-space with
a gradient-recalled echo is shown in Fig. 6.
Starting with stage (1), waveforms are played out on the
z-direction gradient, GZ, and the RF transmit channel to excite a
slice of proton spins. During stage (2), the readout gradient (Gx)
prewind and phase-encode gradient selection (Gy) is performed
while rephasing spins across the slice/slab with GZ. During stage
(3), the readout gradient is switched on while the emitted RF sig-
nal from the sample is recorded, denoted by the block of dotted
lines. The diagrams shown are simplifications, where the timing
and form of the gradients are changed according to various design
considerations. The corresponding traversal of k-space for the
pulse sequence in Fig. 6 is shown in Fig. 7.
Typical conventional (i.e., not single-shot) sequences repeat
this process, for different lines of kY, or phase-encoding positions.
This is shown in Fig. 6, step 2 with the GY gradient at multiple
Fig 6 Pulse sequence diagram for reading one line of k-space, time increases
from left to right. The proton echo from the sample is shown here in light gray
during and under the readout window, which will be sampled by a receive coil
78 Mark J. Lowe and Erik B. Beall
Fig 7 k-space diagram showing 16x16 matrix of data sampling points and trajectory of pulse sequence dia-
grammed in Fig. 6. k-space is first pre-wound in step 2 (GX moves position in kX from 0 to −8, GY moves posi-
tion in kY from 0 down to line −3), then kX is traversed in step 3 while sampling from –8 to +8
Fig 8 Generalized diagram for either gradient-echo (GE) or spin-echo (SE) ultrafast pulse sequences specific to
ultrafast imaging. A fat saturation pulse (described later), followed by excitation, then a refocus pulse (if spin-
echo only), prewind gradients to set position in k-space, then the readout and spatial encoding gradients. Finally
there may be postwind spoiler/crusher gradients (RF may also be used at the end) to dephase residual signal
3.2 Echo-Planar EPI follows the basic strategy of excitation of a slice or slab fol-
Imaging lowed by readout of one line (in the read-out direction, or kx) in
k-space. The GRE sequence was shown in the pulse sequence tim-
ing diagram in Fig. 8 without the SE refocusing pulse, which was
also shown in parts in Fig. 6. With the fast gradient switching
speeds available in recent years, it has become possible to spatially
encode an entire slice in one echo by performing multiple readouts
and phase-encoding steps after a single excitation. The most com-
mon implementation, known as “Blipped EPI”, involves excitation
of a slice followed by readout of kx line like the GRE sequence. The
sequence continues however, after an increment, or “blip”, of the
position in k-space in the other dimension using a short duration
gradient pulse (in the phase-encode direction, or ky). Readout con-
tinues when the read-out gradient is reversed to read another kx
line in k-space immediately adjacent to the first line sampled but in
the opposite direction. This is shown in Fig. 9.
These reversals and blips are repeated to adequately sample
k-space and the resulting data can be treated in the same manner as
multi-shot imaging, with the full readout of the slice or slab cen-
tered on the TE. The trajectory in k-space is shown in Fig. 10.
Three-dimensional acquisitions can be performed using an
additional increment in the perpendicular dimension, or kz,
although most blipped-EPI sequences are two dimensional only
due to the constraint of a shorter TE required. There exist many
modifications to this basic structure, but all EPI strategies contain
Fig 10 Cartesian trajectory in k-space for one-shot blipped-EPI sequence shown in Fig. 5. One-shot means full
coverage of k-space is accomplished during echo of one excitation
3.3 Spiral Imaging Another common strategy for single shot imaging is spiral imaging
[9]. In this scheme, k-space is sampled in a spiral or circular man-
ner, such as in Fig. 11, with less asymmetry between the rate of
sampling in kx- and ky-space.
By applying sinusoidal gradients 90° out of phase to the read-
and phase-encode gradients, k-space can be traversed in a circular
manner by increasing the amplitude of the sinusoidal gradients. A
typical sequence for spiral acquisitions is shown in Fig. 12.
82 Mark J. Lowe and Erik B. Beall
Fig 12 Pulse sequence timing diagram for spiral acquisition. Gradients during
readout window are 90° phase-offset ramped sinusoids
3.4 Parallel Imaging Multiple receive coils have become a popular and widely available
means to increase image SNR by providing multiple samples of a
k-space trajectory. Because the coils cannot be located in the same
place, they have varying spatial sensitivities to the tissue, which is
maximal at the tissue nearest each coil element. This provides an
alternative spatial encoding mechanism, where one sample of sev-
eral parallel coil elements provides information about the magneti-
zation density over several regions of tissue. Parallel imaging
combines this spatial encoding with the gradient-mediated spatial
encoding to skip some gradient-encoded lines in k-space and
replace those gaps with information derived from the parallel coil
elements [15]. The skipped lines in k-space reduce the field of view
(FOV) seen by the coils by a reduction factor. The individual coils,
if reconstructed with only the data acquired, would see the nearest
portion of tissue inside that coil’s FOV, but with aliased image
overlap with other portions of tissue further away from the coil.
3
Eddy currents are currents induced in gradient coils and other scanner com-
ponents from the rapidly changing fields generated by the gradient coils.
84 Mark J. Lowe and Erik B. Beall
Fig 13 Spiral-in/out sequence acquires full k-space data prior to echo time and
a second acquisition of k-space after echo time. The data from these echoes are
combined in reconstruction
The methods used to un-alias the data can be separated into two
strategies: image-space unfolding and k-space interpolation. While
there are many methods, and more than a few hybrids, the most
common implementations of each (sensitivity encoding, SENSE
[16] and generalized autocalibrating partially parallel acquisitions,
GRAPPA [17]) will be discussed, along with benefits/drawbacks.
Fig 14 SENSE k-space traversal for acceleration factor 2. Odd lines of k-space are missed, even lines acquired.
Acceleration factor equals acquired plus nonacquired number of lines, divided by acquired lines, in this case
the full k-space matrix would have twice the number of lines as were actually acquired
each ACS line, for each coil. So for N coils, there will be N2 weights
resulting from the fitting procedure to use in interpolating the non-
acquired lines. A particular coil’s matrix is based on all coil signals,
but masks out, or de-weights, signal from other regions outside the
FOV of that coil, in k-space. The matrix weighting removes the alias-
ing seen in the original, undersampled images. The final, unaliased
image data for each coil is combined by sum-of-square. The greatest
advantage of GRAPPA over image-space methods is the determina-
tion of sensitivity from the k-space data itself, which is useful in images
containing regions with poor homogeneity or low signal, both of
which are the case with ultrafast imaging [18].
3.4.3 Tradeoffs The primary benefit of parallel imaging is a reduction of the time
spent spatially encoding (the readout window), but at a cost of
SNR compared to the same sequence with a fully gradient-based
spatially encoded image using the average signal from the parallel
86 Mark J. Lowe and Erik B. Beall
Fig 15 GRAPPA k-space traversal. Central k-space is fully sampled to provide ACS lines. Outer k-space is
undersampled in phase-encoding direction by acceleration factor. Acceleration factor here is 2, so every other
line is acquired. Same sampling is acquired for other coils
3.4.4 Artifacts Residual alias in the image is a common artifact seen with parallel
imaging. The SENSE method requires the full image FOV to be
greater than the object of interest in any accelerated directions; other-
wise reconstruction will fail resulting in considerably aliased images.
The only current solution with SENSE is to expand the FOV so there
is no image wrapping [26]. This is not an issue with GRAPPA, because
Optimal fMRI Pulse Sequences 87
the spacing of k-space lines determines the FOV, but not the signal of
a particular line. Therefore, k-space fitting under GRAPPA is not
compromised by a smaller FOV, while image space fitting would be
compromised by the aliasing image. With an ideal sensitivity map, it
was believed SENSE could theoretically give better results than
GRAPPA, however the accuracy of the maps are highly dependent on
local field homogeneity and subject motion can invalidate them.
There are now several sensitivity map methods, including ones based
in part on GRAPPA autocalibrating methods that derive the maps
from the data to get around these problems [27]. Finally, there is the
issue of fitting the systems of equations in the presence of incomplete
coil coverage. The SENSE method requires the solution of an inverse
problem, but if there are regions of tissue that no coil has adequate
sensitivity to, this is an ill-conditioned problem that cannot be exactly
solved. All implementations of SENSE regularize or condition the
data to work around this, but it means that the final reconstructed
image may have local noise enhancements [28–30]. GRAPPA is also
sensitive to this problem, but because fitting is done with k-space data,
the noise enhancement is global rather than local [21].
3.4.5 Limitations There is a practical limit to the number of coils and the acceleration
factor, because adjacent coils will overlap spatially in their sensitivity
and coverage of magnetized spins, reducing the ability to separate
aliased signals. In typical applications of ultrafast imaging, the use of
2D EPI sequences leads to a limit on the acceleration factor of
between 4 and 5 [19]. While not every scanner has multiple chan-
nels, it is becoming the standard for vendors to offer such capability.
Many sites do not use parallel imaging for BOLD at 1.5 and 3 T due
to higher than expected SNR loss at even the lowest acceleration
factor [31], but at 7 T and higher field strengths parallel imaging is
necessary in most BOLD acquisitions to reduce echo time and image
warping. Future implementations of parallel imaging promise higher
acceleration factors with less loss of SNR using hybrids of k-space
and image-space methods with dynamically changing undersam-
pling strategies, such as k–t SENSE [32] or k-t GRAPPA [33].
3.5 Partial Fourier Ideal k-space data has complex conjugate symmetry, which can be
Imaging exploited to reduce the acquisition time. Up to half of k-space can
be interpolated from symmetry with the other half. This is referred
to as partial Fourier imaging. The symmetry is only approximately
true in real data due to scanner and tissue nonidealities, so algo-
rithms to take advantage of this fact must use low-resolution
approximations to account for nonzero phases in regions breaking
this symmetry [34, 35]. With the use of partial Fourier acquisition,
a higher spatial resolution can be acquired with less signal loss and
blurring, with the result that the SNR does not drop along with
the reduced acquisition time [36, 37].
88 Mark J. Lowe and Erik B. Beall
that had been excited at the same time as that given location. The
level of artifact has been assessed [40–42] using simulation or recon-
struction details not usually available to the average investigator. It
may be possible to determine the level of artifact on time series
image data alone, such as using seed voxel correlation [43] and com-
puting the difference between correlation patterns to aliased slices
and non-aliased slices, but there is at present no established method.
The flexibility afforded by these new techniques increases the
likelihood that investigators will customize the protocol according
to their specific needs. For example, some investigators are focusing
efforts on whole brain BOLD acquisitions with short TR, while oth-
ers focus on whole brain with thinner slices and are less concerned
with TR. At present it is too early to recommend a specific protocol,
but we will here describe two protocols with different goals that
have been used for connectivity and fMRI studies. The first protocol
was designed for a short TR and whole-brain coverage, the second
protocol designed for high spatial resolution and whole-brain cover-
age. Both protocols use a 32 channel head coil.
3 T Short TR protocol: acceleration factor = 8, 2 mm isotropic
voxel size, TR = 720 milliseconds. The short TR was intended both
to directly sample physiologic noise artifacts from the cardiorespi-
ratory cycles and to enhance statistical power using more BOLD
samples in a given scan time [42]. Figure 16 shows an example
functional connectivity study with this acquisition.
7 T high-resolution protocol: acceleration factor = 3, FOV/3
blipped-CAIPI FOV shifting, voxel size = 1.2 × 1.2 × 1.5 mm, 81
1.5 mm thick slices, TR = 2.8 s. The high resolution protocol was
intended to reduce slice thickness and increase resolution.
Figure 17 shows an example functional connectivity study done
using this protocol.
4 Artifacts
4.1 Non Physiologic There are several potential artifacts from single-shot imaging tech-
niques due to hardware realities, such as chemical shift (fat) artifact,
eddy current artifacts induced by the fast gradient switching, imperfec-
tions in gradient ramping waveforms, and both blurring and signal loss
due to nonuniform TEs combined with static field inhomogeneity.
4.1.1 Water-Fat Shift Water-fat shift image artifact is a consequence of the off-resonance
frequency of body fat that shifts the fat signal mostly in the phase-
encode direction, misplacing it across the image. Fat suppression
with an RF pulse at the resonance frequency of fat, often called
chemical saturation, followed by a strong dephasing gradient is the
standard countermeasure on EPI sequences. This RF pulse is done
immediately before the initial excitation pulse and, due to the fact
that the longitudinal signal of the fat is saturated, water protons in
fat will experience no excitation. One immediate consequence of
90 Mark J. Lowe and Erik B. Beall
Fig 16 (a) Single-slice correlation map to seed located in left primary motor cortex, data consists of 132 vol-
umes acquired with 2.8 s repetition time, (b) shows power spectrum of seed timeseries in Hz, (c) correlation
map to same seed location in separate data consisting of 416 volumes acquired with 0.72 s repetition time
and (d) shows power spectrum of seed timeseries of fast sampled data
Fig 17 Correlation map to seed in left primary motor cortex in 7 T data acquired
with multiband EPI acquisition with 1.2 × 1.2 mm × 1.5 mm voxels
Fig 19 Phase ghost typical level on left. Phase ghost correction algorithm failure on right. All measured signal
values normalized to first brain tissue measurement on top left
4.1.3 Echo Shifting The long readout time employed leaves single-shot images highly
sensitive to static field homogeneity, leading to image distortion
artifact in regions of inhomogeneity. The off-resonance frequency
in these regions causes an accumulation of phase errors in those
regions over the readout time. Phase errors specific to a region
result in spatial encoding errors, which manifest as signal misplace-
ment from that region. For blipped-EPI images, this is insignifi-
cant in the read-out, or kx direction because it is sampled so quickly,
but the phase-encode direction is sampled more slowly, resulting in
spatial distortion, or blurring in the phase-encode direction in
those regions. Spiral imaging samples the kx and ky dimensions at
approximately the same rate, but the radial dimension is sampled
more slowly, akin to the phase-encode direction in EPI. Spiral
images are therefore blurred across both dimensions [48]. The
geometric distortion can be “unwarped” from the images using
the calculated pixel shifts from an acquired field map for both
blipped-EPI [7, 49] and spiral imaging [35] (Fig. 20).
4.1.4 Signal Loss Signal loss, or slice dropout, is caused by through-slice dephasing
after the RF excitation. This signal loss cannot be recovered with-
out modifying the pulse sequence. Strategies for overcoming this
include: use of SE to refocus the dephasing effects, reducing the
TE, reducing slice thickness and/or in-plane voxel size, and chang-
ing the scan plane. If hardware permits, the use of high order gra-
dient shims and image-based shimming can help [50–52].
94 Mark J. Lowe and Erik B. Beall
Fig 20 (a) Sagittal, (b) axial views of fieldmap. Other pictures show (c) and (d) blurred blipped-EPI signal and
(e) and (f) unwarped images of the same using pixelshifts calculated from the fieldmap. Anterior regions
shifted roughly 1–2 pixels, but signal loss near the frontal sinuses cannot be recovered, leaving a signal void
4.1.5 Spiral Regridding Apart from lower spatial specificity and increased acquisition
time, problems typically associated with spiral imaging lie in the
regridding of the spiral trajectory to a Cartesian coordinate sys-
tem prior to Fourier transform. Regridding introduces subtle
artifacts and reduces SNR, and is computationally intensive com-
pared with the reconstruction methods used in blipped EPI,
although computational improvements have been made [56, 57].
Optimal fMRI Pulse Sequences 95
4.2 Physiologic There is another class of image artifacts present in functional neuro-
Artifacts imaging that have physiologic origins. Head motion [59–61] and
physiologic noise from the heart and breathing cycles [62–64] are
unavoidable non-neuronal sources of variance, changing underly-
ing statistical distributions and introducing possible systematic
effects in population studies. Accounting for these artifacts requires
care in the acquisition of data and several stages of postprocessing
of data (retrospective correction techniques will not be discussed
here) after collection is complete. Preventive measures include head
restraints or navigator echoes to reduce the effects of head motion,
and routine scanner quality assurance measures to track the stability
of the scanning hardware [65]. In addition, the collection of paral-
lel measures of state during the image acquisition may be useful for
artifact removal during postprocessing. These parallel measure-
ments can include online motion detection parameters from naviga-
tor echo or prospective motion correction along with signals
representing physiologic cardiac and respiratory cycles. Parallel
measurements in general must be acquired as fast or faster than the
slice acquisition sampling rate. It should be noted that there are
post-acquisition methods for obtaining slice-sampling rate motion
parameters [66] and physiologic cardiac and respiratory cycles [67].
4.2.1 Head Motion Due to the fact that it is desired to maintain a high temporal resolu-
tion, the sampling rate in most fMRI acquisitions is fast compared
to the T1 of brain tissue. The consequence of this is that, after equi-
librium is achieved after acquisition of a few volumes, the tissue is in
a saturated state. This means that the magnetization is not com-
pletely recovered between excitations of a given slice. If a subject
moves such that tissue from one slice moves into an adjacent slice,
the tissue will, for the first excitation after the motion, be in a dif-
ferent state of saturation than the rest of the tissue in that slice. This
leads to a signal change that is correlated with the motion, but will
not be corrected by the traditional technique of retrospective
realignment of the images. Prospective motion correction tech-
niques are intended to deal with this problem in real-time.
Navigator-echoes can be used to obtain motion information
during the acquisition of data [68, 69]. This technique uses the
fact that the phase of MR data is sensitive to motion. Typically,
low-power RF pulses are interspersed with the fMRI data acquisi-
tion and the phase information from the readout of the signal from
these pulses is used to infer motion along a given direction. The
96 Mark J. Lowe and Erik B. Beall
4.2.2 Physiologic Noise Ongoing physiologic processes in living subjects present an addi-
tional potential artifact. Effects due to the cardiac and respiratory
cycles have been identified as being significantly coupled to BOLD-
weighted MR signal in voxels in the brain and spinal cord. The
primary effect of the respiratory cycle on blipped-EPI fMRI data is
an apparent shift in image position in the phase-encode direction.
This is due to shifting of the resonant water frequency as the main
field drifts due to chest expansion and contraction [71, 72]. The
primary effect of the cardiac cycle is pulsatility artifact with each
heartbeat, although the structure and timing of the artifact may
vary across the brain due to the range of vessel sizes, stage in the
vessel network, and location in the brain [73, 74]. The cardiac
effects are more pronounced in certain regions such as the insula
and brainstem, while respiratory effects are more global (Fig. 21).
Fig 21 Averaged physiologic coupling maps in blipped-EPI with phase-encode direction in Anterior-Posterior
axis, determined by temporal ICA. Cardiac coupling overlain on anatomy is shown on top row, respiratory
coupling is shown on bottom row
Optimal fMRI Pulse Sequences 97
5.1.1 Field Strength The effect of static field strength on T2 relaxation in brain tissue, as
and Relaxation Parameters evidenced from examination of Table 1, is generally that it is
reduced. The effect of field strength on the BOLD signal is com-
plex, and depends on the nature of the proton transport mecha-
nism in the presence of the field defects introduced by the
deoxygenated hemoglobin. Recent studies suggest that this mech-
anism is largely diffusive in nature at clinical field strengths, which
would suggest a linear dependence of the BOLD signal on field
strength. Experimental data bear out the linearity of the depen-
dence of BOLD contrast on field strength [78]. Thus, BOLD con-
trast from extravascular protons can be taken to increase
approximately linearly in the regime used by most commonly avail-
able MRI scanners (i.e., 0.3 T to 3.0 T).
The intravascular contribution to BOLD signal stems from the
impact of the change in oxygenated hemoglobin concentration
98 Mark J. Lowe and Erik B. Beall
5.2 Experimental The goal in experimental design of fMRI studies is to take advan-
Design tage of the fact that regional changes in blood flow and oxygen-
ation result proximal to regions of increased neuronal activation.
Historically, the basic methodology has been to acquire properly
weighted MRI data of the brain regions of interest while a subject
repeatedly performs tasks related to the brain function of interest.
Initial methodology took advantage of the observation that con-
tinuously repeating a task during short intervals leads to an accu-
mulation of signal from the overlapping of events in a time short
compared to the hemodynamic response. This is a feature of the
general linear model (GLM) of functional imaging [79].
Blocking activation in bursts of extended activity over many
seconds, interleaved with long period of rest, leads to up to a much
higher increase in hemodynamic response than short, isolated
events. This fact makes it desirable, when possible, to use what is
typically referred to as a block design.
In 1997, Josephs and colleagues proposed an alternative
experimental design, intended to more specifically detect the MRI
signal associated with neuronal events [80]. This experimental
design takes advantage of the fact that, through synchronization of
the time of stimuli and measurement of behavioral responses, func-
tional imaging data can be analyzed for signal fluctuations corre-
lated with brief, temporally separated neuronal events. This type of
experimental design is referred to as event-related fMRI. Due to its
suitability to address more complex neuroscience questions regard-
ing brain activation and interactions, this has become a preferred
experimental design among neuroscience researchers.
Since these two experimental approaches have different analy-
sis strategies, the issues with regard to optimizing pulse sequences
are different between them. In the sections below, we separately
discuss these issues.
5.2.1 Block Design fMRI As stated above, block design fMRI experiments are designed to
create a large aggregate signal from activated neurons extended in
time, interspersed with long periods of rest, or alternate task per-
formance. Analysis of this type of data is typically performed with
what is referred to as a reference function. The simplest method for
analyzing these data is simply to calculate the cross correlation of
the experimental reference function with the timeseries at each
voxel [81]. Although more sophisticated methods have been
developed that allow more complex analyses, accounting for
nuisance effects and systematic effects of no-interest, for purposes
of pulse sequence optimization, a correlation approach is sufficient
to illustrate the issues.
Optimal fMRI Pulse Sequences 99
5.2.2 Event-Related fMRI Event-related fMRI relies on a different analysis strategy than block
design fMRI. The principal difference as it relates to choice of pulse
sequence is temporal resolution. A typical method for analyzing
event-related fMRI is deconvolution. Deconvolution is an analysis
method where rapidly repeated, although temporally separated,
events can be extracted if the timing of the onset of the signal and
either the duration or the hemodynamic response function is
known. A detailed discussion of deconvolution techniques is beyond
the scope of this chapter. The reader is referred to the chapter on
statistical analysis in this volume for a more complete treatment.
Figure 24 shows a typical timing and signal response for a rap-
idly presented event-related fMRI experiment. Studies on the abil-
ity of deconvolution techniques to resolve neuronal timing shifts
indicate that volume sampling rates (i.e., TR) of up to 3 s permit
identification of neuronal timing shifts of order 100 ms [82].
100 Mark J. Lowe and Erik B. Beall
Fig 23 Probability of rejecting a true event (type II error) as a function of the number of samples at false positive
rate of 0.01 for a two-cycle block design fMRI experiment. Result is from simulating image SNR = 50 with a
2 % BOLD signal change
a
1
0.5
0
0 20 40 60 80 100 120 140 160 180 200
Seconds
0.5 b
MR signal (a.u.)
-0.5
1 c
0.5
0 5 10 15 20 25 30
Seconds
Fig 24 Example of event-related fMRI experiment. Rapid presented stimuli (a) results in single-voxel time-
series shown in (b). hemodynamic response (c) is deconvolved from signal average over 100 voxels with
temporal resolution 3 s. is shown in (c)
Optimal fMRI Pulse Sequences 101
6.1 Pulse Sequence Various pulse sequences that have been proposed for fMRI acquisi-
tion were outlined in Sect. 2 of this chapter. Issues with regard to
optimization of fMRI experiments are discussed here.
6.1.1 Echoplanar This is now the most commonly available single shot imaging
Imaging sequence available on MRI scanners. Data can be acquired in GRE
mode and SE mode4. The issues with regard to optimization are
4
In addition, a mixed mode EPI has been used in the literature known as ASE,
or asymmetric spin echo. This is a SE EPI with the acquisition window shifted to
be centered on a time point early on in the SE evolution. The result is an acquisi-
tion that has, in a sense, adjustable sensitivity to capillary and venous signal.
102 Mark J. Lowe and Erik B. Beall
Table 2
Basic sequence parameters for BOLD fMRI acquisition on most clinical
MRI scanners
Field strength
6.1.2 Spiral Imaging Spiral imaging is becoming more common and is available as a prod-
uct sequence on some clinical MRI scanners. As outlined above, the
major advantage of spiral imaging with respect to EPI is that it is less
demanding on the imaging gradients. Thus, there will be reduced
image warping from eddy current effects. In addition, the nonuni-
form sampling of the Fourier domain image will lead to a different,
possibly lower, sensitivity to motion effects and even some types of
physiologic noise. Variants of the spiral technique have been pro-
posed that are specifically designed to be more sensitive to the char-
acteristics of the BOLD signal. This sequence is recommended for
researchers employing systems with underpowered gradient systems
and for situations where motion and/or physiologic noise or other
Optimal fMRI Pulse Sequences 103
6.1.3 Parallel Imaging MRI scanner manufacturers are increasingly moving to the use of
head array RF coils in lieu of circularly polarized quadrature head
RF coils for MRI. The cost, particularly at 1.5 T, is uniformity of
SNR, and thus fMRI signal detection efficiency across the brain.
This is less of an issue at high field strength, since dielectric effects
in this frequency regime reduce the uniformity of the quadrature
coil anyway. The advantage of head arrays is that parallel imaging
methods can be used to accelerate spatial encoding of images. The
result is dramatically increased image quality in regions where sin-
gle shot imaging methods have historically been very poor in qual-
ity (e.g., orbitofrontal regions, mesial temporal lobe, brainstem).
Some of these brain regions have important and interesting func-
tions. Parallel imaging techniques with the head arrays available to
most researchers will typically result in lower SNR throughout
most of the brain, but these can still be effectively employed in
situations where image artifact severely limits experimental options.
6.2 Scan Plane Issues with regard to scan plane are largely esthetic. However,
there are some technical issues that are worth discussing here.
Perhaps the most important issue is brain tissue coverage. The
scan plane of choice can affect the coverage of brain tissue. Given
a TE, receiver bandwidth, and TR, the number of slices available to
be acquired in one TR is fixed on a given scanner.5 The most effi-
cient scan plane for acquiring most human brains is the sagittal
plane. The brain in most adults is shortest in the right/left dimen-
sion, and so fewer slices will be required to cover the entire brain.
An axial acquisition plane is recommended in Table 2 due to
the fact that it is a more intuitive scan plane to work in, both ana-
tomically and from a physics perspective. Eddy current and higher
order artifacts stemming from gradient and shim coil interactions,
that are essentially related to coil geometry, are more easily under-
stood in the axial plane. With that said, it is a simple extension to
understand these effects in other scan planes. Axial imaging has the
added advantage over sagittal and coronal imaging planes in that
lateral, frontal, prefrontal, and posterior regions of the brain can be
imaged entirely within a relatively few slices (i.e., the very top and
very bottom of the brain are considered by many to be more
“expendable” than these other regions). The axial plane is a very
common imaging choice in fMRI and thus is listed in Table 2.
6.3 Field-of-View FOV has an impact on fMRI signal optimization in three ways: (1)
together with image matrix, it determines the in-plane voxel size
and there are a number of issues with regard to this that will be
5
There are, of course, other parameters that can affect this, such as gradient
slew rate, partial fourier and/or field-of-view acquisition, etc.
104 Mark J. Lowe and Erik B. Beall
6.3.1 In-Plane Voxel Size In-plane voxel size affects fMRI acquisition SNR (and thus fMRI
signal detection efficiency) and image quality. At lower field
strengths, the SNR issue will dominate and thus it is recommended
to use a larger voxel size at the expense of spatial resolution in
order to enhance signal detection efficiency. Further signal
enhancement can be attained with minimal loss of spatial resolu-
tion at 1.5 T through special spatial filtering techniques [83, 84].
At field strengths of 3.0 T and higher, voxel size has an inter-
action with physiologic noise from cardiac and respiratory sources
that can be detrimental to fMRI signal detection [64]. It is rec-
ommended that smaller voxels be employed at 3 T and higher to
limit the impact on BOLD CNR from physiologic noise. If spatial
resolution is not a concern, it is still recommended that data be
acquired at a higher spatial resolution (i.e., smaller voxel size)
and retrospective spatial filtering be employed to further increase
the CNR [84].
6.3.2 Image Artifact Due to the fact that the spatial encoding in the phase direction
Reduction (i.e., the direction encoded using, for instance, the phase blip-
ping described in Sect. 2.2 above) is not bandwidth limited, tis-
sue outside of the FOV in the phase direction that experiences
RF excitation, will be appear wrapped into the FOV with a signal
intensity related to the leakage RF experienced by that tissue.
For that reason, it is important for most acquisitions that the
field of view in the phase direction is adequate to contain the
entire brain volume and is oriented such that other tissue is not
proximal to the FOV. An example would be a coronal plane
acquisition with the phase encode direction in the inferior/supe-
rior direction. In this acquisition, even if the entire brain volume
is within the FOV, tissue from the neck and trunk of the body
that is within the sensitive volume of the transmit and receive RF
coils will appear aliased into the top of the FOV. A more com-
mon problem is that the field of view is chosen too small and one
side of the brain is wrapped into the other side of the brain (see
Fig. 25). This is avoided most simply by adopting a large enough
FOV in the phase direction. The recommendation in Table 2 is
sufficient for most adults.
6.4 Image Matrix As stated above, the principal impact of image matrix is on voxel
size and these issues are outlined above. However, it will also affect
the duration of the data acquistion for a single slice in single shot
imaging. This duration, as discussed in Sect. 3.2 above, can have a
detrimental effect on image quality in ultrafast imaging. Generally,
the total readout time should be much less than T2 or T2*. Typical
methods of decreasing the scan duration while maintaining good
Optimal fMRI Pulse Sequences 105
Fig 25 Sagittal EPI image with phase direction too small for the brain dimension
in the anterior–posterior direction. The anterior portion is phase-wrapped into
the tissue at the posterior part of the brain
6.5 Echo Time (TE) The TE is probably the most important consideration in optimiz-
ing a pulse sequence for BOLD contrast (or any T2 or T2* contrast
for that matter). As discussed in Sect. 3.1, optimal BOLD contrast
is obtained by selecting a TE that is the average of the T2 or T2* of
the tissue in the active and inactive states. This will necessarily
depend on the tissue characteristics and the field strength. The TEs
recommended in Table 2 are typical echo times for a GRE-EPI
acquisition that have a reasonable balance between tissue sensitiv-
ity and specificity. Adopting a longer TE can result in less sensitiv-
ity to intravascular signal, especially at 3.0 T and higher, while a
shorter TE can result in higher SNR. Ranges of TE for T2* BOLD
imaging at 1.5 T include 40–65 ms, while ranges used experimen-
tally at 3.0 T can range from 25 to 40 ms. One should be careful
when adopting TEs outside of these ranges as BOLD contrast can
be severely attenuated.
106 Mark J. Lowe and Erik B. Beall
6.6 Repetition Time With regard to BOLD contrast, TR has the fairly simple effect of
(TR) increasing or lowering SNR based on the T1 of the tissue. For a TR
short with respect to the T1 of the tissue of interest, the MR signal
will be saturated. This will be discussed in some detail in the sec-
tion regarding the flip angle. Here, we will simply point out that
very short TRs can result a significant reduction in SNR, and can
subsequently also result in a significant contribution from flow
contrast, depending on the saturated state and whether a slice gap
is included in the prescription.
Increased blood flow results in an apparent shortening of the
T1 relaxation time due to the effect of infusing blood on the net
saturated state of the protons in a given voxel. Generally, the effect
of increased flow on the T1 of a given voxel can be expressed as:
1 1 f
= + (5.1)
T1eff T1 l
6.7 Flip Angle Technically, the flip angle relates to the amount of RF power applied
at the excitation stage of a pulse sequence. For a given tissue type
(i.e., T1) and TE, MR signal is optimized at a flip angle referred to
as the Ernst angle. The formula for the Ernst angle is given by:
ìï æ -TR ö üï
a E = cos -1 íexp ç ÷ý (5.2)
îï è T1 ø þï
6
More accurately, retrospective motion correction techniques will be more
effective.
Optimal fMRI Pulse Sequences 107
6.9 Slice Thickness Choice of slice thickness has generally the same effect as spatial res-
olution mentioned above. Principal effects are brain coverage, SNR,
and image quality. The recommended slice thickness in Table 2
should give nearly whole-brain coverage in most adults, with accept-
able SNR and image artifact given field strength limitations.
As a further note, there is no mention of a slice gap in Table 2.
A slice gap is not recommended in fMRI studies where the entire
brain is desired. The RF excitation for a given slice will not be
perfect, so if a small gap (<10 % of slice thickness) between slices
were permitted, the tissue in this gap would still be sampled,
although less than if the slices were simply made thicker.
Historically, slice gaps were included to improve SNR in 2D
acquisition by reducing RF crosstalk between adjacent slices. The
longer TR recommended in Table 2, along with an interleaved
style pattern of slice excitation, should be sufficient to make this a
negligible effect in most clinical scanners.
6.10 Number The desired number of slices in an fMRI acquisition will affect
of Slices temporal resolution and brain coverage. The recommended num-
ber of slices in Table 2 should permit whole brain coverage in most
situations. Issues with regard to TR are discussed above.
7 Concluding Remarks
References
1. Belliveau JW, Kennedy DN Jr, McKinstry RC 11. Pipe JG, Duerk JL (1995) Analytical resolution
et al (1991) Functional mapping of the human and noise characteristics of linearly reconstructed
visual cortex by magnetic resonance imaging. magnetic resonance data with arbitrary k-space
Science 254:716–719 sampling. Magn Reson Med 34:170–178
2. Ogawa S, Lee TM, Nayak AS, Glynn P (1990) 12. Bornert P, Schomberg H, Aldefeld B, Groen
Oxygenation-sensitive contrast in magnetic J (1999) Improvements in spiral MR imaging.
resonance image of rodent brain at high Magma 9:29–41
magnetic fields. Magn Reson Med 14:68–78 13. Glover GH, Law CS (2001) Spiral-in/out
3. Bandettini PA, Wong EC, Hinks RS, Tikofsky BOLD fMRI for increased SNR and reduced
RS, Hyde JS (1992) Time course EPI of susceptibility artifacts. Magn Reson Med
human brain function during task activation. 46:515–522
Magn Reson Med 25:390–397 14. Preston AR, Thomason ME, Ochsner KN,
4. Kwong KK, Belliveau JW, Chesler DA et al (1992) Cooper JC, Glover GH (2004) Comparison of
Dynamic magnetic resonance imaging of human spiral-in/out and spiral-out BOLD fMRI at
brain activity during primary sensory stimulation. 1.5 and 3 T. Neuroimage 21:291–301
Proc Natl Acad Sci U S A 89:5675–5679 15. Sodickson DK, Manning WJ (1997)
5. Ogawa S, Tank DW, Menon R et al (1992) Simultaneous acquisition of spatial harmonics
Intrinsic signal changes accompanying sensory (SMASH): fast imaging with radiofrequency
stimulation: functional brain mapping with coil arrays. Magn Reson Med 38:591–603
magnetic resonance imaging. Proc Natl Acad 16. Pruessmann KP, Weiger M, Scheidegger MB,
Sci U S A 89:5951–5955 Boesiger P (1999) SENSE: sensitivity encoding
6. Frahm J, Bruhn H, Merboldt KD, Hanicke W for fast MRI. Magn Reson Med 42:952–962
(1992) Dynamic MR imaging of human brain 17. Griswold MA, Jakob PM, Heidemann RM
oxygenation during rest and photic stimula- et al (2002) Generalized autocalibrating par-
tion. J Magn Reson Imaging 2:501–505 tially parallel acquisitions (GRAPPA). Magn
7. Weisskoff RM, Davis TL (1992) Correcting Reson Med 47:1202–1210
gross distortion on echo planar images, Society 18. Heidemann RM, Griswold MA, Kiefer B
of Magnetic Resonance in Medicine 11th et al (2003) Resolution enhancement in lung
annual meeting, Berlin 1H imaging using parallel imaging methods.
8. Foerster BU, Tomasi D, Caparelli EC (2005) Magn Reson Med 49:391–394
Magnetic field shift due to mechanical vibra- 19. Ohliger MA, Grant AK, Sodickson DK (2003)
tion in functional magnetic resonance imag- Ultimate intrinsic signal-to-noise ratio for
ing. Magn Reson Med 54:1261–1267 parallel MRI: electromagnetic field consider-
9. Ahn CB, Kim JH, Cho ZH (1986) High- ations. Magn Reson Med 50:1018–1030
speed spiral-scan echo planar NMR imaging- 20. Wiesinger F, Boesiger P, Pruessmann KP
I. IEEE Trans Med Imaging 5:2–7 (2004) Electrodynamics and ultimate SNR
10. Bruder H, Fischer H, Reinfelder HE, Schmitt in parallel MR imaging. Magn Reson Med
F (1992) Image reconstruction for echo planar 52:376–390
imaging with nonequidistant k-space sam- 21. Blaimer M, Breuer F, Mueller M, Heidemann
pling. Magn Reson Med 23:311–323 RM, Griswold MA, Jakob PM (2004) SMASH,
Optimal fMRI Pulse Sequences 109
SENSE, PILS, GRAPPA: how to choose the 36. Jesmanowicz A, Bandettini PA, Hyde JS (1998)
optimal method. Top Magn Reson Imaging Single-shot half k-space high-resolution gradi-
15:223–236 ent-recalled EPI for fMRI at 3 Tesla. Magn
22. Ohliger MA, Sodickson DK (2006) An Reson Med 40:754–762
introduction to coil array design for parallel 37. Hyde JS, Biswal BB, Jesmanowicz A (2001)
MRI. NMR Biomed 19:300–315 High-resolution fMRI using multislice par-
23. Pruessmann KP, Weiger M, Bornert P, Boesiger tial k-space GR-EPI with cubic voxels. Magn
P (2001) Advances in sensitivity encoding with Reson Med 46:114–125
arbitrary k-space trajectories. Magn Reson 38. Moeller S, Yacoub E, Olman CA et al (2010)
Med 46:638–651 Multiband multislice GE-EPI at 7 tesla, with
24. Weiger M, Pruessmann KP, Osterbauer R, 16-fold acceleration using partial parallel
Bornert P, Boesiger P, Jezzard P (2002) imaging with application to high spatial and
Sensitivity-encoded single-shot spiral imaging temporal whole-brain fMRI. Magn Reson
for reduced susceptibility artifacts in BOLD Med 63:1144–1153
fMRI. Magn Reson Med 48:860–866 39. Setsompop K, Gagoski BA, Polimeni JR,
25. Heidemann RM, Griswold MA, Seiberlich Witzel T, Wedeen VJ, Wald LL (2012)
N et al (2006) Direct parallel image recon- Blipped-controlled aliasing in parallel imag-
structions for spiral trajectories using ing for simultaneous multislice echo planar
GRAPPA. Magn Reson Med 56:317–326 imaging with reduced g-factor penalty. Magn
26. Griswold MA, Kannengiesser S, Heidemann Reson Med 67:1210–1224
RM, Wang J, Jakob PM (2004) Field-of-view 40. Cauley SF, Polimeni JR, Bhat H, Wald LL,
limitations in parallel imaging. Magn Reson (2014) Setsompop K Interslice leakage arti-
Med 52:1118–1126 fact reduction technique for simultaneous
27. Griswold MA, Breuer F, Blaimer M et al multislice acquisitions. Magn Reson Med
(2006) Autocalibrated coil sensitivity esti- 72:93–102
mation for parallel imaging. NMR Biomed 41. Xu J, Moeller S, Auerbach EJ, et al. (2013)
19:316–324 Evaluation of slice accelerations using multi-
28. Sodickson DK (2000) Tailored SMASH image band echo planar imaging at 3 T. Neuroimage
reconstructions for robust in vivo parallel MR 83:991–1001
imaging. Magn Reson Med 44:243–251 42. Ugurbil K, Xu J, Auerbach EJ, et al. (2013)
29. Sanchez-Gonzalez J, Tsao J, Dydak U, Desco Pushing spatial and temporal resolution for
M, Boesiger P, Paul PK (2006) Minimum- functional and diffusion MRI in the Human
norm reconstruction for sensitivity-encoded Connectome Project. Neuroimage 80:80–104
magnetic resonance spectroscopic imaging. 43. Jo HJ, Saad ZS, Simmons WK, Milbury LA,
Magn Reson Med 55:287–295 Cox RW (2010) Mapping sources of correla-
30. Lin FH, Kwong KK, Belliveau JW, Wald LL (2004) tion in resting state FMRI, with artifact detec-
Parallel imaging reconstruction using automatic tion and removal. Neuroimage 52: 571–582
regularization. Magn Reson Med 51:559–567 44. Meyer CH, Pauly JM, Macovski A, Nishimura
31. Block KT, Frahm J (2005) Spiral imaging: DG (1990) Simultaneous spatial and spec-
a critical appraisal. J Magn Reson Imaging tral selective excitation. Magn Reson Med
21:657–668 15:287–304
32. Tsao J, Boesiger P, Pruessmann KP (2003) k-t 45. Zhou XJ, Du YP, Bernstein MA, Reynolds
BLAST and k-t SENSE: dynamic MRI with HG, Maier JK, Polzin JA (1998) Concomitant
high frame rate exploiting spatiotemporal cor- magnetic-field-induced artifacts in axial
relations. Magn Reson Med 50:1031–1042 echo planar imaging. Magn Reson Med
39:596–605
33. Huang F, Akao J, Vijayakumar S, Duensing
GR, Limkeman M (2005) k-t GRAPPA: a 46. Reeder SB, Atalar E, Faranesh AZ, McVeigh
k-space implementation for dynamic MRI ER (1999) Referenceless interleaved echo-
with high reduction factor. Magn Reson Med planar imaging. Magn Reson Med 41:87–94
54:1172–1184 47. Duyn JH, Yang Y, Frank JA, van der Veen JW
34. Cuppen JJ, Groen JP, Konijn J (1986) (1998) Simple correction method for k-space
Magnetic resonance fast Fourier imaging. Med trajectory deviations in MRI. J Magn Reson
Phys 13:248–253 132:150–153
35. Noll DC, Nishimura DG, Macovski A (1991) 48. Yudilevich E, Stark H (1987) Spiral sampling
Homodyne detection in magnetic reso- in magnetic resonance imaging-the effect of
nance imaging. IEEE Trans Med Imaging inhomogeneities. IEEE Trans Med Imaging
10:154–163 6:337–345
110 Mark J. Lowe and Erik B. Beall
49. Jezzard P, Balaban RS (1995) Correction for 62. Jezzard P, LeBihan D, Cuenod D, Pannier L,
geometric distortion in echo planar images Prinster A, Turner R (1992) An investigation
from B0 field variations. Magn Reson Med of the contribution of physiological noise in
34:65–73 human functional MRI studies at 1.5 Tesla
50. Blamire AM, Rothman DL, Nixon T (1996) and 4 Tesla, Society of Magnetic Resonance in
Dynamic shim updating: a new approach Medicine 12th annual meeting, New York
towards optimized whole brain shimming. 63. Lowe MJ, Mock BJ, Sorenson JA (1998)
Magn Reson Med 36:159–165 Functional connectivity in single and multislice
51. Wilson JL, Jenkinson M, de Araujo I, echoplanar imaging using resting-state fluctua-
Kringelbach ML, Rolls ET, Jezzard P (2002) tions. Neuroimage 7:119–132
Fast, fully automated global and local mag- 64. Triantafyllou C, Hoge RD, Krueger G et al
netic field optimization for fMRI of the human (2005) Comparison of physiological noise at
brain. Neuroimage 17:967–976 1.5 T, 3 T and 7 T and optimization of fMRI
52. Ward HA, Riederer SJ, Jack CR Jr (2002) Real- acquisition parameters. Neuroimage 26:243–250
time autoshimming for echo planar timecourse 65. Friedman L, Glover GH (2006) Report on a
imaging. Magn Reson Med 48:771–780 multicenter fMRI quality assurance protocol.
53. Yang QX, Williams GD, Demeure RJ, Mosher J Magn Reson Imaging 23:827–839
TJ, Smith MB (1998) Removal of local field 66. Beall EB, Lowe MJ (2014) SimPACE: gener-
gradient artifacts in T2*-weighted images ating simulated motion corrupted BOLD data
at high fields by gradient-echo slice exci- with synthetic-navigated acquisition for the
tation profile imaging. Magn Reson Med development and evaluation of SLOMOCO:
39:402–409 a new, highly effective slicewise motion correc-
54. Constable RT, Spencer DD (1999) Composite tion. Neuroimage 101:21–34
image formation in z-shimmed functional MR 67. Beall EB, Lowe MJ (2007) Isolating physi-
imaging. Magn Reson Med 42:110–117 ologic noise sources with independently
55. Chen N, Wyrwicz AM (1999) Removal of determined spatial measures. Neuroimage
intravoxel dephasing artifact in gradient-echo 37:1286–1300
images using a field-map based RF refocusing 68. Fu ZW, Wang Y, Grimm RC et al (1995)
technique. Magn Reson Med 42:807–812 Orbital navigator echoes for motion measure-
56. Oesterle C, Markl M, Strecker R, Kraemer ments in magnetic resonance imaging. Magn
FM, Hennig J (1999) Spiral reconstruction Reson Med 34:746–753
by regridding to a large rectilinear matrix: a 69. Lee CC, Jack CR Jr, Grimm RC et al (1996)
practical solution for routine systems. J Magn Real-time adaptive motion correction in func-
Reson Imaging 10:84–92 tional MRI. Magn Reson Med 36:436–444
57. Moriguchi H, Duerk JL (2001) Modified 70. Thesen S, Heid O, Mueller E, Schad LR (2000)
block uniform resampling (BURS) algorithm Prospective acquisition correction for head
using truncated singular value decomposition: motion with image-based tracking for real-
fast accurate gridding with noise and artifact time fMRI. Magn Reson Med 44:457–465
reduction. Magn Reson Med 46:1189–1201 71. Zhao X, Bodurka J, Jesmanowicz A, Li SJ
58. Nehrke K, Bornert P (2005) Prospective (2000) B(0)-fluctuation-induced temporal
correction of affine motion for arbitrary MR variation in EPI image series due to the dis-
sequences on a clinical scanner. Magn Reson turbance of steady-state free precession. Magn
Med 54:1130–1138 Reson Med 44:758–765
59. Hajnal JV, Myers R, Oatridge A, Schwieso 72. Raj D, Anderson AW, Gore JC (2001)
JE, Young IR, Bydder GM (1994) Artifacts Respiratory effects in human functional mag-
due to stimulus correlated motion in func- netic resonance imaging due to bulk suscepti-
tional imaging of the brain. Magn Reson Med bility changes. Phys Med Biol 46:3331–3340
31:283–291 73. Dagli MS, Ingeholm JE, Haxby JV (1999)
60. Friston KJ, Williams S, Howard R, Frackowiak Localization of cardiac-induced signal change
RS, Turner R (1996) Movement-related in fMRI. Neuroimage 9:407–415
effects in fMRI time-series. Magn Reson Med 74. Bhattacharyya PK, Lowe MJ (2004) Cardiac-
35:346–355 induced physiologic noise in tissue is a direct
61. Bullmore ET, Brammer MJ, Rabe-Hesketh S observation of cardiac-induced fluctuations.
et al (1999) Methods for diagnosis and treat- Magn Reson Imaging 22:9–13
ment of stimulus-correlated motion in generic 75. Guimaraes AR, Melcher JR, Talavage TM et al
brain activation studies using fMRI. Hum (1998) Imaging subcortical auditory activity
Brain Mapp 7:38–48 in humans. Hum Brain Mapp 6:33–41
Optimal fMRI Pulse Sequences 111
76. Hu X, Le TH, Parrish T, Erhard P (1995) 81. Bandettini PA, Jesmanowicz A, Wong EC,
Retrospective estimation and correction of Hyde JS (1993) Processing strategies for
physiological fluctuation in functional time-course data sets in functional MRI of
MRI. Magn Reson Med 34:201–212 the human brain. Magn Reson Med
77. Glover GH, Li TQ, Ress D (2000) Image- 30:161–173
based method for retrospective correction of 82. Miezin FM, Maccotta L, Ollinger JM, Petersen
physiological motion effects in fMRI: SE, Buckner RL (2000) Characterizing the
RETROICOR. Magn Reson Med 44:162–167 hemodynamic response: effects of presentation
78. Stefanovic B, Pike GB (2004) Human whole- rate, sampling procedure, and the possibility of
blood relaxometry at 1.5 T: assessment of dif- ordering brain activity based on relative tim-
fusion and exchange models. Magn Reson ing. Neuroimage 11:735–759
Med 52:716–723 83. Lowe MJ, Sorenson JA (1997) Spatially filter-
79. Friston KJ, Holmes AP, Worsley KJ, Poline J-B, ing functional magnetic resonance imaging
Frith CD, Frackowiak R (1995) Statistical para- data. Magn Reson Med 37:723–729
metric mapping in functional imaging: a general 84. Triantafyllou C, Hoge RD, Wald LL (2006)
linear approach. Hum Brain Mapp 2:189–210 Effect of spatial smoothing on physiological
80. Josephs O, Turner R, Friston KJ (1997) Event- noise in high-resolution fMRI. Neuroimage
related fMRI. Hum Brain Mapp 5:243–248 32:551–557
Chapter 4
High-Field fMRI
Alayar Kangarlu
Abstract
Magnetic resonance imaging (MRI) allows detection of signal from constituent of biological tissues.
Hydrogen (1H) is the most widely used element from which spectra and images are detected due to its
abundance and high sensitivity manifested in its gyromagnetic ratio. The high contrast for soft tissue have
afforded scientists invaluable information about brain structure and function. Among many parameters
determining quality of MRI images, field strength is the most decisive one as it determines signal strength
in fMRI images. Considering the low inherent sensitivity of fMRI, high magnetic field are the only way
that activation contrast of neurofunctional studies could be increased. This is why there has been a relent-
less drive towards higher field strength in human imaging raising it up to 11.7 T to date. Technology of
7-T has become more widely available in scanners with fMRI capability. Development of many technolo-
gies such as multichannel RF coils, strong and fast gradients, simultaneous slice excitation, and brain-
stimulation protocols have contributed to the expansion of fMRI as the method of choice for study of
whole brain function. In this chapter, challenges of high-field fMRI in human studies are discussed among
which signal to noise, susceptibility artifacts, multichannel RF coil designs are highlighted.
Key words High field, fMRI, Neuroimaging, Magnetic field, High resolution
1 Introduction
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_4, © Springer Science+Business Media New York 2016
113
114 Alayar Kangarlu
2 MR Signal
3 B0 Effects
4 Relaxation Effects
1.08 7T
3T
normalised signal intensity
1.06 1.5 T
1.04
1.02
0.98
0 50 100 150
time (s)
Fig. 2 fMRI study at 1.5, 3.0, and 7.0 T performed on Philips Achieva. Courtesy of University of Nottingham, UK
High-Field fMRI 119
The best way to image brain function would be to directly detect neu-
ronal activities. Absence of such a technique has provided an opportu-
nity for indirect observation of brain function through BOLD, which
can measure changes in hemodynamics as a result of controlled neu-
ronal activation. The coupling of neuronal activity and vascular hemo-
dynamics makes BOLD dependents on the details of communication
between neurons, glia, and blood vessels. Furthermore, BOLD is an
indication of the existence of a tight level of vascular reactivity between
neuronal network and vascular system.
Accurate fMRI representation of the brain function depends
on the understanding of mechanism of neurovascular coupling,
i.e., how neuronal activity affects hemodynamic response and its
MR signal. Such relationship will enable us to account for pharma-
cological or disease-induced modulations of neurovascular cou-
pling that could use BOLD signal changes, or perfusion, for the
assessment of drug efficacy and pathological modification of physi-
ology. In addition, fMRI has become an important tool in studies
in psychology, psychiatry research, and basic cognitive neurosci-
ence research [11]. This is primarily due to the fact that fMRI has
proven to be able to provide a veritable readout of mental
contents.
Organization of the brain allows the functional studies to iden-
tify the neuronal basis of behavior, or at least the hemodynamic
manifestation of that. The measure of neuronal activity is obtained
by constructing activity maps from the functional units involved in
various brain networks [4, 22, 23]. The functional units, commonly
called cortical columns are made up of neuronal networks involved
in the implementation of a specific function. They form an orga-
nized structure that interacts with other units of the system that
repeatedly occur in the cortex. This organized structure and its
columnar activation contribute to elucidate the function of specific
cortical areas [24]. Imaging with a resolution allowing to detect the
simultaneous activation of these units, i.e., the collective response of
all columns involved in a specific external stimulation, would greatly
High-Field fMRI 121
enhance the credibility of fMRI studies. This is due to the fact that
such resolution can establish the spatial localization of functional
units. Resting-state fMRI, which can detect spatially dispersed but
functionally connected regions that share information with each
other, offers information on functional connectivity (FC) of the
brain. Such FC maps are best utilized at high field, when they are
capable of offering the temporal dependency of neuronal activation
patterns of anatomically separate brain regions with high temporal
resolution. High-field FC offers a measure of interaction between
isolated clusters of columns involved in implementation of a func-
tion that will elucidate functional specialization of units and local
networks at columnar level, as well as new insights in the overall
organization of functional communication of brain networks. High-
field fMRI is capable of whole brain imaging at both high temporal
and spatial resolution, which together offer valuable information
about the core aspects of the human brain, providing an overview of
these novel imaging techniques and their implication to neurosci-
ence. High-field fMRI offers the opportunity of the (1) use of spon-
taneous resting-state fMRI in determining functional connectivity,
(2) to investigate the origin of these signals, (3) how functional con-
nections are related to structural connections in the brain network
and (4) how functional brain communication relate to cognitive
performance. Analysis of functional connectivity patterns using
graph theory, focusing on the overall organization of the functional
brain network, is also a promising technique that takes advantage of
these new functional connectivity tools in examining connectivity
diseases, like multiple sclerosis, dementia, schizophrenia, and
Alzheimer’s disease. The potential to further empower FC fMRI
with high-resolution maps based on functional units of the brain
[5–8, 22] is another reason that makes high field an exciting tech-
nology for brain studies.
The primary advantage of high-field fMRI, however, remains
the possibility of studying the brain physiology noninvasively with
a high spatial and temporal resolution at the same time. FC fMRI
reveals brain networks in resting state or based on experiments that
measure brain activation due to the execution of a specific task.
This is a unique capability that avails the entire brain for investiga-
tion at once. As such, it is critical that this capability is not compro-
mised as field strength increases. Diagnoses based on functional
mapping require high spatiotemporal resolution over the whole
brain as field strength increases. The heterogeneity of the brain
causes susceptibility-induced signal dropouts that worsen as the
field strength increases. Unfortunately, this is the same mechanism
that makes functional measurement of hemodynamics possible. So,
we must develop reliable techniques to suppress signal dropouts
while keeping susceptibility based CNR high. These conflicting
needs may provide new incentives to move fMRI toward direct
detection of neuronal correlates rather than the present
122 Alayar Kangarlu
5.1 Fast Imaging Image acquisition in MRI is slower than in other techniques such as
computed tomography and positron emission tomography. This is
mostly due to the relaxation phenomenon. Fast imaging techniques
are not widely popular for structural imaging due to the poor image
qualities and technical limitations. Relaxations and dephasing
requires refocusing of signals in the intervals of the order of TE and
realignment of spins with the main magnetic field every TR seconds,
High-Field fMRI 123
5.1.1 Echo-Planar Fast imaging techniques achieve their speed by multiple refocusing
Imaging of the spin ensemble during one TR. EPI as a GE-based technique
is the fastest sequence and has a very low RF power content [27].
This aspect of EPI makes it suitable for high-field applications as
RF absorption increases at high fields increasing the RF require-
ments. On the other hand, other aspects of EPI such as geometric
distortion, blurring artifacts, and T2* signal loss are aggravated at
higher fields [28, 29]. For instance, the geometric distortion that
is caused by off-resonance effects will be further aggravated by
long readout train of EPI. A phase offset that increases with TE
will be created that will establish a linear phase gradient over
k-space in the phase-encoding direction [30]. The image signal
from these spins will get shifted as image is reconstructed. At high
fields, this effect is proportionally stronger resulting in larger
frequency shifts. However, the effect of long readouts can be dras-
tically reduced by using parallel imaging. This will reduce geomet-
rical distortions, but the T2* signal decay and blurring on the
images will still remain. Other techniques have been introduced to
deal with T2* relaxation causing distortion in images due to the
decay in the signal along the k-space trajectory. Minimization of
magnetic field inhomogeneities and susceptibility-induced effects
requires the choice of TE close to T2*. As B0 increases, T2* decreases
and hardware and safety considerations often makes the minimum
TE of single-shot EPI longer compared to T2*, which causes signal
loss due to the phase dispersion caused by such choices of
TE. Higher bandwidth could alleviate this problem but possibility
of peripheral nerve stimulation will limit the use of much stronger
gradients to achieve this. Other techniques have been proposed
that will effectively restore T2* relaxation-induced signal loss and
blurring. GE slice excitation profile imaging (GESEPI) is one such
method that, combined with multichannel parallel receiver tech-
nology, such as sensitivity encoding (SENSE), will significantly
enhance high-field EPI image qualities [31, 32].
124 Alayar Kangarlu
of spin coherence which will reduce the signal causing dark regions
on T2*-weighted EPI images. Even spin-echo sequence will bear
reminiscences of such susceptibility-induced signal loss near the
veins. While for stationary tissues RF does refocus the resulted
dephasing of spins, for moving water molecules in veins protons
rephrasing is not complete, making BOLD effective as a T2 as well
as T2* effect.
fMRI signal is believed to largely originate from BOLD effect
around small vessels, i.e., arterioles, capillaries, and venules [25].
The extravascular areas surrounding the small vessels represent
loci of neuronal activity. But, there are contributions from large
vessels to the BOLD signal as well. Such contribution must be
quantified to ensure an accurate account of the role of small ves-
sels vs. large vessels in fMRI. High magnetic fields provide a pow-
erful tool in this regard. A known magnetic field at any position
puts spins in a well-defined precession whose frequency provides
knowledge of its location to produce a map of proton density.
Spatial homogeneity and temporal stability of the field are impor-
tant requirements for creating images faithful to the structures
being studied. B0 field homogeneity of high-field magnets is
around 0.5 ppm that using high-order shimming could improve it
to about 0.1 ppm over the head. Beyond this, as it was discussed,
dHb produces high magnetic susceptibilities leading into compa-
rable local inhomogeneities in the static field within the brain. At
high field, regions in the brain, such as temporal lobes and basal
ganglia, demonstrate high magnetic susceptibility providing a
high contrast from the surrounding tissues [40]. Different sce-
narios for change in T2* are possible depending on the occupation
of the voxel by capillaries, large vessels, and extravascular and
intravascular BOLD [8]. In general, it can be stated that T2* signal
differential between activation and rest period from these regions
increases as a function of magnetic field. For example, if typical
acquisition parameters for fMRI studies are receiver bandwidth of
2 kHz/pixel; TR 4000 ms; TE 40 ms; FOV 190 × 190 mm2;
30–40 slices; slice thickness, 5 mm; then implications of these
parameters at 7.0 T can be contrasted to 1.5 T through a simple
frequency shift. A typical BOLD effect of 0.5 ppm or 150 Hz
frequency shift at 7.0 T could result in as high as 7 % change in
signal. Considering that BOLD has typically produced SNR,
∆R/R, of the order of 1 % at 1.5 T, this fact indicates that a linear
increase in ∆R/R with B0 is possible.
BOLD contrast acts as a change in T2* rate, ΔR2*. What are the
factors affecting ΔR2*? First, ΔR2* is directly influenced by the
change in concentration of dHb. In fact, the volume susceptibility
is directly proportional to volume of dHb and as such on ΔR2*
[34]. Assuming that dHb is proportional to blood volume, the
fraction of the blood volume fdHb occupied by dHb will have
direct effect on the signal. Models have been proposed that assign
128 Alayar Kangarlu
5.4 Physiological In the absence of physiological noise, fMRI at high field could
Noise produce functional maps of the brain with even higher resolution
[43]. Scanners that already acquire submillimeter images at 7.0 T
High-Field fMRI 129
6 High-Resolution fMRI
6.1 RF and Gradient Gradients and RF coils are the two components of MRI scanners at
Coil Technology the forefronts of signal generation and detection. As such, their less
than ideal performance is the source of great many nuisances col-
lectively referred to as artifacts [51]. To eliminate the high-field dis-
tortions of fMRI images, a variety of solutions are available [52].
Postprocessing techniques are proposed to correct for some of the
distortions with known origins. Strong gradients also help reduce
High-Field fMRI 131
7 Conclusion
References
29. Goense JB, Ku SP, Merkle H, Tolias AS, 41. Davis TL, Kwong KK, Weisskopff RM, Rosen
Logothetis NK (2008) fMRI of the temporal BR (1998) Calibrated functional MRI: map-
lobe of the awake monkey at 7 T. Neuroimage ping the dynamics of oxidative metabolism
39:1081–1093 (hypercapniaycerebrovascular reactivity). Proc
30. Farzaneh F, Riederer SJ, Pelc NJ (1990) Natl Acad Sci U S A 95:1834–1839
Analysis of T2 limitations and off-resonance 42. Yacoub E, Shmuel A, Pfeuffer J, Van De
effects on spatial resolution and artifacts in Moortele PF, Adriany G, Ugurbil K, Hu X
echo-planar imaging. Magn Reson Med (2001) Investigation of the initial dip in fMRI
14:123–139 at 7 Tesla. NMR Biomed 14:408–412
31. Yang QX, Smith MB, Briggs RW, Rycyna RE 43. Krüger G, Glover GH (2001) Physiological
(1999) Microimaging at 14 Tesla using noise in oxygenation-sensitive magnetic reso-
GESEPI for removal of magnetic susceptibility nance imaging. Magn Reson Med 46:631–637
artifacts in T(2)(*)-weighted image contrast. 44. Wang SJ, Luo LM, Liang XY, Gui ZG, Chen
J Magn Reson 141:1–6 CX (2005) Estimation and removal of physio-
32. Yang QX, Wang J, Smith MB, Meadowcroft logical noise from undersampled multi-slice
M, Sun X, Eslinger PJ, Golay X (2004) fMRI data in image space. IEEE EMBS
Reduction of magnetic field inhomogeneity 27:1371–1373
artifacts in echo planar imaging with SENSE 45. Hyde JS, Biswal BB, Jesmanowicz A (2001)
and GESEPI at high field. Magn Reson Med High-resolution fMRI using multislice partial
52:1418–1423 k-space GR-EPI with cubic voxels. Magn
33. Chen NK, Wyrwicz AM (2004) Removal of Reson Med 46:114–125
EPI Nyquist ghost artifacts with two- 46. Glover GH, Krüger G (2002) Optimum voxel
dimensional phase correction. Magn Reson size in BOLD fMRI. Proc Int Soc Magn Reson
Med 51:1247–1253 Med 10:1395
34. Schenck JF (1996) The role of magnetic sus- 47. Mountscale VB (1997) The columnar organi-
ceptibility in magnetic resonance imaging: zation of the neocortex. Brain 120:701–722
MRI magnetic compatibility of the first and 48. Triantafylloua C, Hogea RD, Wald LL (2006)
second kinds. Med Phys 23:815–850 Effect of spatial smoothing on physiological
35. Callaghan PT (1990) Susceptibility-limited noise in high-resolution fMRI. Neuroimage
resolution in nuclear magnetic resonance 32:551–557
microscopy. J Magn Reson 87:304–318 49. Duong TQ, Kim DS, Ugurbil K, Kim SG
36. Kangarlu A, Bourekas EC, Ray-Chaudhury A, (2001) Localized cerebral blood flow response
Rammohan KW (2007) Cerebral cortical at submillimeter columnar resolution. Proc
lesions in multiple sclerosis detected by MR Natl Acad Sci U S A 98:10904–10909
imaging at 8 Tesla. AJNR Am J Neuroradiol 50. Kim DS, Duong TQ, Kim SG (2000) High-
28:262–266 resolution mapping of isoorientation columns
37. Filippi M, Rocca MA (2007) Conventional by fMRI. Nat Neurosci 3:164–169
MRI in multiple sclerosis. J Neuroimaging 51. Jezzard P, Clare S (1999) Sources of distortion
17(Suppl 1):3S–9S in functional MRI data. Hum Brain Mapp
38. Fazekas F, Soelberg-Sorensen P, Comi G, 8:80–85
Filippi M (2007) MRI to monitor treatment 52. Speck O, Stadler J, Zaitsev M (2008) High
efficacy in multiple sclerosis. J Neuroimaging resolution single-shot EPI at 7T. MAGMA
17(Suppl 1):50S–55S Magn Reson Mater in Phys Biol Med
39. Christoforidis GA, Bourekas EC, Baujan M, 21:73–86
Abduljalil AM, Kangarlu A, Spigos DG, 53. Baertlein BA, Ozbay O, Ibrahim T, Lee R, Yu
Chakeres DW, Robitaille PM (1999) High Y, Kangarlu A, Robitaille PM (2000)
resolution MRI of the deep brain vascular Theoretical model for an MRI radio frequency
anatomy at 8 Tesla: susceptibility-based resonator. IEEE Trans Biomed Eng
enhancement of the venous structures. 47:535–546
J Comput Assist Tomogr 23:857–866
54. Ibrahim TS, Lee R, Baertlein BA, Kangarlu A,
40. Bourekas EC, Christoforidis GA, Abduljalil Robitaille PL (2000) Application of finite dif-
AM, Kangarlu A, Chakeres DW, Spigos DG, ference time domain method for the design of
Robitaille PM (1999) High resolution MRI of birdcage RF head coils using multi-port excita-
the deep gray nuclei at 8 Tesla. J Comput tions. Magn Reson Imaging 18:733–742
Assist Tomogr 23:867–874
136 Alayar Kangarlu
55. Ibrahim TS, Kangarlu A, Chakeress DW 57. Pruessmann KP, Weiger M, Scheidegger MB,
(2005) Design and performance issues of RF Boesiger P (1999) SENSE: sensitivity encoding
coils utilized in ultra high field MRI: experi- for fast MRI. Magn Reson Med 42:952–962
mental and numerical evaluations. IEEE Trans 58. Sodickson DK, Manning WJ (1997)
Biomed Eng 52:1278–1284 Simultaneous acquisition of spatial harmonics
56. Kangarlu A, Baertlein BA, Lee R, Ibrahim T, (SMASH): fast imaging with radiofrequency
Yang L, Abduljalil AM, Robitaille PM (1999) coil arrays. Magn Reson Med 38:591–603
Dielectric resonance phenomena in ultra high 59. Katscher U, Börnert P, Leussler C, van den
field MRI. J Comput Assist Tomogr Brink JS (2003) Transmit SENSE. Magn
23:821–831 Reson Med 49:144–150
Chapter 5
Experimental Design
Hugh Garavan and Kevin Murphy
Abstract
This chapter addresses issues particular to the optimal design of fMRI experiments. It describes procedures
for isolating the psychological process of interest and gives an overview of block, event-related and
participant-response-dependent designs. An additional focus is placed on data analysis with emphasis on
optimizing and isolating the neuroimaging signal in activated brain regions. Finally, the chapter addresses
a number of practical matters including optimal sample sizes and trial durations that confront all research-
ers when designing their experiments.
Key words Task design, Sample size, Scan durations, Analysis, Regression, Efficiency, Frequency
1 Overview
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_5, © Springer Science+Business Media New York 2016
137
138 Hugh Garavan and Kevin Murphy
2 Task Design
interest amidst other control trials and to categorize the trials after
the participant has completed the experiment. Error trials can be
excluded (or averaged separately) and events can be coded by
whether a participant detected a target or not, responded relatively
fast or not, produced a subsequent behavior or not and so on [4].
This affords the researcher increased flexibility in probing the data-
set and has obvious advantages over the block design in circum-
stances in which the psychological process cannot be presented in
blocks as in, for example, an oddball paradigm in which the nature
of the phenomenon mandates that events are infrequent and
unpredictable. The block and event-related designs can also be
combined such that events of interest during an active task period
can be isolated while the task period itself can be simultaneously
contrasted against a control period [5, 6]. This type of mixed
design provides additional information in that one can determine
the inter-relationships between tonic activity levels (e.g., sustained
attention or an induced emotional state) and the processing of a
discrete trial (e.g., detection of a fearful face).
A final category of experimental design is what we have labeled
participant-response dependent. Here, the participant provides a
continuous measure that can, for example, be used to generate a
regressor to correlate against brain activity measures. Despite a loss
of experimental control over the participant’s behavior, this cate-
gory of design affords much flexibility when the phenomenon of
interest is either not strictly task-dependent or is difficult to experi-
mentally induce. Examples include resting state acquisitions (in
which correlated patterns of brain activity can be detected while
the participant simply rests) [7], biofeedback (in which, for exam-
ple, a participant learns to control their level of brain activity) [8],
passive viewing of a movie clip (in which there may be multiple
sources of stimulation with each varying with a different time-
course) [9] or in which performance varies in an unpredictable
manner [10]. Performance modulations for which one could assess
brain activation changes can be quite wide-ranging including
response times or response time variability on a continuous perfor-
mance task [11], frequently sampled self-report measures of mood
[12] and physiological measures such as heart rate or pupil-
diameter [13]. In these examples the discrete measurements can be
interpolated to provide a continuous time-series that can be cor-
related with the brain activation time-series data.
A related approach, albeit one that is manipulated to a degree
by the experimenter, is one in which the researcher derives a com-
putational model of a subject’s performance. Here, a formal model
of the processes hypothesized to underlie task performance is
developed by the experimenter. For example, on a forced-choice
reward task in which the subject attempts to maximize their win-
nings, one might hypothesize that subjects develop expectations of
reward based on previous trial outcomes, experience prediction
Experimental Design 141
2.1 Choosing The choice of which design to employ will be dictated by the par-
an Experimental ticulars of the psychological process to be investigated and how easy
Design it is to isolate. Block designs can be employed if the psychological
function is easy to isolate or if it is of particular interest to compare
two tasks. A simple example would be a contrast between unilateral
and bilateral finger movements. Here, blocks of finger movement in
just one hand could be alternated with blocks of finger movements
in two hands. Rest periods might also be included in order to pro-
vide a low-level baseline against which any task- related activity
could be assessed. The inclusion of a resting state baseline is gener-
ally advantageous as contrasts between two task-active periods can
often be ambiguous in that greater activation in condition A relative
to condition B could result from either more positive activation in
A or a greater deactivation in B. A resting state baseline allows one
to resolve this ambiguity by showing if activation increases or
decreases in any one condition relative to the resting baseline.
However, care must be taken when comparing task to rest as in the
last few years a large body of literature has demonstrated that rest
itself is not the absence of neural activity. For example, it has been
shown that the default mode network is more active during “rest”
than during a task (see further description of resting-state phenom-
ena and analyses in the chapter by Fornito—Chap. 10).
If the psychological function is not easily isolated then a conjunc-
tion analysis may be useful. As can be seen in the verbal working
memory example given above, the conjunction design enables the
researcher to identify the core functional neuroanatomy that is com-
mon across different operationalizations of a psychological process.
In addition, it can also reveal task-specific activations enabling, for
example, one to determine how verbal working memory rehearsal for
linguistic information differs to that of nonlinguistic information. An
142 Hugh Garavan and Kevin Murphy
3.1 Practical Issues There has been increased concern in recent years about the power
of many (most?) fMRI studies and a number of cogent critiques
suggest that low sample sizes have generated a disconcerting num-
ber of false positive results [18, 19]. However, only a handful of
studies have addressed how many participants are required to yield
stable activation maps. The first paper addressing this issue showed
144 Hugh Garavan and Kevin Murphy
3.2 Analytic Issues The purpose of an experimental design is to alter neural activity,
and hence the blood oxygen level dependent (BOLD) signal, as
effectively as possible and in a predicted way thereby enabling the
researcher to detect the resulting brain changes. Using this predic-
tion, one looks for corresponding patterns in the fMRI time series
to determine which voxels were engaged in the task. A simple ref-
erence time series can be produced by convolving the stimulus
timing function (which is equal to 0 when no stimulus is applied
and 1 when a stimulus is presented) with a hemodynamic response
function (HRF) that accurately represents the shape of the BOLD
response after a single event. The gamma variate function, y(t) = t r
e−t/b, has been shown to effectively model the hemodynamic
response to brief stimuli [28], with parameters r = 8.6 and c = 0.51,
and is a popular choice for modelling the hemodynamic shape. The
difference between two gamma-variates is also used in order to
model the post-stimulus undershoot. It is important that the cho-
sen HRF model accurately reflects the true shape of the response.
If, for some reason, the hemodynamic shape of a participant is
atypical (e.g., following treatment with a substance that directly
affects the vasculature or a patient group with vascular damage),
then the results of the analysis could be confounded by this differ-
ence in shape. (It should be noted that there are more advanced
approaches to reference time-series formation, such as ones that
use basis functions rather than a predetermined HRF shape and
these are addressed in a later chapter).
Experimental Design 147
y (t ) = b × x (t ) + a + e (t ) (2)
where y(t) is the voxel time-series data, x(t) is the reference func-
tion (i.e., the expected BOLD response to the stimulus) with β its
scaling factor, α is a constant, and ε(t) is a random Gaussian white-
noise term. Both β and α are unknown parameters that are fit by
the linear regression method. This equation can be extended to
include extra regressors to remove unwanted trends in the data,
such as baseline drift and physiological nuisance regressors, whilst
simultaneously computing the scaling factor. This scaling factor, β,
can then be used as an activation measure for each voxel.
Multiple reference waveforms can easily be included in this
type of analysis, denoted by the term multiple linear regression. In
an experiment with two active conditions [1, 2] the equation:
y ( t ) = b1 × x1 ( t ) + b 2 × x2 ( t ) + a + d × t + e ( t ) (3)
is fitted to the data. For this model, four parameters are estimated,
the two scaling factors β1 and β2 and the baseline α and also a base-
line drift rate δ which accommodates for linear changes in the base-
line over time. It is possible to investigate whether β1 or β2 are
nonzero and whether β1 is different from β2 with statistical signifi-
cance calculated using F-tests. This method allows one to identify
active areas in the brain and calculate if an area is more active in one
condition than another, thereby satisfying the two primary pur-
poses of fMRI. This equation can be further generalized to Y = X
β + ε, where Y is a column vector of the voxel’s time-series data, X
is the design matrix, β is a column vector of scaling factors and ε is
a column vector of Gaussian white noise terms [29, 30]. This equa-
tion is called the General Linear Model (GLM) and is the basis for
most fMRI analytic techniques. The columns of the design matrix
X model the effects of interest and also confounding variables and
are, in essence, the reference waveforms mentioned above.
When designing an experimental task, one is essentially specify-
ing these reference waveforms/regressors. To maximize statistical
power, these regressors must be chosen wisely. For example, F-tests
are used to determine if there are significant differences between
the regressors. To increase statistical power, one can increase the df
by lengthening the task (for long TRs, each additional timepoint
adds a new df). It might seem like a good idea to use extremely
short TRs to increase the number of time points and hence the
statistical power. However, to gain an extra df for each additional
timepoint, each timepoint must be statistically independent from
every other. Unfortunately, due to autocorrelations introduced into
148 Hugh Garavan and Kevin Murphy
the fMRI data by physiological noise and scanner drifts, this is not
the case. This example demonstrates that knowledge of the under-
lying mechanisms of fMRI along with the analysis methods is
required when choosing even the simplest variables (such as the
number of time points and the TR) for experimental design.
Efficiency of a task design is a measure of how accurately the
GLM can estimate the β weights for each of the regressors, that is,
how small the predicted variance of the β estimates will be. For
example, assume a block design task that induces exactly a 2 % signal
change in hundreds of voxels, all with different noise properties but
with the same noise variance. If a GLM analysis is performed, the β
estimate for every voxel will be approximately 2 % for all voxels with
very little deviation. Since the variance of the estimates does not
differ widely with noise distributions, this would be considered an
efficient task. However, matters become more complicated when
there is more than one regressor. Assume that there are two block
conditions, A and B, where A and B are identical with the exception
that B is delayed with respect to A by one TR. These two regressors
are highly correlated so if a voxel responds only to condition A, it
will be extremely difficult for the GLM to distinguish this from a
voxel that responds only to B. For this reason, the β estimates for
each of the conditions will vary substantially and this would be con-
sidered an inefficient design. If the conditions are designed so that
they have zero correlation (this is achieved by delaying B by half a
block length relative to A), it would be very easy to distinguish
voxels that respond to each of the conditions individually or both of
the conditions together. Therefore, the variance of the β estimates
would be quite small and so the design is efficient. These simple
examples show that efficient task designs come from regressors that
are not correlated with each other. This can be slightly complicated
by the contrasts of interest. For example, say we have a jittered
event-related design where conditions A and B are randomly pre-
sented. If we want to find voxels that respond only to A (i.e., a
contrast matrix of C = [1 0]), only to B (i.e., C = [0 1]) or differ in
their response from A to B (i.e., C = [1 −1]), this design is very effi-
cient. However, if we want to determine voxels that respond equally
to both A and B (i.e., C = [1 1]), then the design is very inefficient
because such a voxel will always have an elevated activation level
and therefore will be indistinguishable from a voxel that does not
respond to either task. The simple idea that regressors must be min-
imally correlated becomes more complicated when multiple condi-
tions, nuisance regressors, and contrasts are placed into a GLM
analysis. The efficiency of a task is related to the covariance of the
design matrix X (i.e., all regressors expressed as columns of a matrix)
and is given the formula:
( )
-1
e = trace C ¢ ´ ( X ¢X ) ´ C
-1
(4)
Experimental Design 149
is the goal. So, despite the HRF being nonlinear at fast presentation
rates, the mismatch with the regressor is compensated by the
increase in trial numbers. Dale also demonstrated that if the ISI var-
ies, the statistical efficiency improves monotonically with decreasing
mean ISI and that the efficiency can be up to ten times greater than
that of a fixed ISI design [34]. These lessons on stimulus timing
suggest that even though the HRF is nonlinear at short ISIs, closely
packed, randomly presented events produce highly efficient designs.
There are two fundamentally different goals when analyzing
event-related fMRI tasks: detection of signal change (which has
been the focus thus far) and estimation of the HRF. Detection of the
signal change involves determining one variable: the amplitude of
the hemodynamic response. More information can be gleaned by
estimating the HRF (e.g., time to onset, rise time, fall time, area
under the curve) which can be used to determine subtle differences
between groups or conditions that may not show up in an amplitude
measure. However, this information comes at a cost: the experimen-
tal task can be optimized for either detection or estimation but not
both. Birn and colleagues showed that the estimation of the HRF is
optimized when stimuli are frequently alternated between task and
control states, whereas detection of activated areas is optimized by
block designs [35]. Liu and colleagues have developed a method
that can simultaneously achieve the estimation efficiency of random-
ized designs and the detection power of block designs at a cost of
increasing the length of the experiment by less than a factor of two
[36]. There are many programs that allow one to randomly (or not
so randomly) generate thousands of task designs in order to choose
the most efficient for the task at hand, be it detection or estimation.
Genetic algorithms (optimization algorithms that code different
designs like chromosomes and allow them to “crossover” and “point
mutate” as they “replicate”) that can produce designs that outper-
form random designs on estimation efficiency, detection efficiency,
and design counterbalancing have also been developed [37]. Further
work has also shown that using advanced mathematical techniques,
block designs, rapid event-related designs, m-sequence designs (ref-
erence time series with an autocorrelation of zero) and mixed designs
can nearly achieve their theoretically predicted efficiency and can be
used in practice to obtain advantageous trade-offs between efficiency
and detection power [38]. It is important when using programs to
design experiments to realize that they may converge on a structure
that may be problematic for the psychological process under investi-
gation (e.g., the most efficient task for detecting activation is a block
design, however, as noted above, if we want to design an oddball
study the oddball events of interest should not occur in a block).
When designing a task, one must also consider the frequencies
at which the events of interest are presented. Analysis packages
often perform high pass filtering to remove low-frequency drifts
from the data. If all frequencies below the limit of, say, 0.01 Hz are
Experimental Design 151
4 Conclusions
References
25. Huettel SA, McCarthy G (2001) The effects ordering brain activity based on relative timing.
of single-trial averaging upon the spatial extent Neuroimage 11:735–759
of fMRI activation. Neuroreport 34. Dale AM (1999) Optimal experimental design
12:2411–2416 for event-related fMRI. Hum Brain Mapp
26. Murphy K, Garavan H (2005) Deriving the 8:109–114
optimal number of events for an event-related 35. Birn RM, Cox RW, Bandettini PA (2002)
fMRI study based on the spatial extent of acti- Detection versus estimation in event-related
vation. Neuroimage 27:771–777 fMRI: choosing the optimal stimulus timing.
27. Murphy K, Bodurka J, Bandettini PA (2007) Neuroimage 15:252–264
How long to scan? The relationship between 36. Liu TT, Frank LR, Wong EC, Buxton RB
fMRI temporal signal to noise ratio and neces- (2001) Detection power, estimation efficiency,
sary scan duration. Neuroimage 34:565–574 and predictability in event-related
28. Cohen MS (1997) Parametric analysis of fMRI fMRI. Neuroimage 13:759–773
data using linear systems methods. Neuroimage 37. Wager TD, Nichols TE (2003) Optimization
6:93–103 of experimental design in fMRI: a general
29. Friston KJ, Holmes AP, Poline JB et al (1995) framework using a genetic algorithm.
Analysis of fMRI time-series revisited. Neuroimage 18:293–309
Neuroimage 2:45–53 38. Liu TT (2004) Efficiency, power, and entropy
30. Worsley KJ, Friston KJ (1995) Analysis of in event-related fMRI with multiple trial types.
fMRI time-series revisited–again. Neuroimage Part II: design of experiments. Neuroimage
2:173–181 21:401–413
31. Smith S, Jenkinson M, Beckmann C, Miller K, 39. Glover GH, Li TQ, Ress D (2000) Image-based
Woolrich M (2007) Meaningful design and method for retrospective correction of physio-
contrast estimability in FMRI. Neuroimage logical motion effects in fMRI:
34:127–136 RETROICOR. Magn Reson Med 44:162–167
32. Bandettini PA, Cox RW (2000) Event-related 40. Birn RM, Diamond JB, Smith MA, Bandettini PA
fMRI contrast when using constant interstimu- (2006) Separating respiratory-variation- related
lus interval: theory and experiment. Magn fluctuations from neuronal-activity-related fluctu-
Reson Med 43:540–548 ations in fMRI. Neuroimage 31:1536–1548
33. Miezin FM, Maccotta L, Ollinger JM, Petersen 41. Birn RM, Murphy K, Handwerker DA,
SE, Buckner RL (2000) Characterizing the Bandettini PA (2009) fMRI in the presence of
hemodynamic response: effects of presentation task-correlated breathing variations.
rate, sampling procedure, and the possibility of NeuroImage 47(3):1092–1104
Chapter 6
Abstract
This chapter describes the procedures applied to fMRI data prior to their statistical analysis. This usually
begins with converting the data from original MR format to a form that can be used by the analysis soft-
ware. The data are then motion corrected. If an anatomical scan is collected for the subject, then it would
be coregistered with the fMRI, and may serve to estimate the warps needed to spatially normalize the
fMRI to some standard space. The final processing step is usually to smooth the data.
Key words Generative model, fMRI, Registration, Artifact correction, Spatial normalization,
Smoothing
1 Introduction
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_6, © Springer Science+Business Media New York 2016
155
156 John Ashburner
1
A voxel is a three dimensional pixel, and can be thought of as a “volume ele-
ment”, as opposed to a “picture element”.
2
See http://nifti.nimh.nih.gov/.
Processing fMRI Data 157
the NIfTI-2 format. Standards have been agreed on how the data
format should be used, with the aim of making it easier to mix dif-
ferent software packages. Providing that only NIfTI compliant
software is used, there should no longer be any confusion about
the orientation of the brains within the images.
Images are usually treated as an array of voxels. For example,
an anatomical image is generally treated as a 3D array, and most
packages require this volume to be stored in a single file. DICOM
usually stores each slice separately (as a series of 2D arrays), but
most file format conversion routines will stack these slices together
into a 3D volume. A run of fMRI data is usually considered as a 4D
array, although for many procedures, it is often convenient to treat
it as a time-series of 3D arrays. Some packages assume that the
entire run is saved in a single file (e.g., FSL), whereas others treat
the data as a series of files containing 3D volumes (e.g., SPM).
The NIfTI format allows storage on disk to be in either a left- or
right-handed coordinate system. However, the format includes an
implicit spatial transformation into a right-handed coordinate system.
This transform maps from data coordinates (e.g., column i, row j,
slice k), into some real world (x,y,z) positions in space. These posi-
tions could relate to Talairach & Tournoux (T&T) space [1],
Montreal Neurological Institute (MNI) space [2, 3], or patient-
based scanner coordinates. For T&T and MNI coordinates, x
increases from left to right, y increases from posterior to anterior, and
z increases in the inferior to superior direction. MRI data are usually
exported from the scanner as DICOM format, which encodes the
positions and orientations of the slices. When data are converted
from DICOM to NIfTI format, the relevant position and orientation
information can be determined from the “Pixel Spacing”, “Image
Orientation” and “Image Position” fields of the DICOM files.
Terms such as “neurological” and “radiological convention”
relate only to visualization of axial images. They are unrelated to
how the data are stored on disk, or even how the real-world coor-
dinates are represented. It is more appropriate to consider whether
the real-world coordinates system is left- or right-handed. T&T
use a right-handed system, whereas the storage convention of
ANALYZE files is usually considered as left-handed (x increases
from right to left). These coordinate systems are mirror images of
each other, so transforming between left- and right-handed sys-
tems involves flipping, and cannot be done by rotations alone.
3.1 Artifact There are a number of image artifacts that result from the very fast
Correction acquisition times required for fMRI. Many of these have detrimental
effects if not properly modelled. Some groups have developed in-
house software to improve on the algorithms supplied by scanner
manufacturers for reconstructing images from the original K-space
data. The objectives of these custom reconstruction algorithms
include reducing Nyquist ghosting artifacts and ensuring that the
model uses a better trajectory through K-space. Sites that perform
their own image reconstruction require the original complex K-space
data, for which there is no clearly defined DICOM standard.
The introduction of a subject into the scanner causes distortions
of the magnetic field [4]. The field may be uniform when there is no
subject present; but with a subject in the scanner, the field is influ-
enced (through Maxwell’s equations) by the varying magnetic suscep-
tibilities of different tissues. These effects are especially prominent at
the interface between tissue and air, resulting in (for example) drop-
out and distortion in the frontal lobe in regions close to the nasal
sinuses. For echo-planar images (EPI), the main effects of magnetic
field inhomogeneity are spatial distortions in the phase-encoding
direction of the images, and dropouts (signal loss) that arise because
of through-plane de-phasing. Some of the distortions can be reduced
by active shimming (changing the field of the scanner via the shim
coils), or by passive shimming (introducing diamagnetic material into
the orifices of the subject), but these measures only reduce the effects
of distortions and dropouts, and cannot completely counteract them.
The models used for intra-subject registration of the head
often assume rigid-body movement. Obtaining accurate alignment
of a relatively distortion-free anatomical image with highly dis-
torted fMRI data is not possible, unless the geometric distortions
are corrected. Therefore, one of the first steps is often a correction
for these distortions. Retrieving signal that is lost in dropouts is
not possible, but there are a number of possible post-hoc approaches
for correcting geometric distortions in the images.
● It is possible to compute field maps from additional scans that
are normally collected just prior to the fMRI runs [4, 5] (see
Fig. 1). This involves processing complex data (i.e., real and
imaginary, or phase and magnitude) from these measurements
in order to compute an unwrapped version of the phase. Phase
measurements are in the range of −π to π radians, or are from
0 to 2π radians. Phase unwrapping [6, 7] involves trying to add
or subtract multiples of 2π to the values, such that the result is
as spatially smooth as possible. With appropriate rescaling, this
unwrapped phase map is converted into a voxel-displacement
map for correcting the fMRI.
Processing fMRI Data 159
● If the air, bone and other tissue can be segmented from the ana-
tomical images, then it becomes possible to simulate field maps
[8] by solving Maxwell’s equations. Separating air from bone is
quite difficult from MRI scans, as both generally appear dark.
Air appears dark because of its low proton density, whereas hard
tissue such as bone has a very short T2 relaxation time, so most
of the signal has decayed before it is detected. Additional prior
information generated from computed tomography (CT) scans
is generally needed in order to attempt such segmentation [9,
10]. For this reason, the approach has not been widely adopted.
● Image registration procedures can also be used to estimate the
warps that align a distorted fMRI scan with a (relatively)
distortion-free anatomical image [11, 12]. Contrast differ-
ences between the images mean that some form of
information-theoretic objective function is usually required,
and the effects of signal dropout in the fMRI should also be
taken into consideration [13].
The field map approach is generally the most accurate way to
correct geometric distortions, although a correction that combines
all of the above strategies into a single model is likely to be the
most accurate.
160 John Ashburner
3
These would actually be Type III errors, rather than Type I, because the null
hypothesis would be correctly rejected (there is a statistically significant effect
in the data) but for the wrong reason (the effect is due to motion, rather than
BOLD signal).
Processing fMRI Data 163
4 Inter-Modality Registration
5 Spatial Normalization
Fig. 2 The top row shows orthogonal sections of two MR images of different contrasts. Below this are joint
intensity histograms of the image pair, both before and after image registration (note that the pictures show
log(1 + N), where N is the count in each histogram bin)
p ( q | D) = p ( D | q) p ( q) / p ( D)
5.1 Matching Criteria The matching criterion is often based upon minimizing the
mean-squared differences or maximizing the correlation between
the image and template. For this criterion to be successful, it
requires the individual’s image to have the visual appearance of a
warped version of the template. In other words, there must be
correspondence in the gray levels of the different tissue types
between the images. The mean-squared difference objective
function makes a number of assumptions. If the image data do
not meet these assumptions, then the objective function may not
accurately reflect the goodness of fit, and the estimated deforma-
tions will be suboptimal. Under some circumstances, it may be
better to weight different regions to a greater or lesser extent.
For example, when spatially normalizing a brain image contain-
ing a lesion, the mean squared difference around the lesion
should contribute little or nothing to the objective function [42].
This is currently achieved by assigned lower weights for the
matching criterion in these regions, so that they have much less
influence on the final solution.
In addition to modeling geometric deformations of the tem-
plate, there may also be extra parameters within the model that
describe intensity variability. A very simple example would be the
inclusion of an additional intensity scaling parameter, but the mod-
els can be much more complicated. There are many possible objec-
tive functions, each making a different assumption about the data
and requiring different parameterizations of the template intensity
distribution. For example, matching can be based on feature vec-
tors derived from the images [43], or can rely on some information
theoretic model [44]. There is no single universally best criterion
to use for all data.
Often, it is necessary to process the anatomical images prior to
any attempt to align them with a template. This may involve strip-
ping off non-brain tissue from the image, which can improve the
accuracy with which the brains themselves are registered. Because
the interesting signal predominantly arises in gray matter, another
strategy for increasing spatial normalization accuracy is to simply
spatially normalize the data by aligning gray matter with a gray
matter template image. A slightly better approach would involve
simultaneously matching gray with gray and white with white.
There are a number of readily available tissue segmentation algo-
rithms that can be used for identifying gray matter in brain MRI.
Another strategy that can be of great benefit for increasing
registration accuracy is the correction of intensity nonunifor-
mity artifact [33, 34], which would otherwise prevent accurate
Processing fMRI Data 169
Fig. 3 2D displacements generated from two scalar fields. The first two panels pf the top row show displace-
ments represented as images. Below these are different representations of these components. The deforma-
tion field resulting from combining the components is overlaid on the top-right image, in order to deform the
image as shown at the bottom-right
Processing fMRI Data 171
4
The word “topology” is used in the same sense as in “Topological Properties
of Smooth Anatomical Maps” [15]. If spatial transformations are not one-to-
one and continuous, then the topological properties of different structures
can change.
172 John Ashburner
Fig. 4 Polynomial basis functions (left) and Cosine transform basis functions (right). Horizontal and vertical (and
through-plane for 3D) displacement fields may be modeled by linear combinations of such functions (see Fig. 3)
Fig. 5 B-spline basis functions allow more detailed warps to be estimated than polynomial or cosine transform
basis functions
5
A diffeomorphism is a globally one-to-one (bijective) smooth and continu-
ous mapping with derivatives that are invertible (i.e., non-zero Jacobian
determinant).
174 John Ashburner
Fig. 6 This figure illustrates the effect of different types of regularization. The top row on the left shows simu-
lated 2D images of a circle and a square. Below these is the circle after it has been warped to match the
square, using both membrane and bending energy priors. These warped images are almost visually indistin-
guishable, but the resulting deformation fields using these different priors are quite different. These are shown
on the right, with the deformation generated with the membrane energy prior shown above the deformation
that used the bending energy prior
6 Smoothing
Fig. 8 This figure shows the effect of convolving the image on the left with different kernels. In the center, the
image has been convolved with a circular kernel. The result is an image in which each pixel is the average of
the values from the original image within the radius of the kernel. At the right is a result from convolving with
a Gaussian kernel. Pixels in this image are weighted averages, where the weights depend on the distance from
the center of the kernel
Processing fMRI Data 177
References
1. Talairach J, Tournoux P (1988) Coplanar 11. Studholme C, Constable RT, Duncan JS
stereotaxic atlas of the human brain. Thieme (2000) Accurate alignment of functional EPI
Medical, New York data to anatomical MRI using a physics-based
2. Evans AC, Collins DL, Milner B (1992) An distortion model. IEEE Trans Med Imaging
MRI-based stereotactic atlas from 250 young 19(11):1115–1127
normal subjects. Soc Neurosci Abstr 18:408 12. Kybic J, Thévenaz P, Nirkko A, Unser M (2000)
3. Evans AC, Collins DL, Mills SR, Brown ED, Unwarping of unidirectionally distorted EPI
Kelly RL, Peters TM. 3D statistical neuroana- images. IEEE Trans Med Imaging 19(2):80–93
tomical models from 305 MRI volumes. In: Proc 13. Li Y, Xu N, Fitzpatrick JM, Morgan
IEEE Nuclear Science Symposium and Medical VL, Pickens DR, Dawant BM (2007)
Imaging Conference. 1993. pp 1813–1817 Accounting for signal loss due to dephas-
4. Jezzard P, Clare S (1999) Sources of distor- ing in the correction of distortions in
tion in functional MRI data. Hum Brain Mapp gradient-echo EPI via nonrigid reg-
8(2):80–85 istration. IEEE Trans Med Imaging
5. Hutton C, Bork A, Josephs O, Deichmann R, 26(12):1698–1707
Ashburner J, Turner R (2002) Image distor- 14. Hajnal JV, Mayers R, Oatridge A, Schwieso JE,
tion correction in fMRI: a quantitative evalua- Young JR, Bydder GM (1994) Artifacts due to
tion. Neuroimage 16(1):217–240 stimulus correlated motion in functional imag-
6. Cusack R, Papadakis N (2002) New robust 3-D ing of the brain. Magn Reson Med 31:289–291
phase unwrapping algorithms: application to 15. Thévenaz P, Blu T, Unser M (2000)
magnetic field mapping and undistorting echo- Interpolation revisited. IEEE Trans Med
planar images. Neuroimage 16(3):754–764 Imaging 19(7):739–758
7. Jenkinson M (2003) Fast, automated, N‐ 16. Eddy WF, Fitzgerald M, Noll DC (1996)
dimensional phase‐unwrapping algorithm. Improved image registration by using Fourier
Magn Reson Med 49(1):193–197 interpolation. Magn Reson Med 36:923–931
8. Jenkinson M, Wilson J, Jezzard P (2004) A 17. Cox RW, Jesmanowicz A (1999) Real-time 3D
perturbation method for magnetic field calcu- image registration for functional MRI. Magn
lations of non-conductive objects. Magn Reson Reson Med 42:1014–1018
Med 52(3):471–477 18. Noll DC, Boada FE, Eddy WF (1997) A
9. Poynton C, Jenkinson M, Whalen S, Golby AJ, spectral approach to analyzing slice selection
Wells W III (2008) Fieldmap-free retrospective in planar imaging: optimization for through-
registration and distortion correction for EPI- plane interpolation. Magn Reson Med
based functional imaging. In: Proc Medical 38:151–160
Image Computing and Computer-Assisted 19. Andersson JLR, Hutton C, Ashburner J,
Intervention (MICCAI). Springer, Berlin- Turner R, Friston KJ (2001) Modeling geo-
Heidelberg, pp 271–279 metric deformations in EPI time series.
10. Poynton C, Jenkinson M, Wells W III (2009) NeuroImage 13:903–919
Atlas-based improved prediction of magnetic 20. Bannister PR, Brady JM, Jenkinson M (2007)
field inhomogeneity for distortion correction of Integrating temporal information with a non-
EPI data. In: Proc Medical Image Computing rigid method of motion correction for func-
and Computer-Assisted Intervention (MICCAI). tional magnetic resonance images. Image Vis
Springer, Berlin, pp 951–959 Comput 25(3):311–320
Processing fMRI Data 179
21. Friston KJ, Williams S, Howard R, Frackowiak image contrast on functional MRI image regis-
RSJ, Turner R (1996) Movement-related tration. NeuroImage 67:163–174
effects in fMRI time-series. Magn Reson Med 33. Hou Z (2006) A review on MR image inten-
35:346–355 sity inhomogeneity correction. Int J Biomed
22. Freire L, Mangin JF (2001) Motion correc- Imag 2006, Article ID 49515, 11 pages.
tion algorithms of the brain mapping commu- doi:10.1155/IJBI/2006/49515
nity create spurious functional activations. In: 34. Vovk U, Pernus F, Likar B (2007) A review of
Insana MF, Leahy RM (eds) Proc Information methods for correction of intensity in homo-
Processing in Medical Imaging (IPMI), vol geneity in MRI. IEEE Trans Med Imaging
2082, Lecture Notes in Computer Science. 26(3):405–421
Springer, Berlin, pp 246–258 35. Fox PT (1995) Spatial normalization: origins,
23. Maclaren J, Herbst M, Speck O, Zaitsev M (2013) objectives, applications, and alternatives. Hum
Prospective motion correction in brain imaging: Brain Mapp 3:161–164
a review. Magn Reson Med 69(3):621–636 36. Mazziotta JC, Toga AW, Evans A, Fox P,
24. Woods RP, Mazziotta JC, Cherry SR (1993) Lancaster J (1995) A probablistic atlas of the
MRI-PET registration with automated algo- human brain: theory and rationale for its devel-
rithm. J Comput Assist Tomogr 17:536–546 opment. NeuroImage 2:89–101
25. Holden M, Hill DLG, Denton ERE, Jarosz 37. Sotiras A, Davatzikos C, Paragios N
JM, Cox TCS, Rohlfing T, Goodey J, Hawkes (2013) Deformable medical image registra-
DJ (2000) Voxel similarity measures for 3D tion: a survey. IEEE Trans Med Imaging
serial MR brain image registration. IEEE Trans 32(7):1153–1190
Med Imaging 19(2):94–102 38. Oliveira FP, Tavares JMR (2014) Medical
26. Collignon A, Maes F, Delaere D, Vandermeulen image registration: a review. Comput Methods
D, Suetens P, Marchal G (1995) Automated Biomech Biomed Engin 17(2):73–93
multi-modality image registration based on 39. Rohlfing T (2012) Image similarity and tissue
information theory. In: Bizais Y, Barillot C, Di overlaps as surrogates for image registration
Paola R (eds) Proc Information Processing in accuracy: widely used but unreliable. IEEE
Medical Imaging (IPMI). Kluwer Academic, Trans Med Imaging 31(2):153–163
Dordrecht, pp 263–274
40. Auzias G, Colliot O, Glaunes JA, Perrot
27. Wells WM III, Viola P, Atsumi H, Nakajima S, M, Mangin JF, Trouvé A, Baillet S (2011)
Kikinis R (1996) Multi-modal volume registra- Diffeomorphic brain registration under
tion by maximisation of mutual information. exhaustive sulcal constraints. IEEE Trans Med
Med Image Anal 1(1):35–51 Imaging 30(6):1214–1227
28. Maes F, Collignon A, Vandermeulen D, Marchal 41. Du J, Younes L, Qiu A (2011) Whole brain
G, Seutens P (1997) Multimodality image reg- diffeomorphic metric mapping via integration
istration by maximisation of mutual informa- of sulcal and gyral curves, cortical surfaces, and
tion. IEEE Trans Image Process 16:187–197 images. NeuroImage 56(1):162–173
29. Studholme C, Hill DLG, Hawkes DJ (1999) An 42. Brett M, Leff AP, Rorden C, Ashburner
overlap invariant entropy measure of 3D medi- J (2001) Spatial normalization of brain images
cal image alignment. Pattern Recogn 32:71–86 with focal lesions using cost function masking.
30. Pluim JPW, Maintz JBA, Viergever MA (2003) NeuroImage 14(2):486–500
Mutual-information-based registration of med- 43. Shen D, Davatzikos C (2002) HAMMER:
ical images: a survey. IEEE Trans Med Imaging hierarchical attribute matching mechanism for
22(8):986–1004 elastic registration. IEEE Trans Image Process
31. West J, Fitzpatrick JM, Wang MY, Dawant BM, 21(11):1421–1439
Maurer CR, Kessler RM, Maciunas RJ, Barillot 44. D’Agostino E, Maes F, Vandermeulen D, Suetens
C, Lemoine D, Collignon A, Maes F, Suetens P (2004) Non-rigid atlas-to-image registra-
P, Vandermeulen D, van den Elsen PA, Napel S, tion by minimization of class-conditional image
Sumanaweera TS, Harkness B, Hemler PF, Hill entropy. In: Barillot C, Haynor DR, Hellier
DLG, Hawkes DJ, Studholme C, Maintz JBA, P (eds) Proc Medical Image Computing and
Viergever MA, Malandain G, Pennec X, Noz Computer-Assisted Intervention (MICCAI),
ME, Maguire GQ, Pollack M, Pelizzari CA, vol 3216, Lecture notes in computer science.
Robb RA, Hanson D, Woods RP. Comparison Springer, Berlin-Heidelberg, pp 745–753
and evaluation of retrospective intermodality
brain image registration techniques. J Comput 45. Sled JG, Zijdenbos AP, Evans AC (1998) A
Assist Tomo 1997;21:554–566. non-parametric method for automatic correc-
tion of intensity non-uniformity in MRI data.
32. Gonzalez-Castillo J, Duthie KN, Saad ZS, Chu IEEE Trans Med Imaging 17(1):87–97
C, Bandettini PA, Luh WM (2013) Effects of
180 John Ashburner
46. Wells WM III, Grimson WEL, Kikinis R, Jolesz template estimation for computational anat-
FA (1996) Adaptive segmentation of MRI data. omy. In: Proc IEEE International Symposium
IEEE Trans Med Imaging 15(4):429–442 on Biomedical Imaging (ISBI). pp 173–176
47. van Leemput K, Maes F, Vandermeulen D, 58. Lorenzen P, Davis B, Gerig G, Bullitt E, Joshi
Suetens P (1999) Automated model-based S (2004) Multi-class posterior atlas formation
bias field correction of MR images of the brain. via unbiased Kullback-Leibler template esti-
IEEE Trans Med Imaging 18(10):885–896 mation. In: Barillot C, Haynor DR, Hellier
48. Studholme C, Cardenas V, Song E, Ezekiel P (eds) Proc Medical Image Computing and
F, Maudsley A, Weiner M (2004) Accurate Computer-Assisted Intervention (MICCAI),
template-based correction of brain MRI inten- vol 3216, Lecture notes in computer science.
sity distortion with application to dementia and Springer, Berlin, pp 95–102
aging. IEEE Trans Med Imaging 23(1):99–110 59. Miller MI (2004) Computational anatomy:
49. Fischl B, Salat DH, Busa E, Albert M, Dieterich shape, growth, and atrophy comparison via dif-
M, Haselgrove C, van der Kouwe A, Killiany R, feomorphisms. NeuroImage 23:S19–S33
Kennedy D, Klaveness S, Montillo A, Makris 60. Christensen GE, Rabbitt RD, Miller MI, Joshi
N, Rosen B, Dale AM (2002) Whole brain seg- SC, Grenander U, Coogan TA, Van Essen
mentation: automated labeling of neuroana- DC (1995) Topological properties of smooth
tomical structures in the human brain. Neuron anatomic maps. In: Bizais Y, Barillot C, Di
33:341–355 Paola R (eds) Proc Information Processing in
50. Fischl B, Salat DH, van der Kouwe AJW, Medical Imaging (IPMI). Kluwer Academic,
Makris N, Ségonne F, Quinn BT, Dale AM Dordrecht, pp 101–112
(2004) Sequence-independent segmentation 61. Woods RP, Grafton ST, Holmes CJ, Cherry
of magnetic resonance images. NeuroImage SR, Mazziotta JC (1998) Automated image
23:S69–S84 registration: I. General methods and intra-
51. Ashburner J, Friston KJ (2005) Unified seg- subject, intramodality validation. J Comput
mentation. NeuroImage 26:839–851 Assist Tomogr 22(1):139–152
52. Pohl KM, Fisher J, Levitt JJ, Shenton ME, 62. Woods RP, Grafton ST, Watson JDG, Sicotte
Kikinis R, Grimson WEL, Wells WM II (2005) NL, Mazziotta JC (1998) Automated image
A unifying approach to registration, segmenta- registration: II. Intersubject validation of lin-
tion, and intensity correction. In: Medical Image ear and nonlinear models. J Comput Assist
Computing and Computer-Assisted Intervention Tomogr 22(1):153–165
(MICCAI). Springer, Berlin, pp 310–318 63. Christensen GE (1999) Consistent linear elastic
53. D’Agostino E, Maes F, Vandermeulen D, transformations for image matching. In: Kuba
Suetens P (2006) A unified framework for A, Sámal M, Todd-Pokropek A (eds) Proc
atlas based brain image segmentation and reg- Information Processing in Medical Imaging
istration. In: Pluim JPW, Likar B, Gerritsen (IPMI), vol 1613, Lecture notes in computer
FA (eds) Proc Third International Workshop science. Springer, Berlin, pp 224–237
on Biomedical Image Registration (WBIR), 64. Ashburner J, Friston KJ (1999) Nonlinear spa-
vol 4057, Lecture notes in computer science. tial normalization using basis functions. Hum
Springer, Berlin-Heidelberg, pp 136–143 Brain Mapp 7:254–266
54. Zhang H, Yushkevich PA, Gee JC (2004) 65. Bookstein FL (1989) Principal warps: thin-
Registration of diffusion tensor images. In: plate splines and the decomposition of defor-
Proc IEEE Computer Society conference on mations. IEEE Trans Pattern Anal Mach Intell
Computer Vision and Pattern Recognition 11(6):567–585
(CVPR).pp 842–847. 66. Bookstein FL (1997) Quadratic variation of
55. Khullar S, Michael AM, Cahill ND, Kiehl KA, deformations. In: Duncan J, Gindi G (eds)
Pearlson G, Baum SA, Calhoun VD (2011) Proc Information Processing in Medical
ICA-fNORM: Spatial normalization of fMRI Imaging (IPMI), vol 1230, Lecture notes in
data using intrinsic group-ICA networks. computer science. Springer, Berlin-Heidelberg,
Front Sys Neurosci 5(93):1–18. pp 15–28
56. Joshi S, Davis B, Jomier M, Gerig G (2004) 67. Bro-Nielsen M, Gramkow G (1996) Fast fluid
Unbiased diffeomorphic atlas construction registration of medical images. In: Höhne
for computational anatomy. NeuroImage KH, Kikinis R (eds) Proc Visualization in
23:S151–S160 Biomedical Computing (VBC), vol 1131,
57. Davis B, Lorenzen P, Joshi S (2004) Large Lecture notes in computer science. Springer,
deformation minimum mean squared error Berlin, pp 267–276
Processing fMRI Data 181
68. Thirion JP (1995) Fast non-rigid match- non-rigid registration using a stationary
ing of 3D medical images. Technical report velocity field. In: Proc IEEE Workshop on
no 2547. Institut National de Recherche en Mathematical Methods in Biomedical Image
Informatique et en Automatique Analysis (MMBIA). IEEE. pp 145–150
69. Rueckert D, Sonoda LI, Hayes C, Hill DLG, 79. Miller MI, Trouvé A, Younes L (2006)
Leachand MO, Hawkes DJ (1999) Nonrigid Geodesic shooting for computational anatomy.
registration using free-form deformations: J Math Imag Vis 24(2):209–228
application to breast MR images. IEEE Trans 80. Ashburner J, Friston KJ (2011) Diffeomorphic
Image Process 18(8):712–721 registration using geodesic shooting and
70. Thévenaz P, Unser M (2000) Optimization Gauss–Newton optimisation. NeuroImage
of mutual information for multiresolution 55(3):954–967
image registration. IEEE Trans Image Process 81. Vialard FX, Risser L, Rueckert D, Cotter CJ
9(12):2083–2099 (2012) Diffeomorphic 3D image registration
71. Christensen GE (1994) Deformable shape mod- via geodesic shooting using an efficient adjoint
els for anatomy. Doctoral Thesis. Washington calculation. Int J Comput Vis 97(2):229–241
University, Sever Institute of Technology 82. Lester H, Arridge SR (1999) A survey of hier-
72. Christensen GE, Rabbitt RD, Miller MI (1996) archical non-linear medical image registration.
Deformable templates using large deforma- Pattern Recogn 32:129–149
tion kinematics. IEEE Trans Image Process 83. Klein A, Andersson J, Ardekani BA, Ashburner
5:1435–1447 J, Avants B, Chiang MC, Christensen
73. Beg MF, Miller MI, Trouvé A, Younes L GE, Collins DL, Gee J, Hellier P, Song
(2005) Computing large deformation metric JH, Jenkinson M, Lepage C, Rueckert D,
mappings via geodesic flows of diffeomor- Thompson P, Vercauteren T, Woods RP,
phisms. Int J Comput Vis 61(2):139–157 Mann JJ, Parsey RV (2009) Evaluation of 14
74. Avants B, Gee JC (2004) Geodesic estima- nonlinear deformation algorithms applied to
tion for large deformation anatomical shape human brain MRI registration. Neuroimage
averaging and interpolation. NeuroImage 46(3):786–802
23:S139–S150 84. Fissell K, Tseytlin E, Cunningham D, Carter
75. Ashburner J (2007) A fast diffeomorphic CS, Schneider W, Cohen JD (2003) Fiswidgets:
image registration algorithm. Neuroimage a graphical computing environment for
38(1):95–113 neuroimaging analysis. Neuroinformatics
76. Hernandez M, Bossa MN, Olmos S (2007) 1(1):111–125
Registration of anatomical images using geo- 85. Rex DE, Maa JQ, Toga AW (2003) The LONI
desic paths of diffeomorphisms parameterized pipeline processing environment. NeuroImage
with stationary vector fields. In: Proc IEEE 19(3):1033–1048
11th International Conference on Computer 86. Zijdenbos AP, Forghani R, Evans AC (2002)
Vision (ICCV). IEEE. pp. 1–8. Automatic ‘pipeline’ analysis of 3-D MRI
77. Vercauteren T, Pennec X, Perchant A, Ayache data for clinical trials: application to mul-
N (2008) Symmetric log-domain diffeomor- tiple sclerosis. IEEE Trans Med Imaging
phic registration: a demons-based approach. 21(10):1280–1291
In: Proc Medical Image Computing and 87. Drobnjak I, Gavaghan D, Suli E, Pitt-Francis
Computer-Assisted Intervention (MICCAI). J, Jenkinson M (2006) Development of a
Springer, Berlin, pp 754–761 fMRI simulator for modelling realistic rigid-
78. Modat M, Daga P, Cardoso MJ, Ourselin S, body motion artifacts. Magn Reson Med
Ridgway GR, Ashburner J (2012) Parametric 56(2):364–380
Chapter 7
Abstract
fMRI is a powerful tool used in the study of brain function. It can noninvasively detect signal changes in
areas of the brain where neuronal activity is varying. This chapter is a comprehensive description of the
various steps in the statistical analysis of fMRI data. This will cover topics such as the general linear model
(including orthogonality, hemodynamic variability, noise modeling, and the use of contrasts), multisubject
statistics, and statistical thresholding (including random field theory and permutation methods).
Key words fMRI, Analysis, Statistics, General linear model, Multisubject statistics, Statistical
thresholding
1 Introduction
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_7, © Springer Science+Business Media New York 2016
183
184 Mark W. Woolrich et al.
2.1 fMRI Data In a typical fMRI imaging session, a low-resolution functional vol-
ume is acquired every few seconds (MR volumes are often also
referred to as “images” or “scans”). Over the course of the experi-
ment, 100 volumes or more are typically recorded. In the simplest
possible experiment, some images will be taken while stimulation1 is
applied, and some will be taken with the subject at rest. Because the
images are taken using an MR sequence which is sensitive to changes
1
For the remainder of this chapter, reference to “stimulation” should be taken
to include also the carrying out of physical or cognitive activity.
Statistical Analysis of fMRI Data 185
Fig. 1 Illustration of the analysis steps carried out in a typical fMRI group study. There are three main compo-
nents in the process. First, the individual subjects’ fMRI data must be processed (top). The information gleaned
from this about the effect sizes (the size of the fMRI signal change in response to the experimental task) for
each subject are then combined in a group analysis (middle). The group effect sizes statistic images are then
statistically thresholded to find significant brain areas (bottom)
186 Mark W. Woolrich et al.
Fig. 2 What are voxels? Shown here are surface renderings of 3D brain images. On the left is a high-resolution
image, with small (0.5 × 0.5 × 0.5 mm) voxels; the voxels are too small to see. On the right is a low-resolution
image of the same brain, with large (5 × 5 × 5 mm) voxels, clearly showing the voxels making up the image
Fig. 3 An example time series at a strongly activated voxel from a visual stimulation experiment. Here the
signal is significantly larger than the noise level. Periods of stimulation are alternated with periods of rest—a
complete stimulation—rest cycle lasts 20 scans
images that actually look like brains. The data is then motion cor-
rected, where each volume is transformed (using rotation and
translation) so that the image of the brain within each volume is
aligned with that in every other volume. Spatial smoothing is then
carried out, principally to reduce noise, hopefully without signifi-
cantly affecting the activation signal. Finally, each voxel’s time
series is temporally high-pass filtered with a filter designed to remove
the large amount of low-frequency temporal noise found in FMRI
data, without removing the signal of interest.
Chapter 6 has already covered fMRI pre-processing in much
more detail, including other optional steps that have not been
mentioned here.
2.3 Predicting The first step in the statistical analysis is to come up with a good
the Response prediction, or model, of what we think the measured fMRI signal
response will look like in voxels that are active. In the simplest type
of fMRI experiment, we alternate periods of stimulation with peri-
ods of rest, in what we will refer to as a square-wave block design,
as shown in Fig. 4 (left). We expect that a voxel which is active in
response to the stimulus will contain an fMRI signal that generally
fluctuates up and down with a time course that is similar to the
stimulus time course (Fig. 4 left), whereas an inactive voxel will not.
2.4 Hemodynamic However, can we come up with a better prediction of the fMRI
Response Function signal than Fig. 4 (left)? In particular, we know, from experiment,
that the response to a very short stimulus looks like the curve
shown in Fig. 5. We refer to this response to an impulse of stimulus
188 Mark W. Woolrich et al.
Fig. 4 Predicting the response using the known stimulus timings. This example is a square-wave block design
where blocks of stimulation are alternated with blocks of rest. The square-waveform (left) describes the input
stimulus timing; the predicted response (middle) results from convolving the stimulus time course with the
hemodynamic response function and then sampling it at the temporal resolution of the experiment. This
experiment has a repetition time (TR) of 2 s. This process produces a model, or predicted response, that looks
much more like the data measured in voxels that are responding to the stimulus (right)
This essentially assumes that the effects of the different impulses that
make up the stimulus time course add together in an additive, linear
fashion [1–4]. Figure 4 (middle) shows the result of convolving the
HRF in Fig. 5 with the square-wave stimulus in Fig. 4 (left) to form
our new improved predicted response. Strictly speaking, making the
assumption of linearity is incorrect. We will see later how we can
address this issue, and also discuss how to tell if that is necessary.
For now, using the HRF and convolving it with the stimulus
time course provides us with a way in which we can come up with a
reasonable prediction of the response for any general stimulus type.
For example, Fig. 6 shows the predicted response for a sparse single-
event design and a dense randomized single-event design. Armed
with our predicted response, we can then look to find those voxels
that have fMRI time courses that match the predicted response well
and, if they pass a statistical test, label them as being active voxels.
2
The particular HRF in Fig. 5 has parameter values μ1 = 6 s, σ1 = 2.45 s,
μ2 = 16 s, σ2 = 4 s, and ρ = 6.
190 Mark W. Woolrich et al.
Fig. 6 Using the hemodynamic response function (HRF) and convolving it with the stimulus time course pro-
vides a way in which we can predict the response for any general stimulus type. Here we can see the predicted
response for a sparse single-event design, where short bursts of 0.1 s stimulation are 25 s apart (top), and for
a randomized single-event design, where short bursts of 0.1 s stimulation are separated with inter-stimulus
intervals (ISIs) sampled from a Poisson distribution with mean of 7 s (bottom). Randomized single-event
designs are an excellent way of working with stimuli that by their nature need to be single events, as they
generally have better sensitivity than sparse single-event designs [5]
y ( t ) = b1 x1 ( t ) + b 2 x2 ( t ) + c + e ( t ) (3)
where y(t) is the data in one voxel, and is a 1D vector (time course)
of intensity values with one value for each time point. x1(t) and
x2(t) are the predicted responses for our auditory and visual experi-
mental stimuli, respectively, and both are also 1D time courses
with one value for each time point. c is a constant and would cor-
respond to the mean intensity value in the data. The linear combi-
nation of the predicted responses needed to explain the data in a
particular voxel is described by the parameters β1 and β2. e(t) mod-
els the noise that is present in fMRI data.
Model fitting involves adjusting the mean level, c, and the
parameters β1 and β2, to best fit the data. For example, if a particu-
lar voxel responds strongly to model x1, the model-fitting will find
a large value for β1; if the data instead looks more like the second
Statistical Analysis of fMRI Data 191
model time course, x2, then the model-fitting will give β2 a large
value. The GLM is used to analyze each voxel’s time series inde-
pendently. This is often referred to as a mass univariate analysis,
and outputs statistics independently at each voxel.
There are a number of different of names that get used to
describe the different components of the GLM. The predicted
responses within a GLM are often referred to as explanatory vari-
ables (EVs), as they explain different processes in the data. They can
also be referred to as regressors, and the βs as regression parameters,
as we are performing what is also known as a multiple regression.
The regression parameters, β, are also sometimes referred to as
effect sizes, as they describe the size of the response to the corre-
sponding underlying experimental stimuli.3
3.1 Design Matrix The GLM is often formulated in matrix notation. All of the param-
eters are grouped together into a P × 1 vector β (where P is the
number of EVs) and all of the EVs are grouped together into an
N × P matrix X, often referred to as the design matrix, where N is
the number of time points in the experiment. This gives us the
GLM in the following form:
¡ = Xb +e (4)
3.2 Fitting the GLM The GLM is fit to the data at each voxel separately. This is achieved
to the Data by adjusting the estimates of the regression parameters to find the
best fit of the model to the data. Typically, it is assumed that the
fMRI noise, e, is well modeled by a Gaussian distribution with a
3
Note that this common usage is slightly different from the definition some-
times used in the statistics literature, where effect size means β divided by the
noise level.
192 Mark W. Woolrich et al.
Fig. 7 Example of the general linear model (GLM) for an experiment containing
auditory and visual stimuli which have different stimulus timings. The design
matrix, X, contains two predicted responses (also known as regressors or
explanatory variables): x1 for the auditory stimulus and x2 for the visual stimu-
lus—see Eq. (4). In this, visualization time is running downwards. It can be seen
that the particular voxel data shown, Y, is from a voxel that is strongly activating
in response to the visual stimulus modeled by x2, but not to the auditory stimulus
modeled by x1
å(¡ - Xt b )
2
t (5)
t
b = ( X T X ) X T ¡
-1
(6)
3.3 Temporal Until now we have considered a rather simple approach to dealing
Autocorrelation with the noise that is present in fMRI, by assuming that it is well
Statistical Analysis of fMRI Data 193
Fig. 8 A summary of the process of pre-whitening. (a) Plot of the power spectrum of the residuals from an
initial general linear model (GLM) fit to the data from one voxel. (b) Estimated power spectrum (representing
the temporal autocorrelation estimate) obtained from fitting an autocorrelation model. (c) This spectrum is
inverted to create the frequency characteristics of a temporal filter designed to “undo” the autocorrelation. (d)
The pre-whitening temporal filter is applied to both the data and the explanatory variables (EVs) in the design
matrix, and then this pre-whitened GLM is refit to the pre-whitened data. The residuals that result from this
refit of the GLM should now be approximately white, that is have a flat power spectrum
194 Mark W. Woolrich et al.
3.4 Inferring Neural When we fit a particular GLM to a particular voxel’s data, we get
Activity regression parameter estimates that indicate how much of each EV is
needed to explain what we see in the data. If the parameter estimate
of β for any particular EV is nonzero, then it might seem reasonable
to assume that the voxel in question is neuronally responding to the
stimulus that the EV represents. However, we only have estimates/
approximations of the true β obtained from a limited amount of
noisy fMRI data. So, given the amount of noise and the estimate of β
obtained, how much can we trust that any particular β is nonzero?
Statistical Analysis of fMRI Data 195
b
t= (7)
( )
std b
P-Value=0.98
P-Value=0.02
P-Value=0.2
−4 −2 0 t 4 −4 −2 0 t 2 4 −4 t 0 2 4
T T T
Fig. 9 Performing a T-test. The T-statistic we calculate for each regression parameter estimate is compared
with the distribution of T-statistics we would expect if the true regression parameter were zero. This “null
distribution” is a standard T-distribution. Here we are using a T-distribution with 100 DOF (this would corre-
spond to about 100 time points in the data). We calculate a probability, or P-value, as the proportion of the area
under the curve in the positive tail of the distribution defined by our T-statistic, t. A low probability, or P-value,
(left plot) of the null hypothesis being true means that we can more confidently reject the null hypothesis and
label the voxel as having a nonzero β, and therefore as being neuronally activated by the stimulus that the β
in question represents. T-statistics that have larger P-values (middle plot) or are deep into the negative tail of
the distribution (right plot) have high P-values. In the latter case, this might seem a bit counterintuitive as we
have a T-statistic that is in the extremities of the null distribution. However, this is because the test is direc-
tional and so we calculate the P-value by looking in the positive tail of the distribution only
196 Mark W. Woolrich et al.
cT b
t= (8)
( )
var c T b
( ) = s c T ( X T X ) c
-1
2
var c T b (9)
rTr
s =
2
(10)
( N - P)
where r is the residual, that is, an estimate of the error, e, and is what
.
is left over after the model is fit to the data, and is given by r = ¡ - X b
Statistical Analysis of fMRI Data 197
Table 1
Examples of contrasts that might be used in the two stimulus auditory/visual experiment
Contrast,
cT COPE, cTβ Meaning
[1 0] β1 Where is there significant auditory activation?
[0 1] β2 Where is there significant visual activation?
[−1 0] −β1 Where is there significant negative auditory activation?
[0 −1] −β2 Where is there significant negative visual activation?
[1 −1] β1 − β2 Where is there auditory activation significantly greater than visual activation?
[−1 1] β2 − β1 Where is there visual activation significantly greater than auditory activation?
[1 1] β1 + β2 Where is there significant activation averaged across both conditions?
æ1 0ö
cT = ç ÷.
è0 1ø
( )
b c var c T b c T b
T
f = . (11)
K
æ 1 0 ö æ -1 0 ö æ -1 0 ö æ 1 0 ö
ç ÷,ç ÷,ç ÷,ç ÷.
è 0 1 ø è 0 -1ø è 0 1 ø è 0 -1ø
3.6 Interaction It is possible that the response to two different stimuli, when
Example applied simultaneously, is greater than that predicted by adding up
the responses to the stimuli when applied separately. If this is the
case, then such “nonlinear interactions” may need to be allowed
for in the model. The simplest way of doing this is to set up the
P-Value=0.2
P-Value=0.2
0 1 f 2 3 4 5 6 −4 −2 -t 0 t 2 4
F T
Fig. 10 Here we are considering an F -test that consists of just one contrast. In this case, the F-test (left) is
equivalent to a two-tailed T-test (right) on the same contrast, with the relationship f = t2
200 Mark W. Woolrich et al.
two originals EVs, and then add an interaction EV, which will only
be “up” when both of the original EVs are “up” and “down” oth-
erwise. In Fig. 11, EV1 could represent the application of a drug
and EV2 could represent visual stimulation. EV3 will model the
extent to which the response to drug + visual is greater than the
sum of drug-only and visual-only. A contrast of [0 0 1] will show
this measure, whereas a contrast of [0 0 − 1] shows where negative
interaction is occurring. An F-contrast of
æ1 0 0 ö
ç ÷
è0 1 0ø
3.7 Converting T- and F-statistics can be converted to Z-statistics, that is, statistics
T- and F-Statistics that are distributed as a standardized Normal (Gaussian) d istribution.
into Z-Statistics This is simply achieved by ensuring that the P-value is the same
regardless of which statistic is used, so to convert from a T- to
Z-statistic, we calculate the P-value for the given T-statistic and then
Statistical Analysis of fMRI Data 201
determine the Z-statistic as being the one that gives the same P-value.
One reason for doing this is so that we can compare statistics using
a common currency. Another reason is so that we can perform
generic thresholding techniques (such as described in Sect. 7), using
Z-statistic maps, regardless of whether we have done T- or F-tests.
Hb
%change = 100 . (12)
C
3.9 Issues We described in Sect. 3.2 how we obtain regression parameter esti-
with Orthogonality mates by finding the best fit of the GLM to the data (by using Eq.
and Estimating (6). The regression parameter estimates describe how much we
Contrasts need of each EV to explain what we see in the data. However, con-
sider what would happen in a poorly designed experiment, where
two different stimuli are switched on and off at very similar times.
The resulting predicted responses (EVs) are highly correlated.
Note that we use the terms correlated and non-orthogonal (simi-
larly uncorrelated and orthogonal) interchangeably. The top of
Fig.13 shows a design matrix containing our two very similar EVs.
The model fitting will determine the regression parameter esti-
mates b and b and describe how much we need of each EV to
1 2
explain what we see in the data. However, because EV1 and EV2
are so similar, we can equally well use either EV1 or EV2 to explain
what we see in the data. The result is that we cannot estimate either
b 1 or b 2 , separately from each other, very well. Mathematically,
we say that the design matrix is not of “full rank,” or that it is “rank
deficient.”
To understand this, consider solving two simultaneous equa-
tions. If we have two unknowns to solve for, then we need two
equations to solve for them. However, if the two equations are the
same, then we really have only one equation and we cannot solve
for the two unknowns.
The good news is that our statistical tests (T- and F-tests) take
this all into account. When two EVs are highly correlated, the
appropriate variances of the regression parameter estimates (see Eq.
(9)) are automatically increased—acknowledging the fact that we
cannot determine which EV should explain what in the data. So,
even though the statistics accounts for orthogonality, it is clear that
when we design our experiments, we want to avoid this being an
issue whenever possible. This can be achieved by using approaches
that assess the efficiency of experimental designs such as [16–18].
In Fig. 14, we can see three different design matrices. The first
contains two EVs that are highly correlated, the second shows two
EVs that are partly correlated, and the third shows two EVs that
are completely uncorrelated. As discussed, the first design matrix is
“rank deficient.” The third design matrix is the ideal scenario in
Statistical Analysis of fMRI Data 203
Fig. 13 Rank-deficient design matrices. At the top, we can see a general linear model (GLM) with a design
matrix containing two identical explanatory variables (EVs). Since EV1 and EV2 are so similar, we can equally
well use either EV1 or EV2 to explain what we see in the data. At the bottom of the figure, we can see examples
+b
of identically good model fits with any linear combination of the EVs as long as b = 0.9. The result is
1 2
that we cannot estimate either b 1 or b 2 with any certainty (corresponding to [1 0] or [0 1] contrasts), but we
+b
can estimate b (corresponding to a [1 1] contrast)
1 2
Fig. 14 Examples of different design matrices. Design matrix with two explana-
tory variables (EVs) that are highly correlated (left). Design matrix with two EVs
that are partially correlated (middle). Design matrix with two EVs that are uncor-
related (right)
Fig. 15 Effects of orthogonalization. We, first, fit the design matrix (containing two
partially correlated explanatory variables, EVs) on the left to the data from a voxel,
and obtain regression parameter estimates b and b . We then construct a new
1 2
design matrix, shown on the right, where EV1 is unchanged and EV2 is the old EV2
orthogonalized with respect to EV1. We then fit this new design matrix to the same
data and obtain new regression parameter estimates b and b . The counterin-
1 2
tuitive result is that even though it is EV2 that has been changed and EV1 that has
remained the same, it is b that has changed and b that remains the same. The
1 2
underlying reason for this is that the model fitting can only ever be driven by the
orthogonal components of the EVs
but it turns out that if we take two times Eq. (1) and subtract Eq.
(2), then we get exactly Eq. (3). In that case, we actually only really
have two equations to solve for our three unknowns, and so we are
in trouble. In the GLM, the EVs are analogous to the equations, and
the regression parameters are the unknowns. And so we have a prob-
lem if any EV is the same (or close to being the same) as a weighted
sum of the other EVs in the design matrix. Again, we describe such
a design matrix as being (or close to being) “rank deficient.”
Even if we do have a design matrix that is close to being rank
deficient, and therefore, there are some regression parameters that
cannot be very well estimated, there may well be other regression
parameters that can be estimated. There may even be contrasts that
actually include the hard-to-estimate regression parameters, but
that can still be estimated [16]. At first glance, this may seem a
little counterintuitive. However, things should become clear if we
consider a simple example of this. This occurs when we have the
situation shown in Fig. 13 where we had a design matrix with two
very similar EVs. As already discussed, we cannot estimate at all
well the individual parameters β1 and β2 with [1 0] and [0 1] con-
trasts. However, we can estimate a [1 1] contrast since this does
206 Mark W. Woolrich et al.
4.1 HRF Basis Sets A popular alternative is to use the approach of basis functions.
These allow HRF modeling flexibility but within the computation-
ally undemanding GLM framework [19]. Figure 16a shows just
one example of an HRF basis set that contains three basis func-
tions. The choice of basis set is clearly important and we will come
to that later. Whatever basis set is used, the principle is the same:
different linear combinations of the basis sets can be used to give
different HRF shapes. This is illustrated in Fig. 16b.
But how do we use these HRF basis functions in combination
with our known stimulus timings to create predicted responses
that can be used in the GLM? The answer is to separately convolve
each of the HRF basis functions with the stimulus function to cre-
ate an EV for each of the basis functions, as shown in Fig. 17.
When the resulting design matrix is fit to the fMRI data, the
required linear combination of these EVs is determined. If desired,
the same linear combination can then be applied to the HRF basis
functions to show the implied HRF shape.
The question remains as to how we set up a statistical test to
ask, for example, “Where is there significant activation due to
Statistical Analysis of fMRI Data 207
Fig. 16 (a) Example of a hemodynamic response function (HRF) basis set con-
taining three basis functions. (b) Different linear combinations of the basis func-
tions in the basis set can be used to obtain different of HRF shapes
æ1 0 0ö
ç ÷
c = ç0 1 0÷
T
ç0 0 1÷
è ø.
Fig. 17 Setting up design matrix explanatory variables (EVs) using a hemodynamic response function (HRF)
basis set. Each HRF basis function is separately convolved with the stimulus function to create an EV for each
of the basis functions
æ1 0 0 -1 0 0 ö
ç ÷
c = ç0 1 0
T
0 -1 0 ÷
ç0 0 1 0 0 -1÷ø.
è
4.1.1 Choosing Figure 16a showed an example basis set that had been derived
a Basis Set from a parameterized HRF that was made up of a series of half-
cosine functions [15]. This was obtained by sampling thousands of
example HRFs from within a range of plausible parameter values
for the HRF model (Fig. 18), and then a principal component
analysis was carried out on these samples to determine the principal
modes/components of variation in the HRF shape. The three
highest principle components were then used as the basis set. Note
that this can equally well be done on any form of parameterized
HRF, for example, a double gamma HRF or a biophysical model
such as the balloon model [25].
When this approach is taken with any plausible HRF model, it
is typical for the first basis function to turn out to be the mean HRF
shape, or a “canonical” HRF, for the second to approximate the
temporal derivative (i.e., linear combinations of the first and second
basis functions result in versions of the canonical HRF shifted in
time), and for the third to approximate a dispersion derivative (i.e.,
linear combinations of the first and third basis functions result in
versions of the HRF with different widths of the main positive
response). Although other basis sets that have been proposed (e.g.,
sets of Gamma functions and finite impulse response functions),
this kind of basis set is highly recommended since it parsimoniously
captures shape variations. It also has the advantage that the first
(canonical) basis function will tend to dominate the fit to the data
and can then be used to determine the positivity or negativity of the
HRF. Note that it is quite common for people to neglect the dis-
persion derivative and use just a temporal derivative since temporal
shifts represent the most important variation in the HRF shape,
particularly when working with boxcar stimuli.
Thus far, we have considered basis sets made up of two or
three basis functions. But why not use more? The reason for this,
a b c
m2
m4
m1 m3 0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30
seconds seconds
Fig. 18 (a) Parameterized hemodynamic response function (HRF) model. (b) Example HRFs sampled from this
parameterized model of the HRF for plausible parameter values. (c) Samples of the HRF obtained from random
linear combinations of the basis set shown in Fig. 16a
210 Mark W. Woolrich et al.
and in general the reason why basis sets with large numbers of basis
functions (e.g., finite impulse response basis sets) are suboptimal,
is that the GLM becomes unrealistically flexible. This problem is
evident even with just three basis functions. Figure 18c shows pos-
sible HRFs that result from random linear combinations of the
basis set in Fig. 16a. Clearly, many of these HRFs are nonsensical.
The problem is that random fluctuations in the fMRI noise can, by
chance, look like these nonsensical HRFs and so we “over-fit” the
model to the noise. The statistical inference (e.g., via F-tests across
the basis functions) is still valid, but we lose sensitivity; it becomes
harder to detect genuine activations. On the contrary, if we use too
few basis functions then we can fail to estimate the true HRF and
our model is then a poor match to the data and again we suffer a
reduction in sensitivity. So there is a trade-off between providing
enough basis functions to provide enough HRF variability, while
not having so many that over-fitting becomes a problem. A general
rule of thumb is that three basis functions are good for single-event
designs and two basis functions are good for boxcar designs.
An increasingly used approach to overcome this problem is to
infer on models that incorporate HRF variability using a Bayesian
framework. One advantage of a Bayesian approach is that prior
information can be included. Priors can be used that prohibit non-
sensical HRFs. Subsequently, more flexibility can be allowed while
protecting against over-fitting [15, 22–24].
4.1.2 Basis Functions In Sect. 6, we will discuss how we model multisession/subject fMRI
and Group Inference data. However, it is worth mentioning how basis functions are best
used when we are ultimately doing a group analysis. In particular, we
consider this in the context of the simple case of inferring a popula-
tion group mean. One option might be to pass up the regression
parameter estimates for all basis functions into the higher-level group
analysis, obtain the group average for each basis function separately,
and then perform an F-test across them at the group level (in the
same manner as we would do in a single-session analysis). However,
it is not clear what benefit there would be of doing this. When our
basis set contains a “canonical” HRF, the other basis functions, such
as the temporal and dispersion derivatives, tend to average out to
zero at the group level due to the different subject HRF shape varia-
tions. Subsequently, an often-recommended approach is to only pass
up to the group level the canonical HRF regression parameter esti-
mates. This makes for a simple group analysis, and the benefits of
including the basis function at the first level are still felt in terms of
accounting for HRF variability that would otherwise cause increased
noise in the first-level analysis.
Another option that can be taken is to calculate a size summary
statistic from the single-session analyses (e.g., the root mean square
of the basis function regression parameter estimates), and pass that
up to the group level. However, it is important to note that this
Statistical Analysis of fMRI Data 211
4.2 Nonlinearities So far we have assumed linearity of the HRF. That is, we have
assumed that the response to a stimulus is well modeled by (linear)
convolution of the stimulus with the HRF. Typically, this is the
approach that people take in the majority of fMRI analyses.
However, it has been shown that this assumption is poor in certain
situations. For example, it can be shown that the response to a
prolonged stimulus is not as large as the one we would predict
from extrapolating results from applying a short stimulus [26, 27],
and nonlinearities are predominant when there are short separa-
tions (less than ∼3 s) between stimuli [28]. Normally, these situa-
tions are intentionally avoided by designing experiments
appropriately. For example, we avoid experiments where single
events are occurring less than approximately 3 s apart, or experi-
ments that require comparisons between a mix of short (e.g.,
single-event) and prolonged (e.g., boxcar) stimuli. However, if
these situations are unavoidable, then it becomes necessary to
model the nonlinearities.
Such nonlinearities are predicted by nonlinear biophysical
models, for example, the balloon model [25]. Hence, one solution
is to model fMRI data using these nonlinear biophysical models
[22]. Another approach that can be used in the GLM setting is to
extend the idea of convolution to include second-order nonlinear
terms using Volterra kernels [28].
where s(t) is the stimulus, the first term contains the traditional
linear HRF first-order kernel, h1(τ), and the second term includes
the second order kernel, h2(τ1, τ2).
Many of the issues surrounding the use of second-order
Volterra kernel basis functions are the same as they are for linear
basis functions. For example, Volterra kernels can be determined
empirically [25, 28], or derived from nonlinear biophysical models
[22]. Either way, as with linear basis functions, there is variability
212 Mark W. Woolrich et al.
5.1 Structured Noise The main problem with such structured noise artifacts is that they
and the GLM can severely impact our GLM-based analysis. The part of the arti-
fact that is orthogonal (uncorrelated) with all of the EVs will sim-
ply not be modeled by the design matrix regressors, and therefore,
the presence of the structured noise effect will be reflected by an
increase in the residual GLM noise variance. This, in turn, will
decrease any T- or F-statistics value, making it harder for us to
detect true activations. The non-orthogonal (correlated) part of
such an artifact, however, will result in wrong parameter estimates
for those EVs that correlate with the artifact. If the correlation is
Statistical Analysis of fMRI Data 213
a b c
0 20 40 60 80 100 120 140 160 180 0 20 40 60 80 100 120 140 160 180 0 20 40 60 80 100 120 140 160 180
d e f
0 20 40 60 80 100 120 140 160 180 0 20 40 60 80 100 120 140 160 180 0 20 40 60 80 100 120 140 160 180
Fig. 19 Examples of “structured noise” identified in a single fMRI dataset using independent component analy-
sis [29]: (a) residual head motion, (b) signal fluctuations in the ventricles, (c) spontaneous fluctuations in the
bilateral sensory motor cortex, (d) fluctuations close to the sinuses (likely due to interactions between B0 field
inhomogeneities and head motion), (e) high-frequency image ghosting, and (f) more spontaneous low-
frequency fluctuations
5.2 Nuisance One possible way of correcting for the negative impact on GLM
Regressors in the GLM statistics is to introduce additional “nuisance” or “confound”
regressors in the GLM design matrix. Remember from Sect. 3.9
that in the case of multiple regressors, the parameter estimates for
each of the EVs can only be driven by the uncorrelated (orthogo-
nal) component of an EV. If we can find a suitable characterization
of the temporal structure of an artifact, we can add this as a new
regressor to the design matrix in order to use this to “explain” some
of the measured variation in the data. The parameter estimate for
EVs of interest will then only reflect the amount of variation that
the EV can explain over and above what can already be explained by
214 Mark W. Woolrich et al.
Fig. 20 Example of the utility of nuisance regressors in the general linear model (GLM): (a) data without artifact
regressed against a single explanatory variables (EVs), (b) data with confound analyzed in a GLM without
nuisance regressor, and (c) confounded data analyzed using both the EV of interest and a nuisance regressor
5.2.1 Deriving Nuisance There are various ways of deriving useful nuisance variables. In
Regressors general, these additional EVs should reflect the temporal charac-
teristics of structured noise that is thought to exist in the data.
Motion of the subject in the scanner is a typical problem in
fMRI, and there are often intensity fluctuations related to head
motion still present in the data even after alignment-based motion
Statistical Analysis of fMRI Data 215
5.2.2 ICA-Based The ability to correct for additive structured noise depends on the
De-noising ability to characterize these noise components in terms of their
temporal evolution. In the previous two examples, this was
obtained by accurately estimating rigid-body motion or by using
secondary measurements of physiological processes. For other
types of noise, it is often not easy to predict such nuisance regres-
sors based on the understanding of the biophysics and of the imag-
ing process. One possibility is to use a model-free data analysis
approach, such as ICA, in order to identify structured noise effects
in the data prior to the GLM analysis. ICA and related techniques
decompose the fMRI data into modes of variation that define the
spatial and temporal extent of underlying fluctuations [29]. The
estimated time courses of a component can then be used as nui-
sance regressors as part of a GLM analysis. An alternative is to
explicitly regress out such effects prior to a GLM-based analysis,
effectively running the model-based analysis on the residuals of a
prior linear regression model designed to de-noise the data.
Currently, as no well-established techniques exist for automatically
identifying such noise components, such an approach relies on the
experimenter to visually inspect all components. Further research
is required to integrate such a model-free identification (e.g., using
ICA) into the standard GLM in an unbiased objective way.
a b c
Z-stats Z-stats
350 350
300 300
250 250
200 200
150 150
100 100
50 50
0 0
–10 –5 0 5 10 –10 –5 0 5 10
Fig. 21 Example of the effect of fMRI de-noising in a simple finger tapping experiment: (a) histogram of the
Z-statistic image for the differential left- versus right-hand finger tapping contrast. Because of the presence of
structured noise, the histogram is far from being Gaussian distributed; (b) after regressing out a variety of
structured noise effects, the histogram of Z-statistic values becomes unimodal and much closer to a Gaussian
distribution; and (c) map of significant voxels after regressing nuisance effects out of the data
6 Multisubject Statistics
Fig. 22 Hierarchical general linear model (GLM) for the analysis of group fMRI data. Within a summary statistics
approach, the GLMs are estimated one level at a time and summary statistics are passed up to the next level
of the hierarchy
6.1 Brain Atlases Registration (aligning different brain images) is typically used
when combining fMRI data from different sessions or subjects in a
multisubject analysis. This allows us to assume that the data we are
comparing across sessions or subjects come from approximately
corresponding areas of the brain. In doing this, it is typical to
transform the data into a common “standard brain space,” for
example, the co-ordinate system specified by Talairach and
Tournoux [32]. These standard spaces can be either what are
known as templates or atlases.
A template is typically an average of many brains, all registered
into a common co-ordinate system. An example is the MNI 305
average [33]. An atlas is also based in a common co-ordinate sys-
tem, but contains richer information about the brain at each voxel,
for example, information about tissue type, local brain structure,
or functional area. Atlases can inform interpretation of fMRI
experiments in a variety of ways, helping the experimenter gain the
maximum value from the data.
6.2 Fixed- An important question is that of how to model and estimate effects
Versus Mixed-Effects at the intermediate and higher level of the hierarchy. If we were
Models only concerned about the particular set of subjects in our study,
then we would use a fixed-effects model. More typically, however,
we would want to generate results that extend beyond the particu-
lar population of subjects scanned as part of the study, into the
wider population. In this case, we also need to account for the fact
that the individual subjects themselves are sampled from the wider
population and thus are random quantities with associated vari-
ances. It is exactly this step that marks the transition from a simple
fixed-effects model to a mixed-effects model and it is imperative to
formulate a model at the group level that allows for the explicit
modeling and estimation of these additional variance terms.
As an example, consider the simplest case of estimating the
effect size of a group of M subjects, where for each subject k, the
pre-processed fMRI data is ϒk, the design matrix is Xk and the
218 Mark W. Woolrich et al.
bk = X g bg (16)
é b1 ù
êb ú
bk = ê 2 ú .
ê ú
ê ú
ëb2 û
bk = X g b g + e g (17)
Fig. 23 Illustration of a simple group analysis using the fixed-effect and the mixed-effects models: (a) in the
fixed-effects analysis, the only variance contribution to consider is the lower-level within-subject variance; (b)
in the mixed-effects analysis, the between-subject random-effects variance, σ2, accounts for the random
sampling of the subjects themselves and contributes to the overall mixed-effects variance
6.3 Summary We now come to the question of how we infer on the multilevel hier-
Statistics Approach archy of a group study (an example of which was shown in Fig. 22),
in order to ask questions such as “Where is there significant activity in
response to the experimental task for the population?” or “Where is
there significant differences between populations (e.g., controls ver-
sus patients)?” Recall that each level in the hierarchy is represented by
its own GLM. Hence, one approach is to formulate a single complete
GLM that combines together the first-level and higher-level GLMs.
An example of such an approach is presented in Friston et al. [34]
where the group analysis is carried out “all-in-one” using the within-
session fMRI time series data as input. However, in fMRI, where the
human and computational costs involved in data analysis are rela-
tively high, it is desirable to be able to make group-level inferences
using the results of separate first-level analyses. This approach is com-
monly referred to as the “summary statistics” approach to fMRI anal-
ysis [31]. Within such an approach, group parameters of interest can
easily be refined as more data become available.
In Holmes and Friston [31], the regression parameter estimates
were used as summary statistics. The regression parameter estimates
from the lower level are used as the “data” at the next level up. For
example, the estimates of βk are used in place of βk in Eq. (17). This
approach was shown to be equivalent to inferring all-in-one under
certain conditions [31]. For example, it requires balanced designs,
that is, all lower-level design matrices need to be identical, preventing
the use of behavioral scores or subject-specific confound regressors.
However, top-level inference using the summary statistics
approach can be made equivalent to the all-in-one approach with-
out such restrictions [35, 36], if we pass up the correct summary
statistics. In particular, it is important to pass up information about
not only the effect sizes from the lower levels, but also their vari-
ances. We shall explore the benefits of doing this in Sect. 6.4.
220 Mark W. Woolrich et al.
6.4 Estimation When we use a summary statistic approach, we infer on each level
of the Mixed-Effects of the hierarchy one at a time. The first-level inference is as we
Model described earlier in the chapter. At higher levels in the hierarchy,
however, estimating the regression parameters and variances of
group-level GLMs offers a different set of challenges. At the first
level, there typically exists a large number of observations (typically
more than 100), so that relevant parameters and variances can be
estimated with high DOF. In contrast, group-level variance com-
ponent estimation is typically troubled by having very few observa-
tions (i.e., low DOF).
A key issue in estimating mixed-effects models within the
“summary statistics” approach is whether the variance information
from the lower levels (e.g., first level) is used at the higher levels
(e.g., group level). Approaches that use the variance information
from the lower levels have a number of substantial advantages.
First, such approaches do not require balanced designs and so per-
mit the analysis of fMRI data where the first-level design matrices
have different structure from each other (e.g., contain behavioral
scores as regressors) or where the data contains different numbers
of observations (e.g., different numbers of sessions for each sub-
ject). Second, they provide more accurate variance estimation (and
therefore more accurate inference) by ensuring that at every level,
only positive estimates of the random-effects variances contribute
to the overall mixed-effects variance. Finally, such approaches
increase the ability to detect real activation, by weighting the dif-
ferent contributions from the lower levels by using the lower-level
variance information. For example, effect sizes from subjects with
high first-level variance get down-weighted compared with those
with low first-level variance, when inferring at the group level.
Estimating mixed-effects models when the lower-level variance
information is ignored can be carried out easily using ordinary least
squares [31]. Approaches that use the lower-level variance
information and provide the advantages described above are a little
more involved. For example, Worsley et al. [12] used an expecta-
tion maximization approach. Woolrich et al. [36] used a fully
Bayesian framework using appropriate noninformative priors. This
approach had the added advantage that one can model different
variance components for different groups. For example, one can
contrast effect sizes in a population of patients relative to a popula-
tion of controls under the assumption that these two groups have
different within-group variance. This is important as, empirically,
patient populations exhibit larger within-group variability than a
carefully selected population of controls.
6.5 Handling Outlier The approaches described so far assume that the population distri-
Subjects butions of the effect sizes are well modeled using a Gaussian distri-
bution. However, in practice group studies can include “outlier”
subjects whose effect sizes are completely at odds with the general
Statistical Analysis of fMRI Data 221
Fig. 24 Typical higher-level general linear model (GLM) design matrices: (a)
group mean effect size over eight subjects, (b) group mean and confounds over
eight subjects, (c) unpaired group difference over two groups of four subjects, (d)
paired group difference test over two conditions for four subjects. Note that for
the sake of space, the number of subjects assumed here is lower than what
would be typically expected in a group study
7 Inference (“Thresholding”)
7.1 Inference The natural questions that any user of FMRI has of their data are:
with the Mass “What is the location of my signal?,” “What is the extent of my sig-
Univariate Model nal about that location?,” “What is the magnitude of my signal?.”
For each of these, our statistical model should provide an estimate, a
measure of uncertainty of the estimate (i.e., a standard error or a
confidence interval), and a significance measure, like a P-value (i.e.,
could our result be explained by chance alone). For example, if a
visual effect produced a cluster (a contiguous groups of supra-
threshold voxels, more on this in Sect. 7.2) with a peak at a certain
location, how certain can I be that the true center of activation is
near that location? Or, if a cluster had a volume of 500 voxels, what
is my confidence that the true signal extent is 500 voxels?
Surprisingly, such basic questions cannot be answered with
standard fMRI methods. In fMRI, we generally use a mass univari-
ate model, where a GLM is fit independently at each voxel. No
information is shared over space, and, specifically, no explicit spatial
model is used to express the extended signals that we expect. While
more advanced methods that address these issues exist [45], the
only inferential questions that a mass univariate model can answer
are (a) “What is the signal magnitude at each voxel (with standard
errors and P-values)?” and (b) “What is the signal extent for a given
cluster-defining threshold (P-values only)?.” However, standard
errors and P-values on locations (e.g., confidence intervals on local
maxima or center of mass of a cluster) are not available.
The remainder of Sect. 7.2 focuses on these two types of infer-
ences: voxel-wise and cluster-wise.
7.2 Voxel-Wise The result of applying a contrast to the GLM fit at each voxel is a
Versus Cluster-Wise statistic image. This is anywhere from I = 20,000–100,000 brain
Inference voxels in a statistic image. The value in the image at each voxel is a
T-, F-, or Z-statistic that measures the evidence for an effect defined
by the contrast. The process of applying threshold u, and retaining
all voxels with statistic value greater than u is known as voxel-wise
inference. Precisely, we are performing I statistical tests of signifi-
cance, rejecting the null hypothesis at voxel i if ti ≥ u, where ti is
the statistic value at voxel i.
Alternatively, we can apply a cluster-forming threshold, uclus,
create a binary image of voxels ti ≥ uclus, and identify clusters, that is,
224 Mark W. Woolrich et al.
Statistic Image
statistic value
space
Voxel-wise Inference
statistic value
uα
space
Cluster-wise Inference
statistic value
uclus
space
Non-Significant kα kα Significant
Cluster Cluster
Fig. 25 Voxel-wise versus cluster-wise inference. Inferences on fMRI statistic images are made either through
voxel-wise or through cluster-wise methods. The top panel illustrates the values in a statistic image through
one line of space, where large values indicate evidence for an experimental effect. The middle panel illustrates
voxel-wise inference, where a significance threshold uα is applied to the image, and voxels above that thresh-
old are labeled as significant. The advantage of voxel-wise inference is that individual voxels are marked as
significant, but the disadvantage is no spatial information is considered, and unusually expansive effects may
be missed. The bottom panel illustrates cluster-wise inference, where a cluster-forming threshold uclus is
applied to the image, and contiguous voxels are formed into clusters. Clusters that exceed a significance
threshold kα in size are marked as significant. The advantage of cluster-wise inference is that low, spatially
extended signals can be detected. The disadvantage is that clusters as a whole are marked as significant, and
individual voxels within a cluster cannot be marked individually as significant
226 Mark W. Woolrich et al.
Table 2
Cross-tabulation of the number voxels in difference inference categories
Fig. 26 Comparison of uncorrected versus family-wise error (FWE)-corrected versus false discovery rate (FDR)-
corrected inferences. The top rows shows ten realizations of a central, circular signal added to smooth noise, and
the next three rows show different possible voxel-wise thresholding methods applied to each realization. The
dashed circles indicate the extent of the signal. The second row shows the result of an α = 10 % uncorrected
threshold; most of the signal is correctly detected, but much of the background is also incorrectly detected. On
average, 10 % of the background consists of false positives, but notice that the exact proportion of false positives
varies from realization to realization. The third row shows the result of using an αFWE = 10 % threshold (e.g., a
threshold from Bonferroni or random field theory). Much less of the signal is detected, but there are many fewer
false positives, only 1-in-10 of the datasets considered had any false positives, this is a FWE. Of course in prac-
tice, we never know if the dataset in our hands is the 1-in-10 (or 20) that contains a family-wise error. The bottom
row shows the result of using an αFDR = 10 % threshold. While we are guaranteed that the percentage of detected
voxels that are false positive does not exceed 10 % on average, the actual percentage can vary considerably
228 Mark W. Woolrich et al.
7.4 Corrected So far we have only defined measure of false positives, but we have
Inference Methods not described how we obtain thresholds that control these false
positive measures. In addition to Bonferroni, there are two further
methods that are commonly used in fMRI for controlling FWE;
these are RFT and permutation, whereas there is generally just a
single method for FDR.
7.4.1 Controlling FWE The Bonferroni method for controlling FWE uses a significance
with RFT threshold corresponding to α = αFWE/I, the nominal FWE test level
divided by the number of tests. Bonferroni becomes quite conserva-
tive when the data is smooth, and has no way to adapt to the data in
anyway. For example, imagine an extreme case where FMRI data is
smoothed with a 1 m wide Gaussian smoothing kernel; such data will
produce a statistic image with essentially a single constant value,
meaning there is no multiple testing problem anymore; however, the
Bonferroni threshold will still prescribe dividing αFWE by, say, 10,000.
RFT uses the smoothness of the data to adjust the significance
threshold while still controlling FWE. The mathematics involved are
elegant yet quite involved, and in what follows, we only give the most
cursory review. For a more detailed review, see [48], or for a more
technical overview, see [49]. The original Gaussian RFT paper for
PET imaging remains a useful introduction [50], though also see [51]
for more up-to-date results including T- and F-statistic RFT results.
To use RFT results, we must know the smoothness of the data,
precisely the smoothness of the standardized noise images (e/σ in
the notation from Sect. 3). Smoothness is parameterized by the full
width at half maximum (FWHM) of the Gaussian kernel required
to simulate images with the same apparent spatial smoothness as
our data (Fig. 27). For example, if we say that our data has 6-mm
FWHM smoothness, it means that if we were to simulate our data,
we would generate noise data with no correlation and then con-
volve it with a Gaussian kernel with FHWM of 6 mm. The exact
Statistical Analysis of fMRI Data 229
Full
Width
Half
Maximum
Fig. 27 Full width at half maximum (FWHM) is a generic way to describe the
spread of a distribution, and is the way that smoothness is measured for random
field theory
form of the spatial dependence of our data does not have to follow
a Gaussian kernel, but for convenience, we describe the strength of
the dependence in terms of the size a Gaussian kernel.
It may seem that if we take our fMRI data fresh from the scan-
ner and convolve it with a 6-mm Gaussian kernel, our FWHM for
RFT would be 6 mm. This is incorrect, however, as the noise
smoothness includes both intrinsic sources of smoothness (imper-
fect MRI resolution, physiological artifacts, etc.) and smoothness
induced by the applied smoothing. As a result, the smoothness for
RFT is not a user-specified parameter, but rather estimated from
(
the residuals of the GLM ¡ - X b .
)
7.4.2 RESELs The definition of FWHM smoothness creates a notion of a
smoothness-equivalent volume, a resolution element or RESEL. If
the smoothness of the data is FWHMx, FWHMy, FWHMz in each
of the principal directions, then a volume of space with dimensions
FWHMx × FWHMy × FWHMz is one RESEL. In very approximate
terms, the RESEL count captures the amount of independent
information in the image; fewer RESELs = smoother data = less
information = less severe multiple testing problem; more
RESELS = rougher data = more severe multiple testing problem.
The total RESEL count for a search volume is:
7.4.3 RFT-Corrected RFT can be used with any type of statistic image a GLM can create,
P-Values including T-, F-, and Z-statistic images [51]. The formulas provide
FWE-corrected P-values for voxel-wise thresholds and cluster sizes
230 Mark W. Woolrich et al.
æ z2 ö
FWE
Pvox ( z ) = RESELcount ( 2p )
-2
(z 2
- 1) exp ç - ÷
è 2ø
This shows that as z grows, the corrected P-value shrinks (the expo-
nential term dominates), which of course makes sense, as larger sta-
tistic values should produce smaller P-values. As the RESEL count
grows, the corrected P-value grows. The RESEL count can increase
because the search volume increases, which again is sensible, as a
greater search volume demands a greater correction for multiple
testing, and hence a less significant P-value. The RESEL count can
also increase if the smoothness decreases, as per [18], which
increases the amount of information in the image, again demanding
a greater correction for multiple testing. For cluster-wise inference,
the equation for the FWE-corrected P-value is more involved, but
it also accounts for the image search space and smoothness.
7.4.4 Small-Volume The first RFT results published (and the equation shown above)
Correction assumed that the search region was large relative to the smooth-
ness of the image. This assumption was needed to avoid dealing
with the case when a cluster touches the boundary of the search
region. To see why this could be a problem, imagine two statistic
images with the same smoothness, one the size and shape of the
brain, the other with the same total volume but the shape of a
long, narrow sausage. In the latter case, it is more likely that clus-
ters will touch the edge of the image, and, relative to other clus-
ters, have smaller volume. The results in [51] can be used to
produce P-values that are accurate even with small search regions.
When these results are used, they are sometimes referred to as
“small volume correction.”
7.4.5 RFT Assumptions The use of RFT results is based on several assumptions and approx-
imations. The essential assumptions are:
1. Gaussian data. For any collection of voxels, the distribution of
the data is multivariate Gaussian.
2. Sufficient smoothness. The data must be sufficiently smooth to
approximate continuous random fields (upon which the the-
ory is based).
3. Known smoothness. The results assume that the FWHM
smoothness parameters are exact and contain at most negligi-
ble error.
Statistical Analysis of fMRI Data 231
7.4.6 Controlling FWE Nonparametric methods are generally used when the standard
with Permutation parametric assumptions are known to be false or cannot be verified.
In the case of RFT, the assumptions are nearly impossible to verify,
but more importantly, it has been found that voxel-wise RFT
results are quite conservative for small group studies (e.g., when
the number of subjects is less than about 40; [48, 53]). Hence,
there has been interest in using alternative methods.
Instead of making assumptions about the distribution of the
data, permutation testing uses the distribution of the data itself to
find P-values and thresholds. Figure 28 illustrates the reasoning of
the permutation test in the two-group setting.
While nonparametric tests are sometimes referred to as
assumption- free, in fact they also have assumptions, just much
weaker ones than standard parametric methods. The essential
assumption for the permutation test is exchangeability under the
null hypothesis. Exchangeability means that the data can be per-
muted (relative to the model) without altering its joint distribution.
fMRI data presents a challenge for permutation testing. At the first
level, temporal autocorrelation renders the data nonexchangeable
and permutation methods cannot be directly applied (the data must
be de-correlated, then permuted, and then re-correlated (see [6,
54] for more details). However, in second-level analyses, exchange-
ability is generally not a problem. In the example in the figure, we
assume that, under the null hypothesis, all six subjects are exchange-
able—this is very reasonable because if there is no group effect (this
is the null hypothesis), then there is nothing special about the first
three subjects versus the last three. For this example, there are 20
possible ways of permuting the groups (including the correct label-
ing). For an arbitrary dataset with group sizes n1 and n2, the num-
ber of possible permutations is (n1 + n2)!/(n1!n2!).
By permuting the data many times, and for each permutation,
assuming that the resulting test statistic (in this case, the
group-difference T-statistic) is an sample of what we would see if
there were no real effect present, we build up a histogram of test
232 Mark W. Woolrich et al.
Fig. 28 Permutation test applied to data from a single voxel for a hypothetical two-group fMRI second-level
analysis
statistic values that will serve as the null distribution of that test
statistic. We can then look to see how much “area under the tail”
lies to the right of the actual test statistic value that we originally
observed (under the correct labeling of the data), and hence esti-
mate our P-value; this is the same principle for relating the null
distribution of the test statistic to the P-value as we saw in Fig. 9,
but in this case, the null distribution has been generated via a com-
pletely different methodology.
The permutation test for the two group case can be generalized
to three or more groups. In that case, we are testing the null
Statistical Analysis of fMRI Data 233
hypothesis that all groups are the same, and use an F-test to measure
the evidence of any difference. Under the null hypothesis, all sub-
jects can be freely permuted. Likewise for a simple correlation
model, the null hypothesis of no association justifies the free per-
mutation of all of the subjects. The permutation test for the one
group case, however, is problematic without further assumptions: If
all we have is a single group, what is there to permute? Shuffling the
order of subjects will not change the value of a one-sample T-test.
In a second-level single group mean, the COPE images are
always created as relative differences between baseline and active
data. Even if an event-related design is used, and a contrast selects a
single predictor, the effective predictor is a subtraction of event and
baseline data. This is the case due to the relative, nonquantitative
nature of the BOLD signal. As a result, we use here a slightly differ-
ent assumption to generate “permutations.” Under the null hypoth-
esis, we assume that each individual’s second-level COPE data are
mean zero and have a symmetric distribution. Assuming mean zero
data is reasonable, as, under the null, we expect no activation, posi-
tive or negative. Assuming a symmetric distribution is a weakened
form of normality, and is exactly satisfied for any balanced effect
(i.e., a COPE constructed as difference of two averages, where an
equal number of scans contributed to each average).
The one-sample, group-level fMRI permutation thus works as
follows. Assuming mean zero, symmetrically distributed COPE data,
we randomly multiply each subject’s data by 1 or −1, or, equivalently,
randomly flip the signs of each subject’s COPE image. Since the data
are symmetrically distributed about zero, multiplication by −1 does
not alter the distribution, and we generate a realization that is equiva-
lent to the original data. If there are n subjects in the analysis, there
are 2n possible ways to flip the signs of the group-level data.
The permutation methods described so far will create uncor-
rected P-values at each voxel. Control of the FWE rate is easily
obtained with permutation testing via the following observation: In
complete-null data, an FWE occurs whenever one or more voxels
exceed the threshold, which occurs exactly when the voxel with the
largest statistic exceeds the threshold. Hence, inference based on the
permutation distribution of the largest (maximum) statistic provides
valid FWE inferences. Specifically, at each permutation, the maxi-
mum statistic value (across all voxels in the brain) is noted, creating
a null distribution of the maximum-across-space test statistic. The
95th percentile of that distribution gives an FWE-corrected thresh-
old (“p < 0.05, corrected”), and any particular statistic value can be
compared to the maximum permutation distribution to obtain an
FWE-corrected P-value. Similarly, the maximal cluster size distribu-
tion can be created to provide FWE cluster-wise inferences.
It is important to note that, while permutation may appear to
be a completely different approach than those we described earlier,
in fact all pre-processing and modeling are generally the same, and
234 Mark W. Woolrich et al.
7.4.7 Controlling FDR The method for finding a threshold that controls FDR is surpris-
ingly simple. It is based only on the uncorrected voxel-wise P-values
in the statistic image. Let Pi be the P-value at voxel i, and P(i) be
the ordered P-values, P(1) ≤ P(2) ≤ · · · ≤ P(I). Then the largest index i
that satisfies
i
P(i ) £ a FDR (19)
I
defines the FDR threshold as P(i) [55]. This method works even
when there is positive dependence between voxels [56, 57].
7.4.8 Controlling False So far we have considered techniques that control the rate of false
Positives and True positives. This depends on knowing the null distribution (or non-
Negatives: Mixture activation distribution) for relevant statistics under the null hypoth-
Modeling esis. In contrast, mixture modeling provides us with a way of
estimating the “activating” and “nonactivating” distributions from
Statistical Analysis of fMRI Data 235
Fig. 29 Mixture modeling of a Z-statistic image. Top-left: Four example slices of a Z-statistic image obtained
from fitting a general linear model (GLM) to the fMRI data at each voxel from an individual subject. The experi-
ment was a pain stimulus applied using a sparse single-event design. Right: Mixture model fit to the histogram
of Z-statistics. Nonactivating voxel are modeled as coming from a close-to-zero mean Gaussian distribution,
activating voxels as coming from a Gamma distribution, and de-activating voxels as coming from a negative
Gamma distribution. However, note that there were found to be no de-activating voxels in this case. Bottom-
left: Image showing the probability that a voxel is activating—this information that can be extracted from the
mixture model fit and can be used in thresholding to approximately control the true positive rate (TPR) as an
alternative to null hypothesis testing
236 Mark W. Woolrich et al.
7.5 Enhancing Aside from thresholding the final statistic images (either voxel-wise
Statistic Images or cluster-wise), it is not advisable to make image-processing
adjustments to statistic images. For example, smoothing a T-statistic
image would be disastrous: While a T-statistic has approximately
unit variance and follows a particular null distribution, a smoothed
T-statistic image will have dramatically reduced variance with no
particular distribution. Two exceptions to this are wavelet de-
noising methods and a recently proposed threshold-free cluster
enhancement (TFCE) method.
Wavelet methods transform the data in a scale-dependent fash-
ion, so that all of the large-scale information is segregated from the
fine-scale information. Since we generally expect the signals of
interest to be spatially extended, wavelet methods can be used to
“shrink” variation associated with the finest scales, “de-noising”
the image, while preserving the large-scale structure. For an over-
view of wavelet methods applied to fMRI, see [62].
Cluster-wise inference also tries to capture spatially extended
signals, but requires the specification of an arbitrary cluster-forming
threshold uclus. TFCE removes this dependence by, in essence,
using all possible uclus values and then merging all the results into a
single image. Specifically, at each voxel, let ei(h) be the extent of
the cluster that voxel i belongs to with cluster-forming threshold h
åe ( h )
E
(or 0 if ti < h). Then, the TFCE image is defined by i hH ,
h>0
where the sum is computed for a discrete set of h values, from 0 to
the maximum statistic value, and E and H are tuning parameters.
In [63], E = 0.5 and H = 2 were found to give generally good per-
formance for a range of classes of signals. TFCE seems to succeed
Statistical Analysis of fMRI Data 237
References
1. Friston K, Worsley K, Frackowiak R, Mazziotta 12. Worsley K, Liao C, Aston J et al (2002) A
J, Evans A (1994) Assessing the significance general statistical analysis for fMRI data.
of focal activations using their spatial extent. NeuroImage 15:1–15
Hum Brain Mapp 1:214–220 13. Gautama T, Van Hulle MM (2004) Optimal spa-
2. Hykin J, Bowtell R, Glover P, Coxon R, tial regularisation of autocorrelation estimates in
Blumhardt L, Mansfield P (1995) Investigation fMRI analysis. Neuroimage 23:1203–1216
of the linearity of functional activation signal 14. Penny W, Kiebel S, Friston K (2003)
changes in the brain using echo planar imag- Variational Bayesian inference for fMRI time
ing (EPI) at 3.0 T. In: Proc of the SMR and series. NeuroImage 19:1477–1491
ESMRB Joint Meeting. p 795 15. Woolrich M, Behrens T, Smith S (2004)
3. Cohen M (1997) Parametric analysis of fMRI Constrained linear basis sets for HRF mod-
data using linear systems methods. NeuroImage elling using Variational Bayes. NeuroImage
6:93–103 21:1748–1761
4. Dale A, Buckner R (1997) Selective averag- 16. Smith S, Jenkinson M, Beckmann C, Miller
ing of rapidly presented individual trials using K, Woolrich M (2007) Meaningful design and
fMRI. Hum Brain Mapp 5:329–340 contrast estimability in fMRI. NeuroImage
5. Burock MA, Buckner RL, Woldorff MG, 34:127–136
Rosen BR, Dale AM (1998) Randomized 17. Dale A, Greve D, Burock M (1999) Optimal
event-related experimental designs allow for stimulus sequences for event-related
extremely rapid presentation rates using func- fMRI. NeuroImage 9:S33
tional MRI. NeuroReport 9:3735–3739 18. Wager T, Nichols T (2003) Optimization of
6. Bullmore E, Brammer M, Williams S et al experimental design in fMRI: a general frame-
(1996) Statistical methods of estimation and work using a genetic algorithm. Neuroimage
inference for functional MR image analysis. 18:293–309
Magn Reson Med 35:261–277 19. Josephs O, Turner R, Friston K (1997) Event-
7. Friston K, Josephs O, Zarahn E, Holmes A, related fMRI. Hum Brain Mapp 5:1–7
Rouquette S, Poline J-B (2000) To smooth or 20. Lange N, Zeger S (1997) Non-linear Fourier
not to smooth? NeuroImage 12:196–208 time series analysis for human brain mapping
8. Woolrich M, Ripley B, Brady J, Smith S (2001) by functional magnetic resonance imaging.
Temporal autocorrelation in univariate lin- Appl Stat 46:1–29
ear modelling of FMRI data. NeuroImage 21. Genovese C (2000) A Bayesian time-course model
14:1370–1386 for functional magnetic resonance imaging data
9. Locascio J, Jennings P, Moore C, Corkin S (with discussion). J Am Stat Assoc 95:691–703
(1997) Time series analysis in the time domain 22. Friston KJ (2002) Bayesian estimation
and resampling methods for studies of func- of dynamical systems: an application to
tional magnetic resonance brain imaging. Hum fMRI. NeuroImage 16:513–530
Brain Mapp 5:168–193
23. Marrelec G, Benali H, Ciuciu P, Pélégrini-Issac
10. Purdon P, Weisskoff R (1998) Effect of tem- M, Poline J-B (2003) Robust Bayesian estima-
poral autocorrelation due to physiological tion of the hemodynamic response function in
noise and stimulus paradigm on voxel-level event-related BOLD MRI using basic physio-
false-positive rates in fMRI. Hum Brain Mapp logical information. Hum Brain Mapp 19:1–17
6:239–249
24. Woolrich M, Jenkinson M, Brady J, Smith S
11. Marchini J, Ripley B (2000) A new statistical (2004) Fully Bayesian spatio-temporal model-
approach to detecting significant activation in ling of FMRI data. IEEE Trans Med Imaging
functional MRI. NeuroImage 12:366–380 23:213–231
238 Mark W. Woolrich et al.
25. Buxton R, Uludag K, Dubowitz D, Liu T ses using robust regression. NeuroImage
(2004) Modeling the hemodynamic response to 26:99–113
brain activation. NeuroImage 23(S1):220–233 41. Woolrich M (2008) Robust group analysis using
26. Boynton G, Engel S, Glover G, Heeger D outlier inference. NeuroImage 41:286–301
(1996) Linear systems analysis of functional 42. Meriaux S, Roche A, Dehaene-Lambertz G,
magnetic resonance imaging in human V1. Thirion B, Poline J (2006) Combined permuta-
J Neurosci 16:4207–4221 tion test and mixed-effect model for group average
27. Glover G (1999) Deconvolution of analysis in fMRI. Hum Brain Mapp 27:402–410
impulse response in event-related BOLD 43. Roche A, Meriaux S, Keller M, Thirion B
fMRI. NeuroImage 9:416–429 (2007) Mixed-effect statistics for group analy-
28. Friston K, Josephs O, Rees G, Turner R sis in fMRI: a nonpara-metric maximum likeli-
(1998) Nonlinear event-related responses in hood approach. Neuroimage 38:501–510
fMRI. Magn Reson Med 39:41–52 44. Thirion B, Pinel P, Mriaux S, Roche A, Dehaene
29. Beckmann C, Smith S (2004) Probabilistic S, Poline J (2007) Analysis of a large fMRI
independent component analysis for functional cohort: statistical and methodological issues
magnetic resonance imaging. IEEE Trans Med for group analyses. Neuroimage 35:105–120
Imaging 23:137–152 45. Hartvig NV, Jensen JL (2000) Spatial mixture
30. Glover G, Li T, Ress D (2000) Image-based modeling of fMRI data. Hum Brain Mapp
method for retrospective correction of physi- 11:233–248
ological motion effects in fMRI: Retroicor. 46. Hayasaka S, Nichols TE (2003) Validating
Magn Reson Med 44:162–167 cluster size inference: random field and permu-
31. Holmes A, Friston K (1998) Generalisability, tation methods. NeuroImage 20:2343–2356
random effects & population inference. Fourth 47. Friston KJ, Holmes A, Poline J-B, Price CJ,
Int Conf on Functional Mapping of the Human Frith CD (1996) Detecting activations in
Brain. NeuroImage 7:S754 PET and fMRI: levels of inference and power.
32. Talairach J, Tournoux P (1988) Co-planar NeuroImage 4:223–235
stereotaxic atlas of the human brain. Thieme, 48. Nichols TE, Hayasaka S (2003) Controlling
New York the familywise error rate in functional neuro-
33. Collins D, Neelin P, Peters T, Evans A imaging: a comparative review. Stat Methods
(1994) Automatic 3D intersubject regis- Med Res 12:419–446
tration of MR volumetric data in standard- 49. Cao J, Worsley KJ (2001) Applications of ran-
ized Talairach space. J Comput Assist Tomo dom fields in human brain mapping. In: Moore
18:192–205 M, (ed) Spatial statistics: methodological
34. Friston KJ, Penny W, Phillips C, Kiebel S, aspects and applications, vol 159, Springer lec-
Hinton G, Ashburner J (2002) Classical and ture notes in statistics. Springer. pp 169–182
Bayesian inference in neuroimaging: theory. 50. Worsley KJ, Evans AC, Marrett S, Neelin P
NeuroImage 16:465–483 (1992) Three-dimensional statistical analy-
35. Beckmann C, Jenkinson M, Smith S (2003) sis for cbf activation studies in human brain.
General multi-level linear modelling for group J Cerebr Blood F Met 12:900–918
analysis in FMRI. NeuroImage 20:1052–1063 51. Worsley KJ, Marrett S, Neelin P, Vandal AC,
36. Woolrich M, Behrens T, Beckmann C, Jenkinson Friston KJ, Evans AC (1996) A unified statisti-
M, Smith S (2004) Multi-level linear modelling cal approach for determining significant signals
for FMRI group analysis using Bayesian infer- in images of cerebral activation. Hum Brain
ence. NeuroImage 21:1732–1747 Mapp 4:58–73
37. Kherif F, Poline J-B, Meriaux S, Benali H, Flandin 52. Hayasaka S, Luan Phan K, Liberzon I, Worsley
G, Brett M (2003) Group analysis in functional KJ, Nichols TE (2004) Nonstationary cluster-
neuroimaging: selecting subjects using similarity size inference with random field and permuta-
measures. Neuroimage 20:2197–2208 tion methods. NeuroImage 22:676–687
38. Luo W-L, Nichols TE (2003) Diagnosis and 53. Nichols T, Holmes A (2001) Nonparametric
exploration of massively univariate neuroimag- permutation tests for functional neuroimaging: a
ing models. Neuroimage 19:1014–1032 primer with examples. Hum Brain Mapp 15:1–25
39. Seghier M, Friston K, Price C (2007) Detecting 54. Bullmore E, Long C, Suckling J et al (2001)
subject-specific activations using fuzzy cluster- Colored noise and computational inference in
ing. Neuroimage 36:594–605 neurophysiological (fMRI) time series analysis:
40. Wager T, Keller M, Lacey S, Jonides J (2005) resampling methods in time and wavelet
Increased sensitivity in neuroimaging analy- domains. Hum Brain Mapp 12:61–78
Statistical Analysis of fMRI Data 239
55. Benjamini Y, Hochberg Y (1995) Controlling 60. Woolrich M, Behrens T (2006) Variational
the false discovery rate: a practical and power- Bayes inference of spatial mixture models for
ful approach to multiple testing. J R Stat Soc segmentation. IEEE Trans Med Imaging
Ser B Methodol 57:289–300 25:1380–1391
56. Genovese C, Lazar N, Nichols T (2002) 61. Bartsch A, Homola G, Biller A, Solymosi L,
Thresholding of statistical maps in functional Bendszus M (2006) Diagnostic functional
neuroimaging using the false discovery rate. MRI: illustrated clinical applications and
NeuroImage 15:870–878 decision-
making. J Magn Reson Imaging
57. Benjamini Y, Yekutieli D (2001) The control of 23:921–932
the false discovery rate in multiple testing 62. Van De Ville D, Blu T, Unser M (2006) Surfing
under dependency. Ann Stat 29:1165–1188 the brain – an overview of wavelet-based tech-
58. Everitt B, Bullmore E (1999) Mixture model niques for fMRI data analysis. IEEE Eng Med
mapping of brain activation in functional mag- Biol 25:65–78
netic resonance images. Hum Brain Mapp 63. Smith SM, Nichols TE (2008) Threshold-free
7:1–14 cluster enhancement: addressing problems of
59. Hartvig N (2000) A stochastic geometry smoothing, threshold dependence and localisa-
model for fMRI data. Technical Report 410. tion in cluster inference. NeuroImage.
Department of Theoretical Statistics, University doi:10.1016/j.neuroimage.2008.03.061, In
of Aarhus press; Epub ahead of print April 11, 2008
Chapter 8
Abstract
This chapter is about modeling-distributed brain responses and, in particular, the functional integration
among neuronal systems. Inferences about the functional organization of the brain rest on models of how
measurements of evoked responses are caused. These models can be quite diverse, ranging from concep-
tual models of functional anatomy to mathematical models of neuronal and hemodynamics. The aim of
this chapter is to introduce dynamic causal models. These models can be regarded as generalizations of the
simple models employed in conventional analyses of regionally specific brain responses. In what follows,
we will start with anatomical models of functional brain architectures, which motivate some of the basic
principles of neuroimaging. We then review briefly statistical models (e.g., the general linear model) used
for making classical and Bayesian inferences about where neuronal responses are expressed. By incorporat-
ing biophysical constraints, these basic models can be finessed and, in a dynamic setting, rendered causal.
This allows us to infer how interactions among brain regions are mediated. This chapter focuses on causal
models for distributed responses measured with fMRI and electroencephalography. The latter is based on
neural-mass models and affords mechanistic inferences about how evoked responses are caused, at the level
of neuronal subpopulations and the coupling among them.
Key words Functional connectivity, Effective connectivity, Dynamic causal modeling, Causal,
Dynamic, Nonlinear
1 Introduction
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_8, © Springer Science+Business Media New York 2016
241
242 Karl J. Friston
2 Anatomical Models
2.2 Functional The functional role of any component (e.g., cortical area, sub-area,
Specialization or neuronal population) of the brain is defined largely by its con-
and Segregation nections. Certain patterns of cortical projections are so common
that they could amount to rules of cortical connectivity. “These
rules revolve around one, apparently, over-riding strategy that the
cerebral cortex uses—that of functional segregation” [5].
Functional segregation demands that cells with common func-
tional properties be grouped together. This architectural constraint
necessitates both convergence and divergence of cortical connec-
tions. Extrinsic connections among cortical regions are not
continuous but occur in patches or clusters. This patchiness has, in
some instances, a clear relationship to functional segregation. For
example, when recordings are made in V2, directionally selective
(but not wavelength or color selective) cells are found exclusively
in its thick stripes. Retrograde (i.e., backward) labeling of cells in
V5 is limited to these thick stripes; all the available physiological
244 Karl J. Friston
3 Statistical Models
3.1 Statistical Functional mapping studies are usually analyzed with some form of
Parametric Mapping statistical parametric mapping (SPM). SPM entails the construc-
tion of continuous statistical maps (e.g., t-maps) to test hypotheses
about regionally specific effects [6]. SPM uses the GLM and ran-
dom field theory (RFT) to analyze and make classical inferences
about brain responses. Parameters of the GLM are estimated in
exactly the same way as in conventional analysis of discrete data.
RFT is used to resolve the multiple-comparisons problem induced
by making inferences over a volume of the brain. RFT provides a
method for adjusting p-values for the search volume of an SPM to
control false positive rates. It plays the same role for continuous
data (i.e., images or time series) as the Bonferroni correction for a
family of discontinuous or discrete statistical tests.
There is a Bayesian alternative to classical inference with SPMs.
This rests on conditional inferences about an effect, given the data,
as opposed to classical inferences about the data, given the effect is
zero. Bayesian inferences about effects that are continuous in space
use posterior probability maps (PPMs). Although less established
than SPMs, PPMs are potentially useful, not least because they do
not have to contend with the multiple-comparisons problem
induced by classical inference (see Ref. [7]). In contradistinction to
SPM, this means that inferences about a given regional response
do not depend on inferences about responses elsewhere. Bayesian
inference is particularly relevant to dynamic casual modeling
because the Bayesian formulation is an essential part of model
specification and inversion. Before looking at the models underly-
ing Bayesian inference, we briefly review estimation and classical
inference in the context of the GLM and show how this can be
generalized to give a Bayesian approach.
Dynamic Causal Modeling of Brain Responses 245
y = X ( ) b ( )e ( )
1 1 1
This is exactly the same as Eq. (1) but now the parameters of the
first level are generated by a supra-ordinate linear model and so on
to any hierarchical depth required. These hierarchical observation
models are an important extension of the GLM and are usually
estimated using expectation maximization (EM) [9]. In the pres-
ent context, the response variables comprise the responses at all
voxels and β(1)s are the treatment effects we want to make an infer-
ence about. Because we have invoked a second level, the first-level
parameters embody random effects and are generated by a second-
level linear model. At the second level, β(2) is the average effect over
voxels and ε(2) is its voxel-to-voxel variation. By estimating the vari-
ance of ε(2), one is implicitly estimating an empirical prior on the
first-level parameters at each voxel. This prior can then be used to
estimate the posterior probability of β(1) being greater than some
threshold at each voxel. An example of the ensuing PPM is pro-
vided in Fig. 1 along with the classical SPM.
In summary, we have seen how the GLM can be used to test
hypotheses about brain responses and how, in a hierarchical form,
it enables empirical Bayesian or conditional inference. Then, we
deal with the dynamic systems and how they can be formulated as
GLMs. These dynamic models take us closer to how brain responses
are actually caused by experimental manipulations and represent
the next step towards dynamic causal models of brain responses.
3.4 Dynamic Models In Friston et al. [10], the form of the impulsed hemodynamic
response function (HRF) was estimated using a least squares de-
3.4.1 Convolution
convolution and a linear time invariant model, where evoked neu-
Models and Temporal
ronal responses are convolved or smoothed with an HRF to give the
Basis Functions
measured hemodynamic response (see also Ref. [11]). This simple
linear convolution model is the cornerstone for making statistical
inferences about activations in fMRI with the GLM. An impulse
response function is the response to a single impulse, measured at
a series of times after the input. It characterizes the input–output
behavior of the system (i.e., voxel) and places important constraints
on the sorts of inputs that will excite a response.
Knowing the form of the HRF is important for several reasons,
not least because it furnishes better statistical models of the data.
248 Karl J. Friston
contrast
100
SPM PPM
200
300
1 2 3 4
Design matrix
z = 3mm z = 3mm
Fig. 1 Statistical parametric mapping (SPM) and posterior probability map (PPM) for an fMRI study of attention
to visual motion. The display format (lower panel) uses an axial slice through extra-striate regions but the
thresholds are the same as employed the in maximum-intensity projections (upper panels). Upper right: The
activation threshold for the PPM was 0.7 au, meaning that all voxels shown had a 90 % chance of an activation
of 0.7 % or more. Upper left: The corresponding SPM using an adjusted threshold at p = 0.05. Note the bilateral
foci of motion-related responses in the PPM that are not seen in the SPM (gray arrows). As can be imputed
from the design matrix (upper-middle panel), the statistical model of evoked responses comprised boxcar
regressors convolved with a canonical hemodynamic response function. The middle column corresponds to
the presentation of moving dots and was the stimulus attribute tested by the contrast
The HRF may vary from voxel to voxel and this has to be accom-
modated in the GLM. To allow for different HRFs in different
brain regions, temporal basis functions were introduced [12] to
model evoked responses in fMRI and applied to event-related
responses in Josephs et al. [13] (see also Ref. [14]). The basic idea
behind temporal basis functions is that the hemodynamic response,
induced by any given trial type, can be expressed as the linear com-
bination of (basis) functions of peri-stimulus time. The convolution
model for fMRI responses takes a stimulus function encoding the
neuronal responses and convolves it with an HRF to give a regres-
sor that enters the design matrix. When using basis functions, the
stimulus function is convolved with each basis function to give a
series of regressors. Mathematically, we can express this model as
Dynamic Causal Modeling of Brain Responses 249
y (t ) = X b + e y (t ) = (t ) Ä h (t )
Û (3)
X i = Ti ( t ) Ä u ( t ) h ( t ) = b1T1 ( t ) + b 2T2 ( t ) + ¼
Fig. 2 Temporal basis functions offer useful constraints on the form of the estimated response that retain the
flexibility of finite impulse response (FIR) models and the efficiency of single regressor models. The specifica-
tion of these constrained FIR models involves setting up stimulus functions u(t) that model expected neuronal
changes, for example, boxcar-functions of epoch-related responses or spike-(δ)-functions at the onset of
specific events or trials. These stimulus functions are then convolved with a set of basis functions Ti(t) of peri-
stimulus time that, in some linear combination, model the HRF. The ensuing regressors are assembled into the
design matrix. The basis functions can be as simple as a single canonical HRF (middle), through to a series of
top-hat-functions δi(t) (bottom). The latter case corresponds to an FIR model and the coefficients constitute
estimates of the impulse response function at a finite number of discrete sampling times. Selective averaging
in event-related fMRI [39] is mathematically equivalent to this limiting case
250 Karl J. Friston
neuronal input
u(t)
induced signal v&q
0.4
1.1
s, f, v, q 0.3
activity-dependent signal 0.2 1.05
0.1
s˙ = u − KS − ?(f −1) 1
0
0.95
f -0.1
-0.2 0.0
flow induction 0 10 20 30 0 10 20 30
f˙ = s
f hemodynamics
1.3
v q 1.2 0.5
1.1
0
1
0.9 0.5
0 10 20 30 0 10 20 30
time secs time secs
y = l(v.q)
Fig. 3 Right: Hemodynamics elicited by an impulse of neuronal activity as predicted by a dynamical biophysical
model (left). A burst of neuronal activity causes an increase in flow-inducing signal that decays with first-order
kinetics and is downregulated by local flow. This signal increases regional cerebral blood flow (rCBF), which
dilates the venous capillaries, increasing volume v. Concurrently, venous blood is expelled from the venous
pool decreasing deoxyhemoglobin content q. The resulting fall in deoxyhemoglobin concentration leads to a
transient increases in blood oxygen level dependent (BOLD) signal and a subsequent undershoot. Left:
Hemodynamic model; on which these simulations were based
Dynamic Causal Modeling of Brain Responses 251
x ( t ) = f ( x,,u ,,q )
(4)
y ( t ) = g ( x,,u ,,q ) + e
¶k (s )1
A ( s )i = , (6)
¶qi
3.5.2 Nonlinear System Once a suitable causal model has been established (e.g., Fig. 3), we
Identification can estimate second-order kernels. These kernels represent a non-
linear characterization of the HRF that can model interactions
among stimuli in causing responses. One important manifestation
of the nonlinear effects, captured by the second-order kernels, is a
modulation of stimulus-specific responses by preceding stimuli
that are proximate in time. This means that responses at high-
stimulus presentation rates saturate and, in some instances, show
an inverted U behavior. This behavior appears to be specific to
BOLD effects (as distinct from evoked changes in CBF) and may
represent a hemodynamic refractoriness. This effect has important
implications for event-related fMRI, where one may want to pres-
ent trials in quick succession.
In summary, we started with models of regionally specific
responses, framed in terms of the GLM, in which responses were
modeled as linear mixtures of designed changes in explanatory vari-
ables. Hierarchical extensions to linear observation models enable
random-effects analyses and, in particular, empirical Bayes. The
mechanistic utility of these models is realized though the use of for-
ward models that embody causal dynamics. Simple variants of these
are the linear convolution models used to construct explanatory
variables in conventional analyses of fMRI data. These are a special
case of generalized convolution models that are mathematically
equivalent to input-state-output systems comprising hidden states.
Estimation and inference with these dynamic models tells us some-
thing about how the response was caused, but only at the level of a
single voxel. Section 4 retains the same perspective on models, but
in the context of distributed responses and functional integration.
4.2 Dynamic Causal This section is about modeling interactions among neuronal popu-
Modeling lations, at a cortical level, using neuroimaging time series and
with Bilinear Models dynamic causal models that are informed by the biophysics of the
system studied. The aim of DCM [31] is to estimate, and make
inferences about, the coupling among brain areas and how that
coupling is influenced by experimental changes (e.g., time or cog-
nitive set). The basic idea is to construct a reasonably realistic neu-
ronal model of interacting cortical regions or nodes. This model is
then supplemented with a forward model of how neuronal or syn-
aptic activity translates into a measured response (see previous sec-
tion). This enables the parameters of the neuronal model (i.e.,
effective connectivity) to be estimated from observed data.
Dynamic Causal Modeling of Brain Responses 255
x = f ( x,u ) = Ax + uBx + Cu
y = g ( x) + e (7)
¶f ( 0,0 ) ¶ f ( 0,0 )
2
¶f ( 0,0 )
A= B= C=
¶x ¶x¶u ¶u
where x = ¶x / ¶t . This is an approximation to any model of how
changes in neuronal activity in one region xi are caused by activity
in the other regions. Here the output function g(x) embodies a
hemodynamic convolution, linking neuronal activity to BOLD, for
each region (e.g., that in Fig. 3). The matrix A represents the cou-
pling among the regions in the absence of input u(t). This can be
thought of as the endogenous coupling in the absence of experi-
mental perturbations. The matrix B is effectively the change in
coupling induced by the input. It encodes the input-sensitive
changes in A or, equivalently, the modulation of coupling by
experimental manipulations. Because B is a second-order deriva-
tive, it is referred to as bilinear. Finally, the matrix C embodies the
exogenous influences of inputs on neuronal activity. The parame-
ters θ = A,B, and C are the connectivity or coupling matrices that
we wish to identify and define the functional architecture and
interactions among brain regions at a neuronal level. They play the
same role as rate constant in kinetic models and therefore have
units of Hertz or per second.
256 Karl J. Friston
Because Eq. (7) has exactly the same form as Eq. (4), we can
express it as a GLM and estimate the parameters using EM in the
usual way (see Ref. [31]). Generally, estimation in the context of
highly parameterized models like DCMs requires constraints in the
form of priors. These priors enable conditional inference about the
connectivity estimates. The sorts of questions that can be addressed
with DCMs are now illustrated by looking at how attentional mod-
ulation is mediated in sensory processing hierarchies in the brain.
4.2.1 DCM and It has been established that the superior parietal cortex (SPC) exerts
Attentional Modulation a modulatory role on V5 responses using Volterra-based regression
models [34] and that the inferior frontal gyrus (IFG) exerts a simi-
lar influence on SPC using structural equation modeling [32]. The
example here shows that DCM leads to the same conclusions but
starting from a completely different construct. The experimental
paradigm and data acquisition are described in the legend to Fig. 4.
This figure also shows the location of the regions that entered the
DCM. These regions were based on maxima from conventional
SPMs testing for the effects of photic stimulation, motion, and
Fig. 4 Results of a dynamic causal modeling (DCM) analysis of attention to visual motion with fMRI. Right
panel: Functional architecture based upon the conditional estimates shown alongside their connections, with
the percent confidence that they exceeded threshold in brackets. The most interesting aspects of this archi-
tecture involve the role of motion and attention in exerting bilinear effects. Critically, the influence of motion is
to enable connections from V1 to the motion-sensitive area V5. The influence of attention is to enable back-
ward connections from the inferior frontal gyrus (IFG) to the superior parietal cortex (SPC). Furthermore, atten-
tion increases the influence of SPC on V5. Dotted arrows connecting regions represent significant bilinear
effects in the absence of a significant intrinsic coupling. Left panel: Fitted responses based upon the condi-
tional estimates and the adjusted data are shown for each region in the DCM. The insert (upper left) shows the
location of the regions
Dynamic Causal Modeling of Brain Responses 257
4.2.2 Structural Equation The central idea, behind DCM, is to treat the brain as a determin-
Modeling as a Special istic nonlinear dynamic system that is subject to inputs and pro-
Case of DCM duces outputs. Effective connectivity is parameterized in terms of
coupling among unobserved brain states (e.g., neuronal activity in
different regions). The objective is to estimate these parameters by
perturbing the system and measuring the response. This is in con-
tradistinction to established methods for estimating effective con-
nectivity from neurophysiological time series, which include SEM
and models based on multivariate auto-regressive processes. In
these models, there is no designed perturbation and the inputs are
treated as unknown and stochastic. Furthermore, the inputs are
often assumed to express themselves instantaneously such that, at
the point of observation, the change in states is zero. From Eq.
(7), in the absence of bilinear effects, we have
x = 0 = Ax + Cu
(8)
x = - A-1Cu
4.3 Dynamic Causal ERPs have been used for decades as electrophysiological correlates
Modeling with Neural of perceptual and cognitive operations. However, the exact neuro-
Mass Models biological mechanisms underlying their generation are largely
unknown. In this section, we use neuronally plausible models to
understand event-related responses. Our example shows that
changes in connectivity are sufficient to explain certain ERP com-
ponents. Specifically, we will look at the MMN, a component asso-
ciated with rare or unexpected events. If the unexpected nature of
rare stimuli depends on learning which stimuli are frequent, then
the MMN must be due to plastic changes in connectivity that
mediate perceptual learning. We conclude by showing that advances
in the modeling of evoked responses now afford measures of con-
nectivity among cortical sources that can be used to quantify the
effects of perceptual learning.
4.3.1 Neural Mass The minimal model we have developed [35] uses the connectivity
Models rules described by Felleman and Van Essen [36] to assemble a net-
work of coupled sources. These rules are based on a partitioning of
the cortical sheet into supra-, infra-granular, and granular layer
(layer 4). Bottom-up or forward connections originate in agranular
layers and terminate in layer 4. Top-down or backward connections
target agranular layers. Lateral connections originate in agranular
layers and target all layers. These long-range or extrinsic cortico-
cortical connections are excitatory and arise from pyramidal cells.
Each region or source is modeled using a neural mass model
described by David and Friston [35], based on the model of Jansen
and Rit [37]. This model emulates the activity of a cortical area
using three neuronal subpopulations, assigned to granular and
agranular layers. A population of excitatory pyramidal (output)
cells receives inputs from inhibitory and excitatory populations of
inter-neurons, via intrinsic connections (intrinsic connections are
confined to the cortical sheet). Within this model, excitatory inter-
neurons can be regarded as spiny stellate cells found predominantly
in layer 4 and in receipt of forward connections. Excitatory pyra-
midal cells and inhibitory inter-neurons are considered to occupy
agranular layers and receive backward and lateral inputs (Fig. 5).
To model event-related responses, the network receives inputs
via input connections. These connections are exactly the same as
forward connections and deliver inputs to the spiny stellate cells in
layer 4. In the present context, inputs u(t) model subcortical audi-
tory inputs. The vector C controls the influence of the input on
each source. The lower, upper, and leading diagonal matrices
AF,AB, AL encode forward, backward, and lateral connections,
respectively. The DCM here is specified in terms of the state equa-
tions shown in Fig. 5 and a linear output equation
260 Karl J. Friston
Fig. 5 Schematic of the dynamic causal modeling (DCM) used to model electrical responses. This schematic shows
the state equations describing the dynamics of sources or regions. Each source is modeled with three subpopula-
tions (pyramidal, spiny stellate, and inhibitory inter-neurons) as described in the main text. These have been
assigned to granular and agranular cortical layers that receive forward and backward connections, respectively
x = f ( x,u )
(9)
y = Lx0 + e
2
Propagation delays on the extrinsic connections have been omitted for clarity
here and in Fig. 5.
Dynamic Causal Modeling of Brain Responses 261
x7 = x8
H 2x x (10)
x8 = e éë( AB + AL + g 3 I ) S ( x0 ) ùû - 8 - 72
te te te
4.3.2 Perceptual The example shown in Fig. 6 is an attempt to model the MMN in
Learning and the MMN terms of changes in backward and lateral connections among corti-
cal sources. In this example, two (averaged) channels of EEG data
were modeled with three cortical sources. Using this generative or
forward model, we estimated differences in the strength of these
connections for rare and frequent stimuli. As expected, we could
account for detailed differences in the ERPs (the MMN) by
changes in connectivity (see figure legend for details). Interestingly,
these differences were expressed selectively in the lateral connec-
tions. If this model is a sufficient approximation to the real sources,
these changes are a noninvasive measure of plasticity, mediating
perceptual learning, in the human brain.
5 Conclusion
Fig. 6 Summary of a dynamic causal modeling (DCM) analysis of event-related potentials (ERPs) elicited during
an auditory oddball paradigm, employing rare and frequent pure tones. Upper panel: Schematic showing the
architecture of the neuronal model used to explain the empirical data. Sources were coupled with extrinsic cor-
tico-cortical connections following the rules of Felleman and van Essen. The free parameters of this model
included intrinsic and extrinsic connection strengths that were adjusted to best explain the data. In this example,
the lead field was also estimated, with no spatial constraints. The parameters were estimated for ERPs recorded
during the presentation of rare and frequent tones and are reported beside their corresponding connection (fre-
quent/rare). The most notable finding was that the mismatch response could be explained by a selective increase
in lateral connection strength from 0.1 to 3.68 Hz (highlighted in bold). Lower panel: The channel positions (left)
and ERPs (right) averaged over two subsets of channels (circled on the left). Note the correspondence between
the measured ERPs and those generated by the model. Auditory stimuli, 1,000 or 2,000 Hz tones with 5 ms rise
and fall times and 80 ms duration, were presented binaurally. The tones were presented for 15 min, every 2 s in
a pseudo-random sequence with 2000-Hz tones occurring 20 % of the time and 1,000-Hz tones occurring 80 %
of the time. The subject was instructed to keep a mental record of the number of 2000-Hz tones (nonfrequent
target tones). Data were acquired using 128 EEG electrodes with 1,000 Hz sample frequency. Before averaging,
data were referenced to mean earlobe activity and band-pass filtered between 1 and 30 Hz. Trials showing ocular
artifacts and bad channels were removed from further analysis
Dynamic Causal Modeling of Brain Responses 263
References
1. Staum M (1995) Physiognomy and phrenol- magnetic resonance imaging in human V1.
ogy at the Paris Athénée. J Hist Ideas J Neurosci 16:4207–4221
6:443–462 12. Friston KJ, Frith CD, Turner R, Frackowiak
2. Phillips CG, Zeki S, Barlow HB (1984) RSJ (1995) Characterising evoked hemody-
Localisation of function in the cerebral cortex: namics with fMRI. NeuroImage 2:157–165
past present and future. Brain 107:327–361 13. Josephs O, Turner R, Friston KJ (1997) Event-
3. Goltz F (1881) Transactions of the 7th interna- related fMRI Hum. Brain Mapp 5:243–248
tional medical congress (W. MacCormac, Ed.), 14. Lange N, Zeger SL (1997) Non-linear Fourier
Vol. I, JW Kolkmann: London, 1881:218–228 time series analysis for human brain mapping by
4. Absher JR, Benson DF (1993) Disconnection functional magnetic resonance imaging (with
syndromes: an overview of Geschwind’s contri- discussion). J Roy Stat Soc Ser C 46:1–29
butions. Neurology 43:862–867 15. Buxton RB, Frank LR (1997) A model for the
5. Zeki S (1990) The motion pathways of the coupling between cerebral blood flow and oxy-
visual cortex. Vision: coding and efficiency gen metabolism during neural stimulation.
(C. Blakemore, Ed.). Cambridge University J Cereb Blood Flow Metab 17:64–72
Press, pp 321–345 16. Mandeville JB, Marota JJ, Ayata C, Zararchuk
6. Friston KJ, Frith CD, Liddle PF, Frackowiak G, Moskowitz MA, Rosen B, Weisskoff RM
RSJ (1991) Comparing functional (PET) (1999) Evidence of a cerebrovascular postarte-
images: the assessment of significant change. riole Windkessel with delayed compliance.
J Cereb Blood Flow Metab 11:690–699 J Cereb Blood Flow Metab 19:679–689
7. Berry DA, Hochberg Y (1999) Bayesian per- 17. Hoge RD, Atkinson J, Gill B, Crelier GR,
spectives on multiple comparisons. J Stat Plan Marrett S, Pike GB (1999) Linear coupling
Infer 82:215–227 between cerebral blood flow and oxygen
8. Holmes A, Ford I (1993) A Bayesian approach consumption in activated human cortex. Proc
to significance testing for statistic images from Natl Acad Sci 96:9403–9408
PET. In: Uemura K, Lassen NA, Jones T, 18. Friston KJ, Mechelli A, Turner R, Price CJ
Kanno I (eds) Quantification of brain function, (2000) Nonlinear responses in fMRI: the
tracer kinetics and image analysis in brain Balloon model, Volterra kernels, and other
PET. Excerpta Medica, Int. Cong. Series No. hemodynamics. NeuroImage 12:466–477
1993. 1030:521–534 19. Fliess M, Lamnabhi M, Lamnabhi-Lagarrigue
9. Dempster AP, Laird NM, Rubin (1977) F (1983) An algebraic approach to nonlinear
Maximum likelihood from incomplete data via functional expansions. IEEE Trans Circuits
the EM algorithm. J Roy Stat Soc B 39:1–38 Syst 30:554–570
10. Friston KJ, Jezzard P, Turner R (1994) Analysis 20. Bendat JS (1990) Nonlinear system analysis
of functional MRI time series. Human Brain and identification from random data. John
Map 1:153–171 Wiley, New York
11. Boynton GM, Engel SA, Glover GH, Heeger 21. Gerstein GL, Perkel DH (1969) Simultaneously
DJ (1996) Linear systems analysis of functional recorded trains of action potentials: analysis
264 Karl J. Friston
Abstract
Imparting functional meaning to neuroanatomical location has been among the greatest challenges to
neuroscientists. The characterization of the brain architecture responsible in human cognition received a
boost in momentum with the emergence of in vivo functional and structural neuroimaging technology
over the past 30 years. Yet, individual variability in cortical gyrification as well as the patterns of blood flow-
related activity measured using fMRI and positron emission tomography complicated direct comparisons
across subjects without spatially accounting for overall brain size and shape. This realization resulted in
considerable effort now involving the collective efforts of neuroscientists, computer scientists, and math-
ematicians to develop common brain atlas spaces against which the regions of activity may be accurately
referenced. We examine recent developments in brain imaging and computational anatomy that have
greatly expanded our ability to analyze brain structure and function. The enormous diversity of brain maps
and imaging methods has spurred the development of population-based digital brain atlases. Atlases store
information on how the brain varies across age and gender, across time, in health and disease, and in large
human populations. We describe how brain atlases, and the computational tools that align new datasets
with them, facilitate comparison of brain data across experiments, laboratories, and from different imaging
devices. The major philosophies are presented that underlie the construction of probabilistic atlases, which
store information on anatomic and functional variability in a population. Algorithms which create compos-
ite brain maps and atlases based on multiple subjects are examined. We show that group patterns of cortical
organization, asymmetry, and disease-specific trends can be resolved that may not be apparent in individual
brain maps. Finally, we describe the development of four-dimensional maps that store information on the
dynamics of brain change in development and disease.
Key words Brain atlases, Neuroanatomy, Diffeomorphism, Warping, Functional activity, Inference
1 Introduction
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_9, © Springer Science+Business Media New York 2016
265
266 John Darrell Van Horn and Arthur W. Toga
Fig. 1 A variety of neuroimaging methods permit the acquisition of brain data over time and space having a
range of resolution granularity. Moreover, variation across individuals and how this changes over the lifespan
must be accounted for in statistical examination of the data. Mapping these data to known spatial coordinate
systems enables highly accurate inference concerning the brain’s structural change over time, between popu-
lations, or in terms of localizing functional change. Atlases denoting this variation after spatial warping, the
characterization of shape, three-dimensional (3D) distortion, etc. will be essential in describing structural and
functional alteration associated with normal aging as well as in disease
268 John Darrell Van Horn and Arthur W. Toga
2.1 Basic Image Image registration is an elemental step in many of the analytic
Registration strategies involving brain imaging today [13]. Initially developed
as an image processing technique to spatially align one image to
match another, image registration now has a vast range of applica-
tions, such as automated image labeling and for pathology detec-
tion in individuals or groups [14]. Registration algorithms can
encode patterns of anatomic variability in large human popula-
tions, and can use this information to create disease-specific,
population-based brain atlases [15]. They may also blend data
from multiple imaging devices to correlate different measures of
brain structure and function. Finally, modern registration algo-
rithms and workflows can serve as a basic measure for patterns of
structural change during brain development, tumor growth, or
degenerative disease processes [16].
2.2 Geodesic The objective in geodesic approaches has been to encourage varia-
Averaging of Brain tional methods for anatomical averaging that operate within the
Shape space of the underlying image registration problem [17]. This
approach is effective when using a large deformation viscous frame-
work, where linear averaging might not be appropriate. The theory
behind it is similar to registration-based techniques but with single
image force replaced by the average forces from multiple sources.
These group forces drive an average transport ordinary differential
equation allowing one to estimate the geodesic that moves an
image toward the mean shape configuration. This model provides
large deformation atlases that are optimal with respect to the shape
manifold as defined by the data and the image registration assump-
tions. These procedures generate refined average representations
of highly variable anatomy from distinct populations. For example,
the population statistics have been used to show a significant dou-
bling of the relative prefrontal lobe size in humans, as compared to
nonhuman primates [18].
2.4 Label-Based In label-based approaches, large ensembles of brain data are labeled
Atlases or “segmented” by a human operator or algorithmically into
Brain Atlases: Their Development and Role in Functional Inference 269
2.5 Encoding Brain Measuring and accounting for the considerable variability in brain
Variation shape across human populations necessitates realistically complex
mathematical strategies to encode comprehensive information on
structural variability [20]. Particularly relevant is a three-
dimensional (3D) statistical information on group-specific patterns
of variation and how these patterns are altered in disease. This
information can be represented such that it can be exploited by
expert diagnostic systems, whose goal is to detect subtle or diffuse
structural alterations in disease [21]. Strategies for detecting struc-
tural anomalies can leverage information in anatomical databases
by invoking encoded knowledge on the variations in geometry and
location of neuroanatomic regions and critical functional inter-
faces, especially at the cortex.
2.7 Deformation When applied to two different 3D brain scans, a nonlinear registra-
Atlases tion or warping algorithm calculates a deformation map that
matches up brain structures in one scan with their counterparts in
the other. The deformation map indicates 3D patterns of anatomic
differences between the two subjects or populations [26]. In prob-
abilistic atlases based on deformation maps, statistical properties of
these deformation maps are encoded locally to determine the mag-
nitude and directional biases of anatomic variation [27]. Encoding
of local variation can then be used to assess the severity of struc-
tural variants outside of the normal range, which may be a sign of
270 John Darrell Van Horn and Arthur W. Toga
Normal Controls
Males Females
Schizophrenic Patients
Males Females
Millimeters of Asymmetry
0 16
2.8 Disease-Specific Disease-specific atlases are designed to reflect the unique anatomy
Atlases and physiology of a particular clinical subpopulation. Based on
well-characterized patient groups, these atlases contain thousands
of structure models, as well as composite maps, average templates,
and visualizations of structural variability, asymmetry, and group-
specific differences. They act as a quantitative framework that cor-
relates the structural, metabolic, molecular, and histologic hallmarks
Brain Atlases: Their Development and Role in Functional Inference 271
2.9 Genetic Atlases Inclusion of genetic data in an atlas makes it possible to go beyond
simply describing the effects of a disease on the brain to investigating
its fundamental causes. This not only allows the direct mapping of
genetic influences on brain structure, but also allows us to quantify
heritability for different features of the brain. Familial, twin, and
genetic linkage studies have recently begun to expand the atlas con-
cept to tie together genetic and imaging studies of disease [7, 29].
Atlases that contain genetic brain maps, and a means to analyze them,
can help screen relatives for inherited disease. They also offer a frame-
work to mine large imaging databases for risk genes and quantitative
trait loci, as well as genetic and environmental triggers of disease.
2.10 Age The brain changes remarkably in its size and complexity over the lifes-
and Developmental pan. There is considerable need to account for the age of particular
Stratification populations in the context of brain maturation and the development of
age-stratified normal brain atlas spaces [30]. People who are mildly
cognitively impaired, for instance, are at a fivefold increased risk of
imminent conversion to dementia, and present specific structural brain
changes that are predictive of imminent disease onset [31, 32].
Language impairment in AD patients is also correlated with cortical
atrophy in the left temporal and parietal lobes, bilateral frontal lobes,
and the right temporal pole [33]. However, characterizing such change
presents particular computational challenges. The fitting of brain anat-
omy to a single template of undetermined age specification may lead to
errors in inference about brain morphometry of function relative to an
inappropriate underlying template. Alternative approaches can also be
fruitful and metrics, such as shape [34], cortical thickness mapping,
tensor-based morphometry (TBM), may be better suited for shedding
light on the neuroscience of aging and brain degeneration in AD and
mild cognitive impairment (MCI) [35].
6 Conclusions
References
1. Haas LF (2001) Phineas Gage and the science 9. Davatzikos C (1996) Spatial normalization of
of brain localisation. J Neurol Neurosurg 3D brain images using deformable models.
Psychiatry 71:761 J Comput Assist Tomogr 20:656–665
2. Cowie SE (2000) A place in history: Paul Broca 10. Davatzikos C (1997) Spatial transformation
and cerebral localization. J Invest Surg and registration of brain images using elasti-
13:297–298 cally deformable models. Comput Vis Image
3. Goedert M, Ghetti B (2007) Alois Alzheimer: Underst 66:207–222
his life and times. Brain Pathol 17:57–62 11. Thompson PM, Woods RP, Mega MS, Toga
4. Roland PE, Zilles K (1994) Brain atlases – a AW (2000) Mathematical/computational
new research tool. Trends Neurosci challenges in creating deformable and probabi-
17:458–467 listic atlases of the human brain. Hum Brain
5. Toga AW, Thompson PM (2001) Maps of the Mapp 9:81–92
brain. Anat Rec 265:37–53 12. Weaver JB, Healy DM Jr, Periaswamy S,
6. Toga AW, Thompson PM (2002) New Kostelec PJ (1998) Elastic image registration
approaches in brain morphometry. Am using correlations. J Digit Imaging 11:59–65
J Geriatr Psychiatry 10:13–23 13. Barillot C, Lemoine D, Le Briquer L,
7. Thompson P, Cannon TD, Toga AW (2002) Lachmann F, Gibaud B (1993) Data fusion in
Mapping genetic influences on human brain medical imaging: merging multimodal and
structure. Ann Med 34:523–536 multipatient images, identification of struc-
tures and 3D display aspects. Eur J Radiol
8. Narr KL, Thompson PM, Sharma T, Moussai 17:22–27
J, Cannestra AF, Toga AW (2000) Mapping
morphology of the corpus callosum in schizo- 14. Woods RP, Grafton ST, Holmes CJ, Cherry
phrenia. Cereb Cortex 10:40–49 SR, Mazziotta JC (1998) Automated image
280 John Darrell Van Horn and Arthur W. Toga
registration. I. General methods and intrasu- 29. Toga AW, Thompson PM (2005) Genetics of
bject, intramodality validation. J Comput Assist brain structure and intelligence. Annu Rev
Tomogr 22:139–152 Neurosci 28:1–23
15. Toga AW, Thompson PM, Mori S, Amunts K, 30. Toga AW, Thompson PM, Sowell ER (2006)
Zilles K (2006) Towards multimodal atlases of Mapping brain maturation. Trends Neurosci
the human brain. Nat Rev Neurosci 7:952–966 29:148–159
16. Woods RP (2003) Characterizing volume and 31. Apostolova LG, Thompson PM (2007) Brain
surface deformations in an atlas framework: mapping as a tool to study neurodegeneration.
theory, applications, and implementation. Neurotherapeutics 4(3):387–400
Neuroimage 18:769–788 32. Apostolova LG, Akopyan GG, Partiali N et al
17. Avants B, Gee JC (2004) Geodesic estimation (2007) Structural correlates of apathy in
for large deformation anatomical shape averag- Alzheimer’s disease. Dement Geriatr Cogn
ing and interpolation. Neuroimage 23(Suppl Disord 24:91–97
1):S139–S150 33. Apostolova LG, Lu P, Rogers S et al (2008) 3D
18. Avants BB, Schoenemann PT, Gee JC (2006) mapping of language networks in clinical and
Lagrangian frame diffeomorphic image regis- pre-clinical Alzheimer’s disease. Brain Lang
tration: morphometric comparison of human 104:33–41
and chimpanzee cortex. Med Image Anal 34. Scher AI, Xu Y, Korf ES et al (2007)
10:397–412 Hippocampal shape analysis in Alzheimer’s dis-
19. Evans AC, Collins DL, Milner B (1992) An ease: a population-based study. Neuroimage
MRI-based stereotactic atlas from 250 young 36:8–18
normal subjects. J Neurosci Abstr 18:408 35. Thompson PM, Hayashi KM, Dutton RA et al
20. Durrleman S, Pennec X, Trouve A, Ayache N (2007) Tracking Alzheimer’s disease. Ann N Y
(2007) Measuring brain variability via sulcal Acad Sci 1097:183–214
lines registration: a diffeomorphic approach. 36. Mazziotta JC, Toga AW, Evans AC, Fox PT,
Med Image Comput Comput Assist Interv Lancaster JL (1995) Digital brain atlases.
10(Pt 1):675–682 Trends Neurosci 18:210–211
21. Alayon S, Robertson R, Warfield SK, Ruiz- 37. Toga AW, Thompson PM, Mega MS, Narr KL,
Alzola J (2007) A fuzzy system for helping Blanton RE (2001) Probabilistic approaches
medical diagnosis of malformations of cortical for atlasing normal and disease-specific brain
development. J Biomed Inform 40:221–235 variability. Anat Embryol (Berl) 204:267–282
22. Rohlfing T, Maurer CR Jr (2007) Shape-based 38. Wakana S, Jiang H, Nagae-Poetscher LM, van
averaging. IEEE Trans Image Process Zijl PC, Mori S (2004) Fiber tract-based atlas
16:153–161 of human white matter anatomy. Radiology
23. Narr KL, Bilder RM, Luders E et al (2007) 230:77–87
Asymmetries of cortical shape: effects of hand- 39. Van Essen DC (2005) A Population-Average,
edness, sex and schizophrenia. Neuroimage Landmark- and Surface-based (PALS) atlas of
34:939–948 human cerebral cortex. Neuroimage
24. Thompson PM, Giedd JN, Woods RP, 15:635–662
MacDonald D, Evans AC, Toga AW (2000) 40. Fox PT, Perlmutter JS, Raichle ME (1984)
Growth patterns in the developing brain Stereotactic method for determining anatomi-
detected by using continuum mechanical ten- cal localization in physiological brain images.
sor maps. Nature 404:190–193 J Cereb Blood Flow Metab 4:634
25. Corouge I, Dojat M, Barillot C (2004) 41. Evans AC, Marrett S, Neelin P et al (1992)
Statistical shape modeling of low level visual Anatomical mapping of functional activation in
area borders. Med Image Anal 8:353–360 stereotactic coordinate space. Neuroimage
26. Cardenas VA, Boxer AL, Chao LL et al (2007) 1:43–53
Deformation-based morphometry reveals brain 42. Nowinski WL, Thirunavuukarasuu A (2001)
atrophy in frontotemporal dementia. Arch Atlas-assisted localization analysis of functional
Neurol 64:873–877 images. Med Image Anal 5:207–220
27. Leow AD, Klunder AD, Jack CR Jr et al (2006) 43. Crivello F, Schormann T, Tzourio-Mazoyer N,
Longitudinal stability of MRI for mapping Roland PE, Zilles K, Mazoyer BM (2002)
brain change using tensor-based morphometry. Comparison of spatial normalization proce-
Neuroimage 31:627–640 dures and their impact on functional maps.
28. Diedrichsen J (2006) A spatially unbiased atlas Hum Brain Mapp 16:228–250
template of the human cerebellum. Neuroimage 44. Swallow KM, Braver TS, Snyder AZ, Speer
33:127–138 NK, Zacks JM (2003) Reliability of functional
Brain Atlases: Their Development and Role in Functional Inference 281
localization using fMRI. Neuroimage and nonlinear models. J Comput Assist Tomogr
20:1561–1577 22:153–165
45. Tu Z, Zheng S, Yuille AL et al (2007) 60. Rex DE, Ma JQ, The TAW, LONI (2003)
Automated extraction of the cortical sulci Pipeline processing environment. Neuroimage
based on a supervised learning approach. IEEE 19:1033–1048
Trans Med Imaging 26:541–552 61. Smith SM, Jenkinson M, Woolrich MW et al
46. Luders E, Thompson PM, Narr KL, Toga AW, (2004) Advances in functional and structural
Jancke L, Gaser C (2006) A curvature-based MR image analysis and implementation as
approach to estimate local gyrification on the FSL. Neuroimage 23(Suppl 1):S208–S219
cortical surface. Neuroimage 29:1224–1230 62. Ashburner J, Friston KJ (2005) Unified seg-
47. Ashburner J, Friston KJ (1999) Nonlinear spa- mentation. Neuroimage 26:839–851
tial normalization using basis functions. Hum 63. Van Essen DC (2002) Windows on the brain:
Brain Mapp 7:254–266 the emerging role of atlases and databases in
48. Friston KJ, Stephan KE, Lund TE, Morcom A, neuroscience. Curr Opin Neurobiol
Kiebel S (2005) Mixed-effects and fMRI stud- 12:574–579
ies. Neuroimage 24:244–252 64. Mazziotta J, Toga A, Evans A et al (2001) A
49. Miller MB, Van Horn JD, Wolford GL et al four-dimensional probabilistic atlas of the
(2002) Extensive individual differences in human brain. J Am Med Inform Assoc
brain activations associated with episodic 8:401–430
retrieval are reliable over time. J Cogn Neurosci 65. Mega MS, Dinov ID, Mazziotta JC et al
14:1200–1214 (2005) Automated brain tissue assessment in
50. Fox PT, Parsons LM, Lancaster JL (1998) the elderly and demented population: con-
Beyond the single study: function/location struction and validation of a sub-volume prob-
metanalysis in cognitive neuroimaging. Curr abilistic brain atlas. Neuroimage
Opin Neurobiol 8:178–187 26:1009–1018
51. Nowinski WL (2005) The cerefy brain atlases: 66. Yoon U, Lee JM, Koo BB et al (2005)
continuous enhancement of the electronic Quantitative analysis of group-specific brain
talairach-tournoux brain atlas. tissue probability map for schizophrenic
Neuroinformatics 3:293–300 patients. Neuroimage 26:502–512
52. Amunts K, Schleicher A, Zilles K (2007) 67. Cannon TD, Thompson PM, van Erp TG et al
Cytoarchitecture of the cerebral cortex – more (2006) Mapping heritability and molecular
than localization. Neuroimage 37:1061–1065, genetic associations with cortical features using
discussion 6–8 probabilistic brain atlases: methods and applica-
53. Mazziotta J, Toga AW, Evans A et al (2001) A tions to schizophrenia. Neuroinformatics 4:5–19
probabilistic atlas and reference system for the 68. Wilke M, Schmithorst VJ, Holland SK (2002)
human brain: International Consortium for Assessment of spatial normalization of whole-
Brain Mapping (ICBM). Philos Trans R Soc brain magnetic resonance images in children.
Lond B Biol Sci 356:1293–1322 Hum Brain Mapp 17:48–60
54. Nowinski WL (2001) Modified Talairach land- 69. Jelacic S, de Regt D, Weinberger E (2006)
marks. Acta Neurochir (Wien) 143:1045–1057 Interactive digital MR atlas of the pediatric
55. Bookstein FL (2001) Voxel-based morphome- brain. Radiographics 26:497–501
try” should not be used with imperfectly regis- 70. Joshi S, Davis B, Jomier M, Gerig G (2004)
tered images. Neuroimage 14:1454–1462 Unbiased diffeomorphic atlas construction for
56. Talairach J, Tournoux P (1988) Co-planar stereo- computational anatomy. Neuroimage 23(Suppl
tactic atlas of the human brain. Tieme, New York 1):S151–S160
57. Maldjian JA, Laurienti PJ, Burdette JH (2004) 71. Mazziotta J, Toga A, Evans A et al (2001) A
Precentral gyrus discrepancy in electronic ver- four-dimensional probabilistic atlas of the
sions of the Talairach atlas. Neuroimage human brain. J Am Med Inform Assoc
21:450–455 8:401–430
58. Shattuck DW, Mirza M, Adisetiyo V et al 72. Narr K, Thompson P, Sharma T et al (2001)
(2008) Construction of a 3D probabilistic atlas Three-dimensional mapping of gyral shape and
of human cortical structures. Neuroimage cortical surface asymmetries in schizophrenia:
39:1064–1080 gender effects. Am J Psychiatry 158:244–255
59. Woods RP, Grafton ST, Watson JD, Sicotte 73. Amunts K, Hawrylycz MJ, Van Essen DC et al
NL, Mazziotta JC (1998) Automated image (2014) Interoperable atlases of the human
registration. II. Intersubject validation of linear brain. Neuroimage 99:525–532
Chapter 10
Abstract
The human brain is a highly interconnected network. It is thus suitable for investigation with graph theory,
a branch of mathematics concerned with understanding systems of interacting elements. Graph theory has
become a popular tool for analyzing human MRI data. In this work, brain networks are modeled as graphs
of nodes connected by edges. The nodes represent distinct brain regions and the edges represent some
measure of structural or functional interaction between regions. This representation enables the computa-
tion of a broad range of metrics that quantify diverse aspects of network organization, thus offering a pow-
erful framework for understanding brain structure and function in both health and disease. This chapter
overviews the principles and methods involved in building and analyzing graph theoretic models of the
brain using MRI. It explains basic concepts, provides examples of how graph theory has shed new light on
brain organization, and considers some limitations of current applications.
Key words Connectome, Connectivity, Graph analysis, Network, Complexity, MRI, DTI, fMRI
1 Introduction
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_10, © Springer Science+Business Media New York 2016
283
284 Alex Fornito
Fig. 1 Matrix and graph-based representations of brain networks. In matrix form (top row), each row and col-
umn represents a different region and each element represents the connectivity between region pairs. Matrices
can be either binary, representing only the presence or absence of a connection (left and middle), or weighted
to represent variations in the strength of inter-regional connectivity (right). The matrices can also be symmet-
ric, representing an undirected network (left; note how the top-right and bottom left triangles of this matrix are
mirror images of each other), or asymmetric, to represent a directed network (middle and right). In graph form
(bottom row), brain regions are represented as nodes or circles and connectivity as edges. Arrowheads can be
used to represent the directionality of connectivity in directed networks (middle, right). Edge thickness can be
used to represent variations in edge weight (right). These graphs are used for illustrative purposes and do not
directly map onto the matrices shown in the top row. Weighted, undirected networks are not shown
Graph Theoretic Analysis of Human Brain Networks 285
Fig. 2 Basic processing pipeline for graph theoretic analysis of MRI data. (a) Imaging data are first acquired.
Structural connectivity is often assessed using either diffusion MRI (top) or T1-weighted MRI (middle).
Functional and effective connectivity are typically investigated using fMRI (bottom). (b) Once the imaging data
have been acquired, the brain must be parcellated into distinct regions, which act as network nodes. Shown
here are examples of an anatomical parcellation (top), a random parcellation (middle), and a functional parcel-
lation (bottom). (c) The next step is to define some measure of connectivity between nodes, which will repre-
sent the edges of our brain graph. With diffusion MRI, inter-regional connectivity is measured using tractography
(top). With T1-weighted MRI, structural connectivity is indirectly measured using inter-subject covariations in
gray matter morphometry (middle). With fMRI, functional connectivity is measured as a statistical dependence
between regional time series (bottom). (d) The connectivity between all pairs of brain regions can be repre-
sented as a connectivity matrix. MRI analyses typically yield weighted and symmetric matrices (left). These
matrices can be thresholded to emphasize the strongest links in the network (right). (e) The connectivity matrix
can be used to generate a graph-based representation of the network (i.e., a brain graph), in which regions are
represented as nodes and connections as edges. See also Figs. 1 and 3. Parts of this figure have been repro-
duced from [109] with permission
286 Alex Fornito
Nodes and edges are the fundamental building block of any net-
work graph. The basic unit of analysis in an MRI experiment is a
voxel, which is typically 1–3 mm3 in volume. Voxels thus represent
an aggregation of large populations of neural elements; on average
an estimated 20,000–30,000 neurons and billions of synapses
[45]. This coarse resolution creates ambiguities when attempting
to define appropriate nodes and edges for network analysis. This
problem is critical as invalid node definitions can alter, distort, or
bias results, and accurate mapping of connectivity (edges) is essen-
tial for any valid analysis of network organization [46–50].
4 Defining Nodes
The central problem for node definition in MRI concerns how vox-
els should be aggregated to define a valid parcellation of the brain.
Individual neurons and neuronal columns have both been proposed
as fundamental units for brain network organization [7, 51], but
cannot be resolved with typical human MRI acquisitions. Similarly,
cytoarchitectonic regions, such as those delineated in Brodmann’s
classic map, cannot be resolved with MRI and the boundaries of
these regions are often poorly correlated with macroscopic land-
marks (e.g., sulci and gyri) [52, 53]. Due to these limitations, several
different heuristic approaches have been used for node definition
with MRI.
To this end, three criteria for an ideal node for a brain graph
have been proposed [18]: (1) spatial embedding; (2) intrinsic
homogeneity; and (3) extrinsic heterogeneity. The first criterion
simply means that spatial relationships between nodes should be
taken into account. The brain is embedded within the
three-dimensional volume of the skull and this embedding places
important constraints on network wiring [54, 55]. Fortunately,
spatial relations between nodes are easily accounted for in MRI
analysis with standard stereotactic mapping techniques, since the
location of each region can be easily represented with reference to
the Montreal Neurological Institute (MNI) or Talairach and
Tourneoux coordinate systems.
The second and third criteria simply mean that each node
should represent a structurally or functionally homogeneous entity
(intrinsic homogeneity) and should be distinguishable from other
nodes (extrinsic heterogeneity). For example, a cytoarchitectonic
region is intrinsically homogeneous, to the extent that it defines a
population of neurons with shared histological properties. Different
regions are extrinsically heterogeneous, to the extent that they
serve different functional roles in the network (e.g., areas of visual
290 Alex Fornito
4.2 Anatomical Anatomical atlases offer an alternative method for brain parcella-
atlases tion that minimizes computational burden. For example, one pop-
ular atlas, the Automated Anatomical Labelling (AAL) atlas [61],
parcellates the brain into 116 regions largely defined according to
the sulcal and gyral landmarks of an individual brain. The general-
izability of this template is thus questionable. Alternative atlases,
based on probabilistic maps of sulcal and gyral regions, such as the
Harvard-Oxford atlas (http://fsl.fmrib.ox.ac.uk/fsl/fslwiki/
Graph Theoretic Analysis of Human Brain Networks 291
4.3 Random One way to ensure that a parcellation comprises regions with homo-
Parcellation geneous volume is to divide the brain into random parcels of similar
size (e.g., Fig. 2b, middle). These parcellations can be performed at
varying resolutions, typically yielding networks between 102 and 104
nodes [12, 48, 50]. This approach ensures homogeneity of regional
volume, but there is no guarantee that the random parcels accurately
capture true functional subdivisions of the brain. Replication of
results across several iterations of a random template may also be
necessary to ensure that any findings are not due to the specific char-
acteristics of any single instance of a random parcellation.
4.7 Summary There is a variety of methods for delineating brain network nodes
in MRI data. Each approach has distinct strengths and weaknesses
and there is no gold standard. Ultimately, investigators must
choose an approach best suited to the specific hypothesis being
tested, and ensure that analyses are interpreted with respect to the
limitations of the specific method employed.
Graph Theoretic Analysis of Human Brain Networks 293
5 Defining edges
5.3 Effective Effective connectivity is the influence of one neural system over
Connectivity another [91]. It thus quantifies causal interactions amongst brain
regions and the resulting connectivity estimates are directed
(Fig. 1). Critically, the causal interactions that define effective con-
nectivity must be specified at the neuronal level. Because neuronal
dynamics are not directly observable with fMRI, estimating effec-
tive connectivity with this imaging technique requires a model that
maps the observed hemodynamic signal changes to the underlying
neuronal dynamics from which they were generated [105].
The most popular framework for effective connectivity analysis
of fMRI data is dynamical causal modeling (DCM) [106]. DCM
uses a model of neurovascular coupling to specify the mapping
between neuronal activity and hemodynamics. Different graph
models of causal interactions (i.e., directed graph configurations)
between neuronal systems are specified and compared to deter-
mine which model best accounts for the observed fMRI data. As
the number of possible graph models rapidly increases with net-
work size, DCM has traditionally only been applied to relatively
small subnetworks comprising a few regions. Work is under way to
scale these methods to larger systems [107]. DCM is applicable to
both task and resting-state fMRI data [108].
In graph form, the rows and columns of the matrix—that is, the brain
regions—are depicted as nodes (often circles), and the matrix ele-
ments Aij determine which pairs of nodes are linked by edges.
Variations in connectivity strength encoded in weighted matrices are
commonly represented as variations in edge thickness (Fig. 1). The
directionality of connectivity that is encoded in asymmetric matrices is
depicted by arrowheads attached to the edges (Fig. 1).
Brain graphs can be projected in different ways, in order to
highlight specific relations between nodes. Common displays in
neuroscience include anatomical projections, in which nodes are
positioned according to their stereotactic coordinates; topological
projections, in which node positions are determined based on
some topological relation between nodes; and circular projections
(also called connectograms [111]), which allow a simplified view
of the connectivity of the entire network. Examples of each of
these projections are presented in Fig. 3.
Once the network has been mapped, it can be analysed with respect
to either its connectivity or topology. Connectivity analysis con-
centrates on variations in the type and strength of connectivity
between brain regions. Topological analysis is concerned with
understanding how connections are arranged with respect to each
other, and provides insight into key organizational principles of the
connectome.
Fig. 3 Different graph projections of a brain network. (a) A projection in anatomical space, where nodes are
positioned according to their stereotactic coordinates. Nodes and edges are colored according to the module
to which they have been assigned. (b) A topological projection, in which nodes are located more closely in
space if they have short path length between them. Nodes near the center of the graph have a short average
path length to other nodes, and thus represent central elements of the network. Nodes colors are the same as
in (a). (c) A ring projection, which is also called a connectogram. Nodes are clustered (and colored) by the
modules to which they belong. Within each grouping, the nodes have been ordered according to their averaged
connectivity strength with other nodes (yellow bars). The nodes have been labeled using arbitrary numbers
Graph Theoretic Analysis of Human Brain Networks 299
8 Connectivity Analysis
9 Topological Analysis
9.1 Hub Dominance Barabási and Albert’s [25] discovery that many real-world net-
works have a heterogeneous distribution of connectivity across
nodes suggests that such networks possess highly connected hubs
that exert a disproportionate influence over the network. In net-
work parlance, the total number of connections attached to a node
is called its connectivity degree, denoted k, and the distribution of
degree values across nodes is the degree distribution of a network.
In the real-world networks they studied, Barabási and Albert found
that the probability of finding a node with increasing k decayed as
a power-law of the form P ( k ) ~ k -a , where α is the scaling expo-
nent that determines the rate of decay. This decay is generally much
slower than the decay observed in homogeneous random networks
such as those studied by Erdős and Rényi, where the degree distri-
bution approximates a Gaussian (more accurately, it conforms to a
binomial distribution); i.e., most nodes have a degree value close
to the mean, and the probability of finding large deviations from
this mean is very low. This clustering around a mean value endows
these networks with a single, characteristic scale. In contrast,
power-law degree distributions are heavily skewed with an extended
tail, pointing to a higher probability of finding nodes with very
high connectivity values despite most nodes having low degree.
There is no meaningful average degree or characteristic scale in
these networks, so they are sometimes called scale-free.
Early work suggested that brain networks, at least when con-
structed at high resolution, conform to a power-law degree distri-
bution [37, 49, 58]. Other studies of both structural and functional
connectivity networks have more commonly reported evidence
Graph Theoretic Analysis of Human Brain Networks 301
Fig. 4 Example of key topological properties of brain networks. (a) illustration of a single network hub (red) with
high degree compared to other nodes. (b) example of the shortest path between two nodes at opposite ends
of the network (red). In this case, the shortest path traverses five edges, so the path length between these two
nodes is five. The characteristic path length of a network is the average path length between every pair of
nodes. (c) Example of a connected triangle of nodes (red). The clustering coefficient of a node is computed as
the number of such triangles attached to that node, relative to the total possible number of triangles. (d)
Illustration of a modular decomposition of the network. Three modules have been identified, as represented by
the different background colors. Nodes within a module are strongly connected with each other and sparsely
connected with nodes in other modules. Such a decomposition allows analysis of node roles and hub category.
Red highlights a provincial hub which is highly connected within its own module. Orange highlights a connec-
tor hub, which has connections distributed across all modules
to fragment the network. However, since hubs are relatively rare, the
probability that they will be affected by random failure is low; hence
the greater resistance to random node deletion relative to single-
scale systems. Compared to scale-free systems, networks with trun-
cated power-law degree distributions, such as those thought to
characterize human brain networks, show comparable resilience to
failure and better robustness in the face of targeted attack [40]. This
enhanced robustness occurs because the concentration of connectiv-
ity on hub nodes is less extreme than in scale-free systems.
Nonetheless, damage to hub nodes, or the links between them,
exerts a more severe impact on network function than damage to
Graph Theoretic Analysis of Human Brain Networks 303
9.3 Small-Worldness The small-world class of networks discovered by Watts and Strogatz
[26] provides an appealing model for the brain. Clustered connectiv-
ity offers a substrate for functional specialization, whereas short aver-
age path length facilities functional integration. Clustering is formally
quantified using the clustering coefficient, which computes the prob-
ability that two nodes linked to an index node are also connected
with each other. In other words, it counts the number of closed tri-
angles attached to a node (Fig. 4c). The path length between two
nodes is simply the number of edges on the shortest path intersecting
those nodes (Fig. 4b). The characteristic path length of a network is
the average path length computed across all node pairs.
Empirically, the small-worldness of a network can be quanti-
fied by comparison to an ensemble of random graphs matched for
the number of nodes, edges, and degree distribution. Various algo-
rithms are available for constructing such graphs by rewiring the
connections of an observed network [122, 134]. Such surrogate
networks provide a useful baseline for evaluating the degree to
which a particular topological property is expressed in the brain.
Formally, a network is considered small-world if the scalar
quantity s > 1, where s = g / l [135]. The quantity γ is computed
as a ratio of the observed clustering coefficient to the average clus-
tering of an ensemble of matched randomized networks (i.e.,
g = Cobs / Crand ). The quantity λ is the ratio of the observed network
path length to average path length computed in the same ensemble
of randomized graphs (i.e., l = Lobs / Lrand ). Small-world networks
will have greater clustering and similar path length to randomized
surrogates. As such, g > 1 and l ~ 1, yielding s > 1. Since σ is a
ratio, variations in this value may be driven by changes in either γ
or λ. It is therefore often more useful to understand variations in
those parameters before considering σ.
9.4 Cost The characteristic path length of a network is closely related to its
and Efficiency topological efficiency. Communication in a network will be more
efficient when fewer connections are required to transfer informa-
tion between any two nodes; i.e., when the characteristic path
length is low. We can thus define a topological measure of the
304 Alex Fornito
1 1
Eglob = å
N ( N - 1) i , j Lij
,
1 æ ki k j ö
Q= å ç Aij -
2 m i , jÎ N è 2m
÷ d si , s j ,
ø
Graph Theoretic Analysis of Human Brain Networks 305
9.6 Summary This section has presented a brief overview of some of the basic
graph theoretic concepts applied to neuroimaging data, and the
insights they have provided into brain network organization. Many
other metrics are available, enabling a diverse range of analyses (see
[159] for an introduction). It is important to bear in mind that
many such methods were developed in the physical or social sci-
ences with networks other than the brain in mind. As such, they
may not be directly portable to neuroscientific contexts. Indeed,
the first wave of graph theoretic studies of neuroimaging data have
concentrated on measuring canonical topological properties such
as those discussed here, but alternative measures may provide more
appropriate models of brain function. We consider some of these
issues in the next section.
Graph Theoretic Analysis of Human Brain Networks 307
References
1. Cherniak C (1990) The bounded brain: structural and functional systems. Nat Rev
toward quantitative neuroanatomy. J Cogn Neurosci 10:186–198
Neurosci 2:58–68 18. Fornito A, Zalesky A, Breakspear M (2013)
2. Sporns O, Tononi G, Kötter R (2005) The Graph analysis of the human connectome:
human connectome: a structural description promise, progress, and pitfalls. Neuroimage
of the human brain. PLoS Comput Biol 1:e42 80:426–444
3. Van Essen DC et al (2012) The Human 19. Euler L (1736) Solutio problematis ad geome-
Connectome Project: a data acquisition per- triam situs pertinentis. Commentarii Academiae
spective. Neuroimage 62:2222–2231 Scientiarum Imperialis Petropolitanae
4. Bohland JW et al (2009) A proposal for a 8:128–140
coordinated effort for the determination of 20. Luce RD, Perry AD (1949) A method
brainwide neuroanatomical connectivity in of matrix analysis of group structure.
model organisms at a mesoscopic scale. PLoS Psychometrika 14:95–116
Comput Biol 5:e1000334 21. Katz L (1947) On the matric analysis of
5. Kandel ER, Markram H, Matthews PM, sociometric data. Sociometry 10:233–241
Yuste R, Koch C (2013) Neuroscience thinks 22. Forsyth E, Katz L (1946) A matrix approach
big (and collaboratively). Nat Rev Neurosci to the analysis of sociometric data: prelimi-
14:659–664 nary report. Sociometry 9:340–347
6. White JG, Southgate E, Thomson JN, Brenner 23. Harary F, Norman RZ (1953) Graph theory
S (1986) The structure of the nervous sys- as a mathematical model in social science.
tem of the nematode Caenorhabditis elegans. University of Michigan Press
Philos Trans R Soc Lond B Biol Sci 314:1–340 24. Erdos P, Renyi A (1959) On random graphs.
7. Lichtman JW, Pfister H, Shavit N (2014) Publ Math Debrecen 6:290–297
The big data challenges of connectomics. Nat 25. Barabasi A, Albert R (1999) Emergence of scal-
Neurosci 17:1448–1454 ing in random networks. Science 286:509–512
8. Chiang A-S et al (2011) Three-dimensional 26. Watts DJ, Strogatz SH (1998) Collective
reconstructionof brain-wide wiring networks dynamics of ‘small-world’ networks. Nature
in Drosophila at single-cell resolution. Curr 393:440–442
Biol 21:1–11 27. Mesulam MM (1998) From sensation to cog-
9. Scannell JW, Young MP (1993) The connec- nition. Brain 121:1013–1052
tional organization of neural systems in the 28. Tononi G, Sporns O, Edelman GM (1994)
cat cerebral cortex. Curr Biol 3:191–200 A measure for brain complexity: relating
10. Shanahan M, Bingman VP, Shimizu T, functional segregation and integration in the
Gunturkun O (2013) Large-scale network nervous system. Proc Natl Acad Sci U S A
organization in the avian forebrain: a connec- 91:5033–5037
tivity matrix and theoretical analysis. Front 29. Friston KJ (2011) Functional and effec-
Comput Neurosci 7:1–17 tive connectivity: a review. Brain Connect
11. Stephan KE (2013) The history of CoCoMac. 1:13–36
NeuroImage 80:46–52 30. Felleman DJ, Van Essen DC (1991)
12. Hagmann P et al (2007) Mapping human Distributed hierarchical processing in the pri-
whole-brain structural networks with diffu- mate cerebral cortex. Cereb Cortex 1:1–47
sion MRI. PLoS One 2:e597 31. Scannell JW, Blakemore C, Young MP (1995)
13. Newman MJE (2003) The structure and Analysis of connectivity in the cat cerebral
function of complex networks. SIAM Rev cortex. J Neurosci 15:1463–1483
45:167–256 32. Hilgetag CC, Burns GA, O'Neill MA,
14. Boccaletti S, Latora V, Moreno Y, Chavez M, Scannell JW, Young MP (2000) Anatomical
Hwang DU (2006) Complex networks: struc- connectivity defines the organization of clus-
ture and dynamics. Phys Rep 424:175–308 ters of cortical areas in the macaque monkey
15. Axer M et al (2011) A novel approach to and the cat. Philos Trans R Soc Lond B Biol
the human connectome: ultra-high resolu- Sci 355:91–110
tion mapping of fiber tracts in the brain. 33. Sporns O, Tononi G, Edelman GM (2000)
Neuroimage 54:1091–1101 Theoretical neuroanatomy: relating anatomi-
16. Chung K, Deisseroth K (2013) CLARITY cal and functional connectivity in graphs and
for mapping the nervous system. Nat Meth cortical connection matrices. Cereb Cortex
10:508–513 10:127–141
17. Bullmore E, Sporns O (2009) Complex 34. Scannell JW, Burns GA, Hilgetag CC, O'Neil
brain networks: graph theoretical analysis of MA, Young MP (1999) The connectional
310 Alex Fornito
65. Fair DA et al (2007) Development of distinct spectral clustering. Hum Brain Mapp
control networks through segregation and 33:1914–1928
integration. Proc Natl Acad Sci U S A 79. Johansen-Berg H et al (2004) Changes in
104:13507–13512 connectivity profiles define functionally dis-
66. Andrews-Hanna JR, Reidler JS, Sepulcre J, tinct regions in human medial frontal cortex.
Poulin R, Buckner RL (2010) Functional- Proc Natl Acad Sci U S A 101:13335–13340
anatomic fractionation of the brain’s default 80. Behrens TE et al (2003) Non-invasive map-
network. Neuron 65:550–562 ping of connections between human thalamus
67. Dwyer DB et al (2014) Large-scale brain net- and cortex using diffusion imaging. Nat
work dynamics supporting adolescent cogni- Neurosci 6:750–757
tive control. J Neurosci 34:14096–14107 81. Anwander A, Tittgemeyer M, von Cramon
68. Cocchi L et al (2014) Complexity in rela- DY, Friederici AD, Knosche TR (2007)
tional processing predicts changes in func- Connectivity-based parcellation of Broca's
tional brain network dynamics. Cereb Cortex area. Cereb Cortex 17:816–825
24:2283–2296 82. Glasser MF, Van Essen DC (2011) Mapping
69. Fornito A, Harrison BJ, Zalesky A, Simons JS human cortical areas in vivo based on myelin
(2012) Competitive and cooperative dynam- content as revealed by T1- and T2-weighted
ics of large-scale brain functional networks MRI. J Neurosci 31:11597–11616
supporting recollection. Proc Natl Acad Sci 83. Eickhoff SB et al (2005) A new SPM toolbox
U S A 109:12788–12793 for combining probabilistic cytoarchitectonic
70. Beckmann CF, DeLuca M, Devlin JT, Smith maps and functional imaging data.
SM (2005) Investigations into resting-state NeuroImage 25:1325–1335
connectivity using independent component 84. Zilles K et al (2002) Architectonics of the
analysis. Philos Trans R Soc Lond B Biol Sci human cerebral cortex and transmitter recep-
360:1001–1013 tor fingerprints: reconciling functional neuro-
71. Calhoun VD, Adali T, Pekar JJ (2004) A anatomy and neurochemistry. Eur
method for comparing group fMRI data Neuropsychopharmacol 12:587–599
using independent component analysis: appli- 85. Alexander-Bloch A, Giedd JN, Bullmore E
cation to visual, motor and visuomotor tasks. (2013) Imaging structural co-variance
Magn Reson Imaging 22:1181–1191 between human brain regions. Nat Rev
72. Smith SM et al (2009) Correspondence of the Neurosci 14:322–336
brain's functional architecture during activa- 86. Lerch JP et al (2006) Mapping anatomical
tion and rest. Proc Natl Acad Sci U S A correlations across cerebral cortex
106:13040–13045 (MACACC) using cortical thickness from
73. Yu Q et al (2011) Altered topological proper- MRI. NeuroImage 31:993–1003
ties of functional network connectivity in 87. Bastiani M, Shah NJ, Goebel R, Roebroeck A
schizophrenia during resting state: a small- (2012) Human cortical connectome recon-
world brain network study. PLoS One 6:e25423 struction from diffusion weighted MRI: the
74. Kiviniemi V et al (2009) Functional segmen- effect of tractography algorithm. Neuroimage
tation of the brain cortex using high model 62:1732–1749
order group PICA. Hum Brain Mapp 88. Jones DK, Knösche TR, Turner R (2013)
30:3865–3886 White matter integrity, fiber count, and other
75. Nelson SM et al (2010) A parcellation scheme fallacies: the do‘s and dont’s of diffusion
for human left lateral parietal cortex. Neuron MRI. Neuroimage 73:239–254
67:156–170 89. van den Heuvel MP, Mandl RCW, Stam CJ,
76. Cohen AL et al (2008) Defining functional Kahn RS, Hulshoff Pol HE (2010) Aberrant
areas in individual human brains using resting frontal and temporal complex network struc-
functional connectivity MRI. NeuroImage ture in schizophrenia: a graph theoretical
41:45–57 analysis. J Neurosci 30:15915–15926
77. Yeo BT et al (2011) The organization of the 90. Alexander DC et al (2010) Orientationally
human cerebral cortex estimated by intrinsic invariant indices of axon diameter and density
functional connectivity. J Neurophysiol from diffusion MRI. Neuroimage
106:1125–1165 52:1374–1389
78. Craddock RC, James GA, Holtzheimer PE, 91. Friston KJ (1994) Functional and effective
Hu XP, Mayberg HS (2012) A whole brain connectivity in neuroimaging: a synthesis.
fMRI atlas generated via spatially constrained Hum Brain Mapping 2:56–78
312 Alex Fornito
92. Vincent JL et al (2007) Intrinsic functional 109. Fornito A, Zalesky A, Pantelis C, Bullmore
architecture in the anaesthetized monkey ET (2012) Schizophrenia, neuroimaging and
brain. Nature 447:83–86 connectomics. Neuroimage 62:2296–2314
93. Honey CJ et al (2009) Predicting human 110. van Wijk BCM, Stam CJ, Daffertshofer A
resting-state functional connectivity from (2010) Comparing brain networks of differ-
structural connectivity. Proc Natl Acad Sci ent size and connectivity density using graph
U S A 106:2035–2040 theory. Plos One 5:e13701
94. Zalesky A, Fornito A, Bullmore E (2012) On 111. Irimia A, Chambers MC, Torgerson CM,
the use of correlation as a measure of network Van Horn JD (2012) Circular representation
connectivity. Neuroimage 60:2096–2106 of human cortical networks for subject and
95. Fox MD, Raichle ME (2007) Spontaneous population- level connectomic visualization.
fluctuations in brain activity observed with Neuroimage 60:1340–1351
functional magnetic resonance imaging. Nat 112. Fornito A, Bullmore ET (2015) Connectomics:
Rev Neurosci 8:700–711 a new paradigm for understanding brain
96. Fornito A, Bullmore ET (2010) What disease. Eur Neuropsychopharmacol. 25:
can spontaneous fluctuations of the blood 733–748
oxygenation- level-dependent signal tell 113. Fornito A et al (2013) Functional dyscon-
us about psychiatric disorders? Curr Opin nectivity of corticostriatal circuitry as a risk
Psychiatry 23:239–249 phenotype for psychosis. JAMA Psychiatry
97. Zalesky A, Fornito A, Cocchi L, Gollo LL, 70:1143–1151
Breakspear M (2014) Time-resolved resting- 114. Meskaldji DE et al (2011) Adaptive strategy
state brain networks. Proc Natl Acad Sci U S A for the statistical analysis of connectomes.
111:10341–10346 Plos One 6:e23009
98. Smith SM et al (2012) Temporally- 115. Ginestet CE, Simmons A (2011) Statistical
independent functional modes of spontane- parametric network analysis of functional
ous brain activity. Proc Natl Acad Sci U S A connectivity dynamics during a working
109:3131–3136 memory task. Neuroimage 55:688–704
99. Hutchison RM et al (2013) Dynamic func- 116. Zalesky A, Fornito A, Bullmore ET (2010)
tional connectivity: promise, issues, and inter- Network-based statistic: Identifying differences
pretations. NeuroImage 80:360–378 in brain networks. Neuroimage 53:1197–1207
100. Feinberg DA et al (2010) Multiplexed echo pla- 117. Nichols TE, Holmes AP (2002)
nar imaging for sub-second whole brain FMRI Nonparametric permutation tests for func-
and fast diffusion imaging. PLoS One 5:e15710 tional neuroimaging: a primer with examples.
101. Rissman J, Gazzaley A, D'Esposito M (2004) Hum Brain Mapp 15:1–25
Measuring functional connectivity during dis- 118. Benjamini Y, Hochberg Y (1995) Controlling
tinct stages of a cognitive task. NeuroImage the false discovery rate: a practical and power-
23:752–763 ful approach to multiple testing. J R Stat Soc
102. Fornito A, Yoon J, Zalesky A, Bullmore ET, Ser B 57:289–300
Carter CS (2011) General and specific func- 119. Zalesky A, Cocchi L, Fornito A, Murray MM,
tional connectivity disturbances in first-episode Bullmore E (2012) Connectivity differences in
schizophrenia during cognitive control perfor- brain networks. Neuroimage 60:1055–1062
mance. Biol Psychiatry 70:64–72 120. Tomasi D, Volkow VD (2010) Functional
103. Friston KJ et al (1997) Psychophysiological connectivity density mapping. Proc Natl Acad
and modulatory interactions in neuroimag- Sci U S A, 107:9885–9890
ing. NeuroImage 6:218–229 121. Cole MW, Anticevic A, Repovs G, Barch D
104. Cole MW et al (2013) Multi-task connectivity (2011) Variable global dysconnectivity and
reveals flexible hubs for adaptive task control. individual differences in schizophrenia. Biol
Nat Neurosci 16:1348–1355 Psychiatry 70:43–50
105. Friston K, Moran R, Seth AK (2013) 122. Rubinov M, Sporns O (2011) Weight-conserving
Analysing connectivity with Granger causal- characterization of complex functional brain
ity and dynamic causal modelling. Curr Opin networks. Neuroimage 56:2068–2079
Neurobiol 23:172–178 123. Rubinov M, Sporns O (2010) Complex net-
106. Friston KJ, Harrison L, Penny W (2003) Dynamic work measures of brain connectivity: uses and
causal modelling. NeuroImage 19:1273–1302 interpretations. Neuroimage 52:1059–1069
107. Seghier ML, Friston KJ (2013) Network 124. Amaral LA, Scala A, Barthelemy M, Stanley
discovery with large DCMs. Neuroimage HE (2000) Classes of small-world networks.
68:181–191 Proc Natl Acad Sci U S A 97:11149–11152
108. Friston KJ, Kahan J, Biswal B, Razi A (2014) 125. van den Heuvel MP, Sporns O (2011) Rich-
A DCM for resting state fMRI. Neuroimage club organization of the human connectome.
94:396–407 J Neurosci 31:15775–15786
Graph Theoretic Analysis of Human Brain Networks 313
126. Buckner RL et al (2009) Cortical hubs revealed 143. Lancichinetti A, Fortunato S (2009)
by intrinsic functional connectivity: map- Community detection algorithms: a compara-
ping, assessment of stability, and relation to tive analysis. Phys Rev E 80:056117
Alzheimer’s Disease. J Neurosci 29:1860–1873 144. Good BH, de Montjoye YA, Clauset A (2010)
127. Power JD, Schlaggar BL, Lessov-Schlaggar Performance of modularity maximization in
CN, Petersen SE (2013) Evidence for hubs practical contexts. Phys Rev E 81:046106
in human functional brain networks. Neuron 145. Lancichinetti A, Fortunato S (2012) Consensus
79:798–813 clustering in complex networks. Sci Rep 2:336
128. van den Heuvel MP, Kahn RS, Goni J, Sporns 146. Guimerà R, Sales-Pardo M, Amaral L (2004)
O (2012) High-cost, high-capacity backbone Modularity from fluctuations in random
for global brain communication. Proc Natl graphs and complex networks. Phys Rev E
Acad Sci U S A 109:11372–11377 70:025101
129. Mišić B, Sporns O, McIntosh AR (2014) 147. Hagmann P et al (2008) Mapping the struc-
Communication efficiency and congestion tural core of human cerebral cortex. PLoS
of signal traffic in large-scale brain networks. Biol 6:e159
PLoS Comput Biol 10:e1003427 148. Meunier D, Achard S, Morcom A, Bullmore
130. van den Heuvel MP, Sporns O (2013) An ana- E (2009) Age-related changes in modular
tomical substrate for integration among func- organization of human brain functional net-
tional networks in human cortex. J Neurosci works. NeuroImage 44:715–723
33:14489–14500 149. Buzsaki G, Geisler C, Henze DA, Wang
131. Albert R, Jeong H, Barabasi AL (2000) Error XJ (2004) Interneuron diversity series: cir-
and attack tolerance of complex networks. cuit complexity and axon wiring economy
Nature 406:378–382 of cortical interneurons. Trends Neurosci
132. Fornito A, Breakspear M, Zalesky A (2015) 27:186–193
The connectomics of brain disorders. Nat Rev 150. Ercsey-Ravasz M et al (2013) A predictive
Neurosci 16:159–172 network model of cerebral cortical con-
133. Crossley NA et al (2014) The hubs of the nectivity based on a distance rule. Neuron
human connectome are generally implicated 80:184–197
in the anatomy of brain disorders. Brain 151. Crossley NA, Mechelli A, Vertes PE (2013)
137:2382–2395 Cognitive relevance of the community struc-
134. Maslov S, Sneppen K (2002) Specificity and ture of the human brain functional coacti-
stability in topology of protein networks. vation network. Proc Natl Acad Sci U S A
Science 296:910–913 110:11583–11588
135. Humphries MD, Gurney K, Prescott TJ 152. Fodor JA (1983) Modularity of mind: an
(2006) The brainstem reticular formation is essay on faculty psychology. MIT Press
a small-world, not scale-free, network. Proc 153. Simon HA (1962) The architecture of com-
Biol Sci 273:503–511 plexity. Proc Am Philos Soc 106:467–482
136. Latora V, Marchiori M (2001) Efficient 154. Kitano H (2004) Biological robustness. Nat
behavior of small-world networks. Phys Rev Rev Genet 5:826–837
Lett 87:198701 155. Guimera R, Nunes Amaral LA (2005)
137. Latora V, Marchiori M (2003) Economic Functional cartography of complex metabolic
small-world behavior in weighted networks. networks. Nature 433:895–900
Eur Phys J B 32:249–263 156. Xie J, Kelley S, Szymanski BK (2013)
138. Kaiser M, Hilgetag CC (2006) Nonoptimal Overlapping community detection in net-
component placement, but short processing works. ACM Comput Surv 45:1–35
paths, due to long-distance projections in 157. Meunier D, Lambiotte R, Bullmore ET (2011)
neural systems. PLoS Comput Biol 2:e95 Modular and hierarchically modular organiza-
139. Chen Y, Wang S, Hilgetag CC, Zhou C tion of brain networks. Front Neurosci 4:200
(2013) Trade-off between multiple con- 158. Meunier D, Lambiotte R, Fornito A, Ersche
straints enables simultaneous formation of KD, Bullmore ET (2009) Hierarchical mod-
modules and hubs in neural systems. PLoS ularity in human brain functional networks.
Comput Biol 9:e1002937 Front Neuroinform 3:37
140. Fornito A et al (2011) Genetic influences on 159. Newman MEJ (2010) Networks. a Introduction.
cost-efficient organization of human cortical Oxford University Press
functional networks. J Neurosci 31:3261–3270 160. Goñi J et al (2014) Resting-brain functional
141. Fortunato S (2010) Community detection in connectivity predicted by analytic measures of
graphs. Phys Rep 486:75–174 network communication. Proc Natl Acad Sci
142. Newman M, Girvan M (2004) Finding and U S A 111:833–838
evaluating community structure in networks. 161. Betzel RF et al (2014) Multi-scale com-
Phys Rev 69:026113 munity organization of the human struc-
314 Alex Fornito
tural connectome and its relationship with 167. de Haan W, Mott K, van Straaten ECW,
resting-state functional connectivity. Net Sci Scheltens P, Stam CJ (2012) Activity depen-
1:353–373 dent degeneration explains hub vulnerability
162. Bassett DS et al (2011) Dynamic reconfigura- in Alzheimer's disease. PLoS Comput Biol
tion of human brain networks during learning. 8:e1002582
Proc Natl Acad Sci U S A 108:7641–7646 168. Vertes PE et al (2012) Simple models of
163. Deco G, Jirsa VK, McIntosh AR (2011) human brain functional networks. Proc Natl
Emerging concepts for the dynamical orga- Acad Sci U S A 109:5868–5873
nization of resting-state activity in the brain. 169. Song HF, Kennedy H, Wang X-J (2014)
Nat Publ Group 12:43–56 Spatial embedding of structural similarity in
164. Alstott J, Breakspear M, Hagmann P, the cerebral cortex. Proc Natl Acad Sci U S A
Cammoun L, Sporns O (2009) Modeling the 111:16580–16585
impact of lesions in the human brain. PLoS 170. Goni J et al (2013) Exploring the morpho-
Comput Biol 5:e1000408 space of communication efficiency in complex
165. Honey CJ, Sporns O (2008) Dynamical con- networks. PLoS One 8, e58070
sequences of lesions in cortical networks. 171. Avena-Koenigsberger A et al (2014) Using
Hum Brain Mapp 29:802–809 Pareto optimality to explore the topol-
166. Raj A, Kuceyeski A, Weiner M (2012) A net- ogy and dynamics of the human connec-
work diffusion model of disease progression tome. Philos Trans R Soc Lond B Biol Sci
in dementia. Neuron 73:1204–1215 369:20130530
Part II
Abstract
Neuroimaging, in many respects, revolutionized the study of cognitive neuroscience, the discipline that
attempts to determine the neural mechanisms underlying cognitive processes. Early studies of brain–
behavior relationships relied on a precise neurological exam as the basis for hypothesizing the site of brain
damage that was responsible for a given behavioral syndrome. The advent of structural brain imaging, first
with computerized tomography and later with magnetic resonance imaging, paved the way for more pre-
cise anatomical localization of the cognitive deficits that manifest after brain injury. Functional neuroimag-
ing, broadly defined as techniques that provide measures of brain activity, further increased our ability to
study the neural basis of behavior. Functional MRI (fMRI), in particular, is an extremely powerful tech-
nique that affords excellent spatial and temporal resolution. This chapter focuses on the principles underly-
ing fMRI as a cognitive neuroscience tool for exploring brain–behavior relationships.
1 Introduction
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_11, © Springer Science+Business Media New York 2016
317
318 Mark D’Esposito et al.
Insight regarding the link between brain and behavior can be gained
through a variety of approaches. It is unlikely that any single neuro-
science method is sufficient to fully investigate any particular ques-
tion regarding the mechanisms underlying cognitive function.
From a methodological point of view, each method will offer differ-
ent temporal and spatial resolution. From a conceptual point of
view, each method will provide data that will support different types
of inferences that can be drawn from it. Thus, data obtained address-
ing a single question but derived from multiple methods can pro-
vide more comprehensive and inferentially sound conclusions.
Functional neuroimaging studies support inferences about the
association of a particular brain system with a cognitive process.
However, it is difficult to prove in such a study that the observed
activity is necessary for an isolated cognitive process because perfect
Functional MRI: Applications in Cognitive Neuroscience 319
due to the small bore of the MRI scanner and find it difficult to
remain completely motionless for a long duration of time that is
required for most experiments (e.g., usually 60–90 min). These
constraints of the MRI scanner make it especially difficult to scan
children or certain patient populations (e.g., Parkinson’s disease
patients), which has resulted in many fewer fMRI studies involving
children than adults and neurological patients in general. However,
mock scanners have been built in many imaging centers, with
motion devices that acclimate children to the scanner environment
before they participate in an fMRI study. This approach has led to
an increasing number of fMRI studies of children being reported in
the literature that are providing tremendous insight regarding the
mechanisms underlying the developing brain (for review, see [23]).
All sensory systems have been investigated with fMRI includ-
ing the visual, auditory, somatosensory, olfactory, and gustatory
systems. Each system requires different technologies for successful
presentation of relevant stimuli within an MRI environment. At
the time of this writing, there are now many off-the-shelf commer-
cial products that exist that are MRI-compatible. Acquiring ancil-
lary electrophysiological data such as electromyographic recordings
to measure muscle contraction or electrodermal responses to mea-
sure autonomic activity enhances many cognitive neuroscience
experiments. Devices have been developed that are MR compatible
for these types of measurements as well as other physiological mea-
sures such as heart rate, electrocardiography, oxygen saturation,
and respiratory rate. The recording of eye movements is common-
place in MRI scanners predominantly with the use of infrared video
cameras equipped with long range optics. Video images of the
pupil–corneal reflection can be sampled at 500–1000 Hz allowing
for the accurate (<0.5°) localization of gaze within 50 horizontal
and 40 vertical degrees of visual angle.
EEG recordings have also been successfully performed during
MRI scanning [24, 25]. Both measures of event-related potentials
(ERPs) and spectral EEG power in specific frequency bands and
have been successfully recorded and related to variations in under-
lying BOLD activity and behavior [26–29]. However, the record-
ing of low amplitude EEG events, such as ERPs and transient
changes in spectral EEG power, can be more difficult in a magnetic
field due to large artifacts induced by gradient switching and head
movement and voltage changes from cardiac pulsation. The
optimization of data acquisition methods and post-processing
algorithms to remove artifacts have allowed for reliable measure-
ments of ERPs and transient EEG events during fMRI scanning
[30–33]. In summary, most challenges facing cognitive experi-
ments and the study of spontaneous activity within the MRI envi-
ronment have been overcome, creating an environment that is
comparable to standard psychophysical testing labs outside of a
scanner. Recent work has focused on minimizing exacerbated EEG
Functional MRI: Applications in Cognitive Neuroscience 323
3.1 Temporal Two types of temporal resolution need to be considered for cogni-
Resolution tive neuroscience experiments. First, what is the briefest neural
event that can be detected as an fMRI signal? Second, how close
together can two neural events occur and be resolved as separable
fMRI signals?
The time scale on which neural changes occur is quite rapid. For
example, neural activity in the lateral intraparietal area of monkeys
increases within 100 ms of the visual presentation of a saccade target
[39]. In contrast, the fMRI signal gradually increases to its peak
magnitude within 4–6 s after an experimentally induced brief (<1 s)
change in neural activity, and then decays back to baseline after sev-
eral more seconds [40–42]. This slow time course of fMRI signal
change in response to such a brief increase in neural activity is infor-
mally referred to as the blood oxygen level-dependent (BOLD)
fMRI hemodynamic response or simply, the hemodynamic response
(Fig. 1). Thus, neural dynamics and neurally evoked hemodynamics,
as measured with fMRI, are on quite different time scales.
The sluggishness of the hemodynamic response limits the tem-
poral resolution of the fMRI signal to hundreds of milliseconds to
seconds as opposed to the millisecond temporal resolution of elec-
trophysiological recordings of neural activity, such as from single-
unit recording in monkeys and EEG or MEG in humans. However,
it has been clearly demonstrated that brief changes in neural activ-
ity can be detected with reasonable statistical power using
fMRI. For example, appreciable fMRI signal can be observed in
sensorimotor cortex in association with single finger movements
[43] and in visual cortex during very briefly presented (34 ms)
visual stimuli [44]. In contrast, the temporal resolution of fMRI
324 Mark D’Esposito et al.
0.5
0.25
-0.25
0 2 4 6 8 10 12 14
time (secs)
Fig. 2 Effect of fixed vs. randomized intertrial intervals on the blood oxygen level-
dependent (BOLD) fMRI signal [46]
3.2 Spatial As approaches are sought that maximize both BOLD signal
Resolution strength and in-plane resolution, fMRI studies in humans have
recently been extended to higher magnetic field strengths (7.0 T
and 9.4 T) [51–53]. Such studies have the power to potentially
evaluate much finer cortical details, such as the representation of
individual fingertips within primary somatosensory cortex [51].
However, as the field strength increases, factors that are less conse-
quential at 3.0 T—including magnetic field inhomogeneities [52]
and the contribution of macrovascular structures to the typical
gradient-echo signal [53]—become significantly more problematic,
requiring further innovations in pulse sequence development.
Single-echo gradient-echo sequences using echo times (TE) that
exceed the repetition time (TR), for example, take advantage of
reduced distortion relative to single-shot gradient-echo sequences,
while also avoiding the prolonged acquisition times of typical
single-echo sequences. During functional MRI of a simple finger
tapping task at 9.4 T using such a sequence, researchers were able
to obtain 0.4 × 0.4 mm in-plane resolution within presumptive pri-
mary motor cortex [53]. Similarly, spin-echo sequences, which
have a reduced signal-to-noise ratio relative to gradient-echo
sequences but greater spatial specificity, become feasible for use at
9.4 T. Within a finger-tapping paradigm, a study taking this
approach reduced the influence of macrovascular contributions to
326 Mark D’Esposito et al.
Fig. 3 Data derived from the performance of a normal subject on a spatial delayed-response task [64]. This
task comprised both delay trials (circles) as well as trials without a delay period (no-delay trials; diamonds). (a)
Trial averaged fMRI signal from prefrontal cortex that displayed delay-correlated activity. The gray bar along
the x-axis denotes the 12 s delay period during delay trials. The delay trials display a level of fMRI signal
greater than baseline throughout the period of time corresponding to the retention delay (taking into account
the delay and dispersion of the fMRI signal). The peaks seen in the signal correspond to the encoding and
retrieval periods. (b) Trial averaged fMRI signal from a region in prefrontal cortex that did not display the char-
acteristics of delay-correlated activity. This region displays a significant functional change associated with the
no-delay trials, and a significant functional change associated with the encoding and retrieval periods of the
delay trials, but not one associated with the retention delay of delay trials
Functional MRI: Applications in Cognitive Neuroscience 329
described above. The left side of the figure illustrates BOLD signal
consistent with delay period activity whereas the right side of the
figure illustrates BOLD signal from another region of PFC that did
not display sustained activity during the delay yet showed greater
activity in the delay trials as compared to the trials without a delay.
In any blocked functional neuroimaging study that compares delay
vs. no-delay trials with subtraction, such a region would be detected
and likely assumed to be a “memory” region. Thus, this result
provides empirical grounds for adopting a healthy doubt regarding
the inferences drawn from imaging studies that rely exclusively on
cognitive subtraction.
In functional neuroimaging, the transform between the neural
signal and the hemodynamic response (measured by fMRI) must
also be linear for the cognitive subtractive method to yield valid
results. In other words, it is assumed that the BOLD signal being
measured is approximately proportional to the local neural activity
that evokes it. Surprisingly, although thousands of empirical stud-
ies using fMRI to study brain–behavior relationships have been
published, only a handful exist that have explored the neurophysi-
ological basis of the BOLD signal (for reviews see refs. [64, 65]).
In several studies it has been demonstrated that linearity does not
strictly hold for the BOLD fMRI system but a linear transform
model is reasonably consistent with the data. For example, Boynton
et al. tested whether the BOLD signal in response to long duration
stimuli can be predicted by summing the responses to shorter
duration stimuli [42]. Using pulses of flickering checkerboard pat-
terns and measuring within human primary visual cortex, these
investigators found that the BOLD signal response to various
durations of stimulus presentation (6, 12, or 24 s) could be pre-
dicted from the responses they obtained from shorter stimulus pre-
sentations. For example, the BOLD signal response to a 6 s pulse
could be predicted from the summation of the BOLD signal
response to the 3 s pulse with a copy of the same response delayed
by 3 s. However, temporal summation did not always hold, and
there are clearly nonlinear effects in the transform of neural activity
to a hemodynamic response that must be considered [66–69]. If
these nonlinearities lead to saturation of the BOLD effect at a cer-
tain stimulus intensity, erroneous interpretation of particular results
of fMRI experiments may occur.
Another class of experimental designs, called event-related
fMRI, attempts to detect changes associated with individual trials,
as opposed to the larger unit of time comprising a block of trials
[70, 71]. Each individual trial may be composed of one behavioral
“event,” such as the presentation of a single stimulus (e.g., a face
or object to be perceived) or several behavioral events such as in
the delayed-response task described above (e.g., an item to be
remembered, a delay period, and a motor response in a delayed-
response task). For example, with an event-related design, activity
330 Mark D’Esposito et al.
within the PFC has consistently been shown to correlate with the
delay period, supporting the role of the PFC in temporarily main-
taining information [63]. This finding is consistent with single-
neuron recording studies in the PFC of monkeys [7]. An
event-related design offers numerous advantages. For example, it
allows for stimulus or trial randomization avoiding the behavioral
confounds of blocked trials. It also permits the separate analysis of
functional responses that are identified only in retrospect (i.e., tri-
als on which the subject made a correct or incorrect response). Of
course, an experiment does not have to be limited to either a block
or event-related designs—a mixed-type (both event-related and
blocked) design where particular trial types are randomized within
a block is perfectly feasible. In this type of design, both item-related
processes (e.g., transient responses to stimuli) as well as state-
related processes (processes sustained throughout a block of trials
or a task) are perfectly feasible [72, 73].
Overall, much flexibility exists in the type of experimental
design that can be utilized in fMRI experiments and continued
innovation in this area will greatly expand the types of neuroscien-
tific questions that can be addressed.
5.1 Statistics Many statistical techniques are used for analyzing fMRI data, but
no single method has emerged as the ideal or “gold standard.” The
analysis of any fMRI experiment designed to contradict the null
hypothesis (i.e., there is no difference between experimental con-
ditions) requires inferential statistics. If the difference between two
experimental conditions is too large to be reasonably due to chance,
then the null hypothesis is rejected in favor of the alternative
hypothesis, which typically is the experimenter’s hypothesis (e.g.,
the fusiform gyrus is activated to a greater extent by viewing faces
than objects). Unfortunately, since errors can occur in any statisti-
cal test, experimenters will never know when an error is committed
and can only try to minimize them [74]. Knowledge of several
basic statistical issues provides a solid foundation for the correct
interpretation of the data derived from fMRI studies.
Two types of statistical errors can occur. A type I error is commit-
ted when the null hypothesis is falsely rejected, that is, a difference
between experimental conditions is found but a difference does not
truly exist. This type of error is also called a false-positive error. In an
fMRI study, a false-positive error would be finding a brain region
activated during a cognitive task, when actually it is not. A type II
error is committed when the null hypothesis is accepted when it is
false, that is, no difference between experimental conditions exists
when a difference does exist. This type of error is also called a false-
negative error. A false-negative error in an fMRI study would be
Functional MRI: Applications in Cognitive Neuroscience 331
5.2 Altered When comparing changes in fMRI BOLD signal levels within the
Hemodynamic brain of an individual subject across different cognitive tasks and
Response making conclusions regarding changes in neural activity and the
pattern of activity, numerous assumptions are made regarding the
steps comprising neurovascular coupling (stimulus → neural activ-
ity → hemodynamic response → BOLD signal) and the regional
variability of the metabolic and vascular parameters influencing the
BOLD signal. It should be obvious that fMRI studies of cognition
of individuals with local vascular compromise or diffuse vascular
disease (e.g., patients with strokes or normal elderly) are poten-
tially problematic. For example, many fMRI studies have sought to
identify age-related changes in the neural substrates of cognitive
processes. Those studies that directly compare changes in fMRI
BOLD signal intensity across age groups rely upon the assumption
of age-equivalent coupling of neural activity to BOLD signal.
However, there is empirical evidence that suggests that this general
assumption may not hold true. Extensive research on the aging
neurovascular system has revealed that it undergoes significant
changes in multiple domains in a continuum throughout the
human lifespan, probably as early as the fourth decade (for review
see ref. [85]). These changes affect the vascular ultrastructure [86],
the resting cerebral blood flow [87, 88], the vascular responsive-
ness of the vessels [89], and the cerebral metabolic rate of oxygen
consumption [90, 91]. Aging is also frequently associated with
comorbidities such as diabetes, hypertension, and hyperlipidemia,
all of which may affect the fMRI BOLD signal by affecting cerebral
blood flow and neurovascular coupling [92]. Any one of these age-
related differences in the vascular system could conceivably pro-
duce age-related differences in BOLD fMRI signal responsiveness,
greatly affecting the interpretation of results from such studies.
Our laboratory compared the hemodynamic response function
(HRF) characteristics in the sensorimotor cortex of young and
older subjects in response to a simple motor reaction-time task
[70]. The provisional assumption was made that there was identi-
cal neural activity between the two populations based on physio-
logical findings of equivalent movement-related electrical potentials
in subjects under similar conditions [93]. Thus, we presumed that
any changes that we observed in BOLD fMRI signal between
young and older individuals in motor cortex would be due to vas-
cular, and not neural activity changes in normal aging. Several
important similarities and differences were observed between age
groups. Although there was no significant difference in the shape
of the hemodynamic response curve or peak amplitude of the sig-
nal, we found a significantly decreased SNR in the fMRI BOLD
signal in older individuals as compared to young individuals. This
was attributed to a greater level of noise in the older individuals.
We also observed a decrease in the spatial extent of the BOLD
signal in older individuals compared to younger individuals in
334 Mark D’Esposito et al.
6.1 Functional The major focus of fMRI studies of cognition is testing theories on
Specialization functional specialization. The concept of functional specialization
is based on the premise that functional modules exist within the
brain, that is, areas of the cerebral cortex are specialized for a spe-
cific cognitive process. For example, facial recognition is a critical
primary function likely served by a functional module.
Prosopagnosia is the selective inability to recognize faces. Patients
with prosopagnosia, however, can recognize familiar faces, such as
those of relatives, by other means, such as the voice, dress, or
shape. Other types of visual recognition, such as identifying com-
mon objects, are normal. Prosopagnosia arises from lesions of the
inferomedial temporo-occipital lobe, which are usually due to a
stroke within the posterior cerebral artery circulation. No lesion
studies have precisely localized the area crucial for facial percep-
tion. However, they provide strong evidence that a brain area is
specialized for processing faces. Functional imaging studies have
336 Mark D’Esposito et al.
Fig. 4 Network analysis of fMRI data using structural equation modeling during performance of a working
memory task across three different delay periods [111]. Areas of correlated increases in activation (solid lines)
and areas of correlated decreases in activation (dotted lines) are shown. Note the different pattern of interac-
tions among brain regions at short and long delays
6.3 Cognitive Theory Experiments using fMRI can also test theories of the underlying
mechanisms of cognition. For example, an fMRI study [124]
attempted to answer the question, “To what extent does perception
depend on attention?” One hypothesis is that unattended stimuli in
the environment receive very little processing [125], but another
hypothesis is that the processing load in a relevant task determines
the extent to which irrelevant stimuli are processed [126]. These
alternative hypotheses were tested by asking normal individuals to
perform linguistic tasks of low or high load while ignoring irrele-
vant visual motion in the periphery of a display. Visual motion was
used as the distracting stimulus, because it activates a distinct region
of the brain (cortical area MT or V5, another functional module in
the visual system). Activation of area MT would indicate that irrel-
evant visual motion was processed. Although task and irrelevant
stimuli were unrelated, fMRI of motion-related activity in MT
showed a reduction in motion processing during the high-process-
ing load condition in the linguistic task. These findings supported
the hypothesis that perception of irrelevant environmental
340 Mark D’Esposito et al.
Fig. 5 A brain graph derived from resting state fMRI data collected from healthy young subjects illustrating
identified modules, represented as different shades of color. There are four distinct modules identified in this
graph
7.1 Combined fMRI/ The combined use of functional neuroimaging and lesions studies
Lesion Studies can be illustrated with studies of the neural basis of semantic mem-
ory, the cognitive system that represents our knowledge of the
Functional MRI: Applications in Cognitive Neuroscience 341
Fig. 6 Regions of overlap of fMRI activity in healthy human subjects (left side of
figure) during the performance of three semantic memory tasks, with the con-
vergence of activity within the left inferior frontal gyrus (white region) [125].
Regions of overlap of lesion location in patients with selection-related deficits on
a verb generation task (right side of figure) with maximal overlap within the left
inferior frontal gyrus [126]
7.3 Combined fMRI/ The strength of combining these two methods is coupling the
Event-Related superb spatial resolution of fMRI with the superb temporal resolu-
Potential Studies tion of ERP recording. An example of such a study was reported
by Dehaene et al. who asked the question “Does the human capac-
ity for mathematical intuition depend on linguistic competence or
on visuospatial representations?” [135]. In this study, subjects per-
formed two addition tasks—one in which they were instructed to
select the correct sum from two numerically close numbers (exact
condition) and one in which they were instructed to estimate the
result and select the closest number (approximate condition).
During fMRI scanning greater bilateral parietal lobe activation was
observed in the approximation condition as compared to the exact
condition. Since this activation was outside the perisylvian lan-
guage zone, it was taken as support that visuospatial processes were
engaged during the cognitive operations involved in approximate
calculation. Greater left lateralized frontal lobe activation was
observed to be greater in the exact condition as compared to the
approximate condition, which was taken as evidence for language
dependent coding of exact addition facts. In order to consider an
alternative explanation of the fMRI findings, the investigators also
performed an ERP study. The alternative explanation was that in
both the exact and approximate tasks, subjects would compute the
exact result using the same representation for numbers but later
processing, when they had to make a decision as to the correct
choice, was what led to the differences in brain activation. Since
fMRI does not offer adequate temporal resolution to resolve these
two behavioral events on such a brief time scale, ERP was the
appropriate method to test this hypothesis. In the ERP study it was
demonstrated that the evoked neural response during exact and
approximate trials already differed significantly during the first
400 ms of a trial before subjects had to make a decision.
8.1 Use Cognitive neuroscience studies using fMRI may provide an impor-
of Biomarkers Derived tant foundation for clinical studies. A biomarker is an indicator that
from Cognitive reflects a process, event, or condition in a biological system.
Neuroscience Studies Biomarkers may be useful for providing a measure of exposure,
effect, or susceptibility. Reliable biomarkers of a neural system could
reliably quantify how such a neural system is affected by almost any
input. The input may be the effects of a drug, the effects of cogni-
tive therapy, or the effects of a disease process. For a measurement
to be useful as a biomarker in clinical studies, it needs to have well-
defined significance based on preclinical studies. That is, a change in
an fMRI measurement would ideally reflect a change in a well-
understood process, thus providing a clear a priori hypothesis and
interpretation of the findings. Once the processes are established,
fMRI biomarkers may then be useful for addressing a number of
clinical questions. For any neurophysiologic measurement to be a
surrogate marker, a stable, reliable relationship between the fMRI
measurement and a defined clinical outcome needs to be defined.
Only in that scenario would an fMRI measurement provide a suit-
able surrogate for other clinical outcomes. Cognitive neuroscience
studies provide the foundation for fMRI biomarkers, but the stud-
ies necessary for defining fMRI surrogate markers are rarely done.
Questions regarding the mechanisms of brain function dis-
rupted by pathologic states, processes affected by treatment inter-
ventions, or the nature of post-injury reorganization of function
are examples of clinical questions that can be tested with fMRI. For
example, attentional modulation of information processing-related
activity in visual cortex is a well-established phenomenon in cogni-
tive neuroscience studies, with effects measurable using fMRI. For
example, it has been shown that activity in category-selective
regions of inferior temporal cortex is modulated based on the tar-
get of attention, relatively up-modulated if the target is relevant to
the region and down-modulated if not relevant [140, 141]. This
346 Mark D’Esposito et al.
8.2 Functional MRI Functional MRI may be useful not only in defining “static” brain–
for Measuring behavioral relationships, but also may be applied to defining the
the Effect of Clinical neural mechanisms that underlie learning, experience, or injury.
Interventions Two general categories of questions may be investigated. First,
fMRI can be used to examine factors that influence response to the
perturbations of training (learning), experience or injury. Second,
fMRI can be used to examine changes that underlie or are the
result of these various perturbations.
Investigation of baseline factors that may influence response to
training has particular clinical relevance. A better understanding of
pre-training neural characteristics that influence response to reha-
bilitation training could have major clinical value in guiding treat-
ment decisions. fMRI could provide a number of possible
measurements that could mark an important neural process. For
example, certain parameters of brain network organization may be
particularly important in supporting the potential for learning and
plasticity. For example, parameters of the functional organization
of whole brain networks have been shown to predict response to
training of attention regulation after injury [142]. In another
example, a simple measurement of the quantity of activation in
prefrontal cortex has been shown to predict response to training to
use a verbal memory strategy [143]. Such approaches may help
elucidate either personal factors or strategic approaches that under-
lie variations in learning or response to interventions.
Investigation of changes over time is particularly relevant for
understanding neural mechanisms of post-injury rehabilitation. In
order to assess changes with intervention, longitudinal or repeated
measurements are required. Because fMRI involves no
exposure-limiting factors such as radiation, it is suitable for repeated
measurements. However, multi-session studies are also signifi-
cantly more complicated to design, analyze, and interpret due to a
number of issues discussed below.
There are at least two distinct approaches relevant to assessing
changes within an individual. First, fMRI may be used for deter-
mining the after-effects of a learning intervention. Functional MRI
measures pre- and post-intervention may be used to address this
question. For example, after two pieces of information have been
strongly associated over repetitive exposures, one may find reduced
activation in response to presentation of that information, but
increased functional connectivity between regions of the brain that
process the two types of information [115]. Second, fMRI may be
used for determining the processes that occur during an interven-
tion, such as cognitive training. To do this one would need to
Functional MRI: Applications in Cognitive Neuroscience 347
9 Conclusions
References
1. Broca P (1861) Remarques sur le siege de la 7. Fuster JM, Alexander GE (1971) Neuron
faculte du langage articule suivies d’une activity related to short-term memory. Science
observation d’amphemie (perte de al parole). 173:652–654
Bull Mem Soc Anat Paris 36:330–357 8. Funahashi S, Bruce CJ, Goldman-Rakic PS
2. Buckner RL, Raichle ME, Petersen SE (1995) (1989) Mnemonic coding of visual space in
Dissociation of human prefrontal cortical the monkey’s dorsolateral prefrontal cortex.
areas across different speech production tasks J Neurophysiol 61:331–349
and gender groups. J Neurophysiol 9. Funahashi S, Bruce CJ, Goldman-Rakic PS
74(5):2163–2173 (1993) Dorsolateral prefrontal lesions and
3. Sarter M, Bernston G, Cacioppo J (1996) oculomotor delayed-response performance:
Brain imaging and cognitive neuroscience: evidence for mnemonic “scotomas”.
toward strong inference in attributing func- J Neurosci 13:1479–1497
tion to structure. Am Psychol 51:13–21 10. Watanabe T, Niki H (1985) Hippocampal
4. Gaffan D, Gaffan EA (1991) Amnesia in man unit activity and delayed response in the mon-
following transection of the fornix: a review. key. Brain Res 325(1–2):241–254
Brain 114:2611–2618 11. Cahusac PM, Miyashita Y, Rolls ET (1989)
5. Feeney DM, Baron JC (1986) Diaschisis. Responses of hippocampal formation neurons
Stroke 17(5):817–830 in the monkey related to delayed spatial
6. Carrera E, Tononi G (2014) Diaschisis: past, response and object-place memory tasks.
present, future. Brain 137:2408–2422 Behav Brain Res 33(3):229–240
Functional MRI: Applications in Cognitive Neuroscience 349
central executive and default-mode networks 51. Besle J, Sanchez-Panchuelo R, Bowtell R, Francis
in humans. Proc Natl Acad Sci U S A S, Schluppeck D (2014) Event-related fMRI at 7
110(49):19944–19949 T reveals overlapping cortical representations for
38. Yau JM, Hua J, Liao DA, Desmond JE (2013) adjacent fingertips in S1 of individual subjects.
Efficient and robust identification of cortical Hum Brain Mapp 35:2027–2043
targets in concurrent TMS-fMRI experi- 52. Ehses P, Bause J, Shajan G, Scheffler
ments. NeuroImage 76:134–144 K. Efficient generation of T2*-weighted con-
39. Gnadt JW, Andersen RA (1988) Memory trast by interslice echo-shifting for human
related motor planning activity in posterior functional and anatomacil imaging at 9.4
parietal cortex of macaque. Exp Brain Res Tesla. Magn Reson Med (epub ahead of print)
70:216–220 53. Budde J, Shajan G, Zaitsev M, Scheffler K,
40. Aguirre GK, Zarahn E, D’Esposito M (1998) Functional PR, MRI (2014) Human subjects
The variability of human, BOLD hemody- with gradient-echo and spin-echo EPI at 9.4
namic responses. Neuroimage 8(4):360–369 T. Magn Reson Med 2014(71):209–218
41. Handwerker DA, Ollinger JM, D’Esposito M 54. Malonek D, Grinvald A (1996) Interactions
(2004) Variation of BOLD hemodynamic between electrical activity and cortical micro-
responses across brain regions and subjects circulation revealed by imaging spectroscopy:
and their effects on statistical analyses. implications for functional brain mapping.
NeuroImage 21:1639–1651 Science 272:551–554
42. Boynton GM, Engel SA, Glover GH, Heeger 55. Kim SG, Duong TQ (2002) Mapping cortical
DJ (1996) Linear systems analysis of func- columnar structures using fMRI. Physiol
tional magnetic resonance imaging in human Behav 77(4–5):641–644
V1. J Neurosci 16:4207–4221 56. Grill-Spector K, Malach R (2001) fMR-
43. Kim SG, Richter W, Ugurbil K (1997) adaptation: a tool for studying the functional
Limitations of temporal resolution in properties of human cortical neurons. Acta
fMRI. Magn Reson Med 37:631–636 Psychol (Amst) 107(1–3):293–321
44. Savoy RL, Bandettini PA, O’Craven KM et al 57. Grill-Spector K, Kushnir T, Edelman S,
(1995) Pushing the temporal resolution of Avidan G, Itzchak Y, Malach R (1999)
fMRI: studies of very brief stimuli, onset of Differential processing of objects under vari-
variability and asynchrony, and stimulu- ous viewing conditions in the human lateral
correlated changes in noise. Proc Soc Magn occipital complex. Neuron 24(1):187–203
Reson Med 3:450 58. Posner MI, Petersen SE, Fox PT, Raichle ME
45. Zarahn E, Aguirre GK, D’Esposito M (1997) (1988) Localization of cognitive operations
A trial-based experimental design for func- in the human brain. Science 240:1627–1631
tional MRI. NeuroImage 6:122–138 59. Sternberg S (1969) The discovery of process-
46. Burock MA, Buckner RL, Woldorff MG, ing stages: extensions of Donders’ method.
Rosen BR, Dale AM (1998) Randomized Acta Psychol 30:276–315
event-related experimental designs allow for 60. Petersen SE, Fox PT, Posner MI, Mintun M,
extremely rapid presentation rates using func- Raichle ME (1988) Positron emission tomo-
tional MRI. Neuroreport 9(16):3735–3739 graphic studies of the cortical anatomy of sin-
47. Clark VP, Maisog JM, Haxby JV (1997) gle word processing. Nature 331:585–589
fMRI studies of visual perception and recog- 61. Fuster J (1997) The prefrontal cortex: anat-
nition using a random stimulus design. Soc omy, physiology, and neuropsychology of the
Neurosci Abstr 23:301 frontal lobes, 3rd edn. Raven, New York
48. Dale AM, Buckner RL (1997) Selective aver- 62. Jonides J, Smith EE, Koeppe RA, Awh E,
aging of rapidly presented individual trials Minoshima S, Mintun MA (1993) Spatial
using fMRI. Hum Brain Mapp 5:1–12 working memory in humans as revealed by
49. Miezin FM, Maccotta L, Ollinger JM, PET. Nature 363:623–625
Petersen SE, Buckner RL (2000) 63. Sreenivasan KK, Curtis CE, D’Esposito M
Characterizing the hemodynamic response: (2014) Revising the role of persistent neural
effects of presentation rate, sampling proce- activity in working memory. Trends Cogn Sci
dure, and the possibility of ordering brain 18:82–89
activity based on relative timing. Neuroimage 64. Attwell D, Iadecola C (2002) The neural
11(6 Pt 1):735–759 basis of functional brain imaging signals.
50. D’Esposito M, Zarahn E, Aguirre GK (1999) Trends Neurosci 25(12):621–625
Event-related functional MRI: implications 65. Heeger DJ, Ress D (2002) What does fMRI
for cognitive psychology. Psychol Bull tell us about neuronal activity? Nat Rev
125:155–164 Neurosci 3(2):142–151
Functional MRI: Applications in Cognitive Neuroscience 351
66. Friston KJ, Josephs O, Rees G, Turner R 81. Zarahn E, Slifstein M (2001) A reference
(1998) Nonlinear event-related responses in effect approach for power analysis in
fMRI. Magn Reson Med 39(1):41–52 fMRI. Neuroimage 14(3):768–779
67. Glover GH (1999) Deconvolution of impulse 82. Van Horn JD, Ellmore TM, Esposito G,
response in event-related BOLD Berman KF (1998) Mapping voxel-based sta-
fMRI. Neuroimage 9(4):416–429 tistical power on parametric images.
68. Miller KL, Luh WM, Liu TT et al (2001) Neuroimage 7(2):97–107
Nonlinear temporal dynamics of the cerebral 83. Aguirre GK, D’Esposito M (1999)
blood flow response. Hum Brain Mapp Experimental design for brain fMRI. In:
13(1):1–12 Moonen CTW, Bandettini PA (eds) Functional
69. Vazquez AL, Noll DC (1998) Nonlinear MRI. Springer, Berlin, pp 369–380
aspects of the BOLD response in functional 84. Logothetis NK, Pauls J, Augath M, Trinath
MRI. NeuroImage 7(2):108–118 T, Oeltermann A (2001) Neurophysiological
70. D’Esposito M, Zarahn E, Aguirre GK, Rypma investigation of the basis of the fMRI signal.
B (1999) The effect of normal aging on the Nature 412(6843):150–157
coupling of neural activity to the bold hemo- 85. Farkas E, Luiten PG (2001) Cerebral micro-
dynamic response. Neuroimage 10(1):6–14 vascular pathology in aging and Alzheimer’s
71. Rosen BR, Buckner RL, Dale AM (1998) disease. Prog Neurobiol 64(6):575–611
Event-related functional MRI: past, present, 86. Fang HCH (1976) Observations on aging
and future. Proc Natl Acad Sci U S A characteristics of cerebral blood vessels, mac-
95(3):773–780 roscopic and microscopic features. In: Gerson
72. Donaldson DI, Petersen SE, Ollinger JM, S, Terry RD (eds) Neurobiology of aging.
Buckner RL (2001) Dissociating state and Raven, New York
item components of recognition memory 87. Bentourkia M, Bol A, Ivanoiu A et al (2000)
using fMRI. Neuroimage 13(1):129–142 Comparison of regional cerebral blood flow
73. Mitchell KJ, Johnson MK, Raye CL, and glucose metabolism in the normal brain:
D’Esposito M (2000) fMRI evidence of age- effect of aging. J Neurol Sci 181(1–2):19–28
related hippocampal dysfunction in feature 88. Schultz SK, O’Leary DS, Boles Ponto LL,
binding in working memory. Brain Res Cogn Watkins GL, Hichwa RD, Andreasen NC
Brain Res 10(1–2):197–206 (1999) Age-related changes in regional cere-
74. Keppel G, Zedeck S (1989) Data analysis for bral blood flow among young to mid-life
research design. W.H. Freeman, New York adults. Neuroreport 10(12):2493–2496
75. Worsley KJ, Friston KJ (1995) Analysis of 89. Yamamoto M, Meyer JS, Sakai F, Yamaguchi
fMRI time-series revisited – again. F (1980) Aging and cerebral vasodilator
Neuroimage 2:173–182 responses to hypercarbia: responses in normal
76. Nichols T, Hayasaka S (2003) Controlling aging and in persons with risk factors for
the familywise error rate in functional neuro- stroke. Arch Neurol 37(8):489–496
imaging: a comparative review. Stat Methods 90. Yamaguchi T, Kanno I, Uemura K et al (1986)
Med Res 12(5):419–446 Reduction in regional cerebral rate of oxygen
77. Eklund A, Andersson M, Josephson C, during human aging. Stroke 17:1220–1228
Johannesson M, Knutsson H (2012) Does 91. Takada H, Nagata K, Hirata Y et al (1992)
parametric fMRI analysis with SPM yield valid Age-related decline of cerebral oxygen metab-
results? An empirical study of 1484 rest data- olism in normal population detected with
sets. Neuroimage 61(3):565–578 positron emission tomography. Neurol Res
78. Zarahn E, Aguirre GK, D’Esposito M (1997) 14(2 Suppl):128–131
Empirical analyses of BOLD fMRI statistics. I 92. Claus JJ, Breteler MM, Hasan D et al (1998)
Spatially unsmoothed data collected under Regional cerebral blood flow and cerebrovas-
null-hypothesis conditions. NeuroImage cular risk factors in the elderly population.
5:179–197 Neurobiol Aging 19(1):57–64
79. Aguirre GK, Zarahn E, D’Esposito M (1997) 93. Cunnington R, Iansek R, Bradshaw JL,
Empirical analyses of BOLD fMRI statistics. Phillips JG (1995) Movement-related poten-
II Spatially smoothed data collected under tials in Parkinson’s disease. Presence and pre-
null-hypothesis and experimental conditions. dictability of temporal and spatial cues. Brain
NeuroImage 5:199–212 118(Pt 4):935–950
80. D’Esposito M, Ballard D, Zarahn E, Aguirre 94. Buckner RL, Snyder AZ, Sanders AL, Raichle
GK (2000) The role of prefrontal cortex in ME, Morris JC (2000) Functional brain
sensory memory and motor preparation: an imaging of young, nondemented, and
event-related fMRI study. Neuroimage 11(5 demented older adults. J Cogn Neurosci
Pt 1):400–408 12(Suppl 2):24–34
352 Mark D’Esposito et al.
95. Huettel SA, Singerman JD, McCarthy G brain activity associated with memory storage
(2001) The effects of aging upon the hemo- and search. Neuroimage 33(2):794–804
dynamic response measured by functional 109. Kay KN, Naselaris T, Prenger RJ, Gallant JG
MRI. Neuroimage 13(1):161–175 (2008) Identifying natural images from
96. Pineiro R, Pendlebury S, Johansen-Berg H, human brain activity. Nature 452:352–355
Matthews PM (2002) Altered hemodynamic 110. Haxby JV, Connolly AC, Swaroop GJ (2014)
responses in patients after subcortical stroke Decoding neural representational spaces
measured by functional MRI. Stroke using multivariate pattern analysis. Annu Rev
33(1):103–109 Neurosci 37:435–456
97. D’Esposito M, Deouell L, Gazzaley A (2003) 111. Tong F, Pratte MS (2012) Decoding patterns
Alterations in the BOLD fMRI signal with of human brain activity. Annu Rev Psychol
ageing and disease: a challenge for neuroim- 63:483–509
aging. Nat Rev Neurosci 4:863–872 112. Todd MT, Nystrom LE, Cohen JE (2013)
98. Handwerker DA, Gazzaley A, Inglis BA, Confounds in multivariate pattern analysis:
D’Esposito M (2006) Reducing vascular vari- theory and rule representation case study.
ability of fMRI data across aging populations NeuroImage 77:157–165
using a breath holding task. Hum Brain Mapp 113. Chen AJW, Britton MS, Thompson TW,
28:846–859 Turner GR, Vytlacil J, D’Esposito M (2012)
99. Wolf RL, Detre JA (2007) Clinical neuroimag- Goal-directed attention alters the tuning of
ing using arterial spin-labeled perfusion mag- object-based representations in extrastriate
netic resonance imaging. Neurotherapeutics cortex. Front Neurosci 6:187
4(3):346–359 114. Çukur TNishimoto S, Huth A, Gallant
100. Brown GG, Clark C, Liu TT (2007) J (2013) Attention during natural vision
Measurement of cerebral perfusion with arte- warps semantic representation across
rial spin labeling. Part 2. Applications. J Int the human brain. Nat Neurosci
Neuropsychol Soc 13(3):526–538 16:763–770
101. Aguirre GK, Detre JA, Zarahn E, Alsop DC 115. Buchel C, Coull JT, Friston KJ (1999) The
(2002) Experimental design and the relative predictive value of changes in effective con-
sensitivity of BOLD and perfusion nectivity for human learning. Science
fMRI. Neuroimage 15(3):488–500 283(5407):1538–1541
102. Liu TT, Brown GG (2007) Measurement of 116. McIntosh AR, Grady CL, Haxby JV,
cerebral perfusion with arterial spin labeling. Ungerleider LG, Horwitz B (1996) Changes
Part 1. Methods. J Int Neuropsychol Soc in limbic and prefrontal functional interac-
13(3):517–525 tions in a working memory task for faces.
103. Fernandez-Seara MA, Wang J, Wang Z et al Cereb Cortex 6(4):571–584
(2007) Imaging mesial temporal lobe activa- 117. Gerstein GL, Perkel DH, Subramanian KN
tion during scene encoding: comparison of (1978) Identification of functionally related
fMRI using BOLD and arterial spin labeling. neural assemblies. Brain Res 140(1):43–62
Hum Brain Mapp 28(12):1391–1400 118. Penny WD, Stephan KE, Mechelli A, Friston
104. Feinberg DA, Beckett A, Chen L (2013) KJ (2004) Modelling functional integration:
Arterial spin labeling with simultaneous a comparison of structural equation and
multi-slice echo planar imaging. Magn Reson dynamic causal models. Neuroimage 23(Suppl
Med 70(6):1500–1506 1):S264–S274
105. Kanwisher N, McDermott J, Chun MM 119. Sun FT, Miller LM, D’Esposito M (2004)
(1997) The fusiform face area: a module in Measuring interregional functional con-
human extrastriate cortex specialized for face nectivity using coherence and partial coher-
perception. J Neurosci 17:4302–4311 ence analyses of fMRI data. Neuroimage
106. Haxby JV, Gobbini MI, Furey ML, Ishai A, 21(2):647–658
Schouten JL, Pietrini P (2001) Distributed 120. Sun FT, Miller LM, D’Esposito M (2005)
and overlapping representations of faces and Measuring temporal dynamics of functional
objects in ventral temporal cortex. Science networks using phase spectrum of fMRI data.
293(5539):2425–2430 Neuroimage 28(1):227–237
107. Polyn SM, Natu VS, Cohen JD, Norman KA 121. Sun FT, Miller LM, Rao AA, D’Esposito M
(2005) Category-specific cortical activity pre- (2007) Functional connectivity of cortical net-
cedes retrieval during memory search. Science works involved in bimanual motor sequence
310(5756):1963–1966 learning. Cereb Cortex 17(5):1227–1234
108. Zarahn E, Rakitin BC, Abela D, Flynn J, 122. Gazzaley A, Rissman J, D’Esposito M (2004)
Stern Y (2006) Distinct spatial patterns of Functional connectivity during working
Functional MRI: Applications in Cognitive Neuroscience 353
memory maintenance. Cogn Affect Behav 136. Gibbs SE, D’Esposito M (2005) Individual
Neurosci 4(4):580–599 capacity differences predict working memory
123. Fuhrmann Alpert G, Sun FT, Handwerker performance and prefrontal activity following
D, D’Esposito M, Knight RT (2007) Spatio- dopamine receptor stimulation. Cogn Affect
temporal information analysis of event- Behav Neurosci 5(2):212–221
related BOLD responses. Neuroimage 137. Gibbs SE, D’Esposito M (2005) A func-
34(4):1545–1561 tional MRI study of the effects of bro-
124. Rees G, Frith CD, Lavie N (1997) mocriptine, a dopamine receptor agonist, on
Modulating irrelevant motion perception by component processes of working memory.
varying attentional load in an unrelated task. Psychopharmacology (Berl) 180(4):644–653
Science 278(5343):1616–1619 138. Gibbs SE, D’Esposito M (2006) A func-
125. Treisman AM (1969) Strategies and mod- tional magnetic resonance imaging study of
els of selective attention. Psychol Rev the effects of pergolide, a dopamine receptor
76(3):282–299 agonist, on component processes of working
126. Lavie N, Tsal Y (1994) Perceptual load as memory. Neuroscience, 139:359–71
a major determinant of the locus of selec- 139. Cools R, Sheridan M, Jacobs E, D’Esposito M
tion in visual attention. Percept Psychophys (2007) Impulsive personality predicts dopa-
56(2):183–197 mine-dependent changes in frontostriatal
127. McCarthy RA, Warrington EK (1994) activity during component processes of work-
Disorders of semantic memory. Philos Trans ing memory. J Neurosci 27(20):5506–5514
R Soc Lond B Biol Sci 346(1315):89–96 140. Kastner S, Pinsk MA (2004) Visual attention
128. Warrington EST (1984) Category specific as a multilevel selection process. Cogn Affect
semantic impairments. Brain 107:829–854 Behav Neurosci 4(4):483–500
129. Thompson-Schill SL (2003) Neuroimaging 141. Gazzaley A, Cooney JW, McEvoy K, Knight
studies of semantic memory: inferring RT, D’Esposito M (2005) Top-down
“how” from “where”. Neuropsychologia enhancement and suppression of the mag-
41(3):280–292 nitude and speed of neural activity. J Cogn
Neurosci 17(3):507–517
130. Thompson-Schill SL, D’Esposito M, Aguirre
GK, Farah MJ (1997) Role of left inferior pre- 142. Arnemann KL, Chen AJW, Novakovic-
frontal cortex in retrieval of semantic knowl- Agopian T, Gratton C, Nomura EM,
edge: a reevaluation. Proc Natl Acad Sci U S D'Esposito M (2015) Functional brain net-
A 94(26):14792–14797 work modularity predicts response to cogni-
tive training after brain injury. Neurology
131. Thompson-Schill SL, Swick D, Farah MJ, 84:1568–74
D’Esposito M, Kan IP, Knight RT (1998)
Verb generation in patients with focal frontal 143. Chen AJW, Novakovic-Agopian T, Nycum
lesions: a neuropsychological test of neuro- TJ, Song S, Turner G, Rome S, Abrams G,
imaging findings. Proc Natl Acad Sci U S A D’Esposito M (2011) Training of goal-
95(26):15855–15860 directed attention regulation enhances con-
trol over neural processing for individuals
132. Pascual-Leone A, Tarazona F, Keenan J, with brain injury. Brain 134(5):1541–1554
Tormos JM, Hamilton R, Catala MD (1999)
Transcranial magnetic stimulation and neuro- 144. Poldrack RA (2000) Imaging brain plasticity:
plasticity. Neuropsychologia 37(2):207–217 conceptual and methodological issues – a the-
oretical review. Neuroimage 12(1):1–13
133. Rushworth MF, Hadland KA, Paus T, Sipila
PK (2002) Role of the human medial fron- 145. Aron AR, Gluck MA, Poldrack RA (2006)
tal cortex in task switching: a combined Long-term test-retest reliability of functional
fMRI and TMS study. J Neurophysiol MRI in a classification learning task.
87(5):2577–2592 Neuroimage 29(3):1000–1006
134. Ruff CC, Bestmann S, Blankenburg F et al 146. Wei X, Yoo SS, Dickey CC, Zou KH,
(2008) Distinct causal influences of parietal Guttmann CR, Panych LP (2004) Functional
versus frontal areas on human visual cortex: MRI of auditory verbal working memory:
evidence from concurrent TMS fMRI. Cereb long-term reproducibility analysis.
Cortex 18(4):817–827 Neuroimage 21(3):1000–1008
135. Dehaene S, Spelke E, Pinel P, Stanescu R, 147. Yoo SS, Wei X, Dickey CC, Guttmann CR,
Tsivkin S (1999) Sources of mathematical Panych LP (2005) Long-term reproducibility
thinking: behavioral and brain-imaging evi- analysis of fMRI using hand motor task. Int
dence. Science 284(5416):970–974 J Neurosci 115(1):55–77
Chapter 12
Abstract
Language refers to the uniquely human capacity for communication through productive combination of
symbolic representations. Functional neuroimaging studies have in recent decades greatly expanded our
knowledge of the brain systems supporting language, producing a dramatic reawakening of interest in this
topic and a call to revise and extend the nineteenth century neuroanatomical model formulated by Broca,
Wernicke, and others. This chapter presents some theoretical issues regarding functional imaging of lan-
guage systems, a model of the functional neuroanatomy of language based on recent empirical results in
several selected processing domains, and a survey of language mapping paradigms in common clinical use.
A central theme is that interpretation of fMRI language studies depends on an informed analysis of the
cognitive processes engaged during scanning. This analytic approach can help avoid common pitfalls in
task design that limit the sensitivity and specificity of language mapping studies and should encourage the
development of a standardized methodological and conceptual framework for such studies.
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_12, © Springer Science+Business Media New York 2016
355
356 Jeffrey R. Binder
Fig. 1 A schematic model of some major language regions and networks. Yellow indicates a bilaterally repre-
sented phoneme (speech sound) perception system. Blue indicates the posterior perisylvian area (posterior
superior temporal and supramarginal gyri, roughly equivalent to the traditional Wernicke area), which supports
pre-articulatory phonological access. Red indicates the temporal and parietal components of a distributed sys-
tem that stores and processes word meaning (semantic memory) information. ORANGE indicates several pre-
frontal components of the semantic processing network, including the pars orbitalis of the inferior frontal gyrus
(IFGpo) and the dorsomedial prefrontal cortex (DMPFC), which are proposed to control the activation and selec-
tion of information in the posterior semantic memory store. GREEN indicates a more general language control
system, made up of the pars triangularis and opercularis of the IFG (roughly equivalent to the traditional Broca
area) and adjacent cortex in the inferior frontal sulcus, which is proposed to control the retrieval and mainte-
nance of phonological information, a process that is critical for word retrieval, verbal working memory, and
sentence production. Speech repetition requires the pathway designated A in the figure, linking phoneme per-
ception with phonological access systems, as well as more anterior sensorimotor regions (not shown) that
support articulatory preparation and execution. Spoken word comprehension involves pathway B in the figure,
which maps perceived phoneme sequences to word concepts. Communicative speech production, in which the
speaker retrieves words and formulates sentences to express concepts, requires control of the semantic system
by the pathways marked D, as well as pathway C, which maps concept representations onto phonological rep-
resentations, and pathway E, which controls and maintains the activation of phonological codes. Pathway F
indicates a direct mapping from visual word forms to phonological representations, required for reading aloud
the left and right STS, though usually with leftward lateralization.
These activations are identical whether the stimuli are words or
word-like pseudowords, thus they reflect processing of phonemes
and not word meaning [54]. Though some of the activation in this
region could be explained by the fact that speech sounds are more
acoustically complex than the tones and noises used as nonspeech
controls, more recent experiments using acoustically matched
speech and nonspeech sounds (e.g., rotated speech, sinewave
speech) have shown convincingly that at least some of the activa-
tion in this region is due specifically to activation of phoneme codes
[58–63].
These observations are fully consistent with localization data
from patients with pure word deafness, who typically have lesions
restricted to the STG and STS [64–70]. Most of these cases have
bilateral lesions, though rarely a large left temporal lobe lesion can
produce the syndrome [71, 72]. These patients show impairments
in recognizing speech phonemes and may have other deficits of
higher-order auditory perception, especially when the lesions are
bilateral, but they have no deficits in written comprehension, nam-
ing, or propositional speech production that would indicate any loss
of word concepts. Taken together, these functional imaging and
lesion data make it clear that the left STG plays a relatively specific
role in language processing, i.e., that it contains general auditory
systems and specialized networks for recognizing speech phonemes,
regardless of whether these phonemes form words or have meaning.
This conceptualization stands in stark contrast to the traditional clin-
ical model of aphasia, which identifies the left STG as “Wernicke’s
area,” the principal site for “language comprehension.”
2.3 Phonological The term paraphasia refers to speech production that is fluent but
Access and contains errors, such as substitution of incorrect phonemes or words,
Phonological Working or rearrangement of the order of phonemes within a word. Paraphasia
Memory is characteristic of many forms of aphasia, particularly the Wernicke
and conduction syndromes, and typically affects both spoken and
written output. Paraphasia indicates an inability to retrieve or properly
use a mental representation of word sounds—what nineteenth cen-
tury theorists called “sound images” and what in modern parlance are
referred to as phonological representations. Patients who cannot access
(i.e., activate, compute) correct phonological representations show
paraphasic errors on all speech output tasks, including conversing,
naming objects, reading aloud, and repeating, as well as on a variety of
other tasks that require phonological access. For example, patients
with phonological impairments may be unable to determine whether
two printed words rhyme. Writing normally involves a mapping from
phonological (sound-based) to orthographic (grapheme-based) rep-
resentations, which is why patients with impaired phonology also typi-
cally show paraphasic errors in their writing.
The brain regions most strongly implicated in phonological access
(see blue region in Fig. 1) are in the left posterior perisylvian area,
especially the posterior STG, posterior STS, and supramarginal gyrus
(SMG). For example, patients with conduction aphasia—a relatively
isolated disorder of phonological access featuring phonemic parapha-
sia in naming, reading, and repetition tasks—have lesions confined to
this region [97–104], as do patients with phonological deficits in writ-
ten production [105–107]. A recent voxel-based lesion correlation
study linked damage in this posterior perisylvian region with inability
362 Jeffrey R. Binder
2.4 Semantic The human brain has an enormous capacity to acquire knowledge
Memory and Semantic from experience. The characteristic shapes, colors, textures, move-
Processing ments, sounds, smells, and actions associated with objects in the
environment, for example, must all be learned from experience. In
addition, consider the enormous variety of verb concepts (build, cel-
ebrate, discuss, throw, etc.), which depend on knowledge of how par-
ticular kinds of events happen, or social/emotional concepts (anger,
deceit, love, trust, etc.), which depend on knowledge of how human
beings behave and why. Much of this knowledge is represented sym-
bolically in language and underlies our understanding of word
meanings. These relationships between words and the stores of
knowledge they signify are known collectively as the semantics of a
language [127]. The term semantic processing refers to the cognitive
act of accessing stored knowledge about the world through words.
The stored knowledge itself is often called semantic memory.
fMRI of Language Systems 363
2.5 Sentence The work reviewed so far focused on processing of single word
Comprehension structure and meaning, which can be thought of as the basic
and Syntax Processing building blocks of language. Natural language, however, consists
almost entirely of sentences. At least two phenomena distinguish
processing at the sentence level from processing of single words.
First, in sentence processing, the meanings of individual words are
combined to create more complex and context-specific meanings.
For example, consider:
1. The tigers lost their jungle habitat.
2. The tigers lost in extra innings.
It is the combination of words that specifies in each case the
meaning of “tigers” and “lost.” This process of conceptual combi-
nation is a fundamental phenomenon in language production and
comprehension. A second distinguishing feature of sentence pro-
cessing is the use of syntactic information—word order, grammati-
cal function words, and word inflections—to indicate the thematic
roles played by constituent content words. In the two example
sentences above, for example, “the” marks the beginning of a noun
phrase, which can be followed by either a noun or a modifier
phrase. The plural inflection of “tigers” then identifies this second
word as a noun, which because of its position is likely to be the
subject of the sentence, and so on.
One type of neuroimaging study used to examine these pro-
cesses compares processing of sentences with word lists that do not
form a sentence, the latter sometimes created simply by randomly
fMRI of Language Systems 365
Table 1
Effects of auditory and visual stimuli on sensory and linguistic processing systems
Table 2
Effects of task states on some linguistic processing systems
“passively” to the subject [25, 33–35, 37, 208]. Other tasks sup-
press semantic processing by requiring a focusing of attention on
perceptual, orthographic, or phonological properties of stimuli [25,
26, 35, 208]. Examples include “Sensory Discrimination” tasks
(e.g., intensity, size, color, frequency, and other discriminations
based on physical features), “Phonetic Decision” tasks in which the
subject must detect a target phoneme or phonemes, “Phonological
Decision” tasks requiring a decision based on the phonological
properties of a stimulus (e.g., detection of rhymes, judgment of syl-
lable number), and “Orthographic Decision” tasks requiring a deci-
sion based on the letters in the stimulus (e.g., case matching, letter
identification). Other tasks, such as reading and repeating, make no
overt demands on semantic systems but may elicit automatic seman-
tic processing. The extent to which this occurs probably depends
on how meaningful the stimulus is: sentences likely elicit more
semantic processing than isolated words, which in turn elicit more
than pseudowords. Finally, many tasks make overt demands on
retrieval and use of semantic knowledge. These include “Semantic
Decision” tasks requiring a decision based on the meaning of the
stimulus (e.g., “Is it living or nonliving?”), “Word Generation”
tasks requiring retrieval of a word or series of words related in mean-
ing to a cue word, and “Naming” tasks requiring retrieval of a ver-
bal label for an object or object description.
As noted earlier, “Phonological Access” refers to the processes
engaged in retrieving a phonological (sound-based) representation
of a word (or pseudoword). In addition to speech output and pho-
nological tasks, any task using printed words, including ortho-
graphic and semantic tasks, will be accompanied to some degree by
obligatory phonological access [8, 15, 111]. In contrast, “Speech
Articulation” processes are engaged fully only when an overt spo-
ken response is produced [119]. “Verbal Working Memory” is
required whenever a written or spoken stimulus must be held in
memory. Some degree of short-term phonological memory is
needed for most language tasks, and particularly in cases where the
stimulus is relatively long (i.e., sentences more than single words)
or has multiple components, or must be held in memory while a
response is generated (e.g., word generation tasks involving mul-
tiple responses for each cue). Finally, semantic decision, word gen-
eration, and naming tasks make strong demands on frontal
mechanisms involved in searching for and retrieving information
associated with a stimulus [118, 195, 209, 210].
With these somewhat over-simplified stimulus and task charac-
terizations in mind, it is possible to make some general predictions
about the processing systems whose level of activation will differ
when two task conditions are contrasted, and thus the likely pat-
tern of brain activation that will be observed in a simple subtrac-
tion analysis. Some commonly encountered examples are listed
below and in Table 3.
fMRI of Language Systems 369
Table 3
Some task contrasts used for language mapping and the regions in which robust activations are
typically observed
Paradigm 1
Language task: Passively Listening to Words or Sentences
Control task: Rest
As shown in Table 1, auditory words activate early auditory cor-
tices and phoneme perception areas. Since both rest and passive
stimulation are accompanied by spontaneous semantic processes and
make no other overt cognitive demands, no other language-related
activation should appear in this contrast. The resulting activation
pattern involves mainly auditory cortex in the superior temporal gyri
bilaterally (Fig. 2a) [54, 167, 208, 211, 212]. The magnitude and
extent of this activation increase with rate of word presentation
[213, 214]. This STG activation is relatively symmetrical and is not
correlated with language dominance as measured by Wada testing
[215]. Although some authors have equated this STG activation
with “Wernicke’s area for receptive language,” most of this
Fig. 2 Group average fMRI activation patterns in 26 neurologically normal, right-handed volunteers during five
fMRI language paradigms (see [208] for details). Auditory word and tone stimuli were equivalent in each of the
five paradigms. (a) Passive listening to words contrasted with resting. Superior temporal activation occurs
bilaterally. (b) Passive listening to words contrasted with passive listening to tones. A small region in the left
STS shows activation specifically related to speech processing. (c) Semantic decision on words contrasted
with resting. Activation occurs in bilateral auditory (STG) and attentional/working memory (dorsolateral pre-
frontal, anterior cingulate, anterior insula, IPS, and subcortical) networks, with left lateralization in the IFG.
fMRI of Language Systems 371
Fig. 2 (continued) (d) Semantic decision on words contrasted with a tone decision task. Activation is strongly
left-lateralized in prefrontal, lateral and ventral temporal, angular, and posterior cingulate cortices. (e) Semantic
decision on words contrasted with a phoneme decision task on pseudowords. Activation is strongly left-later-
alized in dorsal prefrontal, angular, ventral temporal, and posterior cingulate cortices. Data are displayed as
serial sagittal sections through the brain at 9-mm intervals. X-axis locations for each slice are given in the top
panel. Green lines indicate the stereotaxic Y and Z origin planes. Hot colors (red–yellow) indicate positive
activations and cold colors (blue–cyan) negative activations for each contrast. All maps are thresholded at a
whole-brain corrected P < 0.05 using voxel-wise P < 0.0001 and cluster extent >200 mm3. Adapted, with per-
mission, from [208]
372 Jeffrey R. Binder
Paradigm 4
Language task: Word Generation
Control task: Reading or Repeating
Here we assume that the same stimulus modality (auditory or
visual) is used for both tasks. The stimuli in both cases are single
words, thus no difference in activation of sensory, phoneme per-
ception, or visual word form systems is expected. Both tasks are
accompanied by semantic processing (automatic semantic access in
the case of the control task, effortful semantic retrieval in the case
of word generation) and by phonological access processes. The
word generation task makes greater demands on lexical search and
on working memory; consequently greater activation is expected
in left inferior frontal areas associated with these processes. These
predictions match findings in many studies using this contrast,
which show primarily left-lateralized activation in the IFG [118,
210, 226, 227].
Paradigm 5
Language task: Visual Object Naming
Control task: Rest
Compared to resting, visual object perception activates early visual
sensory cortices and higher-level object recognition systems bilaterally
(Table 1) [228–230]. There may be additional, left-lateralized activa-
tion in semantic systems of the ventrolateral posterior temporal lobe
[231–235]. Unlike resting, naming requires lexical search and phono-
logical access, and, when overt, speech articulation (Table 2). These
predictions match findings in several studies using this contrast, which
show extensive bilateral visual system activation and modest left later-
alized inferior frontal activation [223, 234–236].
Paradigm 6
Language task: Semantic Decision
Control task: Sensory Discrimination
We again assume that the same stimulus modality is used for
both tasks. If the stimuli in the sensory discrimination task are non-
linguistic (e.g., tones or nonsense shapes), then the semantic deci-
sion task will produce relatively greater activation in phoneme
perception or visual wordform systems, depending on the sensory
modality. In addition, there will be greater activation of semantic
memory and semantic search mechanisms in the semantic decision
task. Note that unlike the resting and passive control tasks used in
the protocols described so far, effortful sensory discrimination tasks
interrupt ongoing semantic processes, providing a control state that
is relatively free of conceptual or semantic processing [25, 34, 35,
37, 208]. Working memory systems may or may not be activated in
fMRI of Language Systems 373
this contrast, depending on whether or not the control task also has
a working memory component. These predictions match findings
in studies using this contrast, which show left lateralized activation
of phoneme perception (middle and anterior superior temporal sul-
cus) or visual wordform (mid-fusiform gyrus) regions, and exten-
sive activation of left prefrontal, lateral and ventral left temporal,
and left posterior parietal systems involved in semantic memory and
semantic access (Fig. 2d) [5, 50, 208, 237–240].
Paradigm 7
Language task: Semantic Decision
Control task: Phonological Decision
These tasks can also be given in either the visual or auditory
modality. Stimuli in the phonological decision task can be either
words or pseudowords, and these can be matched to the words
used in the semantic task on all structural (physical, orthographic,
phonological) variables. Thus, there should be no activation of
sensory or wordform systems in this contrast. There will be greater
activation of semantic memory and semantic search systems in the
semantic decision task. These predictions match findings in many
studies using this contrast, which show activation of left prefrontal,
lateral and ventral left temporal, and left posterior parietal systems
believed to be involved in semantic processing (Fig. 2e) [25, 50,
142–144, 146, 148, 208, 241, 242].
Paradigm 8
Language task: Sentence or Word Reading
Control task: Passively Viewing Letterstrings
Compared to letterstrings, sentences engage visual word-form,
syntactic, and phonological access systems, and make variable
demands on working memory. Both reading and passive viewing
probably involve semantic processing. There should be left-lateralized
activation of the fusiform gyrus (visual word-form system), posterior
STG and STS (phonological access), and IFG (orthographic-phono-
logical mapping, working memory, syntax). These predictions are
consistent with several studies using this contrast [111, 243–246].
In many clinical settings, the main goal of language mapping is
simply to identify as many language-related areas as possible and to
assess hemispheric lateralization of language. A review of Table 3
suggests that the “Semantic Decision vs. Sensory Discrimination”
paradigm may offer advantages for this purpose in terms of the sheer
number of regions activated and leftward lateralization of activation.
Binder et al. put this prediction to a quantitative test by comparing
the extent and lateralization of activation produced by five language-
related task contrasts, conducted on the same 26 participants during
a single scanning session [208]. These contrasts included: (1)
374 Jeffrey R. Binder
Fig. 3 Group average activation volumes (top graph) and laterality indexes (bottom graph) for five fMRI lan-
guage paradigms [208]. Laterality indexes can vary from −1 (all activation in the right hemisphere) to +1 (all
activation in the left hemisphere). Error bars represent standard error. The Semantic Decision–Tone Decision
paradigm produces the greatest left hemisphere activation as well as a strongly left-lateralized pattern
region. The reader should appreciate that the review given here is
merely a coarse outline of some of the most commonly used types
of stimuli and tasks. Above all, it is important to note that activa-
tions in a particular part of the language system are seldom “all or
none,” but vary in a graded way depending on the particular stim-
uli and tasks used.
References
1. Broca P (1861) Remarques sur le siège de 17. Glaser WR (1992) Picture naming. Cognition
la faculté du langage articulé; suivies d’une 42:61–105
observation d’aphemie. Bulletin de la Société 18. James W (1890) Principles of psychology.
Anatomique de Paris 6:330–357 Dover, New York
2. Wernicke C (1874) Der aphasische 19. Hebb DO (1954) The problem of con-
Symptomenkomplex. Cohn & Weigert, sciousness and introspection. In: Adrian ED,
Breslau Bremer F, Jasper HH (eds) Brain mechanisms
3. Lichtheim L (1885) On aphasia. Brain 7:433–484 and consciousness. A symposium. Charles
4. Geschwind N (1971) Aphasia. N Engl J Med C. Thomas, Springfield, IL, pp 402–421
284(12):654–656 20. Miller GA, Galanter E, Pribram K (1960)
5. Binder JR, Frost JA, Hammeke TA, Cox RW, Plans and the structure of behavior. Holt,
Rao SM et al (1997) Human brain language New York
areas identified by functional MRI. J Neurosci 21. Pope KS, Singer JL (1976) Regulation of the
17(1):353–362 stream of consciousness: Toward a theory of
6. Démonet J-F, Thierry G, Cardebat D (2005) ongoing thought. In: Schwartz GE, Shapiro
Renewal of the neurophysiology of lan- D (eds) Consciousness and self-regulation.
guage: functional neuroimaging. Physiol Rev Plenum, New York, pp 101–135
85(1):49–95 22. Picton TW, Stuss DT (1994) Neurobiology of
7. Binder JR, Price CJ (2001) Functional imag- conscious experience. Curr Opin Neurobiol
ing of language. In: Cabeza R, Kingstone A 4:256–265
(eds) Handbook of functional neuroimaging 23. Antrobus JS, Singer JL, Greenberg S (1966)
of cognition. MIT Press, Cambridge, MA, Studies in the stream of consciousness:
pp 187–251 experimental enhancement and suppression
8. Macleod CM (1991) Half a century of of spontaneous cognitive processes. Percept
research on the Stroop effect: an integrative Mot Skills 23:399–417
review. Psychol Bull 109:163–203 24. Teasdale JD, Proctor L, Lloyd CA, Baddeley
9. Reicher GM (1969) Perceptual recognition AD (1993) Working memory and stimulus-
as a function of meaningfulness of stimulus independent thought: effects of memory load
material. J Exp Psychol 81:274–280 and presentation rate. Eur J Cogn Psychol
10. Warren RM, Obusek CJ (1971) Speech per- 5(4):417–433
ception and phonemic restorations. Percept 25. Binder JR, Frost JA, Hammeke TA, Bellgowan
Psychophys 9:358–362 PSF, Rao SM et al (1999) Conceptual pro-
11. Ganong WF (1980) Phonetic categorization cessing during the conscious resting state:
in auditory word perception. J Exp Psychol a functional MRI study. J Cogn Neurosci
Hum Percept Perform 6:110–115 11(1):80–93
12. Marslen-Wilson WD, Tyler LK (1981) 26. McKiernan KA, Kaufman JN, Kucera-
Central processes in speech understanding. Thompson J, Binder JR (2003) A parametric
Philos Trans R Soc Lond B 295:317–332 manipulation of factors affecting task-induced
deactivation in functional neuroimaging.
13. Carr TH, McCauley C, Sperber RD, Parmalee J Cogn Neurosci 15(3):394–408
CM (1982) Words, pictures, and priming: on
semantic activation, conscious identification, 27. Révész G (ed) (1954) Thinking and speak-
and the automaticity of information process- ing: a symposium. North Holland Publishing,
ing. J Exp Psychol Hum Percept Perform Amsterdam
8:757–777 28. Weiskrantz L (ed) (1988) Thought without
14. Marcel AJ (1983) Conscious and uncon- language. Clarendon, Oxford
scious perception: Experiments on visual 29. Vygotsky LS (1962) Thought and language.
masking and word recognition. Cogn Psychol Wiley, New York
15:197–237 30. Karmiloff-Smith A (1992) Beyond modular-
15. Van Orden GC (1987) A ROWS is a ROSE: ity: a developmental perspective on cognitive
spelling, sound, and reading. Mem Cogn science. MIT Press, Cambridge, MA
15(3):181–198 31. Andreasen NC, O’Leary DS, Cizadlo T,
16. Burton MW, Baum SR, Blumstein SE (1989) Arndt S, Rezai K et al (1995) Remembering
Lexical effects on phonetic categorization of the past: two facets of episodic memory
speech: the role of acoustic structure. J Exp explored with positron emission tomography.
Psychol Hum Percept Perform 15:567–575 Am J Psychiatry 152:1576–1585
fMRI of Language Systems 377
32. Shulman GL, Fiez JA, Corbetta M, Buckner 47. Buchweitz A, Prat C (2013) The bilingual
RL, Meizin FM et al (1997) Common blood brain: flexibility and control in the human
flow changes across visual tasks: II. Decreases cortex. Phys Life Rev 10(4):428–443
in cerebral cortex. J Cogn Neurosci 48. Li P, Legault J, Litcofsky KA (2014)
9(5):648–663 Neuroplasticity as a function of second lan-
33. Mazoyer B, Zago L, Mellet E, Bricogne S, guage learning: anatomical changes in the
Etard O et al (2001) Cortical networks for human brain. Cortex 58:301–324
working memory and executive functions sus- 49. Bogen JE, Bogen GM (1976) Wernicke’s
tain the conscious resting state in man. Brain region – where is it? Ann NY Acad Sci
Res Bull 54(3):287–298 290:834–843
34. Stark CE, Squire LR (2001) When zero is 50. Démonet JF, Chollet F, Ramsay S, Cardebat
not zero: the problem of ambiguous baseline D, Nespoulous JL et al (1992) The anatomy
conditions in fMRI. Proc Natl Acad Sci U S A of phonological and semantic processing in
98(22):12760–12766 normal subjects. Brain 115:1753–1768
35. McKiernan KA, D’Angelo BR, Kaufman JN, 51. Zatorre RJ, Evans AC, Meyer E, Gjedde A
Binder JR (2006) Interrupting the “stream (1992) Lateralization of phonetic and pitch
of consciousness”: an fMRI investigation. discrimination in speech processing. Science
Neuroimage 29(4):1185–1191 256:846–849
36. Smallwood J, Schooler JW (2006) The rest- 52. Mummery CJ, Ashburner J, Scott SK, Wise
less mind. Psychol Bull 132(6):946–958 RJS (1999) Functional neuroimaging of
37. Mason MF, Norton MI, Van Horn JD, speech perception in six normal and two apha-
Wegner DM, Grafton ST et al (2007) sic subjects. J Acoust Soc Am 106:449–457
Wandering minds: the default network and 53. Belin P, Zatorre RJ, Lafaille P, Ahad P, Pike B
stimulus-independent thought. Science (2000) Voice-selective areas in human audi-
315(5810):393–395 tory cortex. Nature 403:309–312
38. Andrews-Hanna JR (2012) The brain’s 54. Binder JR, Frost JA, Hammeke TA, Bellgowan
default network and its adaptive role in inter- PSF, Springer JA et al (2000) Human tempo-
nal mentation. Neuroscientist 18(3):251–270 ral lobe activation by speech and nonspeech
39. Abutalebi J, Cappa S, Perani D (2005) What sounds. Cereb Cortex 10:512–528
can functional neuroimaging tell us about the 55. Blumstein SE, Myers EB, Rissman J (2005)
bilingual brain? In: Kroll J, de Groot AMB The perception of voice onset time: an fMRI
(eds) Handbook of bilingualism: psycholin- investigation of phonetic category structure.
guistic approaches. Oxford University Press, J Cogn Neurosci 17(9):1353–1366
New York, pp 497–515 56. Desai R, Liebenthal E, Possing ET, Waldron E,
40. Hernandez A, Li P, MacWhinney B (2005) Binder JR (2005) Volumetric vs surface-based
The emergence of competing modules in alignment for localization of auditory cortex
bilingualism. Trends Cogn Sci 9(5):220–225 activation. Neuroimage 26(4):1019–1029
41. Corina DP, Knapp H (2006) Sign language 57. Turkeltaub PE, Coslett HB (2010)
processing and the mirror neuron system. Localization of sublexical speech perception
Cortex 42(4):529–539 components. Brain Lang 114:1–15
42. MacSweeney M, Capek CM, Campbell R, 58. Dehaene-Lambertz G, Pallier C, Serniclaes W,
Woll B (2008) The signing brain: the neu- Sprenger-Charolles L, Jobert A et al (2005)
robiology of sign language. Trends Cogn Sci Neural correlates of switching from auditory
12(11):432–440 to speech perception. Neuroimage 24:21–33
43. Corina D, Singleton J (2009) Developmental 59. Liebenthal E, Binder JR, Spitzer SM, Possing
social cognitive neuroscience: insights from ET, Medler DA (2005) Neural substrates
deafness. Child Dev 80(4):952–967 of phonemic perception. Cerebral Cortex
44. Emmorey K, McCullough S (2009) The 15:1621–1631
bimodal bilingual brain: effects of sign lan- 60. Möttönen R, Calvert GA, Jaaskelainen IP,
guage experience. Brain Lang 109:124–132 Matthews PM, Thesen A et al (2006) Perceiving
45. Kotz SA (2009) A critical review of ERP and identical sounds as speech or non-speech modu-
fMRI evidence on L2 syntactic processing. lates activity in the left posterior superior tem-
Brain Lang 109(2):68–74 poral sulcus. Neuroimage 30:563–569
46. van Heuven WJ, Dijkstra T (2010) Language 61. Obleser J, Zimmerman J, Van Meter J,
comprehension in the bilingual brain: fMRI Rauschecker JP (2007) Multiple stages of
and ERP support for psycholinguistic models. auditory speech perception reflected in event-
Brain Res Rev 64(1):104–122 related fMRI. Cereb Cortex 17:2251–2257
378 Jeffrey R. Binder
62. Desai R, Liebenthal E, Waldron E, Binder JR 78. Vincent FM, Sadowsky CH et al (1977)
(2008) Left posterior temporal regions are Alexia without agraphia, hemianopia, or
sensitive to auditory categorization. J Cogn color-naming defect: a disconnection syn-
Neurosci 20:1174–1188 drome. Neurology 27:689–691
63. Liebenthal E, Desai R, Ellingson MM, 79. Henderson VW (1986) Anatomy of posterior
Ramachandran B, Desai A et al (2010) pathways in reading: a reassessment. Brain
Specialization along the left superior temporal Lang 29:119–133
sulcus for phonemic and non-phonemic cat- 80. Binder JR, Mohr JP (1992) The topography
egorization. Cerebral Cortex 20:2958–2970 of callosal reading pathways: a case-control
64. Barrett AM (1910) A case of pure word- analysis. Brain 115:1807–1826
deafness with autopsy. J Nerv Ment Dis 81. Beversdorf DQ, Ratcliffe NR, Rhodes CH,
37(2):73–92 Reeves AG (1997) Pure alexia: clinical-
65. Henschen SE (1918–1919) On the hearing pathological evidence for a lateralized visual
sphere. Acta Otolaryngol 1:423–486 language association cortex. Clin Neuropathol
66. Wohlfart G, Lindgren A, Jernelius B (1952) 16(6):328–331
Clinical picture and morbid anatomy in a case 82. Sakurai Y, Takeuchi S, Takada T, Horiuchi
of “pure word deafness”. J Nerv Ment Dis E, Nakase H et al (2000) Alexia caused by a
116:818–827 fusiform or posterior inferior temporal lesion.
67. Lhermitte F, Chain F, Escourolle R, Ducarne J Neuro Sci 178:42–51
B, Pillon A et al (1972) Etude des troubles 83. Leff AP, Crewes H, Plant GT, Scott SK,
perceptifs auditifs dans les lésions temporales Kennard C et al (2001) The functional anatomy
bilatérales. Revue Neurologique 24:327–351 of single-word reading in patients with hemi-
68. Kanshepolsky J, Kelley J, Waggener JD anopic and pure alexia. Brain 124:510–521
(1973) A cortical auditory disorder: clinical, 84. Cohen L, Martinaud O, Lemer C, Lehericy S,
audiologic and pathologic aspects. Neurology Samson Y et al (2003) Visual word recogni-
23:699–705 tion in the left and right hemispheres: ana-
69. Buchman AS, Garron DC, Trost-Cardamone tomical and functional correlates of peripheral
JE, Wichter MD, Schwartz D (1986) Word alexias. Cereb Cortex 13:1313–1333
deafness: one hundred years later. J Neurol 85. Tarkiainen A, Helenius P, Hansen PC,
Neurosurg Psychiatry 49:489–499 Cornelissen PL, Salmelin R (1999) Dynamics
70. Poeppel D (2001) Pure word deafness and of letter string perception in the human occipi-
the bilateral processing of the speech code. totemporal cortex. Brain 122:2119–2131
Cogn Sci 25:679–693 86. Cohen L, Dehaene S, Naccache L, Lehéricy
71. Liepmann H, Storch E (1902) Der mikros- S, Dehaene-Lambertz G et al (2000) The
kopische gehirnbefund bei dem fall gorstelle. visual word form area. Spatial and temporal
Monatsschrift fur Psychiatrie und Neurologie characterization of an initial stage of reading
11:115–120 in normal subjects and posterior split-brain
72. Stefanatos GA, Gershkoff A, Madigan S patients. Brain 123:291–307
(2005) On pure word deafness, temporal 87. Dehaene S, Naccache L, Cohen L, Bihan DL,
processing and the left hemisphere. J Int Mangin JF et al (2001) Cerebral mechanisms
Neuropsychol Soc 11(4):456–470 of word masking and unconscious repetition
73. Déjerine J (1891) Sur un cas de cécité verbal priming. Nat Neurosci 4(7):752–758
avec agraphie, suivi d’autopsie. C R Seances 88. Polk TA, Farah MJ (2002) Functional MRI
Soc Biol 3:197–201 evidence for an abstract, not perceptual, word-
74. Déjerine J (1892) Contribution à l’étude form area. J Exp Psychol Gen 131:65–72
anatomo-pathologique et clinique des différ- 89. Cohen L, Jobert A, Le Bihanc D, Dehaene
entes variétés de cécité verbale. C R Seances S (2004) Distinct unimodal and multimodal
Soc Biol 44(2):61–90 regions for word processing in the left tempo-
75. Déjerine J, Vialet N (1893) Contribution a ral cortex. Neuroimage 23:1256–1270
l’étude de la localisation anatomique de la 90. Vinckier F, Dehaene S, Jobert A, Dubus J,
cécité verbale pure. C R Seances Soc Biol Sigman M et al (2007) Hierarchical coding
45:790–793 of letter strings in the ventral stream: dissect-
76. Geschwind N (1965) Disconnection syn- ing the inner organization of the visual word-
dromes in animals and man. Brain 88:237– form system. Neuron 55(1):143–156
294, 585-644 91. Glezer LS, Jiang X, Riesenhuber M (2009)
77. Greenblatt SH (1976) Subangular alexia Evidence for highly selective neuronal tuning
without agraphia or hemianopsia. Brain Lang of whole words in the “Visual Word Form
3:229–245 Area”. Neuron 62:199–204
fMRI of Language Systems 379
92. Mano QR, Humphries CJ, Desai R, 105. Roeltgen DP, Sevush S, Heilman KM (1983)
Seidenberg MS, Osmon DC et al (2013) Phonological agraphia: writing by the lexical-
The role of left occipitotemporal cor- semantic route. Neurology 33:755–765
tex in reading: reconciling stimulus, 106. Alexander MP, Friedman RB, Loverso F,
task, and lexicality effects. Cereb Cortex Fischer RS (1992) Lesion localization of pho-
23(4):988–1001 nological agraphia. Brain Lang 43:83–95
93. Binder JR, Medler DA, Westbury CF, 107. Rapcsak SZ, Beeson PM, Henry ML, Leyden
Liebenthal E, Buchanan L (2006) Tuning A, Kim E et al (2009) Phonological dyslexia
of the human left fusiform gyrus to sub- and dysgraphia: cognitive mechanisms and
lexical orthographic structure. Neuroimage neural substrates. Cortex 45(5):575–591
33:739–748 108. Pillay SB, Stengel BC, Humphries C, Book
94. Patterson KE, Kay J (1982) Letter-by-letter DS, Binder JR (2014) Cerebral localization of
reading: psychological descriptions of a impaired phonological retrieval during rhyme
neurological syndrome. Q J Exp Psychol judgment. Ann Neurol 76:738–746
34A:411–442 109. Howard D, Patterson K, Wise R, Brown WD,
95. Reuter-Lorenz PA, Brunn JL (1990) A prel- Friston K et al (1992) The cortical localiza-
exical basis for letter-by-letter reading: a case tion of the lexicons. Brain 115:1769–1782
study. Cognit Neuropsychol 7:1–20 110. Price CJ, Wise RJS, Watson JDG, Patterson
96. Behrmann M, Plaut DC, Nelson J (1998) K, Howard D et al (1994) Brain activity dur-
A literature review and new data supporting ing reading. The effects of exposure duration
an interactive activation account of letter- and task. Brain 117:1255–1269
by-letter reading. In: Coltheart M (ed) Pure 111. Price CJ, Wise RSJ, Frackowiak RSJ (1996)
alexia (letter-by-letter reading). Pscyhology, Demonstrating the implicit processing of
Hove, UK, pp 7–51 visually presented words and pseudowords.
97. Liepmann H, Pappenheim M (1914) Über Cereb Cortex 6:62–70
einem Fall von sogenannter Leitungsaphasie 112. Hickok G, Erhard P, Kassubek J, Helms-
mit anatomischer Befund. Z Gesamte Neurol Tillery AK, Naeve-Velguth S et al (2000) A
Psychiatr 27:1–41 functional magnetic resonance imaging study
98. Benson DF, Sheremata WA, Bouchard R, of the role of left posterior superior tempo-
Segarra JM, Price D et al (1973) Conduction ral gyrus in speech production: implications
aphasia. A clinicopathological study. Arch for the explanation of conduction aphasia.
Neurol 28:339–346 Neurosci Lett 287:156–160
99. Damasio H, Damasio AR (1980) The ana- 113. Booth JR, Burman DD, Meyer JR, Gitelman
tomical basis of conduction aphasia. Brain DR, Parrish TB et al (2002) Functional anat-
103:337–350 omy of intra- and cross-modal lexical tasks.
100. Anderson JM, Gilmore R, Roper S, Crosson Neuroimage 16:7–22
B, Bauer RM et al (1999) Conduction apha- 114. Indefrey P, Levelt WJM (2004) The spatial
sia and the arcuate fasciculus: a reexamination and temporal signatures of word production
of the Wernicke-Geschwind model. Brain components. Cognition 92(1-2):101–144
Lang 70:1–12 115. Burton MW, Locasto PC, Krebs-Noble D,
101. Quigg M, Fountain NB (1999) Conduction Gullapalli RP (2005) A systematic investi-
aphasia elicited by stimulation of the left gation of the functional neuroanatomy of
posterior superior temporal gyrus. J Neurol auditory and visual phonological processing.
Neurosurg Psychiatry 66:393–396 Neuroimage 26(3):647–661
102. Axer H, Keyserlingk AG, Berks G, Keyserlingk 116. Callan AM, Callan DE, Masaki S (2005)
DF (2001) Supra- and infrasylvian conduc- When meaningless symbols become letters:
tion aphasia. Brain Lang 76:317–331 neural activity change in learning new phono-
103. Fridriksson J, Kjartansson O, Morgan PS, grams. Neuroimage 28:553–562
Hjaltason H, Magnusdottir S et al (2010) 117. Fiez JA, Raichle ME, Balota DA, Tallal P,
Impaired speech repetition and left parietal Petersen SE (1996) PET activation of poste-
lobe damage. J Neurosci 30:11057–11061 rior temporal regions during auditory word
104. Buchsbaum BR, Baldo J, D’Esposito presentation and verb generation. Cereb
M, Dronkers N, Okada K et al (2011) Cortex 6:1–10
Conduction aphasia, sensory-motor integra- 118. Warburton E, Wise RJS, Price CJ, Weiller C,
tion, and phonological short-term memory: Hadar U et al (1996) Noun and verb retrieval
an aggregate analysis of lesion and fMRI data. by normal subjects. Studies with PET. Brain
Brain Lang 119:119–128 119:159–179
380 Jeffrey R. Binder
119. Wise RSJ, Scott SK, Blank SC, Mummery 134. Dronkers NF, Wilkins DP, Van Valin RD,
CJ, Murphy K et al (2001) Separate neural Redfern BB, Jaeger JJ (2004) Lesion analysis
subsystems within ‘Wernicke’s area’. Brain of the brain areas involved in language com-
124:83–95 prehension. Cognition 92:145–177
120. Baddeley AD (1986) Working memory. 135. Patterson K, Nestor PJ, Rogers TT (2007)
Oxford University Press, Oxford Where do you know what you know? The
121. Paulesu E, Frith CD, Frackowiak RSJ (1993) representation of semantic knowledge in the
The neural correlates of the verbal component human brain. Nat Rev Neurosci 8:976–987
of working memory. Nature 362:342–345 136. Thompson-Schill SL (2003) Neuroimaging
122. Hickok G, Buchsbaum B, Humphries C, studies of semantic memory: inferring
Muftuler T (2003) Auditory-motor inter- “how” from “where”. Neuropsychologia
action revealed by fMRI: speech, music, 41:280–292
and working memory in area Spt. J Cognit 137. Martin A (2007) The representation of object
Neurosci 15(5):673–682 concepts in the brain. Annu Rev Psychol
123. Buchsbaum BR, Olsen RK, Koch P, Berman 58:25–45
KF (2005) Human dorsal and ventral audi- 138. Binder JR, Desai R, Conant LL, Graves WW
tory streams subserve rehearsal-based and (2009) Where is the semantic system? A
echoic processes during verbal working mem- critical review and meta-analysis of 120 func-
ory. Neuron 48(4):687–697 tional neuroimaging studies. Cereb Cortex
124. Buchsbaum BR, D’Esposito M (2008) The 19:2767–2796
search for the phonological store: from 139. Kiefer M, Pulvermüller F (2012) Conceptual
loop to convolution. J Cogn Neurosci representations in mind and brain: theoretical
20(5):762–778 developments, current evidence and future
125. Acheson DJ, Hamidi M, Binder JR, Postle directions. Cortex 48:805–825
BR (2011) A common neural substrate for 140. Meteyard L, Rodriguez Cuadrado S, Bahrami
language production and verbal working B, Vigliocco G (2012) Coming of age: a
memory. J Cogn Neurosci 23:1358–1367 review of embodiment and the neuroscience
126. DeWitt I, Rauschecker JP (2012) Phoneme of semantics. Cortex 48:788–804
and word recognition in the auditory ven- 141. Mummery CJ, Patterson K, Hodges JR, Wise
tral stream. Proc Natl Acad Sci U S A RJS (1996) Generating ‘tiger’ as an animal
109:E505–514 name or a word beginning with T: differ-
127. Bréal M (1897) Essai de sémantique (science ences in brain activation. Proc R Soc Lond B
des significations). Librairie Hachette, Paris 263:989–995
128. Alexander MP, Hiltbrunner B, Fischer RS 142. Price CJ, Moore CJ, Humphreys GW, Wise
(1989) Distributed anatomy of transcortical RJS (1997) Segregating semantic from pho-
sensory aphasia. Arch Neurol 46:885–892 nological processes during reading. J Cognit
129. Hart J, Gordon B (1990) Delineation of Neurosci 9(6):727–733
single-word semantic comprehension deficits 143. Cappa SF, Perani D, Schnur T, Tettamanti
in aphasia, with anatomic correlation. Ann M, Fazio F (1998) The effects of seman-
Neurol 27(3):226–231 tic category and knowledge type on lexical-
130. Chertkow H, Bub D, Deaudon C, Whitehead semantic access: a PET study. Neuroimage
V (1997) On the status of object concepts in 8(4):350–359
aphasia. Brain Lang 58:203–232 144. Roskies AL, Fiez JA, Balota DA, Raichle ME,
131. Tranel D, Damasio H, Damasio AR (1997) Petersen SE (2001) Task-dependent modu-
A neural basis for the retrieval of con- lation of regions in the left inferior frontal
ceptual knowledge. Neuropsychologia cortex during semantic processing. J Cognit
35:1319–1327 Neurosci 13(6):829–843
132. Gainotti G (2000) What the locus of brain 145. Binder JR, McKiernan KA, Parsons M,
lesion tells us about the nature of the cogni- Westbury CF, Possing ET et al (2003)
tive defect underlying category-specific disor- Neural correlates of lexical access during
ders: a review. Cortex 36:539–559 visual word recognition. J Cognit Neurosci
15(3):372–393
133. Damasio H, Tranel D, Grabowski T, Adolphs
R, Damasio A (2004) Neural systems behind 146. Devlin JT, Matthews PM, Rushworth MFS
word and concept retrieval. Cognition (2003) Semantic processing in the left infe-
92:179–229 rior prefrontal cortex: a combined functional
magnetic resonance imaging and transcranial
fMRI of Language Systems 381
magnetic stimulation study. J Cogn Neurosci 160. Rohrer JD, Warren JD, Modat M, Ridgway
15(1):71–84 GR, Douiri A et al (2009) Patterns of corti-
147. Rissman J, Eliassen JC, Blumstein SE (2003) cal thinning in the language variants of fron-
An event-related fMRI investigation of totemporal lobar degeneration. Neurology
implicit semantic priming. J Cogn Neurosci 72:1562–1569
15(8):1160–1175 161. Mion M, Patterson K, Acosta-Cabronero J,
148. Scott SK, Leff AP, Wise RJS (2003) Going Pengas G, Izquierdo-Garcia D et al (2010)
beyond the information given: a neural sys- What the left and right anterior fusiform
tem supporting semantic interpretation. gyri tell us about semantic memory. Brain
Neuroimage 19:870–876 133:3256–3268
149. Ischebeck A, Indefrey P, Usui N, Nose I, 162. Warrington EK, Shallice T (1984) Category
Hellwig F et al (2004) Reading in a regular specific semantic impairments. Brain
orthography: an fMRI study investigating the 107:829–854
role of visual familiarity. J Cognit Neurosci 163. Gonnerman LM, Andersen ES, Devlin JT,
16(5):727–741 Kempler D, Seidenberg MS (1997) Double
150. Binder JR, Westbury CF, Possing ET, dissociation of semantic categories in
McKiernan KA, Medler DA (2005) Distinct Alzheimer’s disease. Brain Lang 57:254–279
brain systems for processing concrete 164. Chan AS, Salmon DP, De La Pena J (2001)
and abstract concepts. J Cognit Neurosci Abnormal semantic network for “animals”
17(6):905–917 but not “tools” in patients with Alzheimer’s
151. Binder JR, Medler DA, Desai R, Conant LL, disease. Cortex 37:197–217
Liebenthal E (2005) Some neurophysiologi- 165. Fung TD, Chertkow H, Whatmough C,
cal constraints on models of word naming. Murtha S, Péloquin L et al (2001) The spec-
Neuroimage 27:677–693 trum of category effects in object and action
152. Sabsevitz DS, Medler DA, Seidenberg M, knowledge in dementia of the Alzheimer’s
Binder JR (2005) Modulation of the seman- type. Neuropsychology 15(3):371–379
tic system by word imageability. Neuroimage 166. Alexander MP, Benson DF, Stuss DT (1989)
27:188–200 Frontal lobes and language. Brain Lang
153. Vandenbulcke M, Peeters R, Fannes K, 37:656–691
Vandenberghe R (2006) Knowledge of 167. Mazoyer BM, Tzourio N, Frak V, Syrota A,
visual attributes in the right hemisphere. Nat Murayama N et al (1993) The cortical rep-
Neurosci 9(7):964–970 resentation of speech. J Cognit Neurosci
154. Damasio H (1989) Neuroimaging contri- 5(4):467–479
butions to the understanding of aphasia. In: 168. Stowe LA, Paans AMJ, Wijers AA, Zwarts F,
Boller F, Grafman J (eds) Handbook of neu- Mulder G et al (1999) Sentence comprehen-
ropsychology. Elsevier, Amsterdam, pp 3–46 sion and word repetition: a positron emission
155. Rapcsak SZ, Rubens AB (1994) Localization tomography investigation. Psychophysiology
of lesions in transcortical aphasia. In: Kertesz 36:786–801
A (ed) Localization and neuroimaging in 169. Friederici AD, Meyer M, von Cramon DY
neuropsychology. Academic, San Diego, (2000) Auditory language comprehension:
pp 297–329 an event-related fMRI study on the process-
156. Berthier ML (1999) Transcortical aphasias. ing of syntactic and lexical information. Brain
Psychology, Hove Lang 74:289–300
157. Mummery CJ, Patterson K, Price CJ, Ashburner 170. Vandenberghe R, Nobre AC, Price CJ (2002)
J, Frackowiak RS et al (2000) A voxel-based The response of left temporal cortex to sen-
morphometry study of semantic dementia: rela- tences. J Cogn Neurosci 14(4):550–560
tionship between temporal lobe atrophy and 171. Humphries C, Swinney D, Love T, Hickok G
semantic memory. Ann Neurol 47:36–45 (2005) Response of anterior temporal cortex
158. Rosen HJ, Gorno-Tempini ML, Goldman to syntactic and prosodic manipulations dur-
WP et al (2002) Patterns of brain atrophy ing sentence processing. Hum Brain Mapp
in frontotemporal dementia and semantic 26:128–138
dementia. Neurology 58:198–208 172. Humphries C, Binder JR, Medler DA,
159. Davies RR, Hodges JR, Krill JJ, Patterson Liebenthal E (2006) Syntactic and semantic
K, Halliday GM et al (2005) The patho- modulation of neural activity during auditory
logical basis of semantic dementia. Brain sentence comprehension. J Cognit Neurosci
128:1984–1995 18:665–679
382 Jeffrey R. Binder
173. Pallier C, Devauchelle AD, Devauchelle AD, 186. Wartenburger I, Heekeren HR, Burchert
Dehaene S (2011) Cortical representation of F, Heinemann S, De Bleser R et al (2004)
the constituent structure of sentences. Proc Neural correlates of syntactic transformations.
Natl Acad Sci U S A 108(6):2522–2527 Hum Brain Mapp 22:72–81
174. Humphries C, Binder JR, Medler DA, 187. Fiebach CJ, Schlesewsky M, Lohmann G
Liebenthal E (2007) Time course of semantic (2005) Revisiting the role of Broca’s area
processes during sentence comprehension: an in sentence processing: Syntactic integration
fMRI study. Neuroimage 36(3):924–932 versus syntactic working memory. Hum Brain
175. Kang AM, Constable RT, Gore JC, Avrutin Mapp 24:79–91
S (1999) An event-related fMRI study of 188. Chen E, West WC, Waters G, Caplan D
implicit phrase-level syntactic and semantic (2006) Determinants of BOLD signal cor-
processing. Neuroimage 10(5):98–110 relates of processing object-extracted relative
176. Embick D, Marantz A, Miyashita Y, O’Neil clauses. Cortex 42:591–604
W, Sakai KL (2000) A syntactic specializa- 189. Caplan D, Stanczak L, Waters G (2008)
tion for Broca’s area. Proc Natl Acad Sci USA Syntactic and thematic constraint effects
97:6150–6154 on blood oxygenation level dependent sig-
177. Meyer M, Friederici AD, von Cramon DY nal correlates of comprehension of relative
(2000) Neurocognition of auditory sentence clauses. J Cogn Neurosci 20(4):643–656
comprehension: event related fMRi reveals 190. Just MA, Carpenter PA, Keller TA, Eddy WF,
sensitivity to syntactic violations and task Thulborn KR (1996) Brain activation modu-
demands. Cogn Brain Res 9:19–33 lated by sentence comprehension. Science
178. Ni W, Constable RT, Mencl WE, Pugh KR, 274:114–116
Fullbright RK et al (2000) An event-related 191. Stowe LA, Broere CA, Paans AM, Wijers
neuroimaging study distinguishing form and AA, Mulder G et al (1998) Localizing com-
content in sentence processing. J Cognit ponents of a complex task: Sentence pro-
Neurosci 12:120–133 cessing and working memory. Neuroreport
179. Newman AJ, Pancheva R, Ozawa K, Neville 9:2995–2999
HJ, Ullman MT (2001) An event-related 192. Caplan D, Waters GS (1999) Verbal working
fMRI study of syntactic and semantic viola- memory and sentence comprehension. Behav
tions. J Psycholinguist Res 30:339–364 Brain Sci 22(1):77–94
180. Kuperberg GR, Holcomb PJ, Sitnikova T, 193. Keller TA, Carpenter PA, Just MA (2001)
Greve D, Dale AM et al (2003) Distinct pat- The neural bases of sentence comprehension:
terns of neural modulation during the pro- a fMRI examination of syntactic and lexical
cessing of conceptual and syntactic anomalies. processing. Cereb Cortex 11:223–237
J Cognit Neurosci 15(2):272–293 194. Cooke A, Zurif EB, DeVita C, Alsop D,
181. Caplan D, Alpert N, Waters GS (1998) Effects Koenig P et al (2002) Neural basis for sen-
of syntactic structure and prepositional num- tence comprehension: grammatical and short-
ber on patterns of regional cerebral blood term memory components. Hum Brain Mapp
flow. J Cogn Neurosci 10(4):541–552 15:80–94
182. Fiebach CJ, Schlesewsky M, Friederici 195. Novick JM, Trueswell JC, Thompson-Schill
AD (2001) Syntactic working memory SL (2005) Cognitive control and parsing:
and the establishment of filler-gap depen- reexamining the role of Broca’s area in sen-
dencies: insights from ERPs and fMRI. J tence comprehension. Cognit Affect Behav
Psycholinguist Res 30:321–338 Neurosci 5(3):263–281
183. Ben-Shachar M, Hendler T, Kahn I, Ben- 196. Rogalsky C, Hickok G (2011) The role of
Bashat D, Grodzinsky Y (2003) The neural Broca’s area in sentence comprehension.
reality of syntactic transformations: evidence J Cogn Neurosci 23(7):1664–1680
form fMRI. Psychol Sci 14:433–440 197. Caplan D, Waters G (2013) Memory mecha-
184. Friederici AD, Rüschemeyer SA, Hahne A, nisms supporting syntactic comprehension.
Fiebach CJ (2003) The role of left inferior fron- Psychon Bull Rev 20(2):243–268
tal gyrus and superior temporal cortex in sen- 198. Caplan D (2001) Functional neuroim-
tence comprehension: localizing syntactic and aging studies of syntactic processing.
semantic processes. Cerebr Cortex 13:170–177 J Psycholinguist Res 30(3):297–320
185. Ben-Shachar M, Palti D, Grodzinsky Y (2004) 199. Friederici AD, Kotz SA (2003) The brain basis
The neural correlates of syntactic movement: of syntactic processes: functional imaging and
converging evidence from two fMRI experi- lesion studies. Neuroimage 20:S8–S17
ments. Neuroimage 21:1320–1336
fMRI of Language Systems 383
200. Martin RC (2003) Language processing: within the human auditory cortex when lis-
functional organization and neuroanatomical tening to words. Neurosci Lett 146:179–182
basis. Annu Rev Psychol 54:55–89 214. Binder JR, Rao SM, Hammeke TA, Frost JA,
201. Grodzinsky Y, Friederici AD (2006) Bandettini PA et al (1994) Effects of stimu-
Neuroimaging of syntax and syntactic pro- lus rate on signal response during functional
cessing. Curr Opin Neurobiol 16:240–246 magnetic resonance imaging of auditory cor-
202. Bornkessel-Schlesewsky I, Schlesewsky M tex. Cogn Brain Res 2:31–38
(2013) Reconciling time, space and function: 215. Lehéricy S, Cohen L, Bazin B, Samson S,
a new dorsal-ventral stream model of sentence Giacomini E et al (2000) Functional MR
comprehension. Brain Lang 125(1):60–76 evaluation of temporal and frontal language
203. Badre D, Poldrack RA, Pare-Blagoev EJ, dominance compared with the Wada test.
Insler RZ, Wagner AD (2005) Dissociable Neurology 54:1625–1633
controlled retrieval and generalized selection 216. Humphries C, Willard K, Buchsbaum B,
mechanisms in ventrolateral prefrontal cortex. Hickok G (2001) Role of anterior temporal
Neuron 47:907–918 cortex in auditory sentence comprehension:
204. Scott SK, Blank C, Rosen S, Wise RJS an fMRI study. Neuroreport 12:1749–1752
(2000) Identification of a pathway for intel- 217. Crinion JT, Lambon-Ralph MA, Warburton
ligible speech in the left temporal lobe. Brain EA, Howard D, Wise RJS (2003) Temporal
123:2400–2406 lobe regions engaged during normal speech
205. Davis MH, Johnsrude IS (2003) Hierarchical comprehension. Brain 126:1193–1201
processing in spoken language comprehen- 218. Spitsyna G, Warren JE, Scott SK, Turkheimer
sion. J Neurosci 23(8):3423–3431 FE, Wise RJS (2006) Converging lan-
206. Specht K, Reul J (2003) Functional segrega- guage streams in the human temporal lobe.
tion of the temporal lobes into highly differ- J Neurosci 26(28):7328–7336
entiated subsystems for auditory perception: 219. Awad M, Warren JE, Scott SK, Turkheimer
an auditory rapid event-related fMRI task. FE, Wise RJS (2007) A common system for
Neuroimage 20:1944–1954 the comprehension and production of narra-
207. Uppenkamp S, Johnsrude IS, Norris D, tive speech. J Neurosci 27(43):11455–11464
Marslen-Wilson W, Patterson RD (2006) 220. Eulitz C, Elbert T, Bartenstein P, Weiller C,
Locating the initial stages of speech-sound Müller SP et al (1994) Comparison of mag-
processing in human temporal cortex. netic and metabolic brain activity during a
Neuroimage 31:1284–1296 verb generation task. NeuroReport 6:97–100
208. Binder JR, Swanson SJ, Hammeke TA, 221. Ojemann JG, Buckner RL, Akbudak E, Snyder
Sabsevitz DS (2008) A comparison of five AZ, Ollinger JM et al (1998) Functional MRI
fMRI protocols for mapping speech compre- studies of word-stem completion: reliability
hension systems. Epilepsia 49(12):1980–1997 across laboratories and comparison to blood
209. Thompson-Schill SL, Aguirre GK, D’Esposito flow imaging with PET. Hum Brain Mapp
M, Farah MJ (1997) Role of left inferior pre- 6:203–215
frontal cortex in retrieval of semantic knowl- 222. Yetkin FZ, Swanson S, Fischer M, Akansel
edge: a reevaluation. Proc Natl Acad Sci U S G, Morris G et al (1998) Functional MR
A 94:14792–14797 of frontal lobe activation: comparison with
210. Thompson-Schill SL, D’Esposito M, Kan IP Wada language results. Am J Neuroradiol
(1999) Effects of repetition and competition 19:1095–1098
on activity in left prefrontal cortex during 223. Benson RR, FitzGerald DB, LeSeuer LL,
word generation. Neuron 23:513–522 Kennedy DN, Kwong KK et al (1999)
211. Wise R, Chollet F, Hadar U, Friston K, Language dominance determined by whole
Hoffner E et al (1991) Distribution of cor- brain functional MRI in patients with brain
tical neural networks involved in word lesions. Neurology 52:798–809
comprehension and word retrieval. Brain 224. Palmer ED, Rosen HJ, Ojemann JG, Buckner
114:1803–1817 RL, Kelley WM et al (2001) An event-related
212. Price CJ, Wise RJS, Warburton EA, Moore fMRI study of overt and covert word stem
CJ, Howard D et al (1996) Hearing and say- completion. Neuroimage 14:182–193
ing. The functional neuro-anatomy of audi- 225. Liégois F, Connelly A, Salmond CH, Gadian
tory word processing. Brain 119:919–931 DG, Vargha-Khadem F et al (2002) A direct
213. Price C, Wise R, Ramsay S, Friston K, Howard test for lateralization of language activation
D et al (1992) Regional response differences using fMRI: comparison with invasive assess-
384 Jeffrey R. Binder
ments in children with epilepsy. Neuroimage 239. Devlin JT, Russell RP, Davis MH, Price
17:1861–1867 CJ, Moss HE et al (2002) Is there an ana-
226. Raichle ME, Fiez JA, Videen TO, MacLeod AM, tomical basis for category-specificity?
Pardo JV et al (1994) Practice-related changes Semantic memory studies with PET and
in human brain functional anatomy during non- fMRI. Neuropsychologia 40:54–75
motor learning. Cereb Cortex 4(1):8–26 240. Xu B, Grafman J, Gaillard WD, Spanaki M,
227. Petersen SE, Fox PT, Posner MI, Mintun M, Ishii K et al (2002) Neuroimaging reveals
Raichle ME (1988) Positron emission tomo- automatic speech coding during percep-
graphic studies of the cortical anatomy of tion of written word meaning. Neuroimage
single-word processing. Nature 331:585–589 17(2):859–870
228. Malach R, Reppas JB, Benson RR, Kwong 241. Mummery CJ, Patterson K, Hodges JR, Price
KK, Jiang H et al (1995) Object-related activ- CJ (1998) Functional neuroanatomy of the
ity revealed by functional magnetic resonance semantic system: divisible by what? J Cogn
imaging in human occipital cortex. Proc Natl Neurosci 10(6):766–777
Acad Sci USA 92:8135–8139 242. Miceli G, Turriziani P, Caltagirone C, Capasso
229. Kanwisher N, Woods R, Iacoboni M, R, Tomaiuolo F et al (2002) The neural cor-
Mazziotta J (1996) A locus in human relates of grammatical gender: an fMRI inves-
extrastriate cortex for visual shape analysis. tigation. J Cognit Neurosci 14:618–628
J Cognit Neurosci 91:133–142 243. Bavelier D, Corina D, Jezzard P, Padmanabhan
230. Grill-Spector K, Kushnir T, Edelman S, S, Clark VP et al (1997) Sentence reading:
Avidian-Carmel G, Itzchak Y et al (1999) a functional MRI study at 4 tesla. J Cognit
Differential processing of objects under vari- Neurosci 9(5):664–686
ous viewing conditions in the human lateral 244. Herbster AN, Mintun MA, Nebes RD, Becker
occipital complex. Neuron 24:187–203 JT (1997) Regional cerebral blood flow dur-
231. Martin A, Wiggs CL, Ungerleider LG, ing word and nonword reading. Hum Brain
Haxby JV (1996) Neural correlates of Mapp 5:84–92
category-specific knowledge. Nature 245. Indefrey P, Kleinschmidt A, Merboldt KD,
379(6566):649–652 Krüger G, Brown C et al (1997) Equivalent
232. Price CJ, Moore CJ, Humphreys GW, responses to lexical and nonlexical visual
Frackowiak RSJ, Friston KJ (1996) The stimuli in occipital cortex: a functional mag-
neural regions sustaining object recogni- netic resonance imaging study. Neuroimage
tion and naming. Proc Roy Soc Lond B 5:78–81
263:1501–1507 246. Chee MW, Caplan D, Soon CS, Sriram
233. Zelkowicz BJ, Herbster AN, Nebes RD, N, Tan EWL et al (1993) Processing of
Mintun MA, Becker JT (1998) An examina- visually presented sentences in Mandarin
tion of regional cerebral blood flow during and English studied with fMRI. Neuron
object naming tasks. J Int Neuropsychol Soc 23:127–137
4:160–166 247. Démonet JF, Wise R, Frackowiak RSJ (1993)
234. Murtha S, Chertkow H, Beauregard M, Evans Language functions explored in normal sub-
A (1999) The neural substrate of picture jects by positron emission tomography: a crit-
naming. J Cognit Neurosci 11(4):399–423 ical review. Hum Brain Mapp 1(1):39–47
235. Price CJ, Devlin JT, Moore CJ, Morton C, Laird 248. Fiez JA (1997) Phonology, semantics and
AR (2005) Meta-analyses of object naming: the role of the left inferior prefrontal cortex.
effect of baseline. Hum Brain Mapp 25:70–82 Hum Brain Mapp 5:79–83
236. Kiasawa M, Inoue C, Kawasaki T, Tokoro T, 249. Poldrack RA, Wagner AD, Prull MW,
Ishii K et al (1996) Functional neuroanatomy Desmond JE, Glover GH et al (1999)
of object naming: a PET study. Graefes Arch Functional specialization for semantic and
Clin Exp Ophthalmol 234:110–115 phonological processing in the left inferior
237. Vandenberghe R, Price C, Wise R, Josephs O, prefrontal cortex. Neuroimage 10:15–35
Frackowiak RSJ (1996) Functional anatomy 250. Gold BT, Buckner RL (2002) Common
of a common semantic system for words and prefrontal regions coactivate with disso-
pictures. Nature 383:254–256 ciable posterior regions during controlled
238. Carpentier A, Pugh KR, Westerveld M, semantic and phonological tasks. Neuron
Studholme C, Skrinjar O et al (2001) 35:803–812
Functional MRI of language processing: 251. Henson RNA, Price CJ, Rugg MD, Turner
dependence on input modality and temporal R, Friston KJ (2002) Detecting latency dif-
lobe epilepsy. Epilepsia 42:1241–1254 ferences in event-related BOLD responses:
fMRI of Language Systems 385
application to words versus nonwords and demand in healthy volunteers: An fMRI study.
initial versus repeated face presentations. Synapse 42:266–272
Neuroimage 15(1):83–97 256. Braver TS, Barch DM, Gray JR, Molfese DL,
252. Mechelli A, Gorno-Tempini ML, Price CJ Snyder A (2001) Anterior cingulate cortex and
(2003) Neuroimaging studies of word and response conflict: effects of frequency, inhibi-
pseudoword reading: consistencies, incon- tion and errors. Cereb Cortex 11:825–836
sistencies, and limitations. J Cogn Neurosci 257. Ullsperger M, Von Cramon DY (2001)
15(2):260–271 Subprocesses of performance monitoring: a
253. Braver TS, Cohen JD, Nystrom LE, Jonides dissociation of error processing and response
J, Smith EE et al (1997) A parametric study competition revealed by event-related fMRI
of prefrontal cortex involvement in human and ERPs. Neuroimage 14:1387–1401
working memory. Neuroimage 5:49–62 258. Binder JR, Liebenthal E, Possing ET,
254. Honey GD, Bullmore ET, Sharma T (2000) Medler DA, Ward BD (2004) Neural cor-
Prolonged reaction time to a verbal working relates of sensory and decision processes in
memory task predicts increased power of pos- auditory object identification. Nat Neurosci
terior parietal cortical activation. Neuroimage 7(3):295–301
12(5):495–503 259. Desai R, Conant LL, Waldron E, Binder JR
255. Adler CM, Sax KW, Holland SK, Schmithorst (2006) FMRI of past tense processing: the
V, Rosenberg L et al (2001) Changes in neu- effects of phonological complexity and task
ronal activation with increasing attention difficulty. J Cognit Neurosci 18(2):278–297
Chapter 13
Abstract
Selective attention is a core cognitive ability that enables organisms to effectively process and act upon relevant
information while ignoring distracting events. Elucidating the neural bases of selective attention remains a key
challenge for neuroscience and represents an essential aim in translational efforts to ameliorate attentional deficits
in a wide variety of neurological and psychiatric disorders. Moreover, knowledge about the cognitive and neural
mechanisms of attention is essential for developing and refining brain–machine interfaces, and for advancing
methods for training and education. We will discuss how functional imaging methods have helped us to under-
stand fundamental aspects of attention: How attention is controlled, focused on relevant inputs, and reoriented,
and how this control results in the selection of relevant information. Work from our groups and from others will
be reviewed. We will focus on fMRI methods, but where appropriate will include related discussion of electro-
magnetic recording methods used in conjunction with fMRI, including simultaneous EEG/fMRI methods.
Key words Attention, Control, fMRI, EEG, ERPs, DCM, Human, Vision
1 Introduction
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_13, © Springer Science+Business Media New York 2016
387
388 George R. Mangun et al.
1.2 Early and Late One of the key questions in attention research has been where in
Selection Models information processing attention has its influence. If as Helmholtz
suggested, attention could select some information coming from
Neuroimaging Approaches to the Study of Visual Attention 389
1.3 Methodological The focus of the vast majority of studies of effects of attention on
Issues in Experimental perception involves selective attention, attention to one thing at
Studies of Selective the expense of another. This is to be contrasted with nonselective
Attention attention, which includes generalized behavioral arousal (e.g., the
classic orienting response) that does not necessarily involve attend-
ing one input while ignoring another. These latter nonselective
attention mechanisms are certainly interesting, but selective mech-
anisms have generated the greatest interest and we review only
these studies. Therefore, it is critical to understand the design
parameters that permit selective versus nonselective attention to be
isolated and studied, and to make note of some of the confounding
influences that might contaminate studies of selective attention
when nonselective factors (e.g., arousal) are not properly con-
trolled. We turn then to an example from the early history of the
physiology of attention.
390 George R. Mangun et al.
Fig. 1 Stimuli and task used in a functional imaging and event-related potential
(ERP) study of spatial attention [105]. Two blocked conditions of attention are
shown: Attend left condition (left column) and attend right condition (right col-
umn). Subjects viewed rapid sequences of arrays (about 2.5/s) of nonsense sym-
bols (flashed for 100 ms) while maintaining fixation of their eyes on a central
fixation spot (plus sign). There were always two symbols in the left and two in the
right visual hemifield in locations each demarcated by an outline rectangle. The
task was to detect and press a button to pairs of identical symbols at the attended
location and to ignore all stimuli in the opposite hemifield. In the figure (but not
in the actual experiment) the focus of covert spatial attention is indicated by a
dashed circle
Neuroimaging Approaches to the Study of Visual Attention 393
Fig. 2 fMRI and event-related potential (ERP) data from a study of visual-spatial selective attention. (a) Coronal
structural scan of a single subject showing activations in contralateral visual cortex with spatial attention in
the extrastriate cortex (left hemisphere is on the left). The activations were focused in the lingual gyrus (LG)
and posterior fusiform gyrus (FG) and the middle occipital gyrus (MOG). (b) ERP attention effects shown as
difference waves from a single lateral occipital electrode site in the right hemisphere (attend left minus attend
right). The vertical scale is 2 μV per side (positive plotted downward). The onset of the array is indicated at time
zero (t = 0), and the tick marks are 100 ms. (c) Topographic voltage attention difference map (110-ms latency)
on the scalp surface viewed from the rear (left side of head on left side of figure) (adapted from [105])
394 George R. Mangun et al.
attended side of the stimulus arrays (Fig. 2a). These activations were
highly statistically significant in the posterior fusiform gyrus on the
ventral cortical surface and in lateral-ventral occipital regions. In
these early studies, no effects of visual-spatial attention were observed
in primary visual cortex (area V1) but as we shall see later, such
effects have since been observed using fMRI. The finding of local-
ized brain activations corresponding to the action of spatial atten-
tion alone is in accord with evidence from prior ERP studies [15, 29,
30]. Also in this study, in a separate session, testing the same sub-
jects, we recorded ERPs for the same stimuli and task. The scalp-
recorded ERPs showed the expected P1 attention effects (Fig. 2b,
c), and by using neuroelectric dipole modeling we investigated
whether intracranial neural generators at the loci of functional acti-
vations could produce activity in the model scalp that was similar to
what we actually recorded in the ERPs (not shown in figure). We
showed that the ERP and fMRI attention effects in the posterior
fusiform gyrus (i.e., extrastriate visual cortex) were strongly related.
This combined use of ERPs and functional imaging provided
evidence for short-latency (around 100 ms after stimulus onset)
changes in responses to visual stimuli as a function of spatial attention
that were generated early in extrastriate visual cortex. In several stud-
ies, we and others have followed up these effects [31, 32] and have
observed that these modulations with spatial attention affect multiple
stages of visual cortical processing, from V1 toward inferotemporal
cortex in the ventral visual stream. The next section reviews related
work that also combined structural, functional, and cognitive imag-
ing to detail the structure of spatial attention effects in visual cortex.
2.2 Mapping Spatial A beautiful illustration of how spatial attention influences sensory
Attention in Vision processing in human visual cortex comes from the work of Tootell
et al. [33]. They used fMRI to identify the borders of the first few
visual areas in humans (retinotopic mapping) and then conducted a
spatial attention study similar to what was described earlier. In the
Tootell study, subjects performed a simple spatial attention task that
required subjects to covertly and selectively attend stimuli located in
one visual field quadrant while ignoring those in the other quad-
rants; different quadrants were attended in different conditions.
Attentional activations were then mapped onto the flattened repre-
sentations of the visual cortex, permitting the attention effects to be
related directly to the multiple visual areas of human visual cortex,
showing that spatial attention led to robust modulations of activity
in striate cortex and multiple extrastriate visual areas (Fig. 3).
By combining different methods for recording electrical activ-
ity, imaging brain structure, defining functional anatomy (i.e., reti-
notopic maps), and combining this with functional imaging in
carefully controlled studies of selective attention, we can learn a
great deal about the effects exerted by attention on sensory pro-
cesses in humans. Specifically, for spatial attention we now
Neuroimaging Approaches to the Study of Visual Attention 395
Fig. 3 Spatial attention effects in multiple visual cortical areas in humans as demonstrated by fMRI. Activations
with spatial attention to left-field stimuli are shown in the flattened right visual cortices of two subjects (one in
the left column and the other on the right). The white lines (dotted and solid) indicate the borders of the visual
areas as defined by representations of the horizontal and vertical meridians; each area is labeled from V1 (stri-
ate cortex) through V7, a retinotopic area adjacent to V3A. The solid black line is the representation of the hori-
zontal meridian in V3A. Panels (a) and (b) show the retinotopic mappings of the left visual field for each subject,
with colored activations corresponding to the polar angles shown at right (which represent the left visual field).
Panels (c) and (d) show the attention-related modulations (attended vs. unattended) of sensory responses to
a target in the upper left quadrant (the quadrant of the stimuli is shown at right). Panels (e) and (f) show the
same for stimuli in the lower left quadrant. In (c) through (f), the yellow to red colors indicate areas where
activity was greater when the stimulus was attended to than when it was ignored; the bluish colors represent
the opposite, where the activity was greater when the stimulus was ignored than when attended. The attention
effects in (c) through (f) can be compared to the pure sensory responses to the target bars when passively
viewed [(g) and (h)]. Note the retinotopic pattern of the attention effects in (c) through (f): The attention effects
to targets in the lower left quadrant produced activity in several lower field representations, which included the
appropriate half of V3A (inferior V3A labeled with an “i”) in both subjects, and V3 and V2 in one subject. In
contrast, attention to the upper left quadrant produced activity in the upper field representation of V3A (S) and
in the adjacent upper field representation of area V7 (from [33])
processing would not take place until the target is presented later at
time 2. Hence, by recording brain responses to the cue and target
separately, one can identify the networks for control and selection,
which was first utilized in ERP studies by Russell Harter and his col-
leagues in the late 1980s [39]. Blocked-design fMRI attention stud-
ies make this difficult because the activations produced by attentional
control and selection processes cannot be easily distinguished given
that they are co-occurring in the images. Although functional brain
imaging, which taps hemodynamic changes, is a rather poor method
for tracking the time course of brain activity, when used in an event-
related design, and appropriate analytic strategies are employed to
separate overlapping hemodynamic responses, it is possible to con-
duct an experiment like that just described; we and others have done
so in several studies of selective attention.
In our initial studies [31], we used event-related fMRI to inves-
tigate attentional control mechanisms during visual-spatial
398 George R. Mangun et al.
Fig. 5 Activations during attention control and target selection. Group (N = 6 subjects) average activations
time-locked to the cue (cues > targets contrast) isolating attentional control regions are shown on the left.
Those time locked to the targets (targets > cues contrast), indicating target processing and motor responses,
are on the right. The activated areas are shown in different colors to reflect the statistical contrasts that
revealed the activations: bluish for brain regions that were more activated to cues than targets, and reddish-
yellow to indicate regions that were more active to targets than cues. The top row shows a view of the dorsal
surface of the brain, the middle row shows a lateral view of the left hemisphere, and the bottom row shows its
medial surface; the activity was the same for the right hemisphere, which is not shown. Attentional control
involved the frontal-parietal attention network involving superior and middle frontal gyri (labeled 1–3), and the
regions in and around the IPS (labeled 4–7), the superior temporal cortex (labeled 8), and the posterior cingu-
late cortex (labeled 10). In contrast, during target selection, the main areas of activation were now the supple-
mentary motor area (labeled a), the motor and somatosensory cortex (labeled b–e), posterior superior parietal
lobule (labeled f, g), the ventral-lateral prefrontal cortex (labeled h) and the cuneus (labeled i) and the visual
cortex (labeled j) (adapted from [31])
Fig. 6 Priming of visual cortex by spatial attention. The top row shows increased baseline activity in six sub-
jects to the cues in visual cortex. The bottom row shows the same effects to targets (adapted from [31])
3.2 Specializations One key question about top–down attentional control systems is
in Attentional Control the extent to which such mechanisms are generalized for the con-
trol of attention regardless of the modality or specific stimulus
attributes to which selective attention is directed [46]. For exam-
ple, does preparatory attention for spatial selective attention involve
the same or different control networks as would selective attention
for color or motion? We have addressed this in studies using similar
methods to those described in the foregoing.
In one study [32] we undertook a direct test of whether the neu-
ral mechanisms for the top-down control of both spatial and nonspa-
tial attention were the same. In this study, we compared spatial
selective attention to nonspatial color-selective attention. That is,
subjects were either cued to select the target stimulus based on loca-
tion or color. Our design was as follows. We randomly intermingled
trials in which either the location or the color of an upcoming target
was cued. The cues were letters (e.g., “L” = attend left and “B” = attend
blue) located at fixation in one task or the periphery above fixation in
another version of the task. The stimuli to be discriminated were
rectangles (outlines) that when following spatial cues were located in
the left or right hemifields or when following color cues were over-
lapping outline rectangles located at fixation in one task version or
above fixation in the second task version. The participants were
instructed to covertly direct attention to select an upcoming rectan-
gle stimulus based on the cued feature (location or color) with the
Neuroimaging Approaches to the Study of Visual Attention 401
task of discriminating its orientation. That is, in the spatial cue condi-
tion, if the cue signified left, the subjects were to maintain fixation on
the central fixation point, but focus covert attention to the left and to
discriminate the orientation of the rectangle presented there (vertical
or horizontal), and to press the appropriate response key. When the
cue indicated blue, however, this meant that the subjects should
focus covert attention in preparation for discriminating the blue rect-
angle location in the same place as the cue (i.e., either at fixation or
in the periphery just above fixation).
As in our studies [31] described earlier, event-related fMRI
measures were obtained to the cues and to the targets, and sepa-
rately for spatial location and color attention trials. Because we
jittered the stimulus-onset-asynchrony between cues and targets
from 1000 to 8000 ms, it was possible to deconvolve the overlap-
ping hemodynamic responses [47–49]. This permitted us to utilize
a paradigm that was more similar to cued attention designs in the
cognitive psychology literature where the time between cues and
targets was not too long for subjects to maintain a strong atten-
tional set (i.e., to sustain the covert allocation of attention to the
cued color or location). In our prior work [31] and that of others,
the need to separate the sluggish hemodynamic responses to adja-
cent stimuli typically resulted in long stimulus-onset-asynchronies
in order to avoid overlap of the responses to cues and targets. By
using specially designed stimulus sequences and analytic strategies
it is possible to design cued attention studies that optimize the
design parameters for investigating attentional mechanisms.
Indeed, in cued attention designs, the speed of stimulus presenta-
tion can be even faster than the 1000–8000-ms lag between cues
and target described here [50, 51].
In this study, comparing preparatory attention for locations
and colors we found that large regions of the frontoparietal net-
work were commonly activated by the spatial and nonspatial cues
alike (Fig. 7). Such related patterns of activity reflect those aspects
of the task that the two attentional control conditions shared, such
as low-level sensory processing of the cues, decoding of linguistic
information in the cue letter and matching that to the task instruc-
tion, establishing the appropriate attentional set and holding this
information in working memory during the cue-to-target period,
and finally, preparing to respond.
To test whether any of the areas activated in response to the
cues were selective for spatial or nonspatial orienting, we directly
statistically compared activity in response to location and color
cues. The results of this direct comparison are shown in Fig. 8a for
the activations where locations cues produce more activity than
color cues. We found nonoverlapping regions of superior frontal
and parietal cortex that were activated during orienting to location
versus color. Orienting attention based on stimulus location acti-
vated regions of the dorsal frontal cortex [posterior middle frontal
402 George R. Mangun et al.
gyrus and frontal eye fields (FEF)], posterior parietal cortex [intra-
parietal sulcus (IPS) and precuneus], and supplementary motor
cortex. The inverse statistical contrast (color > location cues)
produced no significant activations in the superior frontal or pari-
etal regions, and only showed activity in the ventral occipital cortex
(OCC) (posterior fusiform, posterior middle temporal cortex, and
left insula) (not shown in the figure).
The pattern of selectivity in the frontoparietal attentional con-
trol network for location cuing suggests that neural specializations
exist for the control of orienting attention to locations or for cod-
ing some aspect of space tapped in spatial attention tasks [52]. But
a question that must be addressed is whether these superior frontal
and parietal regions are sensitive only for spatial orienting, and we
have tested this by investigating specializations in attentional con-
trol for other presumably nonspatial features.
We investigated the idea that specializations in superior frontal
and parietal cortex for preparatory spatial selective attention might
also be involved in other aspects of visual attentional processing. In
one study we compared preparatory attention for stimulus motion, a
dorsal stream process, to the same nonspatial ventral stream feature,
color used in Giesbrecht and colleagues [32]. In this study each trial
began with an auditory word cue that instructed subjects to attend to
a target of a particular stimulus feature (i.e., involving either color or
motion attention). If cued to a color, then they were to prepare for
and detect brief color flashes within a display of randomly moving
dots presented during a subsequent test period. If cued to motion,
then they were to prepare for and detect the brief coherent motion
stimulus in the display of randomly moving dots [53, 54].
Neuroimaging Approaches to the Study of Visual Attention 403
Fig. 8 Attentional control activations for location and motion cues versus color
cues. (Top) Results of the direct comparison between location and color cue
conditions overlaid on an axial slice. Greater activity to location cues than color
cues is seen in the superior frontal gyrus (SFG) and the superior parietal lobule
(SPL) bilaterally (modified from [32]). (Bottom) Results of the direct comparison
between motion and color cue conditions overlaid on an axial slice. Greater activ-
ity to motion cues than color cues is seen in the superior frontal gyrus (SFG)
bilaterally, but the superior parietal lobule (SPL) only on the left (modified from
[54])
404 George R. Mangun et al.
3.3 The Ventral So far we have focused our discussion of attentional control on the
Attention Network dorsal frontal and parietal cortex, the so-called dorsal attention
network. The idea is that top-down signals from this system exert
attentional control over perception based on behavioral goals and
strategies. But when attention must be reoriented from an attended
location to another location or object, a different network of brain
areas has been shown to be activated (Fig. 9). This network has
been called the ventral attention network, and involves the
temporal-parietal junction (TPJ), together with ventral frontal
regions including the insula, the portions of the inferior frontal
gyrus and middle frontal gyrus of the right hemisphere [41, 68,
69]. Evidence from various sources, such as resting-state fMRI
[70] indicates that the dorsal and ventral attention networks are
Neuroimaging Approaches to the Study of Visual Attention 405
Fig. 9 Diagram of the nodes in the dorsal attention network (blue), ventral atten-
tion network (orange), and visual cortex (light blue) in the left and right hemi-
spheres of the human brain. The dorsal system is clearly bilateral in organization,
but the ventral system is believed to be more right lateralized (noted by the dif-
ference in saturation of the orange-colored nodes). Putative connections within
each hemisphere are shown by the arrows. FEF frontal eye fields, IPS intrapari-
etal sulcus, VFC ventral frontal cortex, TPJ temporoparietal junction, V visual
cortex [76]
3.4 Attentional Models of attentional control, as laid out in the foregoing, hold
Control Mechanisms that activity in the sensory cortex is under top-down influences
Revealed from control areas in the frontoparietal cortex during voluntary
by Simultaneous attention [35, 41, 83]. Despite recent advancements in our under-
EEG-fMRI standing of visual selective attention, one unresolved issue about
the nature of top-down attentional control is whether attentional
biasing is achieved primarily by enhancing the sensitivity of task-
relevant sensory cortices or alternatively, via the inhibition of task-
irrelevant areas. Further, a related question is whether areas residing
outside of the sensory cortex but known to mediate other task-
independent processes, such as the default mode network [84], are
also “sites” that receive top–down modulation from the attentional
control system as part of goal-directed behavior.
408 George R. Mangun et al.
Fig. 10 Dynamic causal models of interactions between rTPJ and rFEF showing
driving inputs and intrinsic connections. (a) Model for the condition where the stimu-
lus included the target-colored distractor, which shows positive stimulus-driven
influences on rFEF (0.72) but not rTPJ (-0.22), and positive (.60) intrinsic connectivity
from FEF to TPJ: the only significant modulatory parameter (−0.19) was on the con-
nection from rFEF to rTPJ. (b) Model for the condition where there was a target
present (on the cued side). Here again the pattern is for a positive stimulus driven
input to rFEF, and a positive intrinsic connection from rFEF to rTPJ. These models are
consistent with FEF begin active first, followed by activity in TPJ
Fig. 11 Attentional modulation of EEG alpha and its coupling with attentional control structures. (a) Scalp topog-
raphy of alpha desynchronization showed that relative to a pre-cue baseline, alpha power was more strongly
suppressed over the hemisphere contralateral to the attended hemifield for both the attend-left (top left panel)
and attend-right (top-middle) conditions. The asymmetry in alpha desynchronization was further demonstrated
by the difference topography contrasting attend-left and attend-right (top-right). (b) BOLD activities in bilateral
intraparietal sulci (IPS) showed inverse coupling with alpha power measured on the hemispheres both contra-
lateral (bottom-left) and ipsilateral (bottom-right) to the attended hemifield (adapted from [85])
410 George R. Mangun et al.
Fig. 12 Regions showing positive coupling between BOLD and alpha power mea-
sured on hemispheres contralateral (a) and ipsilateral (b) to the attended hemi-
field. (c) A sagittal slice showing a region in the dorsal anterior cingulate cortex
(dACC) with BOLD being positively correlated with the alpha hemispheric lateral-
ization. MPFC medial prefrontal cortex, MTG middle temporal gyrus, postCG
post-central gyrus (adapted from [85])
4 Conclusions
References
1. James W (1890) Principles of psychology. H 10. Luck SJ, Hillyard SA, Mouloua M, Woldorff
Holt, New York MG, Clark VP, Hawkins HL (1994) Effects
2. Hv H (1867) Handbuch der physiologischen of spatial cuing on luminance detectability:
optik. Voss, Leipzig psychophysical and electrophysiological evi-
3. Posner MI, Cohen Y (1984) Components of dence for early selection. J Exp Psychol Hum
visual orienting. In: Bouma H, Bouwhis D Percept Perform 20:887–904
(eds) Attention and performance. Erlbaum, 11. Palmer J, Ames CT, Lindsey DT (1993)
Hillsdale, NJ, pp 531–556 Measuring the effect of attention on simple
4. Harter MR, Aine CJ (1984) Brain mecha- visual search. J Exp Psychol Hum Percept
nisms of visual selective attention. In: Perform 19:108–130
Parasuraman R, Davies DR (eds) Varieties of 12. Hernández-Peón R, Scherrer R, Jouvet M
attention. Academic, Orlando, pp 293–321 (1956) Modification of electric activity in
5. Malcolm GL, Shomstein S (2015) Object- cochlear nucleus during “attention” in
based attention in real-world scenes. J Exp unanesthetized cats. Science 123:331–332
Psychol Gen 144:257–263 13. Naatanen R (1975) Selective attention and
6. Broadbent DE (1958) Perception and com- evoked potentials in humans–a critical review.
munication. Pergamon, New York Biol Psychol 2:237–307
7. Deutsch JA, Deutsch D (1963) Attention – 14. Hillyard SA, Hink RF, Schwent VL, Picton
some theoretical considerations. Psychol Rev TW (1973) Electrical signs of selective atten-
70:80–90 tion in the human brain. Science
182:177–180
8. Johnston WA, Heinz SP (1979) Depth of
nontarget processing in an attention task. 15. Van Voorhis S, Hillyard SA (1977) Visual
J Exp Psychol Hum Percept Perform evoked potentials and selective attention to
5:168–175 points in space. Percept Psychophys
22:54–62
9. Hawkins HL, Hillyard SA, Luck SJ, Mouloua
M, Downing CJ, Woodward DP (1990) 16. Eason R, Harter M, White C (1969) Effects
Visual attention modulates signal detectabil- of attention and arousal on visually evoked
ity. J Exp Psychol Hum Percept Perform cortical potentials and reaction time in man.
16:802–811 Physiol Behav 4:283–289
414 George R. Mangun et al.
17. Spong P, Haider M, Lindsley DB (1965) 31. Hopfinger JB, Buonocore MH, Mangun GR
Selective attentiveness and cortical evoked (2000) The neural mechanisms of top-down
responses to visual and auditory stimuli. attentional control. Nat Neurosci 3:284–291
Science 148:395–397 32. Giesbrecht B, Woldorff MG, Song AW,
18. Nunez PL, Srinivasan R (2006) Electric Mangun GR (2003) Neural mechanisms of
fields of the brain: the neurophysics of EEG, top-down control during spatial and feature
2nd edn. Oxford University Press, Oxford, attention. Neuroimage 19:496–512
New York 33. Tootell RBH, Hadjikhani N, Hall EK et al
19. Moran J, Desimone R (1985) Selective atten- (1998) The retinotopy of visual spatial atten-
tion gates visual processing in the extrastriate tion. Neuron 21:1409–1422
cortex. Science 229:782–784 34. Desimone R, Duncan J (1995) Neural mech-
20. Chelazzi L, Miller EK, Duncan J, Desimone R anisms of selective visual-attention. Annu Rev
(1993) A neural basis for visual search in infe- Neurosci 18:193–222
rior temporal cortex. Nature 363:345–347 35. Posner MI, Petersen SE (1990) The atten-
21. Chelazzi L, Miller EK, Duncan J, Desimone tion system of the human brain. Annu Rev
R (2001) Responses of neurons in macaque Neurosci 13:25–42
area V4 during memory-guided visual search. 36. Gitelman DR, Nobre AC, Parrish TB et al
Cereb Cortex 11:761–772 (1999) A large-scale distributed network for
22. Luck SJ, Chelazzi L, Hillyard SA, Desimone R covert spatial attention: further anatomi-
(1997) Neural mechanisms of spatial selective cal delineation based on stringent behav-
attention in areas V1, V2, and V4 of macaque ioural and cognitive controls. Brain 122(Pt
visual cortex. J Neurophysiol 77:24–42 6):1093–1106
23. McAdams CJ, Reid RC (2005) Attention 37. Mesulam MM (1981) A cortical network for
modulates the responses of simple cells in directed attention and unilateral neglect. Ann
monkey primary visual cortex. J Neurosci Neurol 10:309–325
25:11023–11033 38. Posner MI, Snyder CR, Davidson BJ (1980)
24. Briggs F, Mangun GR, Usrey WM (2013) Attention and the detection of signals. J Exp
Attention enhances synaptic efficacy and the Psychol 109:160–174
signal-to-noise ratio in neural circuits. Nature 39. Harter MR, Miller SL, Price NJ, Lalonde ME,
499:476–480 Keyes AL (1989) Neural processes involved
25. McAlonan K, Cavanaugh J, Wurtz RH (2008) in directing attention. J Cogn Neurosci
Guarding the gateway to cortex with atten- 1:223–237
tion in visual thalamus. Nature 456:391–394 40. Corbetta M, Kincade JM, Ollinger JM,
26. Corbetta M, Miezin FM, Dobmeyer S, McAvoy MP, Shulman GL (2000) Voluntary
Shulman GL, Petersen SE (1991) Selective orienting is dissociated from target detec-
and divided attention during visual discrimi- tion in human posterior parietal cortex. Nat
nations of shape, color, and speed: functional Neurosci 3:292–297
anatomy by positron emission tomography. 41. Corbetta M, Shulman GL (2002) Control of
J Neurosci 11:2383–2402 goal-directed and stimulus-driven attention in
27. Heinze HJ, Mangun GR, Burchert W et al the brain. Nat Rev Neurosci 3:201–215
(1994) Combined spatial and temporal imag- 42. Corbetta M, Shulman GL (2011) Spatial
ing of brain activity during visual selective neglect and attention networks. Annu Rev
attention in humans. Nature 372:543–546 Neurosci 34(34):569–599
28. Mangun GR, Hopfinger JB, Kussmaul CL, 43. McMains SA, Fehd HM, Emmanouil TA,
Fletcher EM, Heinze HJ (1997) Covariations Kastner S (2007) Mechanisms of feature-
in ERP and PET measures of spatial selective and space-based attention: response modu-
attention in human extrastriate visual cortex. lation and baseline increases. J Neurophysiol
Hum Brain Mapp 5:273–279 98:2110–2121
29. Hillyard SA, Munte TF (1984) Selective 44. Kastner S, Pinsk MA, De Weerd P, Desimone
attention to color and location: an analysis R, Ungerleider LG (1999) Increased activ-
with event-related brain potentials. Percept ity in human visual cortex during directed
Psychophys 36:185–198 attention in the absence of visual stimulation.
30. Mangun GR, Hillyard SA (1991) Modulations Neuron 22:751–761
of sensory-evoked brain potentials indicate 45. Chawla D, Rees G, Friston KJ (1999) The
changes in perceptual processing during physiological basis of attentional modula-
visual-spatial priming. J Exp Psychol Hum tion in extrastriate visual areas. Nat Neurosci
Percept Perform 17:1057–1074 2:671–676
Neuroimaging Approaches to the Study of Visual Attention 415
46. Greenberg AS, Esterman M, Wilson D, 59. Cohen YE, Andersen RA (2002) A common
Serences JT, Yantis S (2010) Control of spa- reference frame for movement plans in the
tial and feature-based attention in frontopari- posterior parietal cortex. Nat Rev Neurosci
etal cortex. J Neurosci 30:14330–14339 3:553–562
47. Burock MA, Buckner RL, Woldorff MG, 60. Goodale MA, Milner AD (1992) Separate
Rosen BR, Dale AM (1998) Randomized visual pathways for perception and action.
event-related experimental designs allow for Trends Neurosci 15:20–25
extremely rapid presentation rates using func- 61. Goodale MA, Westwood DA (2004) An
tional MRI. Neuroreport 9:3735–3739 evolving view of duplex vision: separate
48. Ollinger JM, Corbetta M, Shulman GL but interacting cortical pathways for per-
(2001) Separating processes within a trial in ception and action. Curr Opin Neurobiol
event-related functional MRI – II analysis. 14:203–211
Neuroimage 13:218–229 62. Goodale MA (2014) How (and why) the
49. Ollinger JM, Shulman GL, Corbetta M visual control of action differs from visual per-
(2001) Separating processes within a trial ception. Proc Biol Sci 281:20140337
in event-related functional MRI - I. The 63. Ungerleider LG, Mishkin M (1982) Two cor-
method. Neuroimage 13:210–217 tical visual systems. In: Ingle DJ, Goodale MA,
50. Woldorff MG, Hazlett CJ, Fichtenholtz HM, Mansfield RJW (eds) Analysis of visual behav-
Weissman DH, Dale AM, Song AW (2004) ior. MIT Press, Cambridge, pp 549–586
Functional parcellation of attentional con- 64. Corbetta M, Tansy AP, Stanley CM, Astafiev
trol regions of the brain. J Cogn Neurosci SV, Snyder AZ, Shulman GL (2005) A func-
16:149–165 tional MRI study of preparatory signals for
51. Walsh BJ, Buonocore MH, Carter CS, spatial location and objects. Neuropsychologia
Mangun GR (2011) Integrating conflict 43:2041–2056
detection and attentional control mecha- 65. Moore T, Fallah M (2004) Microstimulation
nisms. J Cogn Neurosci 23:2211–2221 of the frontal eye field and its effects on covert
52. Corbetta M (1998) Frontoparietal cortical spatial attention. J Neurophysiol 91:152–162
networks for directing attention and the eye 66. Nobre AC, Sebestyen GN, Miniussi C
to visual locations: identical, independent, or (2000) The dynamics of shifting visuospatial
overlapping neural systems? Proc Natl Acad attention revealed by event-related potentials.
Sci U S A 95:831–838 Neuropsychologia 38:964–974
53. Fannon SP, Saron CD, Mangun GR (2007) 67. Astafiev SV, Shulman GL, Stanley CM, Snyder
Baseline shifts do not predict attentional modu- AZ, Van Essen DC, Corbetta M (2003)
lation of target processing during feature-based Functional organization of human intrapari-
visual attention. Front Hum Neurosci 1:7 etal and frontal cortex for attending, looking,
54. Mangun GR, Fannon SP (2007) Networks and pointing. J Neurosci 23:4689–4699
for attentional control and selection in spa- 68. Corbetta M, Patel G, Shulman GL (2008)
tial vision. In: Mast F, Jäncke L (eds) Spatial The reorienting system of the human brain:
processing in navigation, imagery and percep- from environment to theory of mind. Neuron
tion. Springer, New York, pp 411–432 58:306–324
55. Kastner S, Ungerleider LG (2000) 69. Shulman GL, Pope DLW, Astafiev SV,
Mechanisms of visual attention in the human McAvoy MP, Snyder AZ, Corbetta M (2010)
cortex. Annu Rev Neurosci 23:315–341 Right hemisphere dominance during spatial
56. Kincade JM, Abrams RA, Astafiev SV, Shulman selective attention and target detection occurs
GL, Corbetta M (2005) An event-related outside the dorsal frontoparietal network.
functional magnetic resonance imaging study J Neurosci 30:3640–3651
of voluntary and stimulus-driven orienting of 70. Fox MD, Corbetta M, Snyder AZ, Vincent
attention. J Neurosci 25:4593–4604 JL, Raichle ME (2006) Spontaneous neuro-
57. Wilson KD, Woldorff MG, Mangun GR nal activity distinguishes human dorsal and
(2005) Control networks and hemispheric ventral attention systems. Proc Natl Acad Sci
asymmetries in parietal cortex during atten- U S A 103:10046–10051
tional orienting in different spatial reference 71. Sestieri C, Pizzella V, Cianflone F, Romani
frames. Neuroimage 25:668–683 GL, Corbetta M (2008) Sequential activation
58. Geng JJ, Mangun GR (2009) Anterior of human oculomotor centers during planning
intraparietal sulcus is sensitive to bottom-up of visually guided eye movements: a combined
attention driven by stimulus salience. J Cogn fMRI-MEG study. Front Hum Neurosci
Neurosci 21:1584–1601 1:10.3389/neuro.3309/3001.2007
416 George R. Mangun et al.
72. Yamaguchi S, Knight RT (1991) Anterior 86. Worden MS, Foxe JJ, Wang N, Simpson GV
and posterior association cortex contribu- (2000) Anticipatory biasing of visuospatial
tions to the somatosensory P300. J Neurosci attention indexed by retinotopically specific
11:2039–2054 alpha-band electroencephalography increases
73. He BJ, Snyder AZ, Vincent JL, Epstein A, over occipital cortex. J Neurosci 20:RC63
Shulman GL, Corbetta M (2007) Breakdown 87. Sauseng P, Klimesch W, Stadler W et al (2005)
of functional connectivity in frontoparietal A shift of visual spatial attention is selectively
networks underlies behavioral deficits in spa- associated with human EEG alpha activity.
tial neglect. Neuron 53:905–918 Eur J Neurosci 22:2917–2926
74. Chica AB, Bartolomeo P, Valero-Cabré A 88. Thut G, Nietzel A, Brandt SA, Pascual-Leone
(2011) Dorsal and ventral parietal contribu- A (2006) Alpha-band electroencephalo-
tions to spatial orienting in the human brain. graphic activity over occipital cortex indexes
J Neurosci 31:8143–8149 visuospatial attention bias and predicts visual
75. Shulman GL, McAvoy MP, Cowan MC et al target detection. J Neurosci 26:9494–9502
(2003) Quantitative analysis of attention 89. Rajagovindan R, Ding M (2011) From pre-
and detection signals during visual search. stimulus alpha oscillation to visual-evoked
J Neurophysiol 90:3384–3397 response: an inverted-U function and its
76. Vossel S, Geng J, Fink GR (2014) Dorsal and attentional modulation. J Cogn Neurosci
ventral attention systems: distinct neural cir- 23:1379–1394
cuits but collaborative roles. Neuroscientist 90. Romei V, Brodbeck V, Michel C, Amedi
20:150–159 A, Pascual-Leone A, Thut G (2008)
77. Geng JJ, Vossel S (2013) Re-evaluating Spontaneous fluctuations in posterior alpha-
the role of TPJ in attentional control: con- band EEG activity reflect variability in excit-
textual updating? Neurosci Biobehav Rev ability of human visual areas. Cereb Cortex
37:2608–2620 18:2010–2018
78. DiQuattro NE, Sawaki R, Geng JJ (2014) 91. Romei V, Gross J, Thut G (2010) On the role
Effective connectivity during feature- of prestimulus alpha rhythms over occipito-
based attentional capture: evidence against parietal areas in visual input regulation: corre-
the attentional reorienting hypothesis of lation or causation? J Neurosci 30:8692–8697
TPJ. Cereb Cortex 24:3131–3141 92. Klimesch W, Sauseng P, Hanslmayr S (2007)
79. Friston KJ, Harrison L, Penny W (2003) EEG alpha oscillations: the inhibition-timing
Dynamic causal modelling. Neuroimage hypothesis. Brain Res Rev 53:63–88
19:1273–1302 93. Jensen O, Mazaheri A (2010) Shaping functional
80. Friston KJ, Li B, Daunizeau J, Stephan architecture by oscillatory alpha activity: gating
KE (2011) Network discovery with by inhibition. Front Hum Neurosci 4:186
DCM. Neuroimage 56:1202–1221 94. Haegens S, Osipova D, Oostenveld R, Jensen
81. Folk CL, Remington RW, Johnston JC (1992) O (2010) Somatosensory working memory
Involuntary covert orienting is contingent on performance in humans depends on both
attentional control settings. J Exp Psychol engagement and disengagement of regions
Hum Percept Perform 18:1030–1044 in a distributed network. Hum Brain Mapp
82. Serences JT, Shomstein S, Leber AB, Golay 31:26–35
X, Egeth HE, Yantis S (2005) Coordination 95. Anderson KL, Ding M (2011) Attentional
of voluntary and stimulus-driven atten- modulation of the somatosensory mu rhythm.
tional control in human cortex. Psychol Sci Neuroscience 180:165–180
16:114–122 96. Foxe JJ, Simpson GV, Ahlfors SP (1998)
83. Petersen SE, Posner MI (2012) The atten- Parieto-occipital approximately 10 Hz activ-
tion system of the human brain: 20 years after. ity reflects anticipatory state of visual atten-
Annu Rev Neurosci 35:73–89 tion mechanisms. Neuroreport 9:3929–3933
84. Buckner RL, Andrews-Hanna JR, Schacter 97. Fu KM, Foxe JJ, Murray MM, Higgins BA,
DL (2008) The brain’s default network: anat- Javitt DC, Schroeder CE (2001) Attention-
omy, function, and relevance to disease. Ann dependent suppression of distracter visual
N Y Acad Sci 1124:1–38 input can be cross-modally cued as indexed
85. Liu Y, Bengson J, Huang H, Mangun GR, by anticipatory parieto-occipital alpha-band
Ding M (2016) Top-down modulation of oscillations. Brain Res Cogn Brain Res
neural activity in anticipatory visual attention: 12:145–152
control mechanisms revealed by simultaneous 98. Bollimunta A, Chen Y, Schroeder CE, Ding
EEG-fMRI. Cereb Cortex 26:517–529 M (2008) Neuronal mechanisms of cortical
Neuroimaging Approaches to the Study of Visual Attention 417
fMRI of Memory
Federica Agosta, Indre V. Viskontas, Maria Luisa Gorno-Tempini,
and Massimo Filippi
Abstract
Numerous fMRI studies have investigated the network of brain regions critical for memory. Whereas
neuropsychological techniques can delineate the brain regions that are necessary for intact memory func-
tion, neuroimaging techniques can be used to investigate which regions are recruited during healthy
memory formation, storage, and retrieval. For example, fMRI studies have shown that lateral prefrontal
cortex (PFC) supports some components of working memory function. However, working memory is not
localized to a single brain region but is likely a property of the functional interaction between the PFC and
posterior brain regions. The medial temporal lobe (MTL) and its connections with neocortical, prefrontal,
and limbic structures are implicated in episodic memory. Semantic memory is mediated by a network of
neocortical structures, including lateral and anterior temporal lobes, and inferior frontal cortex, possibly to
a greater extent in the left hemisphere. Memory for semantic information benefits from the MTL for only
a limited time, and can be acquired, albeit slowly and with difficulty, without it. To date, most of the
emphasis has been on exploring the unique aspects of these different types of memory. Some evidence,
however, of functional overlap in general retrieval processes does exist.
Key words Working memory, Encoding, Retrieval, Episodic memory, Semantic memory, Prefrontal
cortex, Medial temporal lobe
1 Introduction
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_14, © Springer Science+Business Media New York 2016
419
420 Federica Agosta et al.
2 Working Memory
2.1 Organization The extensive reciprocal connections from the PFC to virtually all
of the WM Network cortical and subcortical structures place the PFC in an unique neu-
roanatomical position to monitor and manipulate diverse cognitive
processes [20]. Marklund et al. [28] employed a mixed block and
422 Federica Agosta et al.
2.1.1 Organization First, let us evaluate the evidence supporting anatomical segregation
of WM by Material Type on the basis of stimulus category. Three types of material have been
most commonly studied: verbal, spatial, and object information. A
prevalent theory of material-type segregation in the frontal cortex
suggests that there are dorsal and ventral memory streams for spatial
and object information, respectively, similarly to the “where” and
“what” pathways of the visual system [35]. The dorsal stream proj-
ects from the extrastriate cortex to the inferior parietal lobule (IPL)
and the IPS and is involved in processing spatial information [35].
The ventral stream extends from the extrastriate cortex to the inferior
surface of the frontal pole and processes object information [35].
Within the frontal cortex, WM for spatial information involves the
superior DL PFC or the superior frontal sulcus, whereas object WM
relies upon several mid- and inferior frontal regions (VL PFC) (for
review, see Ref. [3]). Furthermore, there is a tendency for verbal and
object WM to recruit more left-hemisphere areas, and for spatial tasks
to recruit more right-hemisphere areas (for review, see Ref. [3]).
The processing of nonverbal spatial information is right lateral-
ized and associated with the activation of a fronto-parietal network
(e.g., see Ref. [36–41]). Using event-related fMRI, Courtney et al.
[37] demonstrated a neuroanatomical dissociation between delay
period activity during WM maintenance for either the identity
(object memory) or location (spatial memory) of a set of three face
stimuli. Greater activity during the delay period on face identity trials
was observed in the left inferior frontal gyrus (IFG), whereas greater
activity during the delay period of the location task was observed in
dorsal frontal cortex (bilateral superior frontal sulcus) [37] (Fig. 1).
WM for objects, mostly visually presented faces, houses, and line
drawings that are not easily verbalizable, seems to be right-lateralized
and activates the temporal-occipital regions [Brodmann area (BA)
37] (e.g., see Refs. [21, 26, 29, 36, 39, 42–44]).
fMRI of Memory 423
a b Superior frontal
sulcus
L Superior
R frontal
Control stimulus set Memory set sulcus
3 Precentral
sulcus +32 m
4 m
Inferior
frontal
gyrus
5
6
1 unit
0 6 12 18 24 30 36 42 s
+7 mm
Fig. 1 An event-related fMRI study of delay period activity during working memory (WM) maintenance for
either the identity (object memory) or location (spatial memory) of a set of three face stimuli. (a) Schematic
depiction of the fMRI tasks: subjects saw a series of three faces, each presented for 2 s in a different location
on the screen, followed by a 9-s memory delay. Then a single test face appeared in some location on the
screen for 3 s, followed by a 6-s intertrial interval. Before each series, subjects were instructed to remember
the locations or the identities of the three faces in the memory set. For the spatial task, the subject indicated
with a left or right button whether the test location was the same as one of the three locations presented in the
memory set, regardless of the face that marked that location. For the face memory task, the subject indicated
whether the test face was the same as one of the three faces observed in the memory set, regardless of the
location where the face appeared. For the sensorimotor control task, scrambled faces appeared (control stimu-
lus set), and when the fourth scrambled picture appeared after the delay, subjects pressed both buttons
(control response). Contrasts between task components are shown below the task diagram: (1) visual stimula-
tion vs. no visual stimulation, (2) memory stimuli vs. control stimuli, (3) control stimulus set vs. control response,
(4) memory stimulus set vs. test stimulus and response, (5) delays during anticipation of response vs. intertrial
intervals, and (6) memory delays vs. control delays. (b) Areas with significant sustained recruitment in a single
subject during the WM delay for faces (blue outline) and for spatial locations (red outline) overlaid onto the
subject’s Talairach normalized anatomical MR image. Level above the bicommissural plane is indicated for
each axial section (from Courtney et al. [37])
activity in parietal areas, particularly the IPL (i.e., BAs 39 and 40)
[24, 45, 46, 49–53] but also in the superior parietal lobe [47, 49,
52, 53]. For instance, using an event-related design, Chein and Fiez
[45] designed a delayed serial recall task requiring subjects to encode,
maintain, and overtly recall sets of verbal items for which phonologi-
cal similarity, articulatory length, and lexical status were manipulated
and reported that Broca’s area and BA 40 showed patterns of sus-
tained activity during the delay period of a verbal WM task.
More recently, Raye et al. [48] provided evidence that left VL
PFC appears associated with subvocal rehearsal of single words,
whereas left DL PFC activation is associated with participants
“refreshing” (simply thinking about) the visual appearance of a
recently presented word [48].
Despite this growing evidence for segregation by material type,
there are also several human functional imaging studies that have
failed to find evidence for segregation in PFC WM activity (e.g., see
Refs. [16, 54, 55]). For instance, using an event-related design,
Postle and D’Esposito [16] evaluated the organization of WM for
the identity and location of visually presented stimuli (target stimuli
for all object trials were 16 abstract polygon stimuli, determined in
normative testing to be difficult to associate with real-world objects).
Although the task produced considerable delay-period activity in VL
PFC, DL PFC, and superior frontal cortex, in no subject, PFC activ-
ity was greater for one stimulus domain than for the other [16].
Moreover, in a large meta-analysis based on 60 neuroimaging (both
PET and fMRI) studies of WM [55], analyses of material type showed
the expected dorsal–ventral dissociation between spatial and nonspa-
tial storage in the posterior cortex, but not in the frontal lobe. Some
support was found for left frontal dominance in verbal WM, but only
for tasks with low executive demand. Executive demand increased
right lateralization in the frontal cortex for spatial WM [55].
2.1.2 Organization The other axis along which investigators have suggested that human
of WM by Process Type lateral PFC involvement in WM is segregated is according to the type
of operation performed upon the contents of WM, rather than the
type of information being maintained. In particular, several fMRI
studies have focused on the distinction between two fundamental
WM processes, namely the passive maintenance of information in
short-term memory and the active manipulation of this information,
within the PFC (for review, see Ref. [3]). This model received initial
support from a PET study by Owen et al. [56] in which dorsal PFC
activation was found during three spatial WM tasks thought to
require greater monitoring of remembered information (i.e., a mne-
monic variant of modified Tower of London planning task requiring
short-term retention and reproduction of problem solutions) than
two other WM tasks (i.e., the modified Tower of London planning
task, and a control condition that involved identical visual stimuli and
motor responses) that activated only ventral PFC.
fMRI of Memory 425
This model was also tested using fMRI (for review, see Ref. [3]).
The VL PFC has been suggested to be primarily involved only in
WM mechanisms that support simple retrieval of information for
sensory-guided sequential behavior (maintenance) [57–60], whereas
DL PFC (particularly BA 46) has been found to serve mechanisms
of active monitoring and manipulation of information (or general-
ized executive processing) in WM [61–63]. For instance, in an
event-related fMRI study [61], subjects were presented with two
types of trials in random order in which they were required to either
maintain a sequence of letters across a delay period or manipulate
(alphabetize) this sequence during the delay in order to respond cor-
rectly to a probe. The authors found that dorsal PFC activity was
greater in trials during which actively maintained information was
manipulated, providing further support for a process-specific PFC
organization. However, this functional PFC division has not been
consistently replicated and seems to vary with materials and difficulty
[55]. Moreover, this distinction is not compatible with evidence of
continuous DL PFC activity in tasks without any manipulation [32,
64]. A large meta-analysis [55] showed that tasks requiring execu-
tive processing generally produce more dorsal frontal activations
than do storage-only tasks, but not all executive processes show this
pattern. For instance, superior frontal cortex (BAs 6, 8, and 9)
responded most when WM must be continuously updated and when
memory for temporal order has to be maintained [55]. Right ventral
frontal cortex (BAs 10 and 47) responded more frequently with
demand for manipulation (including dual-task requirements or
mental operations) [55]. Posterior parietal cortex (BA 7) was found
to be involved in all types of executive functions [55].
2.2 Medial Temporal fMRI studies of WM have also found that the fronto-parietal net-
Lobe Involvement work is not the only region that is active during the temporary
in WM retention of task-relevant information. PFC and parietal cortex do
not seem to be sufficient to perform WM for novel stimuli when
parahippocampal regions are lesioned [65], although they are suf-
ficient to maintain normal WM for familiar stimuli [66].
Surprisingly, early fMRI studies of WM did not report activity
within parahippocampal regions such as perirhinal (PrC) or ento-
rhinal cortex [67]. An fMRI study by Stern et al. [68] demon-
strated differential activation for novel vs. familiar stimuli during
performance of a 2-back WM task. This study showed that WM for
a highly familiar set of complex visual images primarily activated
prefrontal and parietal cortices, whereas the same task using novel
(trial-unique) visual images strongly activated parahippocampal
structures in addition to prefrontal and parietal cortices. Activation
of parahippocampal structures associated with WM for novel stim-
uli has also been shown in an event-related fMRI study using novel
face stimuli [25, 69].
426 Federica Agosta et al.
a b
Delay Network: Right FFA seed
–12 –6 0 6 12 18
18 Right FFA Correlation Data
Cue
16 Delay
Probe
14
% ROI suprathreshold
12 24 30 36 42 48 54
10
6 L R
4
0
Right MFG Left MFG Right IFG Left IFG R L
Region of Interest left hemisphere rostral right hemisphere
t-value
7.23 18.0
Fig. 2 Functional connectivity analysis between brain regions associated with the maintenance of a represen-
tation of a visual stimulus (face recognition task) over a short delay interval. The fusiform face area (FFA) was
used as the exploratory seed. (a) Quantification of right FFA correlation effects in the prefrontal cortex (PFC).
Bars indicate the number of significant voxels present during each stage (cue, delay, and probe) in four pre-
frontal regions of interest. MFG middle frontal gyrus, IFG inferior frontal gyrus, ROI region of interest. (b) Delay
period right FFA seed correlation map including bilateral regions in the dorsolateral and ventrolateral PFC,
premotor cortex, IPS, caudate nucleus, thalamus, hippocampus, and occipitotemporal regions. Activations are
thresholded at p < 0.05 (corrected) and are shown overlaid on both axial slices and a three-dimensionally
rendered Montreal Neurological Institute (MNI) template brain. The color scale indicates the magnitude of t
values (from Gazzaley et al. [71])
fMRI of Memory 427
4 Episodic Memory
4.1 Distinguishing One of the major advances in the study of EM has been the appli-
Encoding and Retrieval cation of neuroimaging techniques to distinguish the component
Processes processes of encoding and retrieval. In neuropsychological studies,
it is often difficult or even impossible to separate failures of EM
428 Federica Agosta et al.
4.1.1 Encoding Simply perceiving and attending to information in the world is not
sufficient to create a lasting long-term memory trace. Some other
process or processes are engaged in order to bind elements of an
episode into a coherent memory trace. There is some disagreement
concerning whether all memories are initially episodic and later
become semantic, or whether some memories are semantic in
nature from the outset [97, 98]. One might argue that every learn-
ing episode is encoded as such, and as items from an episode get
integrated into the network of semantic information, those items
become disassociated from the episode itself and associated instead
with the information already in semantic memory [97]. In fact,
several memory models suggest that episodic and semantic infor-
mation is learned via different mechanisms: elements of an episode
are rapidly bound together by autoassociative processes in the den-
tate gyrus and CA3 field [99], while semantic information is grad-
ually acquired over many repetitions by reorganization of Hebbian
synapses in the neocortex [91, 100, 101].
Pioneering neuroimaging studies in the 1990s implicated the
PFC, particularly the left lateral PFC, in semantic or associative
encoding, by showing that this region, in addition to the MTL,
shows greater neural activity during semantic encoding than during
more superficial or perceptual encoding [102, 103]. Furthermore,
novel stimuli have been shown to elicit greater neural activity in the
MTL than familiar stimuli [104, 105], providing more evidence
for the involvement of the MTL in encoding processes.
A direct link to episodic encoding processes, however, required
the advent of event-related fMRI designs to investigate these cog-
nitive processes [106]. Encoding trials were binned according to
whether information presented on a given trial is subsequently
remembered or forgotten. Subsequent memory studies have shown
that activation in the left and right PFC, the parahippocampal
gyrus [107, 108], and hippocampus [109] during encoding pre-
dicts successful memory retrieval (Fig. 3). Since these original
studies, many more studies have replicated the finding that MTL
activity correlates with episodic encoding [110–113]. Furthermore,
fMRI studies have shown that hippocampal activity correlates with
the subsequent retrieval of contextual details presented during
encoding, while activity in the PrC correlates with successful
retrieval of an individual item, but not with retrieval of episodic
details [111–115]. Staresina and Davachi [116] have shown that
since the PrC receives inputs mainly from cortical areas devoted to
the processing of visual information, recruitment of this region
also correlates with the retrieval of visual features of items, such as
the color in which they were presented, but not with other
430 Federica Agosta et al.
Signal intensity
Signal intensity
1765
1495
1760
1490
1755 1485
1750 1480
1745 1475
Slice 1 Slice 2 0 5 10 15 20
0 5 10 15 20
Scan within trial Scan within trial
c 1440 Right parahippocampal
d 120 Remembered
Event-related response
Familiar
100 Forgotten
Signal intensity
1435
1430 80
60
1425
40
L Slice 3 Slice 4 R 1420
20
P< 0.01 P< 0.0005 1415 0
0 5 10 15 20 1 2 3 4 5 6
Scan within trial Subject
Fig. 3 Composite statistical activation maps displaying voxels with significant positive correlations between
event-related activations to pictures and subsequent memory for those pictures. Areas activated are right
dorsolateral prefrontal cortex (upper right in slice 1) and bilateral parahippocampal cortex (lower left in slices
1 and 3; left and right in slices 2 and 4). Examples of average signal magnitude during study from six subjects
in (a) right frontal (slice 1), (b) left parahippocampal (slice 4), and (c) right parahippocampal (slice 2) regions
for remembered (red), familiar (green), and forgotten (blue) pictures. Gray block depicts onset and offset of
picture presentation. (d) Mean voxel response in parahippocampal areas showing significant correlation with
subsequent memory in each subject for remembered, familiar, and forgotten pictures (from Brewer et al. [107])
PFC and Posterior Sensory Early neuroimaging work contrasting encoding and retrieval
Cortices processes revealed a consistent asymmetry in the activations of pre-
frontal regions [121, 122]: specifically, greater activation of the left
PFC during encoding, and relatively greater activation of the right
PFC during episodic retrieval [121, 122]. This pattern, termed
“hemispheric encoding-retrieval asymmetry” (HERA), provided
early insights into the neural basis of EM retrieval and a framework
for the design of future studies [122].
Further exploration of this relationship, however, demon-
strated that the HERA model was insufficient. The current view is
that HERA reflects the maintenance of a retrieval mode, or a back-
ground cognitive state in which one is mentally attuned to the
retrieval process, is sensitive to incoming cues, and is capable of
becoming consciously aware of successful retrieval. Lepage et al.
[123] compared data across four PET studies and concluded that
six PFC regions were consistently recruited by episodic recogni-
tion. These included bilateral posterior ventrolateral areas, bilateral
frontopolar regions, right dorsal PFC, and midline cingulate area
near the supplementary motor area. How and when these regions
are recruited in EM retrieval depends on task demands. As pro-
posed in the source-monitoring framework [124], retrieval
attempts differ in the strategies used to access different cortical
representations depending on the nature and source of those represen-
tations. For example, visuospatial, semantic, or emotional cues are
often called upon to aid in episodic retrieval, and each of these cues
stimulates a slightly different set of cortical regions.
Using event-related fMRI, Dobbins and Wagner [125] showed
that the recollection of conceptual or perceptual details of an epi-
sode results in greater activation in the left frontopolar and poste-
rior PFC than the detection of novelty. The authors interpret this
finding as an evidence that a domain-general control network is
engaged during contextual remembering. In contrast, left anterior
VL PFC coactivated with a left middle temporal region associated
with semantic representation, during conceptual recollection,
while right VL PFC and bilateral occipito-temporal cortices were
coactivated during recollection of perceptual details. Therefore,
whereas left frontopolar and posterior PFC may be involved in
domain-general retrieval processes, the middle temporal, right VL
PFC, and occipito-temporal regions may be more
domain-specific.
Interestingly, emerging data suggest that PFC is not involved in
distinguishing correct from incorrect retrieval. Dobbins et al. [126]
found that activation in the left MTL was greater for correct than
for incorrect source attributions, or episodic retrieval, but that
numerous PFC regions that showed increased activation during
source memory retrieval did not distinguish between correct and
incorrect trials. Moreover, activation in some of these regions was
actually numerically greater for failed retrieval attempts. The authors
432 Federica Agosta et al.
MTL Activation in Episodic Investigating MTL activation using fMRI can be difficult not only
Remembering because the region is susceptible to artifacts attributed to the ear
canal, but also because the region seems to be active during a wide
variety of tasks, including undirected “rest” [130]. Therefore,
many studies of episodic remembering do not report greater acti-
vation in the MTL, even though this region is known to be
involved. In order to observe MTL activity, a baseline task such as
an odd/even digit judgment may be used to deactivate the MTL
during the control trials [130].
Unlike data from the PFC, activation in the MTL has been
found to correlate with episodic retrieval success [131, 132]
(Fig. 4). Several neuroimaging studies investigating the role of
MTL subregions in EM have relied on the remember/know
procedure to distinguish episodic retrieval, or the processes of
“remembering or recollecting” (R), from retrieval based on famil-
iarity, called the “knowing or recalling” (K) [133]. This technique
was based on the finding that patients with MTL damage, espe-
cially those showing selective loss of hippocampal function, show
impairments in EM and a comparatively intact SM [80, 134]. In
addition, patients with degeneration of extrahippocampal temporal
cortex show the opposite pattern: impaired SM combined with a
relatively intact EM [78]. Using the remember/know procedure,
Eldridge et al. [131, 132] found that hippocampal activity is pri-
marily associated with “remembering” rather than “knowing”: the
hippocampus was more active during retrieval of “R” items than
fMRI of Memory 433
b
Left hippocampus Right hippocampus
fMRI signal (percent change)
0.4
Correct R
0.3 Correct K
0.2 Correct Rejection
Miss
0.1
0
–0.1
–0.2
–0.3
–0.4
0 5 10 15 0 5 10 15
Time (s)
Fig. 4 Results from anatomically defined hippocampal regions of interest. (a) Sections from the anatomical
template with the left hippocampal region of interest outlined in red. (b) Averaged event-related responses in
the hippocampus from 11 subjects. Error bars represent one standard error (between subjects) of estimated
response amplitudes. Correct R correct remember (R) response (when subjects could recollect the moment the
item was studied), Correct K correct know (K) response (when the word is familiar but unaccompanied by the
recollection of the specific moment the word was presented), Correct rejection correct response for nonrecog-
nized items, Miss miss response (in which subjects did not recognize old items) (from Eldridge et al. [131])
Parietal Cortex The medial parietal cortex exhibits opposite levels of fMRI activity
during encoding and retrieval, a pattern dubbed the encoding/
retrieval (E/R) flip. These opposing effects were originally reported
434 Federica Agosta et al.
by Daselaar et al. [136] who observed this pattern within the same
participants for a variety of stimuli and memory paradigms. Since
then, the E/R-flip pattern has been replicated in several other
studies (for review, see Ref. [137]). It has been shown that this pat-
tern occurs regardless of the type of information (words, faces,
spatial scenes), stimulus modality (auditory or visual), and memory
test (item or relational memory) (Fig. 5). Several hypotheses have
been formulated to explain the E/R-flip pattern (for review, see
Ref. [137]). The internal orienting account asserts that medial
parietal cortex involvement in encoding and retrieval is dependent
on the internal versus external orientation of attentional. The self-
referential processing account explains medial parietal cortex activity
in terms of orienting toward self-relevant thoughts versus the
external environment. The reallocation account states that the acti-
vation differences in medial parietal cortex depend on task-demands
and follows response times. Finally, the bottom-up attention account
asserts that activity within medial parietal regions reflects bottom-
up orienting of attention towards information retrieved from
memory. Yet none of these cognitive accounts seems to provide a
full explanation for the E/R-flip pattern in the PMC [137].
Fig. 5 Figure shows the encoding/retrieval flip in four different fMRI experi-
ments using (1) faces, (2) spatial scenes, (3) word pairs, and (4 and 5) single
words. During encoding, activity in the posterior midline region was greater for
misses (M) than for hits (H), whereas during retrieval, activity was greater for
hits than for misses. Bar graphs indicate mean cluster activity for the compari-
son between hits and misses during encoding and retrieval, respectively (from
Daselaar et al.[136])
436 Federica Agosta et al.
4.3 Summary Prior to the advent of fMRI, declarative memory, particularly EM,
of Episodic Memory was thought to be dependent almost exclusively on the MTL. Both
animal and human lesion studies provided the bulk of this evi-
dence. Neuroimaging techniques have demonstrated, however,
that PFC regions are recruited time and time again to support EM
retrieval. Interestingly, PFC does not seem to be sensitive to the
outcome of retrieval, but rather seems to support retrieval inten-
tion and strategy. The MTL, in contrast, seems to be driven largely
by “bottom-up” processes: being more data-driven and less volun-
tary [142]. Whereas details of an event are thought to be eventu-
ally stored in the same regions that were initially involved in their
perception, the hippocampus creates and stores an index of the
memory that can regenerate the original pattern of activation dur-
ing EM retrieval. Therefore, in healthy adults, EM involves a large
network of prefrontal, neocortical, and MTL regions acting in
concert to support this complex reconstructive process.
4.4 Future Directions The difference between EM and SM is the difference between the
active reconstruction of an event to extract information specific to
that occurrence and the abstraction of statistical regularities and
general properties about the world over multiple experiences. The
constructive nature of memory, therefore, is most easily observed
in EM: the reexperiencing of past events requires the reconstruc-
tion of a narrative sequence via the reactivation of stored sensory
information. A growing interest in the constructive aspects of EM
has led to the postulation of the constructive episodic simulation
hypothesis [143–145], which suggests that the EM system is built,
in part, to enable the simulation or imagination of future events.
Support for this view comes from several neuroimaging studies
demonstrating that the network of brain regions involved in EM
retrieval overlaps substantially with that supporting the imagina-
tion of future events [146–148]. Furthermore, as Hassabis and
Maguire [149, 150] have eloquently described, EM retrieval can
be thought of as relying heavily on scene construction as a key
component process. This approach may help explain why many
other cognitive tasks such as imagining a fictitious event seem to
employ many of the same regions involved in episodic
remembering.
5 Semantic Memory
system for “words and other verbal symbols, their meanings and ref-
erents, about relations among them, and about rules, formulas and
algorithms” [76], SM currently refers to a broader knowledge set
that includes facts, concepts, and beliefs [151].
Reports of patients with focal deficits in SM that were category-
specific [152–155] led to the hypothesis that SM is organized by
taxonomic categories [156]. In addition, a highly influential the-
ory proposed by Allport [157] suggests that the same sensorimo-
tor areas that are involved in the initial processing of experience are
also used to represent abstractions related to the experience. Much
like findings from neuroimaging studies of EM, this theory pre-
dicts that modality-specific information is represented in the cortex
that processes that modality. Finally, studies of patients with pro-
gressive neural degeneration leading to deficits in SM suggest that
information in SM is also organized hierarchically, such that the
more specific a concept is, the more vulnerable it may be to brain
damage [158]. Therefore, we will consider evidence from neuro-
imaging studies that address these three components of SM:
category specificity, reactivation of sensorimotor areas, and hierar-
chical organization.
5.1 Organization Before the advent of functional brain imaging, our knowledge of
of SM Network the neural bases of SM was dependent on studies of patients with
brain injury. Investigations of semantic impairment arising from
brain disease suggested that the anterior temporal lobes (ATLs)
are critical for semantic abilities in humans, across all stimulus
modalities and for all types of conceptual knowledge [154, 155,
159–165]. As mentioned earlier, patients with semantic demen-
tia and progressive degeneration of the anterior temporal cortex
are impaired on all tasks requiring knowledge about the mean-
ings of words, objects, and people, although possibly to different
degrees for each category depending on the lateralization of
atrophy [162, 164, 166, 167]. Other brain diseases that can
affect the ATLs, such as Alzheimer’s disease [168] and herpes
simplex viral encephalitis [155], also often disrupt SM. Finally, it
is worth noting that patients with damage to the left PFC often
have difficulty in retrieving words in response to specific cues,
even in the absence of aphasia [169].
Although ATLs’ activation has been associated with a few
semantic tasks (i.e., sentence comprehension, and famous face
naming or identification, e.g., see Refs. [170, 171]), the vast major-
ity of the functional imaging experiments on SM have reported
posterior temporal, typically stronger in the left than in the right
hemisphere, and/or frontal activations in the left VL PFC, with no
mention to the ATLs (for review, see Refs. [151, 172–175]).
Intersubject variability and the occurrence of fMRI susceptibility
artifacts in the ATLs may be possible reasons for the lack of fMRI
activations detected in ATLs [176].
438 Federica Agosta et al.
Fig. 6 A PET study investigating brain responses to famous and nonfamous faces and buildings. The results
showed category-specific effects in the right fusiform and bilateral parahippocampal/lingual gyri for faces and
buildings, respectively, but no effect of fame. In contrast, the left anterior middle temporal gyrus showed an
effect of fame for both faces and buildings, but no effect of category. (a) Examples of the stimuli used in experi-
ments 1 and 2 for the face conditions and in experiment 2 for the building conditions. (b) From top to bottom,
this figure illustrates areas of activation and parameter estimates for regions that were more activated for (a)
faces than for buildings, (b) buildings than for faces, (c) famous than for nonfamous faces and buildings, and
(d) famous than for nonfamous faces only. In the left column, all activations are superimposed on axial slices
of the mean of the nine subjects’ normalized structural MRIs and thresholded at p < 0.001 (uncorrected). In the
right column, the plots indicate the value of the normalized regional cerebral blood flow at the indicated voxel
(y-axis) for each of the experimental conditions in experiments 1 and 2 (x-axis). EXP experiment, FF famous
faces, NFF nonfamous faces, FB famous buildings, NFB nonfamous buildings (from Gorno-Tempini et al. [196])
5.3 Modality- Several of the early PET studies of semantic processes focused on
Specific the question of whether words (both auditory and visual) and pic-
Organization of SM tures are interpreted by a common semantic system [170, 188,
199, 200], or whether distinct systems are needed to support these
two domains. These studies reached similar conclusions providing
evidence for a distributed semantic system that is shared by visual/
auditory and verbal modalities [170, 188, 199, 200] and that it is
distributed throughout inferior temporal and frontal cortices with
a few areas uniquely activated by pictures only (left posterior
inferior temporal sulcus) or words (left anterior MTG and left infe-
rior frontal sulcus) [200]. More recently, fMRI studies have pro-
vided additional support for this view by demonstrating that
regions of the left posterior temporal cortex, known to be active
during conceptual processing of pictures and words (fusiform
gyrus and inferior and middle temporal gyri), are also active during
auditory sentence comprehension [201–203]: activity was modu-
lated by speech intelligibility [201, 202] and semantic ambiguity
[203]. A fMRI study used the phenomenon of semantic ambiguity
to identify regions within the fronto-temporal language network
that are involved in the processes of activating, selecting, and inte-
grating contextually appropriate word meanings [203]. Subjects
heard sentences containing ambiguous words (e.g., “The shell was
fired towards the tank”) and well-matched low-ambiguity sen-
tences (e.g., “Her secrets were written in her diary”). Although
these sentences had similar acoustic, phonological, syntactic, and
prosodic properties, the high-ambiguity sentences required addi-
tional processing by those brain regions involved in activating and
selecting contextually appropriate word meanings. The ambiguity
in these sentences went largely unnoticed, and yet high-ambiguity
sentences resulted in increased recruitment of the left posterior
inferior temporal cortex and the IFG bilaterally [203].
5.4 Representations While the neural systems involved in SM may be modulated both
of Object Properties by categories and modalities there is also evidence that the senso-
rimotor regions that are involved in the initial processing of par-
ticular information are recruited during SM retrieval (for review,
see Refs. [172, 173, 175, 177, 181, 204]). In particular, semantic
decisions involving object properties suggest a broad relationship
between perceptual knowledge retrieval and sensory brain mecha-
nisms, though recent work also suggests that this observation may
be itself category-specific [165].
Activation of the left or bilateral ventral temporal cortex (fusi-
form gyrus) when retrieving color information, relative to other prop-
erties, has been replicated by several fMRI studies [174, 205, 206].
Beauchamp et al. [205] showed that neural activity is limited to the
occipital lobes when color perception was tested by a passive viewing.
When the task was made more demanding by requiring subjects to
use color information to perform a color-sequencing task, several
442 Federica Agosta et al.
areas in the ventral cortex were identified: the most posterior, located
in the posterior fusiform gyrus, corresponded to an area activated by
passive viewing of colored stimuli, and more anterior and medial
color-selective areas located in the collateral sulcus and fusiform gyrus
[205]. These more anterior areas were also most active when visual
color information was behaviorally relevant, suggesting that atten-
tion influences activity in color-selective areas [205].
Parietal cortex appears to be involved in retrieval of size [174].
For instance, in Oliver and Thompson-Schill [174], seven subjects
made binary decisions about the shape, color, and size of named
objects during fMRI. Bilateral parietal activity was significantly greater
during retrieval of shape and size than during retrieval of color [174].
An area in the posterior superior temporal cortex, adjacent to
the auditory-association cortex, is activated when participants are
asked to judge the sound that an object makes [207]. Action knowl-
edge involves the left lateral temporal cortex, particularly the medial
and superior temporal (MT/MST) regions, anterior to an area
associated with motion perception [208–210]. For instance, in two
fMRI experiments Kourtzi and Kanwisher [209] found stronger
activation of the MT/MST regions during viewing of static photo-
graphs with implied motion compared with viewing of photographs
without it. Taken together, these data provide strong evidence that
information about a particular object property is stored in the same
neural system engaged when the property is perceived.
6 Conclusions
References
1. Baddeley A, Hitch G (1974) Working mem- imaging of human prefrontal cortex activa-
ory. The psychology of learning and motiva- tion during a spatial working memory task.
tion. B. GH Academic, San Diego, pp 47–90 Proc Natl Acad Sci U S A 91:8690–8694
2. Baddeley A (2000) The episodic buffer: a 15. Awh E, Jonides J, Smith EE, Buxton RB, Frank
new component of working memory? Trends LR, Love T, Wong EC et al (1999) Rehearsal
Cogn Sci 4:417–423 in spatial working memory: evidence from neu-
3. Curtis CE, D’Esposito, M (2006) Working roimaging. Psych Sci 10:433–437
memory. In: Cabeza R, Kingstone A (eds) 16. Postle BR, D’Esposito M (1999) “What”-
Handbook of functional neuroimaging of Then-Where” in visual working memory: an
cognition. MIT Press, Cambridge, MA. event-related fMRI study. J Cogn Neurosci
269–306 11:585–597
4. Miyake A, Shah P (1999) Models of work- 17. Rowe JB, Passingham RE (2001) Working
ing memory. Cambridge University Press, memory for location and time: activity in
New York prefrontal area 46 relates to selection rather
5. Miller GA (1956) The magical number seven, than maintenance in memory. Neuroimage
plus or minus two: some limits on our capac- 14:77–86
ity for processing information. Psych Rev 18. Corbetta M, Kincade JM, Shulman GL
63:81–97 (2002) Neural systems for visual orienting
6. Goldman-Rakic PS (1994) Working and their relationships to spatial working
memory dysfunction in schizophrenia. memory. J Cogn Neurosci 14:508–523
J Neuropsychiatry Clin Neurosci 6:348–357 19. Davachi L, Maril A, Wagner AD (2001)
7. Muller NG, Machado L, Knight RT (2002) When keeping in mind supports later bring-
Contributions of subregions of the prefron- ing to mind: neural markers of phonological
tal cortex to working memory: evidence from rehearsal predict subsequent remembering.
brain lesions in humans. J Cogn Neurosci J Cogn Neurosci 13:1059–1070
14:673–686 20. D’Esposito M (2008). Working memory. In:
8. Fuster JM, Alexander GE (1971) Neuron Goldenberg G, Miller B (eds) Handbook
activity related to short-term memory. Science of clinical neurology. Neuropsychology and
173:652–654 Behavioral Neurology, Vol 88. Elsevier,
9. Kubota K, Niki H (1971) Prefrontal cor- Amsterdam, The Netherlands. 237–247
tical unit activity and delayed alternation 21. Courtney SM, Ungerleider LG, Keil K, Haxby
performance in monkeys. J Neurophysiol JV (1997) Transient and sustained activity in
34:337–347 a distributed neural system for human work-
10. Funahashi S, Bruce CJ, Goldman-Rakic PS ing memory. Nature 386:608–611
(1989) Mnemonic coding of visual space in 22. Courtney SM, Petit L, Haxby JV, Ungerleider
the monkey’s dorsolateral prefrontal cortex. LG (1998) The role of prefrontal cortex in
J Neurophysiol 61:331–349 working memory: examining the contents of
11. Owen AM, Herrod NJ, Menon DK, Clark consciousness. Philos Trans R Soc Lond B
JC, Downey SP, Carpenter TA, Minhas Biol Sci 353:1819–1828
PS et al (1999) Redefining the functional 23. Jha AP, McCarthy G (2000) The influence of
organization of working memory processes memory load upon delay-interval activity in a
within human lateral prefrontal cortex. Eur working-memory task: an event-related func-
J Neurosci 11:567–574 tional MRI study. J Cogn Neurosci 12(Suppl
12. Jonides J, Smith EE, Koeppe RA, Awh E, 2):90–105
Minoshima S, Mintun MA (1993) Spatial 24. Gruber O (2001) Effects of domain-specific
working memory in humans as revealed by interference on brain activation associated
PET. Nature 363:623–625 with verbal working memory task perfor-
13. Petrides M, Alivisatos B, Meyer E, Evans AC mance. Cereb Cortex 11:1047–1055
(1993) Functional activation of the human 25. Ranganath C, D’Esposito M (2001) Medial
frontal cortex during the performance of ver- temporal lobe activity associated with active
bal working memory tasks. Proc Natl Acad maintenance of novel information. Neuron
Sci U S A 90:878–882 31:865–873
14. McCarthy G, Blamire AM, Puce A, Nobre 26. Postle BR, Druzgal TJ, D’Esposito M (2003)
AC, Bloch G, Hyder F, Goldman-Rakic P Seeking the neural substrates of visual work-
et al (1994) Functional magnetic resonance ing memory storage. Cortex 39:927–946
444 Federica Agosta et al.
27. Curtis CE, Rao VY, D’Esposito M (2004) 40. Walter H, Wunderlich AP, Blankenhorn M,
Maintenance of spatial and motor codes Schafer S, Tomczak R, Spitzer M, Gron G
during oculomotor delayed response tasks. (2003) No hypofrontality, but absence of
J Neurosci 24:3944–3952 prefrontal lateralization comparing verbal and
28. Marklund P, Fransson P, Cabeza R, Petersson spatial working memory in schizophrenia.
KM, Ingvar M, Nyberg L (2007) Sustained Schizophr Res 61:175–184
and transient neural modulations in prefron- 41. Leung HC, Seelig D, Gore JC (2004) The
tal cortex related to declarative long-term effect of memory load on cortical activity in
memory, working memory, and attention. the spatial working memory circuit. Cogn
Cortex 43:22–37 Affect Behav Neurosci 4:553–563
29. Druzgal TJ, D’Esposito M (2003) Dissecting 42. Courtney SM, Ungerleider LG, Keil K,
contributions of prefrontal cortex and fusi- Haxby JV (1996) Object and spatial visual
form face area to face working memory. working memory activate separate neural sys-
J Cogn Neurosci 15:771–784 tems in human cortex. Cereb Cortex 6:39–49
30. D’Esposito M, Aguirre GK, Zarahn E, Ballard 43. Druzgal TJ, D'Esposito M (2001) Activity in
D, Shin RK, Lease J (1998) Functional MRI fusiform face area modulated as a function of
studies of spatial and nonspatial working working memory load. Brain Res Cogn Brain
memory. Brain Res Cogn Brain Res 7:1–13 Res 10:355–364
31. D’Esposito M, Postle BR, Rypma B (2000) 44. Rama P, Sala JB, Gillen JS, Pekar JJ, Courtney
Prefrontal cortical contributions to working SM (2001) Dissociation of the neural systems
memory: evidence from event-related fMRI for working memory maintenance of ver-
studies. Exp Brain Res 133:3–11 bal and nonspatial visual information. Cogn
32. Curtis CE, D’Esposito M (2003) Persistent Affect Behav Neurosci 1:161–171
activity in the prefrontal cortex during work- 45. Chein JM, Fiez JA (2001) Dissociation of
ing memory. Trends Cogn Sci 7:415–423 verbal working memory system components
33. Passingham D, Sakai K (2004) The prefron- using a delayed serial recall task. Cereb Cortex
tal cortex and working memory: physiology 11:1003–1014
and brain imaging. Curr Opin Neurobiol 46. Gruber O, von Cramon DY (2003) The
14:163–168 functional neuroanatomy of human working
34. Sreenivasan KK, Curtis CE, D'Esposito M memory revisited. Evidence from 3-T fMRI
(2014) Revisiting the role of persistent neu- studies using classical domain-specific inter-
ral activity during working memory. Trends ference tasks. Neuroimage 19:797–809
Cogn Sci 18:82–89 47. Ravizza SM, Delgado MR, Chein JM, Becker
35. Ungerleider LG, Haxby JV (1994) ‘What’ JT, Fiez JA (2004) Functional dissociations
and ‘where’ in the human brain. Curr Opin within the inferior parietal cortex in verbal
Neurobiol 4:157–165 working memory. Neuroimage 22:562–573
36. McCarthy G, Puce A, Constable RT, Krystal 48. Raye CL, Johnson MK, Mitchell KJ, Greene
JH, Gore JC, Goldman-Rakic P (1996) EJ, Johnson MR (2007) Refreshing: a mini-
Activation of human prefrontal cortex during mal executive function. Cortex 43:135–145
spatial and nonspatial working memory tasks 49. Chen SH, Desmond JE (2005)
measured by functional MRI. Cereb Cortex Cerebrocerebellar networks during articula-
6:600–611 tory rehearsal and verbal working memory
37. Courtney SM, Petit L, Maisog JM, Ungerleider tasks. Neuroimage 24:332–338
LG, Haxby JV (1998) An area specialized for 50. Kirschen MP, Chen SH, Schraedley-Desmond
spatial working memory in human frontal P, Desmond JE (2005) Load- and practice-
cortex. Science 279:1347–1351 dependent increases in cerebro-cerebellar
38. Munk MH, Linden DE, Muckli L, activation in verbal working memory: an
Lanfermann H, Zanella FE, Singer W, Goebel fMRI study. Neuroimage 24:462–472
R (2002) Distributed cortical systems in 51. Smith EE, Jonides J (1999) Storage and exec-
visual short-term memory revealed by event- utive processes in the frontal lobes. Science
related functional magnetic resonance imag- 283:1657–1661
ing. Cereb Cortex 12:866–876 52. Henson RN, Burgess N, Frith CD (2000)
39. Sala JB, Rama P, Courtney SM (2003) Recoding, storage, rehearsal and grouping in
Functional topography of a distributed neu- verbal short-term memory: an fMRI study.
ral system for spatial and nonspatial infor- Neuropsychologia 38:426–440
mation maintenance in working memory. 53. Crottaz-Herbette S, Anagnoson RT, Menon
Neuropsychologia 41:341–356 V (2004) Modality effects in verbal work-
fMRI of Memory 445
ing memory: differential prefrontal and pari- cal course and experimental findings in
etal responses to auditory and visual stimuli. H.M. Semin Neurol 4:249–259
Neuroimage 21:340–351 67. Braver TS, Cohen JD, Nystrom LE, Jonides
54. Postle BR (2006) Working memory as an J, Smith EE, Noll DC (1997) A paramet-
emergent property of the mind and brain. ric study of prefrontal cortex involvement
Neuroscience 139:23–38 in human working memory. Neuroimage
55. Wager TD, Smith EE (2003) Neuroimaging 5:49–62
studies of working memory: a meta-analysis. 68. Stern CE, Sherman SJ, Kirchhoff BA,
Cogn Affect Behav Neurosci 3:255–274 Hasselmo ME (2001) Medial temporal and
56. Owen AM, Doyon J, Petrides M, Evans AC prefrontal contributions to working mem-
(1996) Planning and spatial working mem- ory tasks with novel and familiar stimuli.
ory: a positron emission tomography study in Hippocampus 11:337–346
humans. Eur J Neurosci 8:353–364 69. Ranganath C, Rainer G (2003) Neural mech-
57. Thompson-Schill SL, D'Esposito M, Aguirre anisms for detecting and remembering novel
GK, Farah MJ (1997) Role of left inferior pre- events. Nat Rev Neurosci 4:193–202
frontal cortex in retrieval of semantic knowl- 70. Honey GD, Fu CH, Kim J, Brammer MJ,
edge: a reevaluation. Proc Natl Acad Sci U S Croudace TJ, Suckling J, Pich EM et al (2002)
A 94:14792–14797 Effects of verbal working memory load on
58. Thompson-Schill SL, D’Esposito M, Kan IP corticocortical connectivity modeled by path
(1999) Effects of repetition and competition analysis of functional magnetic resonance
on activity in left prefrontal cortex during imaging data. Neuroimage 17:573–582
word generation. Neuron 23:513–522 71. Gazzaley A, Rissman J, D’Esposito M (2004)
59. D’Esposito M, Postle BR, Jonides J, Smith Functional connectivity during working
EE (1999) The neural substrate and tempo- memory maintenance. Cogn Affect Behav
ral dynamics of interference effects in work- Neurosci 4:580–599
ing memory as revealed by event-related 72. Crowe DA, Goodwin SJ, Blackman RK,
functional MRI. Proc Natl Acad Sci U S A Sakellaridi S, Sponheim SR, MacDonald AW
96:7514–7519 3rd, Chafee MV (2013) Prefrontal neurons
60. Jonides J, Smith EE, Marshuetz C, Koeppe transmit signals to parietal neurons that reflect
RA, Reuter-Lorenz PA (1998) Inhibition executive control of cognition. Nat Neurosci
in verbal working memory revealed by 16:1484–1491
brain activation. Proc Natl Acad Sci U S A 73. Edin F, Klingberg T, Johansson P, McNab F,
95:8410–8413 Tegner J, Compte A (2009) Mechanism for
61. D’Esposito M, Postle BR, Ballard D, Lease top-down control of working memory capac-
J (1999) Maintenance versus manipulation ity. Proc Natl Acad Sci U S A 106:6802–6807
of information held in working memory: 74. Chadick JZ, Gazzaley A (2011) Differential
an event-related fMRI study. Brain Cogn coupling of visual cortex with default or
41:66–86 frontal-parietal network based on goals. Nat
62. Postle BR, Berger JS, D'Esposito M (1999) Neurosci 14:830–832
Functional neuroanatomical double disso- 75. Hazy TE, Frank MJ, C O’Reilly R (2007)
ciation of mnemonic and executive control Towards an executive without a homunculus:
processes contributing to working memory computational models of the prefrontal cor-
performance. Proc Natl Acad Sci U S A tex/basal ganglia system. Philos Trans R Soc
96:12959–12964 Lond B Biol Sci 362:1601–1613
63. Bunge SA, Klingberg T, Jacobsen RB, 76. Tulving E (1972) Episodic and semantic
Gabrieli JD (2000) A resource model of the memory. Organisation of memory (Tulving E
neural basis of executive working memory. and Donaldson W (eds). Academic,
Proc Natl Acad Sci U S A 97:3573–3578 New York, pp 381–403
64. Cohen JD, Perlstein WM, Braver TS, 77. Vargha-Khadem F, Gadian DG, Watkins KE,
Nystrom LE, Noll DC, Jonides J, Smith EE Connelly A, Van Paesschen W, Mishkin M
(1997) Temporal dynamics of brain activa- (1997) Differential effects of early hippocam-
tion during a working memory task. Nature pal pathology on episodic and semantic mem-
386:604–608 ory. Science 277:376–380
65. Squire LR, Stark CE, Clark RE (2004) The 78. Hodges JR, Graham KS (2001) Episodic mem-
medial temporal lobe. Annu Rev Neurosci ory: insights from semantic dementia. Philos
27:279–306 Trans R Soc Lond B Biol Sci 356:1423–1434
66. Corkin S (1984) Lasting consequences of 79. Levine B, Black SE, Cabeza R, Sinden M,
bilateral medial temporal lobectomy: clini- McIntosh AR, Toth JP, Tulving E et al (1998)
446 Federica Agosta et al.
Episodic memory and the self in a case of iso- 94. Amaral DG, Insausti R, Cowan WM (1987)
lated retrograde amnesia. Brain 121(Pt The entorhinal cortex of the monkey:
10):1951–1973 I. Cytoarchitectonic organization. J Comp
80. Viskontas IV, McAndrews MP, Moscovitch M Neurol 264:326–355
(2000) Remote episodic memory deficits in 95. Moscovitch M, Rosenbaum RS, Gilboa A,
patients with unilateral temporal lobe epilepsy Addis DR, Westmacott R, Grady C,
and excisions. J Neurosci 20:5853–5857 McAndrews MP et al (2005) Functional neu-
81. Dudukovic NM, Knowlton BJ (2006) roanatomy of remote episodic, semantic and
Remember-know judgments and retrieval of spatial memory: a unified account based on
contextual details. Acta Psychol (Amst) multiple trace theory. J Anat 207:35–66
122:160–173 96. Teyler TJ, Rudy JW (2007) The hippocampal
82. Levine B, Svoboda E, Hay JF, Winocur G, indexing theory and episodic memory: updat-
Moscovitch M (2002) Aging and autobio- ing the index. Hippocampus 17:1158–1169
graphical memory: dissociating episodic from 97. Squire LR, Knowlton B, Musen G (1993)
semantic retrieval. Psychol Aging The structure and organization of memory.
17:677–689 Annu Rev Psychol 44:453–495
83. Bayley PJ, Hopkins RO, Squire LR (2006) 98. Baddeley A, Vargha-Khadem F, Mishkin M
The fate of old memories after medial tempo- (2001) Preserved recognition in a case of
ral lobe damage. J Neurosci developmental amnesia: implications for the
26:13311–13317 acquisition of semantic memory? J Cogn
84. Nolde SF, Johnson MK, D’Esposito M Neurosci 13:357–369
(1998) Left prefrontal activation during epi- 99. Rolls ET (1996) A theory of hippocampal
sodic remembering: an event-related fMRI function in memory. Hippocampus
study. Neuroreport 9:3509–3514 6:601–620
85. Wheeler ME, Petersen SE, Buckner RL 100. Gluck MA, Myers CE (1993) Hippocampal
(2000) Memory’s echo: vivid remembering mediation of stimulus representation: a com-
reactivates sensory-specific cortex. Proc Natl putational theory. Hippocampus 3:491–516
Acad Sci U S A 97:11125–11129 101. Norman KA, O'Reilly RC (2003) Modeling
86. Heil M, Rosler F, Hennighausen E (1996) hippocampal and neocortical contributions to
Topographically distinct cortical activation in recognition memory: a complementary-
episodic long-term memory: the retrieval of learning-systems approach. Psychol Rev
spatial versus verbal information. Mem 110:611–646
Cognit 24:777–795 102. Fletcher PC, Shallice T, Dolan RJ (1998) The
87. O’Craven KM, Kanwisher N (2000) Mental functional roles of prefrontal cortex in epi-
imagery of faces and places activates corre- sodic memory. I. Encoding. Brain 121(Pt
sponding stiimulus-specific brain regions. 7):1239–1248
J Cogn Neurosci 12:1013–1023 103. Shallice T, Fletcher P, Frith CD, Grasby P,
88. Halpern AR, Zatorre RJ (1999) When that Frackowiak RS, Dolan RJ (1994) Brain
tune runs through your head: a PET investi- regions associated with acquisition and
gation of auditory imagery for familiar melo- retrieval of verbal episodic memory. Nature
dies. Cereb Cortex 9:697–704 368:633–635
89. Nyberg L, Petersson KM, Nilsson LG, 104. Dolan RJ, Fletcher PC (1997) Dissociating
Sandblom J, Aberg C, Ingvar M (2001) prefrontal and hippocampal function in epi-
Reactivation of motor brain areas during explicit sodic memory encoding. Nature
memory for actions. Neuroimage 14:521–528 388:582–585
90. Marr D (1971) Simple memory: a theory for 105. Gabrieli JD, Brewer JB, Desmond JE, Glover
archicortex. Philos Trans R Soc Lond B Biol GH (1997) Separate neural bases of two fun-
Sci 262:23–81 damental memory processes in the human
91. Hasselmo ME, McClelland JL (1999) Neural medial temporal lobe. Science 276:264–266
models of memory. Curr Opin Neurobiol 106. Wagner AD, Schacter DL, Rotte M, Koutstaal
9:184–188 W, Maril A, Dale AM, Rosen BR et al (1998)
92. Aggleton JP, Brown MW (1999) Episodic Building memories: remembering and forget-
memory, amnesia, and the hippocampal- ting of verbal experiences as predicted by
anterior thalamic axis. Behav Brain Sci brain activity. Science 281:1188–1191
22:425–444, discussion 444-489 107. Brewer JB, Zhao Z, Desmond JE, Glover
93. Insausti R, Amaral DG, Cowan WM (1987) GH, Gabrieli JD (1998) Making memories:
The entorhinal cortex of the monkey: brain activity that predicts how well visual
II. Cortical afferents. J Comp Neurol experience will be remembered. Science
264:356–395 281:1185–1187
fMRI of Memory 447
108. Wagner AD, Poldrack RA, Eldridge LL, 120. Otten LJ, Rugg MD (2001) Task-dependency
Desmond JE, Glover GH, Gabrieli JD (1998) of the neural correlates of episodic encoding
Material-specific lateralization of prefrontal as measured by fMRI. Cereb Cortex
activation during episodic encoding and 11:1150–1160
retrieval. Neuroreport 9:3711–3717 121. Nyberg L, McIntosh AR, Cabeza R, Habib R,
109. Fernandez G, Weyerts H, Schrader-Bolsche Houle S, Tulving E (1996) General and spe-
M, Tendolkar I, Smid HG, Tempelmann C, cific brain regions involved in encoding and
Hinrichs H et al (1998) Successful verbal retrieval of events: what, where, and when.
encoding into episodic memory engages the Proc Natl Acad Sci U S A 93:11280–11285
posterior hippocampus: a parametrically ana- 122. Tulving E, Kapur S, Craik FI, Moscovitch M,
lyzed functional magnetic resonance imaging Houle S (1994) Hemispheric encoding/
study. J Neurosci 18:1841–1847 retrieval asymmetry in episodic memory: pos-
110. Davachi L, Wagner AD (2002) Hippocampal itron emission tomography findings. Proc
contributions to episodic encoding: insights Natl Acad Sci U S A 91:2016–2020
from relational and item-based learning. 123. Lepage M, Ghaffar O, Nyberg L, Tulving E
J Neurophysiol 88:982–990 (2000) Prefrontal cortex and episodic mem-
111. Davachi L, Mitchell JP, Wagner AD (2003) ory retrieval mode. Proc Natl Acad Sci U S A
Multiple routes to memory: distinct medial 97:506–511
temporal lobe processes build item and source 124. Johnson MK, Hashtroudi S, Lindsay DS
memories. Proc Natl Acad Sci U S A (1993) Source monitoring. Psychol Bull
100:2157–2162 114:3–28
112. Kirwan CB, Stark CE (2004) Medial tempo- 125. Dobbins IG, Wagner AD (2005) Domain-
ral lobe activation during encoding and general and domain-sensitive prefrontal
retrieval of novel face-name pairs. mechanisms for recollecting events and
Hippocampus 14:919–930 detecting novelty. Cereb Cortex
113. Ranganath C, Yonelinas AP, Cohen MX, Dy 15:1768–1778
CJ, Tom SM, D'Esposito M (2004) 126. Dobbins IG, Rice HJ, Wagner AD, Schacter
Dissociable correlates of recollection and DL (2003) Memory orientation and success:
familiarity within the medial temporal lobes. separable neurocognitive components under-
Neuropsychologia 42:2–13 lying episodic recognition. Neuropsychologia
114. Kensinger EA, Schacter DL (2006) Amygdala 41:318–333
activity is associated with the successful 127. Velanova K, Jacoby LL, Wheeler ME, McAvoy
encoding of item, but not source, informa- MP, Petersen SE, Buckner RL (2003)
tion for positive and negative stimuli. Functional-anatomic correlates of sustained
J Neurosci 26:2564–2570 and transient processing components engaged
115. Uncapher MR, Otten LJ, Rugg MD (2006) during controlled retrieval. J Neurosci
Episodic encoding is more than the sum of its 23:8460–8470
parts: an fMRI investigation of multifeatural 128. Kahn I, Davachi L, Wagner AD (2004)
contextual encoding. Neuron 52:547–556 Functional-neuroanatomic correlates of rec-
116. Staresina BP, Davachi L (2008) Selective and ollection: implications for models of recogni-
shared contributions of the hippocampus and tion memory. J Neurosci 24:4172–4180
perirhinal cortex to episodic item and associa- 129. Wheeler ME, Buckner RL (2003) Functional
tive encoding. J Cogn Neurosci dissociation among components of remem-
20:1478–1489 bering: control, perceived oldness, and con-
117. Baker JT, Sanders AL, Maccotta L, Buckner tent. J Neurosci 23:3869–3880
RL (2001) Neural correlates of verbal mem- 130. Stark CE, Squire LR (2001) When zero is not
ory encoding during semantic and structural zero: the problem of ambiguous baseline con-
processing tasks. Neuroreport ditions in fMRI. Proc Natl Acad Sci U S A
12:1251–1256 98:12760–12766
118. Buckner RL, Wheeler ME, Sheridan MA 131. Eldridge LL, Knowlton BJ, Furmanski CS,
(2001) Encoding processes during retrieval Bookheimer SY, Engel SA (2000)
tasks. J Cogn Neurosci 13:406–415 Remembering episodes: a selective role for
119. Henson RN, Hornberger M, Rugg MD the hippocampus during retrieval. Nat
(2005) Further dissociating the processes Neurosci 3:1149–1152
involved in recognition memory: an FMRI 132. Wheeler ME, Buckner RL (2004) Functional-
study. J Cogn Neurosci 17:1058–1073 anatomic correlates of remembering and
knowing. Neuroimage 21:1337–1349
448 Federica Agosta et al.
133. Tulving E (1985) Memory and conscious- future: common and distinct neural substrates
ness. Can Psychol 1:1–12 during event construction and elaboration.
134. Tulving E, Schacter DL, McLachlan DR, Neuropsychologia 45:1363–1377
Moscovitch M (1988) Priming of semantic 148. Szpunar KK, Watson JM, McDermott KB
autobiographical knowledge: a case study of (2007) Neural substrates of envisioning the
retrograde amnesia. Brain Cogn 8:3–20 future. Proc Natl Acad Sci U S A
135. Eldridge LL, Engel SA, Zeineh MM, 104:642–647
Bookheimer SY, Knowlton BJ (2005) A dis- 149. Hassabis D, Maguire EA (2007)
sociation of encoding and retrieval processes Deconstructing episodic memory with con-
in the human hippocampus. J Neurosci struction. Trends Cogn Sci 11:299–306
25:3280–3286 150. Hassabis D, Maguire EA (2009) The con-
136. Daselaar SM, Prince SE, Dennis NA, Hayes struction system of the brain. Philos Trans R
SM, Kim H, Cabeza R (2009) Posterior mid- Soc Lond B Biol Sci 364:1263–1271
line and ventral parietal activity is associated 151. Martin A (2001) Functional neuroimaging of
with retrieval success and encoding failure. semantic memory, Handbook of functional
Front Hum Neurosci 3:13 neuroimaging of cognition (Cabeza R and
137. Huijbers W, Vannini P, Sperling RA, Pennartz Kingstone A (eds) ). The MIT Press,
CM, Cabeza R, Daselaar SM (2012) Cambridge, pp 153–186
Explaining the encoding/retrieval flip: 152. Damasio AR, McKee J, Damasio H (1979)
memory-related deactivations and activations Determinants of performance in color ano-
in the posteromedial cortex. Neuropsychologia mia. Brain Lang 7:74–85
50:3764–3774 153. McKenna P, Warrington EK (1980) Testing
138. Konishi S, Uchida I, Okuaki T, Machida T, for nominal dysphasia. J Neurol Neurosurg
Shirouzu I, Miyashita Y (2002) Neural cor- Psychiatry 43:781–788
relates of recency judgment. J Neurosci 154. Warrington EK, McCarthy R (1983)
22:9549–9555 Category specific access dysphasia. Brain
139. Milner B, Corsi P, Leonard G (1991) Frontal- 106(Pt 4):859–878
lobe contribution to recency judgements. 155. Warrington EK, Shallice T (1984) Category
Neuropsychologia 29:601–618 specific semantic impairments. Brain 107(Pt
140. Petrides M (1991) Functional specialization 3):829–854
within the dorsolateral frontal cortex for serial 156. Caramazza A, Shelton JR (1998) Domain-
order memory. Proc Biol Sci 246:299–306 specific knowledge systems in the brain the
141. Suzuki M, Fujii T, Tsukiura T, Okuda J, animate-inanimate distinction. J Cogn
Umetsu A, Nagasaka T, Mugikura S et al Neurosci 10:1–34
(2002) Neural basis of temporal context 157. Allport DA (1985) Distributed memory,
memory: a functional MRI study. Neuroimage modular systems and dysphagia, Current per-
17:1790–1796 spectives in dysphagia (Newman SK and
142. Moscovitch M (1995) Recovered conscious- Epstein R (eds)). Churchill Livingstone,
ness: a hypothesis concerning modularity and Edinburgh, pp 32–60
episodic memory. J Clin Exp Neuropsychol 158. Hodges JR, Graham N, Patterson K (1995)
17:276–290 Charting the progression in semantic demen-
143. Schacter DL, Addis DR (2007) The cognitive tia: implications for the organisation of
neuroscience of constructive memory: semantic memory. Memory 3:463–495
remembering the past and imagining the 159. Warrington EK (1975) The selective impair-
future. Philos Trans R Soc Lond B Biol Sci ment of semantic memory. Q J Exp Psychol
362:773–786 27:635–657
144. Dudai Y, Carruthers M (2005) The Janus face 160. Warrington EK, McCarthy RA (1987)
of Mnemosyne. Nature 434:567 Categories of knowledge. Further fraction-
145. Suddendorf T, Corballis MC (1997) Mental time ations and an attempted integration. Brain
travel and the evolution of the human mind. 110(Pt 5):1273–1296
Genet Soc Gen Psychol Monogr 123:133–167 161. Hart J Jr, Gordon B (1990) Delineation of
146. Okuda J, Fujii T, Ohtake H, Tsukiura T, Tanji single-word semantic comprehension deficits
K, Suzuki K, Kawashima R et al (2003) in aphasia, with anatomical correlation. Ann
Thinking of the future and past: the roles of Neurol 27:226–231
the frontal pole and the medial temporal 162. Hodges JRPK, Oxbury S, Funnell F (1992)
lobes. Neuroimage 19:1369–1380 Semantic dementia. Progressive fluent aphasia
147. Addis DR, Wong AT, Schacter DL (2007) with temporal lobe atrophy. Brain
Remembering the past and imagining the 115:1783–1806
fMRI of Memory 449
163. Gainotti G (2000) What the locus of brain comparing PET and fMRI on a semantic task.
lesion tells us about the nature of the cogni- Neuroimage 11:589–600
tive defect underlying category-specific disor- 177. Martin A, Chao LL (2001) Semantic memory
ders: a review. Cortex 36:539–559 and the brain: structure and processes. Curr
164. Gorno-Tempini ML, Dronkers NF, Rankin Opin Neurobiol 11:194–201
KP, Ogar JM, Phengrasamy L, Rosen HJ, 178. Caramazza A, Mahon BZ (2003) The organi-
Johnson JK et al (2004) Cognition and anat- zation of conceptual knowledge: the evidence
omy in three variants of primary progressive from category-specific semantic deficits.
aphasia. Ann Neurol 55:335–346 Trends Cogn Sci 7:354–361
165. Brambati SM, Myers D, Wilson A, Rankin 179. Thompson-Schill SL, Aguirre GK, D'Esposito
KP, Allison SC, Rosen HJ, Miller BL et al M, Farah MJ (1999) A neural basis for cate-
(2006) The anatomy of category-specific gory and modality specificity of semantic
object naming in neurodegenerative diseases. knowledge. Neuropsychologia 37:671–676
J Cogn Neurosci 18:1644–1653 180. Damasio AR (1989) Time-locked multire-
166. Snowden JS, Thompson JC, Neary D (2004) gional retroactivation: a systems-level pro-
Knowledge of famous faces and names in posal for the neural substrates of recall and
semantic dementia. Brain 127:860–872 recognition. Cognition 33:25–62
167. Gorno-Tempini ML, Rankin KP, Woolley JD, 181. Thompson-Schill SL (2003) Neuroimaging
Rosen HJ, Phengrasamy L, Miller BL (2004) studies of semantic memory: inferring “how”
Cognitive and behavioral profile in a case of from “where”. Neuropsychologia 41:280–292
right anterior temporal lobe neurodegenera- 182. Tyler LK, Moss HE (2001) Towards a dis-
tion. Cortex 40:631–644 tributed account of conceptual knowledge.
168. Hodges JR, Patterson K (1995) Is semantic Trends Cogn Sci 5:244–252
memory consistently impaired early in the 183. Hillis AE, Caramazza A (1991) Category-
course of Alzheimer's disease? specific naming and comprehension impair-
Neuroanatomical and diagnostic implications. ment: a double dissociation. Brain 114(Pt
Neuropsychologia 33:441–459 5):2081–2094
169. Baldo JV, Shimamura AP (1998) Letter and 184. Tranel D, Damasio H, Damasio AR (1997) A
category fluency in patients with frontal lobe neural basis for the retrieval of conceptual
lesions. Neuropsychology 12:259–267 knowledge. Neuropsychologia 35:1319–1327
170. Gorno-Tempini ML, Price CJ, Josephs O, 185. Martin A, Wiggs CL, Ungerleider LG, Haxby
Vandenberghe R, Cappa SF, Kapur N, JV (1996) Neural correlates of category-
Frackowiak RS (1998) The neural systems specific knowledge. Nature 379:649–652
sustaining face and proper-name processing.
Brain 121(Pt 11):2103–2118 186. Mummery CJ, Patterson K, Hodges JR, Wise
RJ (1996) Generating ‘tiger’ as an animal name
171. Damasio H, Grabowski TJ, Tranel D, Hichwa or a word beginning with T: differences in
RD, Damasio AR (1996) A neural basis for brain activation. Proc Biol Sci 263:989–995
lexical retrieval. Nature 380:499–505
187. Perani D, Cappa SF, Bettinardi V, Bressi S,
172. Bookheimer S (2002) Functional MRI of lan- Gorno-Tempini M, Matarrese M, Fazio F
guage: new approaches to understanding the (1995) Different neural systems for the rec-
cortical organization of semantic processing. ognition of animals and man-made tools.
Annu Rev Neurosci 25:151–188 Neuroreport 6:1637–1641
173. Martin A (2007) The representation of object 188. Perani D, Schnur T, Tettamanti M, Gorno-
concepts in the brain. Annu Rev Psychol Tempini M, Cappa SF, Fazio F (1999) Word
58:25–45 and picture matching: a PET study of seman-
174. Oliver RT, Thompson-Schill SL (2003) tic category effects. Neuropsychologia
Dorsal stream activation during retrieval of 37:293–306
object size and shape. Cogn Affect Behav 189. Devlin JT, Moore CJ, Mummery CJ, Gorno-
Neurosci 3:309–322 Tempini ML, Phillips JA, Noppeney U,
175. Patterson K, Nestor PJ, Rogers TT (2007) Frackowiak RS et al (2002) Anatomic con-
Where do you know what you know? The straints on cognitive theories of category
representation of semantic knowledge in the specificity. Neuroimage 15:675–685
human brain. Nat Rev Neurosci 8:976–987 190. Gorno-Tempini ML, Cipolotti L, Price CJ
176. Devlin JT, Russell RP, Davis MH, Price CJ, (2000) Category differences in brain activa-
Wilson J, Moss HE, Matthews PM et al tion studies: where do they come from? Proc
(2000) Susceptibility-induced loss of signal: Biol Sci 267:1253–1258
450 Federica Agosta et al.
191. Cappa SF, Perani D, Schnur T, Tettamanti M, 201. Davis MH, Johnsrude IS (2003) Hierarchical
Fazio F (1998) The effects of semantic cate- processing in spoken language comprehen-
gory and knowledge type on lexical-semantic sion. J Neurosci 23:3423–3431
access: a PET study. Neuroimage 8:350–359 202. Giraud AL, Kell C, Thierfelder C, Sterzer P,
192. Leube DT, Erb M, Grodd W, Bartels M, Russ MO, Preibisch C, Kleinschmidt A
Kircher TT (2001) Activation of right fronto- (2004) Contributions of sensory input, audi-
temporal cortex characterizes the ‘living’ cat- tory search and verbal comprehension to cor-
egory in semantic processing. Brain Res Cogn tical activity during speech processing. Cereb
Brain Res 12:425–430 Cortex 14:247–255
193. Chao LL, Haxby JV, Martin A (1999) 203. Rodd JM, Davis MH, Johnsrude IS (2005)
Attribute-based neural substrates in temporal The neural mechanisms of speech compre-
cortex for perceiving and knowing about hension: fMRI studies of semantic ambiguity.
objects. Nat Neurosci 2:913–919 Cereb Cortex 15:1261–1269
194. Kanwisher N, McDermott J, Chun MM 204. Kartsounis LD, Shallice T (1996) Modality
(1997) The fusiform face area: a module in specific semantic knowledge loss for unique
human extrastriate cortex specialized for face items. Cortex 32:109–119
perception. J Neurosci 17:4302–4311 205. Beauchamp MS, Haxby JV, Jennings JE,
195. Joseph JE (2001) Functional neuroimaging DeYoe EA (1999) An fMRI version of the
studies of category specificity in object recog- Farnsworth-Munsell 100-Hue test reveals
nition: a critical review and meta-analysis. multiple color-selective areas in human ven-
Cogn Affect Behav Neurosci 1:119–136 tral occipitotemporal cortex. Cereb Cortex
196. Gorno-Tempini ML, Price CJ (2001) 9:257–263
Identification of famous faces and buildings: a 206. Goldberg RF, Perfetti CA, Schneider W
functional neuroimaging study of semanti- (2006) Perceptual knowledge retrieval acti-
cally unique items. Brain 124:2087–2097 vates sensory brain regions. J Neurosci
197. Aguirre GK, Zarahn E, D'Esposito M (1998) 26:4917–4921
An area within human ventral cortex sensitive 207. Kellenbach ML, Brett M, Patterson K (2001)
to “building” stimuli: evidence and implica- Large, colorful, or noisy? Attribute- and
tions. Neuron 21:373–383 modality-specific activations during retrieval
198. Epstein R, Harris A, Stanley D, Kanwisher N of perceptual attribute knowledge. Cogn
(1999) The parahippocampal place area: recog- Affect Behav Neurosci 1:207–221
nition, navigation, or encoding? Neuron 208. Puce A, Allison T, Bentin S, Gore JC,
23:115–125 McCarthy G (1998) Temporal cortex activa-
199. Petersen SE, Fox PT, Posner MI, Mintun M, tion in humans viewing eye and mouth move-
Raichle ME (1988) Positron emission tomo- ments. J Neurosci 18:2188–2199
graphic studies of the cortical anatomy of sin- 209. Kourtzi Z, Kanwisher N (2000) Activation in
gle-word processing. Nature 331:585–589 human MT/MST by static images with
200. Vandenberghe R, Price C, Wise R, Josephs O, implied motion. J Cogn Neurosci 12:48–55
Frackowiak RSJ (1996) Functional anatomy 210. Kable JW, Lease-Spellmeyer J, Chatterjee A
of a common semantic system for words and (2002) Neural substrates of action event
pictures. Nature 383:254–256 knowledge. J Cogn Neurosci 14:795–805
Chapter 15
fMRI of Emotion
Simon Robinson, Ewald Moser, and Martin Peper
Abstract
Recent brain imaging work has expanded our understanding of the mechanisms of perceptual, cognitive, and
motor functions in human subjects, but research into the cerebral control of emotional and motivational
function is at a much earlier stage. Important concepts and theories of emotion are briefly introduced, as are
research designs and multimodal approaches to answering the central questions in the field. We provide a
detailed inspection of the methodological and technical challenges in assessing the cerebral correlates of
emotional activation, perception, learning, memory, and emotional regulation behavior in healthy humans.
fMRI is particularly challenging in structures such as the amygdala as it is affected by susceptibility-related
signal loss, image distortion, physiological and motion artifacts, and colocalized Resting State Networks
(RSNs). We review how these problems can be mitigated by using optimized echo-planar imaging (EPI)
parameters, alternative MR sequences, and correction schemes. High-quality data can be acquired rapidly in
these problematic regions with gradient-compensated multiecho EPI or high-resolution EPI with parallel
imaging and optimum gradient directions, combined with distortion correction. Although neuroimaging
studies of emotion encounter many difficulties regarding the limitations of measurement precision, research
design, and strategies of validating neuropsychological emotion constructs, considerable improvement in
data quality and sensitivity to subtle effects can be achieved. The methods outlined offer the prospect for
fMRI studies of emotion to provide more sensitive, reliable, and representative models of measurement that
systematically relate the dynamics of emotional regulation behavior with topographically distinct patterns of
activity in the brain. This will provide additional information as an aid to assessment, categorization, and
treatment of patients with emotional and personality disorders.
Key words Emotion, fMRI, Research design, Reliability, Validity, Amygdala, Signal loss, Distortion,
Resting state networks
1 Introduction
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_15, © Springer Science+Business Media New York 2016
451
452 Simon Robinson et al.
2 Psychological Methods
2.1 Emotion Theories Emotions have been defined as episodes of temporarily coupled,
and Constructs coordinated changes in component functions as a response of the
organism to external or internal events of major significance. These
2.1.1 Definitions
component functions entail subjective feelings, physiological acti-
vation processes, cognitive processes, motivational changes, motor
expression, and action tendencies [33, 34]. Emotions represent
functions of fast and flexible systems that provide basic response
tendencies for adaptive action [35].
Emotions can be differentiated from mood changes (extended
change in subjective feeling with low intensity), interpersonal
stances (affective positions during interpersonal exchange), atti-
tudes (enduring, affectively colored beliefs, preferences, and pre-
dispositions toward objects or persons), and personality traits
(stable dispositions and behavior tendencies) [29, 34].
The frequently used concept of “emotional activation” charac-
terizes a relatively broad class of physiological or mental phenom-
ena (e.g., strain, stress, physiological activation, arousal, etc.). It
can be specified with respect to a variety of dimensions such as
valence (quality of emotional experience), intensity or arousal
(global organismic change), directedness (motivational and orien-
tating functions), and selectivity (specific patterns of change) [36].
In contrast, the terms emotional reactivity or arousability, and psy-
chophysical reactivity refer to the dispositional variability of the
above activation processes under defined test conditions [37, 38].
Environmental objects possess a latent meaning structure of
emotional information, which is represented by a hierarchy of
454 Simon Robinson et al.
Table 1
Hierarchical organization of emotion concepts (modified from [29])
Emotion concepts or
domains Example constructs Basis for higher-order grouping
Dimensional concepts Valence (positive/negative Conceptual or meaning space for
emotions), approach/withdrawal, subjective experience and verbal
activity (active/passive), control, labels
etc.
Basic, fundamental, Anger, fear, sadness, joy, etc. Similarity of appraisal, motivational
discrete, modal emotions consequences, and response
or emotion families patterns; convenient label for
appropriate description and
communication
Specific appraisal/response Righteous anger, jealousy, mirth, Temporal coordination of different
configurations for fright, etc. response systems for a limited
recurring events/ period of time as produced by a
situations specific appraisal pattern
Continuous adaptational Orienting reflex, defense reflex, Automatic activations and
changes startle, sympathetic arousal, etc. coordination of basic
biobehavioral units
fMRI of Emotion 455
2.2 Research Design The requirements for experimental research [32] are not always
and Validity fulfilled by many early research designs of emotional neuroimaging
work. This is typical for the pilot stage of scientific progress. In
2.2.1 Research Design
many cases, only preliminary or correlational interpretations are
possible due to incomplete or missing control conditions (e.g.,
with respect to the “awareness” of emotional stimuli; [54]). In
fMRI of Emotion 457
2.2.3 Internal Validity FMRI is known to be a highly reactive measure because the scanner
setting (gradient noise and the supine position) causes the subject to
respond to the experimental situation as a stressor. Unless habitua-
tion sessions are included in the procedure, tonic stress and arousal
effects may be induced that modulate responding as discussed above.
For example, a decreasing rate of response of the amygdala to a con-
ditioned stimulus during the late phase of acquisition [10, 24, 26,
65] may also be attributable to testing effects (sensitization to the
setting, acquaintance with the procedure, and type of unconditioned
stimulation) rather than fast amygdala habituation per se (other fac-
tors might also explain reduced amygdala perfusion measures such
as potential ceiling effects, baseline dependencies, and regression to
the mean). In general, familiarity with emotionally activating proce-
dures in the scanner induces states of expectation, sensitizing or
desensitizing effects that may confound follow-up measurement. In
addition to these testing effects, history, that is, occurrences other
than the treatment and individual experiences between a first and a
second measurement are likely to endanger the assessment of emo-
tion (e.g., when assessing psychotherapy effects).
Changes in the observational technique, the measurement
device or sequence and other instrumentation effects may also
obscure emotion-related treatment variance during an fMRI
fMRI of Emotion 459
2.2.4 External Validity The majority of current paradigms have focused on lower-level
and Generalizability perceptual or learning processes pertaining to basic or secondary
emotional categories. Since the results depend on the selected task
Generalization to Other
parameters (degree of induced arousal, hedonic strength, and
Procedures and Paradigms
motivational value; degree of involvement of memory processes;
reinforcement schedule; conditioning to cues or contexts; etc.), a
comparison with and generalization to other operationalizations
remains difficult. Systematic neuroimaging approaches to higher-
level appraisal processes are still sparse. These involve evaluations
of the motivational conditions and coping potential, that is, the
ability to overcome obstructions or to adapt to unavoidable con-
sequences [29]. An expanded range of constructs would involve
an assessment of social communication processes, beliefs, prefer-
ences, predispositions, high-level evaluation checks, as well as
modulating sociocultural influences. Higher-order appraisal pro-
cesses involve the evaluation of whether stimulus events are com-
patible with social standards and values or with the self-concept.
Another function to be explored concerns the degree to which a
stimulus event may increase, decrease, or even block goal attain-
ment or need satisfaction, and activate a reorientation of the indi-
vidual’s goal/need hierarchy and behavioral planning (goal/need
priority setting) [29].
fMRI of Emotion 461
Generalization to Other The study groups of many fMRI studies have been relatively small
Subjects and Populations and poorly described with respect to personality dimensions. Since
several studies provide evidence for trait-dependent differences in
responding [73–76], it remains unclear to what extent the results
may have been influenced by interindividual differences of the par-
ticipating subjects. The representativeness of results is particularly
poor if members of the social circle of the lab serve as participants
instead of independently recruited participants. Thus, when the
effects of an emotional paradigm interact with characteristics of the
study groups (such as a low level of emotionality in subjects willing to
participate in an activating scanning condition), this selection × treat-
ment effect may endanger generalizations to other populations.
462 Simon Robinson et al.
3 fMRI Methods
3.1 Methodological A host of fMRI studies have identified the amygdalae as central
Challenges structures in emotion processing (see Sect. 1 and Zald et al. [5], for
example, for a review). The amygdalae lie in the anterior medial
3.1.1 Introduction
temporal lobe (MTL), bounded ventrolaterally by the lateral ven-
tricles and medially by the sphenoid sinuses (Fig. 1). The differing
magnetic susceptibilities of these tissues cause large deviations in
the static magnetic field, B0. There is also a strong gradient in B0 in
the MTL, and differing precession frequencies lead to dephasing of
Fig. 1 The amygdalae, central brain structures in emotion processing, lie in a region of moderate deviation
from the static magnetic field (left) and very high static magnetic field gradients (right). The planes intersect in
the amygdala at MNI coordinate (18, −2, −18), marked by arrows. Single subject measurement at 4.0 T
fMRI of Emotion 463
10
8
Patient 13
6 Patient 8
Displacement (mm)
2
Patient 1
0 Control 3
Instruction
Fig. 2 Large static magnetic field gradients make the amygdala region prone to the artifactual appearance of
neuronal activation when stimulus-correlated motion (SCM) is present. Left : Observed patterns of SCM of
schizophrenic patients and controls in a 3.0-T experiment with three stimulus blocks (facial emotion and age
discrimination “EMO” and “AGE”). Right : a baseline (no stimulus) study in which a subject executed submil-
limeter SCM similar to that of Patient 1. The contrast corresponds to the “EMO” periods (uncorrected p < 0.0001;
t threshold = 5, Montreal Neurological Institute coordinates 22, −6, −16)
464 Simon Robinson et al.
Fig. 3 Signal changes in the amygdala in emotion experiments have to be measured against a background of
resting state fluctuations. A resting state network recently been reported, covering the amygdala and basal
ganglia (3.0 T, group independent component analysis of 26 young healthy adults). Adapted from [106] with
permission from the ISMRM
3.1.2 Signal Loss It is worthwhile to briefly review the problem of signal loss from an
and BOLD Sensitivity Loss empirical perspective. A temporal resolution of 1–3 s is usually
desirable in fMRI. The whole brain may be covered in this time by
acquiring images with voxels of typically 3-mm size (or 27 μl).
fMRI of Emotion 465
Fig. 4 Effects of voxel size and acceleration factor on T2* and echo-planar imag-
ing (EPI) image quality at high field (4.0 T). Top: T2* in coronal and axial slices
through the amygdala at two voxel sizes. Bottom: corresponding EPI in slices
through the amygdala with acquisition voxel sizes of 4 × 4 × 4 mm, 3 × 3 × 3 mm,
2 × 2 × 2 mm, and 2 × 2 × 2 with GRAPPA acceleration of factor 2, all with echo
time (TE) = 32 ms
466 Simon Robinson et al.
æ -TE ö
S ( TE ) = S ( 0 ) ´ exp ç ÷ ´ F ( TE ) , (1)
è T2 ø
where an approximation to F(TE) for linear field variations over
voxels, ΔBi, in the x, y, and z directions is
æ ‡ Bx TE ö æ ‡ By TE ö æ ‡ Bz TE ö
F ( TE ) = sinc ç ÷ ´ sinc ç ÷ ´ sinc ç ÷ . (2)
è 2 ø è 2 ø è 2 ø
This illustrates that the signal decay rate may be reduced by decreas-
ing the voxel size—to reduce the gradients across voxels, ΔBi—or
by reducing the TE.
The aim of any attempt to optimize an EPI sequence is not just
to maximize signal, described above, but also BOLD sensitivity (BS),
which is equal to the product of image intensity and TE; for magneti-
cally homogeneous regions is a maximum when the EPI effective TE
is equal to the T2* of the target region [83]. In homogeneous regions,
however, the presence of field gradients shifts the location of signal in
k-space, mainly in the phase-encode direction (because of the low
bandwidth), changing the local TE [81]. Through-plane field gradi-
ents lead to signal loss and reduce BS. If the component of the in-
plane susceptibility gradient in the phase- encode direction is
antiparallel to the phase-encode gradient “blip” direction, then the
TE is also reduced, reducing BS further. Conversely, if it is parallel to
the phase-encode “blips” then TE increases. While this increases BS,
to some extent compensating for signal loss, if the shift of TE is too
large the echo will fall outside the acquisition window, leading to
complete signal dropout. This is commonly observed in the anterior
MTL for a negative-going phase-encode scheme.
This description motivates the optimization approaches to EPI
in susceptibility-affected regions which will be outlined later in this
section; compensating through-plane gradients, selecting image
orientation and gradient direction to minimize echo shifts, and
reducing voxel sizes to reduce field gradients. These techniques
will be shown to increase both signal and BS.
3.1.3 Image Distortion Accurate spatial encoding in MRI is founded upon a homogeneous
static magnetic field in the object. The location of signal is deduced
from the local field strength under the application of small orthog-
onal, linear magnetic fields in directions usually referred to as slice
select, readout, and phase-encode. The method is confounded if
there are regional variations in the static magnetic field, which lead
to signal mislocalization (distortion). Typical field offsets are illus-
trated in Fig. 1 (left) and lead to EPI distortions of the image
shown in Fig. 4.
The extent of distortion, expressed as the number of pixels by
which signal is mislocalized, is equal to the local magnetic field
fMRI of Emotion 467
3.1.5 Motion Artifacts Motion artifacts affect all regions of the brain, but are particularly
problematic in emotion studies because the nature of the task
material is prone to induce SCM as a startle, attention, or repulse
response. Patients with disorders with emotional components
(such as schizophrenia and posttraumatic stress disorder) are less
likely to remain still throughout the experiment and the interac-
tion between motion and distortion in regions of high susceptibil-
ity gradient produces nonlinear pixel shifts that are not well
corrected with rigid-body methods. Partial brain coverage proto-
cols, such as those that may be used to allow z-shimming or high
spatial and temporal resolution fMRI in the amygdala, are also
more prone to partial voluming in the outermost slices and spin
history effects, in which motion between the acquisition of adja-
cent slices leads to some spins being excited twice within one rep-
etition time (TR) while others are not excited at all.
Head motion can be minimized using bite bars, vacuum cush-
ions, thermoplastic masks, or plaster head casts. As well as effective
fMRI of Emotion 469
3.2 MR Methods, While the signal to noise ratio (SNR), the magnitude of BOLD
Sequences, signal changes, and the specificity of the BOLD response to micro-
and Protocols vascular contributions all increase with field strength, so do physi-
ological noise, field inhomogeneities, and physiological artifacts
3.2.1 Field Strength
which specifically affect the anterior MTL. The advantages of high
field for emotion studies are therefore restricted to particular
regimes and methods in which these problems are minimized.
Human emotion fMRI studies have been carried out at field
strengths from 1.0 to 7.0 T. In line with the development of
sequences and approaches to EPI in susceptibility-affected area
which are discussed in Sect. 3.2.2–3.2.8 (high-resolution single
and multishot EPI, multiecho and spiral acquisitions, gradient
compensation, and parallel imaging), emotion fMRI in the high
field regime (3.0–4.0 T) has become commonplace, although
applied studies have generally used standard sequences and param-
eters despite the problems which have received attention in the
MR literature [110] and a number of promising remedies (see the
following sections). Ultra-high field strength studies of emotion
are still sparse, however, and it is likely that they will be restricted
to highly specific questions during the next 5–10 years of hardware
and sequence development.
Theoretical gains in SNR at high field are limited by physio-
logical noise, which increases both with field strength and voxel
size, and causes time-series SNR (tSNR) to reach as asymptotic
limit with voxel volume [111]. This limit was found to increase
only modestly with field strength, being 65 at 1.5 T, 75 at 3 T,
and 90 at 7 T, so that for large (5 × 5 × 3 mm) voxels, tSNR was
only 11 % higher at 3 T than at 1.5 T, and only 25 % higher at 7 T
than 1.5 T. The tendency toward asymptotic behavior began at
relatively small volume volumes, with 80 % of the asymptotic
maximum being reached at 28.6, 15.0, and 11.7 mm3 at 1.5, 3,
and 7 T, respectively. For small voxels, however, where thermal
noise dominates, tSNR gains were almost linear with field
strength. In the same study, the authors found that with
1.5 × 1.5 × 3 mm3 voxels, tSNR increased by 110 % at 3 T com-
pared to 1.5 T, and by 245 % at 7 T compared to 1.5 T [111].
This study clearly shows that tSNR gains are to be made at high
field in the small voxel volume regime.
fMRI of Emotion 471
These tSNR results also explain the often modest gains achieved
in fMRI studies at higher field, particularly in regions affected by
signal dropout. Krasnow et al. [112] compared activation in
response to perceptual, cognitive, and affective tasks at 1.5 and 3 T
with a relatively large voxel protocol (3 × 3 × 4 mm) and observed
only moderate increases in activated volume at 3 T for the percep-
tual and cognitive tasks (23 and 36 %, respectively), but no signifi-
cant improvement in the activated amygdala volume due to
increased susceptibility-related signal loss. A high-resolution, high-
field approach has been exemplified in the only human study of
amygdala function at 7 T to date of which we are aware, which was
carried out at submillimeter resolution [113].
These studies define the regime in which field strength gains
are to be made, but it is fair to ask why one should move to high-
resolution measurements if the neuroscience question does not
require, for instance, subnuclei of the amygdala to be resolved,
but—as is more commonly the case—the study of interactions
between the amygdalae and the cortex, for which whole brain cov-
erage is essential. The use of high resolution here is not principally
to distinguish activation in small structures, but to reduce both
physiological noise and susceptibility artifacts. A number of works
have shown the value of averaging thin slices, downsampling, and
smoothing data acquired at high resolution [114–116] and using
multichannel coils [115] to regain losses in SNR inherent to small
voxels generally and yielding net gains in susceptibility affected
areas [115, 117].
3.2.2 z-Shimming, The effect of signal dephasing arising from through-plane gradi-
Gradient Compensation, ents may be reduced by creating a composite image from a number
Tailored RF Pulses of acquisitions in which different slice-select gradients are applied
[118], a process known as z-shimming. In each image the applied
gradient pulse is appropriate to counteract susceptibility gradients
in particular regions. The method is effective in regaining signal in
the anterior MTL, but clearly reduces temporal resolution by a fac-
tor equal to the number of images acquired, usually a minimum of
3. Alternatively, a single, moderate preparation pulse may be used.
This reduces through-plane dephasing in affected areas at limited
cost to BS and signal in homogeneous areas, and allows slices to be
orientated so that TE shifts are small, reducing signal loss due to
in-plane gradients [119]. z-Shimming and other compensation
schemes have been applied in a number of other sequences
described in this section.
Spins may also be refocused using tailored radio frequency
pulses which create uniform in-plane phase but quadratic phase
variation through the slice, allowing dephasing to be “precompen-
sated” [120]. Analogous to z-shimming, in the original implemen-
tation a number of acquisitions with different precompensations
were required, suited to different regions. More recently 3D
472 Simon Robinson et al.
versions have been developed, and while these are promising the
pulse lengths are long, and the distribution of susceptibilities must
be known [121], or calculated iteratively online [122]. These are,
however, important steps toward single-shot compensation of sus-
ceptibility dropout.
3.2.3 Slice Orientation Divergent findings and recommendations for the optimum slice
and Gradient Directions orientation for amygdala fMRI are due to the absence, until rela-
tively recently, of an adequate description of signal loss and BS in
the presence of field gradients [81, 119, 123].
In many early studies, quite nonisotropic voxels were used to
achieve short TR while minimizing demands on scanner hardware,
with slice thickness being substantially larger than the in-plane
voxel size. Gradients across voxels were highest then, and signal
loss most severe, if the direction of strongest field gradient was
along the slice (through-plane) direction [124]. With many studies
finding that the direction of the field vector across the amygdala
was principally superior-inferior [125], this prescription precluded
an axial orientation. As bilateral structures, the amygdala could be
imaged in the same slice in the coronal but not the sagittal planes,
leading to the coronal orientation being preferred by many [110].
The optimum imaging plane is also dependent on whether
gradient compensation is used [81]. If so, through-plane gradients
may be compensated for with a moderate gradient in the slice
direction, although this will lead to a small decrease in BS in unaf-
fected areas. The slice can then be orientated so that in-plane
gradients are below the critical threshold for Type 2 signal loss.
The value of this has been demonstrated in the orbitofrontal cortex
[119] but the approach yields lower rewards in the amygdala
region [126] as gradients are higher (making it more difficult to
find a suitable value for compensation), and are more variable
between subjects.
The simulations of Chen et al. [125] for the amygdala sug-
gested that the maximum BS was to be achieved by orienting the
slice direction perpendicular to the maximum gradient vector and
the readout direction parallel to it, indicating an (oblique) coronal
orientation with superior–inferior readout. The angle between the
gradient vector and the superior–inferior direction was shown to
vary widely between subjects (from −7° to +26° at 1.5 T, from −5°
to +34° at 3 T), meaning that field gradients need to be mapped
for each subject before measurement. This scheme also invokes
distortions which are asymmetric about the midline (left–right). If
erroneous conclusions about lateralization are to be avoided, resid-
ual distortions in the amygdala should be symmetric, requiring the
phase–encode direction to be superior–inferior for coronal slices or
anterior–posterior for axial slices.
As well as the direction of imaging gradients, the sign of phase-
encode blips is important for signal loss and BS [123]. Encoding in
fMRI of Emotion 473
3.2.4 Voxel Size Among many solutions to the problem of signal loss in the anterior
MTL, reduced voxel size was established very early as an effective
means of mitigating susceptibility-related signal loss [127, 128].
Equation (2) describes how the rate of signal decay is reduced with
voxel size by lowering field gradients across voxels. The effective-
ness of this can be seen in the 4-T images of Fig. 4 over a range of
resolutions, with T2* in the amygdala (measured with a multiple
gradient-echo sequence with the same geometry as the EPI)
increasing from 22 to 38 ms when the voxel size is reduced from
64 to 8 mm3, with corresponding EPI signal increase apparent in
the anterior MTL.
Reducing voxel size comes at the expense of temporal resolu-
tion (or brain coverage) and SNR. The relationship between image
SNR and voxel volume, ΔV, is
474 Simon Robinson et al.
Nx N y Nz
SNR = V , (3)
rBW
3.2.5 Echo Time Taking the simplest approach of matching effective echo time
(TEeff) to the T2* of the structures of interest in GE-EPI might be
seen as being problematic in large voxel size acquisitions, with T2*
s varying quite widely (e.g., between the amygdala and the fusi-
form face area). One solution is to use a multiecho sequence, in
which the each time of each image is appropriate for regions with
particular field gradients, as will be described in more detail in
Sect. 3.2.8. A novel solution to matching TEeff to T2* in the amyg-
dala without sacrificing BS in more dorsal slices is to use an axial
acquisition with slice-specific TE, demonstrated at 1.5 T with
TEeff = 60 ms in dorsal slices, TEeff = 40 ms in ventral slices, and a
transition zone with intermediate effective TE [135].
It should be remembered, though, that the maximum of BS is
quite flat as a function of TE, and TE is itself not well defined in
EPI. In the previous sections, we also saw that in-plane susceptibil-
ity gradients change local TE [81]. This exposes the limitation of
the approach of simply reducing the TEeff of the sequence. In the
common, negative blip scheme, signal in the anterior MTL will in
fact be shifted to a longer TE. Using a short TEeff makes the
sequence more prone to complete (type 2) signal loss.
This explains the experimental findings of Gorno-Tempini
et al. [136] and Morawetz et al. [134]. In 2-T dual-echo EPI with
large voxels, Gorno-Tempini et al. found that although signal loss
was reduced at the short TE (26 ms) BOLD activation was signifi-
cantly greater in the hippocampus at the longer TE (40 ms).
Morawetz et al. [134] studied four EPI protocols in their efficacy
at mapping amygdala activation, using variants with two different
TE (27 and 36 ms) and slices thicknesses (2 and 4 mm), all with
high in-plane resolution (2 mm). Activation results were poor in
the 4-mm protocols, even at the shorter TE.
A more effective approach than reducing TEeff is to reduce
susceptibility gradients, and thereby signal dephasing and echo
shifts, using the techniques described earlier; gradient compensa-
tion, selection of appropriate gradient direction and slice orienta-
tion, and the use of smaller voxels. This increases T2* in
susceptibility-affected regions and, by reducing echo shifts, makes
BS more homogeneous throughout the imaging volume.
Conditions then approach those with a homogeneous static field,
where BS is maximized by using TE eff = T2* .
The increase in T2* in the amygdala with reduced voxel size is
illustrated at 4 T in Fig. 4; from 22 ms in a 4 × 4 × 4-mm acquisition
to 38 ms in 2 × 2 × 2-mm data, consistent with previous results at
3 T [117]. Likewise, increase in BS was illustrated in the Morawetz
et al. study [134], in which robust amygdala activation was only
detectable in the high-resolution acquisition.
3.2.6 Parallel Imaging The previous sections have shown that many of the techniques
which mitigate susceptibility-related signal loss in the amygdala,
476 Simon Robinson et al.
3.2.7 Flip Angle The following is a consideration which is common to fMRI studies in
all brain regions. The flip angle that should be used in a sequence is
that which maximizes the signal with a particular experimental TR. In
a spoiled gradient-echo sequence this is the Ernst angle, θE, given by
æ TR
ö
q E = arccos ç e T1
÷.
ç ÷
è ø
3.2.8 Alternatives to 2D, If multiple echo images are acquired following a single excitation, the
Single-Shot, Gradient-Echo range of TEeff in these provides near-optimum BS for a number of
EPI regions [144, 145]. Images acquired at different TEs may be analyzed
separately, or combined to maximize BOLD contrast-to-noise ratio
[145]. Acquiring multiple images in a single shot also allows
fMRI of Emotion 477
3.2.9 Summary In the subsections of Sect. 3.2 we have looked at the influence of
field strength, gradient compensation, slice orientation, voxel size,
TE, and acquisition acceleration factor on susceptibility-related
signal and BS reduction in the anterior MTL, as well as discussing
some variants of multiecho and spiral schemes which have been
tailored for this region. While the interdependent nature of EPI
parameters and changing considerations at different field strength
necessarily make some considerations complex, we would like to
pick out two lines of approach presented here as being particularly
effective, and clarify recommendations.
The first approach is high-field, high-resolution single-shot
EPI with gradient compensation and acceleration. BOLD signal
changes are greater at high field (3.0–4.0 T), and the tSNR advan-
tages of high field strength are capitalized upon by measuring with
small (circa 8-μl voxels), where thermal noise rather than physio-
logical noise dominates. Measuring with small voxels reduces sig-
nal dephasing, making T2* more homogeneous. Shifts in local TE
are also less, reducing Type 2 signal loss and increasing BOLD
sensitivity. Moderate slice select gradient compensation and an
oblique axial acquisition with a tilt between 20 and 45° (anterior
478 Simon Robinson et al.
slice edge toward the head) reduces in-plane gradients and echo
shifts further. With susceptibility gradients reduced—evidenced by
T2* values close to those in magnetically homogeneous regions—
BS can be maximized by setting the TE eff = T2* . The TEeff can be
reached using parallel imaging acceleration (e.g., factor 2), which
further reduces both TE shifts and image distortion. Images
acquired with these parameters have high signal in the anterior
MTL, low distortion, and quite homogenous BS. Time-series SNR
can be increased before statistical analysis by downsampling or
smoothing images. This approach is attractive in that it may be
achieved on most modern high field systems.
Not only the value of gradient compensation was discussed in
Sect. 3.2.2, but also the high cost in temporal resolution, if images
with a number of compensation gradients are acquired. The s econd
approach we wish to highlight involves the application of a range
of compensation gradients to each of a number of echoes acquired
after a single excitation, so reducing the time penalty. Both the
multiecho echo-planar [146] and multiecho spiral acquisitions
[151] described in Sect. 3.2.8 have been shown to be effective in
reducing susceptibility-related signal loss in the anterior MTL.
3.3 Correction The field map (FM) method was first described by Weisskoff and
Methods Davis [153] and developed by Jezzard and Balaban [154]. In
Sect. 3.1.2 we saw that distortion in EPI is only significant in the
3.3.1 Distortion
phase-encode direction and that the number of pixels by which
Correction with the Field
signal is mislocated is equal to the local field offset divided by the
Map and Point-Spread
bandwidth per pixel in the phase-encode direction. In the fieldmap
Function Methods
method, static magnetic field deviations, ΔB, are calculated from
the phase difference, Δϕ, between two scans with TE separated by
ΔTE (or a dual-echo scan), using the relation DB = 2pg TEDj .
This map is distorted (forward-warped) to provide a map of the
voxel shifts required to reverse the distortion at each EPI location.
Gaps in the corrected image are filled by interpolation.
While undemanding from the sequence perspective, considerable
postprocessing is required to produce FMs that do not contain errors.
Phase imaging is only capable of encoding phase values in a 2π range,
with values outside this range being aliased, causing “wraps” in the
image. These can be removed in the spatial domain using a number of
freely available algorithms (e.g., PRELUDE [155] or ΠUN [156]),
or by examining voxel-wise phase evolution in time if three or more
echoes are acquired [157]. If imaging is being carried out with a mul-
tichannel radiofrequency receive coil, phase images created via the
sum-of-squares reconstruction [158] will show nonphysical disconti-
nuities from arbitrary phase offsets between the coil channels (incon-
gruent wraps) unless these offsets are removed [159, 160].
Alternatively, images from channels may be processed separately and
individual FMs, weighted by coil sensitivities, combined. In 2D spatial
unwrapping, additional global, erroneous 2π phase changes are occa-
sionally inferred between TE when the algorithm begins to unwrap
fMRI of Emotion 479
Fig. 5 Distortion correction of echo-planar imaging (EPI) at high field (4.0 T). A comparison of field-map (col-
umn 3) and point-spread function (column 4) correction of distortion in EPI (column 2) at the level of the
amygdala (top row) compared to a more dorsal section (bottom row). Salient features have been copied from
a gradient-echo geometric reference scan (column 1)
480 Simon Robinson et al.
through the amygdala (top row), and comparing this with the
situation in a more dorsal slice (bottom row). Raw and corrected
EPIs are compared to a gradient-echo reference which has the
same (subvoxel) distortion in the readout direction, but no dis-
tortion in the phase- encode direction. The distortion at the
anterior boundary of the amygdala (A) is circa 3 mm—moderate
compared to the displacement of the ventricles (9 mm at B) and
the frontal gray-white matter border indicated at C (12 mm). If
the multiplicity of phase information available from multichan-
nel coils is used in the FM method [161], both FM and PSF
methods perform very well in all areas, with only minor errors at
the periphery of the FM-corrected images due to residual field
map inaccuracies at those locations (at D, not present in the
PSF-corrected images).
The choice of correction method is often a pragmatic one
based on which is more robustly and conveniently implemented.
3.3.3 Correction In patient group studies, Bullmore et al. [179] have shown the
of Stimulus-Correlated need to compare the extent to which SCM explains variance
Motion Artifacts between the groups, and suggest that this be identified using an
analysis of covariance (ANCOVA). Without this approach, differ-
ences between the groups arising from higher SCM in the
schizophrenic group in their study would have been attributed to
differential activation in response to the task.
In the example of Fig. 2 (left), realignment of the time series
in the motion-only replication did not substantially reduce the
amygdala SCM artifact (right), but including identified motion
parameters in the model as NVRs was effective [168, 180].
Alternatively, a boxcar NVR corresponding to presentation and
response periods can be included in the model [181]. This and a
number of other studies [182] have shown that the temporal
shift in response introduced by the hemodynamic response func-
tion (HRF) makes it possible to separate motion from activation
for short presentation periods, making event-related designs less
sensitive to motion than block designs.
482 Simon Robinson et al.
Fig. 6 High-resolution imaging detailed in this chapter allows the acquisition of low-artifact echo-planar imag-
ing (EPI) and allows subtle processing effects to be distinguished. Group results from 29 subjects for the condi-
tions (a) emotion recognition (b) implicit emotion processing (age discrimination) and (c) the difference
between the two conditions (3.0 T). Results, showing activation in the amygdala and fusiform gyrus (as well as
cerebellum and brainstem) are overlaid on mean EPI and thresholded at p = 0.05, family-wise error corrected.
Reprinted from [192], with permission
fMRI of Emotion 485
5 Conclusions
Acknowledgments
References
1. Büchel C, Dolan RJ (2000) Classical fear 17. Adolphs R, Tranel D, Damasio H, Damasio
conditioning in functional neuroimaging. A (1994) Impaired recognition of emotion in
Curr Opin Neurobiol 10:219–223 facial expressions following bilateral damage
2. Phan KL, Wager T, Taylor SF, Liberzon I to the human amygdala. Nature 372:669–672
(2002) Functional neuroanatomy of emotion: 18. Broks P, Young AW, Maratos EJ et al (1998)
a meta-analysis of emotion activation studies Face processing impairments after encephali-
in PET and fMRI. NeuroImage 16:331–348 tis: amygdala damage and recognition of fear.
3. Dolan RJ, Vuilleumier P (2003) Amygdala Neuropsychologia 36:59–70
automaticity in emotional processing. Ann N 19. Sprengelmeyer R, Young AW, Schröder U
Y Acad Sci 985:348–355 et al (1999) Knowing no fear. Proc R Soc
4. Wager TD, Phan KL, Liberzon I, Taylor SF Lond B Biol Sci 266:2451–2456
(2003) Valence, gender, and lateralization 20. Adolphs R, Gosselin F, Buchanan T, Tranel
of functional brain anatomy in emotion: a D, Schyns P, Damasio A (2005) A mechanism
meta-analysis of findings from neuroimaging. for impaired fear recognition after amygdala
NeuroImage 19:513–531 damage. Nature 433:68–72
5. Zald DH (2003) The human amygdala and 21. Adolphs R, Tranel D, Hamann S et al (1999)
the emotional evaluation of sensory stimuli. Recognition of facial emotion in nine indi-
Brain Res Brain Res Rev 41:88–123 viduals with bilateral amygdala damage.
6. Ochsner KN, Gross JJ (2005) The cognitive Neuropsychologia 37:1111–1117
control of emotion. Trends Cogn Sci 9:242–249 22. Schmolck H, Squire LR (2001) Impaired
7. Panksepp J (1998) Affective neuroscience: perception of facial emotions following bilat-
the foundations of human and animal emo- eral damage to the anterior temporal lobe.
tions. Oxford University Press, Oxford Neuropsychology 15:30–38
8. Davidson RJ, Jackson DC, Kalin NH (2000) 23. LeDoux JE (1995) Emotion: clues from the
Emotion, plasticity, context, and regulation: brain. Annu Rev Psychol 46:209–235
perspectives from affective neuroscience. 24. LaBar KS, Gatenby JC, Gore JC, LeDoux JE,
Psychol Bull 126:890–909 Phelps EA (1998) Human amygdala activa-
9. Dolan RJ (2002) Emotion, cognition, and tion during conditioned fear acquisition and
behavior. Science 298:1191–1194 extinction: a mixed-trial fMRI study. Neuron
10. Breiter HC, Rauch SL (1996) Functional 20:937–945
MRI and the study of OCD: from symptom 25. Morris JS, Büchel C, Dolan RJ (2002) Parallel
provocation to cognitive-behavioral probes neural responses in amygdala subregions and
of cortico-striatal systems and the amygdala. sensory cortex during implicit fear. Neurobiol
NeuroImage 4:127–138 12:169–177
11. Johnson PA, Hurley RA, Benkelfat C, Herpertz 26. Büchel C, Morris J, Dolan RJ, Friston KJ
SC, Taber KH (2003) Understanding emo- (1998) Brain systems mediating aversive
tion regulation in borderline personality conditioning: an event-related fMRI study.
disorder: contributions of neuroimaging. Neuron 20:947–957
J Neuropsychiatry Clin Neurosci 15:397–402 27. Krech D (1950) Dynamic systems as open neu-
12. Hamann S, Canli T (2004) Individual dif- rological systems. Psychol Rev 57:345–361
ferences in emotion processing. Curr Opin 28. Cacioppo JT, Tassinary LG, Berntson GG
Neurobiol 14:233–238 (eds) (2000) Handbook of psychophysiology.
13. Paquette V, Levesque J, Mensour B et al Cambridge University Press, Cambridge
(2003) Change the mind and you change the 29. Scherer KR, Peper M (2001) Psychological
brain: effects of cognitive-behavioral therapy theories of emotion and neuropsychological
on the neural correlates of spider phobia. research. In: Gainotti G (ed) Handbook of
Neuroimage 18:401–409 neuropsychology. Vol. 5: Emotional behav-
14. Popma A, Raine A (2006) Will future forensic ior and its disorders, 2nd edn. Elsevier,
assessment be neurobiologic? Child Adolesc Amsterdam, pp 17–48
Psychiatr Clin N Am 15:429–444 30. Coan JA, Allen JJB (eds) (2007) Handbook
15. Bräutigam S (2005) Neuroeconomics – from of emotion elicitation and assessment. Oxford
neural systems to economic behaviour. Brain University Press, New York, NY
Res Bull 67:355–360 31. Peper M, Vauth R (2008) Socio-emotional pro-
16. Walter H, Abler B, Ciaramidaro A, Erk S cessing competences: assessment and clinical
(2005) Motivating forces of human actions: application. In: Vandekerckhove M, von Scheve
neuroimaging reward and social interaction. C, Ismer S, Jung S, Kronast S (eds) Regulating
Brain Res Bull 67:368–381 emotions (ch. 9). Wiley, Hoboken, NJ
fMRI of Emotion 489
32. Kerlinger FN, Lee HB (2000) Foundations 47. Lang P, Rice DG, Sternbach RA (1972) The
of behavioral research (4th edition). Harcourt psychophysiology of emotion. In: Greenfield
College Publishers, Fort Worth, TX NS, Sternbach RA (eds) Handbook of
33. Scherer KR (1993) Neuroscience projections psychophysiology. Holt, New York, NY,
to current debates in emotion psychology. pp 623–643
Cogn Emot 7:1–41 48. Fahrenberg J (1983) Psychophysiologische
34. Scherer KR (2000) Psychological models of Methodik. In: Groffman K-J, Michel
emotion. In: Borod J (ed) The neuropsychol- L (eds) Enzyklopädie der Psychologie:
ogy of emotion. Oxford University Press, Themenbereich B Methodologie und
Oxford, pp 137–162 Methoden, Serie II Psychologische
35. Scherer KR (1999) Appraisal theories. In: Diagnostik, Band 4 Verhaltensdiagnostik.
Dalgleish T, Power M (eds) Handbook of cogni- Hogrefe, Göttingen, pp 1–192
tion and emotion. Wiley, Chichester, pp 637–663 49. Lazarus RS (1991) Emotion and adaptation.
36. Peper M, Fahrenberg J (2008) Oxford University Press, New York, NY
Psychophysiologie. In: Sturm W, Herrmann 50. Parkinson B, Totterdell P (1999) Classifying
M, Münte TF (eds) Lehrbuch der Klinischen affect-regulation strategies. Cogn Emot
Neuropsychologie. Spektrum Akademischer 13:277–303
Verlag, Heidelberg 51. Gross JJ, Levenson RW (1991) Emotional
37. Stemmler G, Fahrenberg J (1989) suppression: physiology, self-report, and
Psychophysiological assessment: conceptual, expressive behavior. J Pers Soc Psychol
psychometric, and statistical issues. In: Turpin 64:970–986
G (ed) Handbook of clinical psychophysiol- 52. Lazarus RS, Folkman S (1984) Stress,
ogy. Wiley, Chichester, pp 71–104 appraisal, and coping. Springer, New York, NY
38. Stemmler G (1992) Differential psycho- 53. Gross JJ (1998) The emerging field of emo-
physiology: persons in situations. Springer, tion regulation: an integrative review. Rev
Heidelberg Gen Psychol 2:271–299
39. Johnsen BH, Thayer JF, Hugdahl K (1995) 54. Whalen PJ, Rauch SL, Etcoff NL, McInerney
Affective judgment of the Ekman faces: SC, Lee MB, Jenike MA (1998) Masked
a dimensional approach. J Psychophysiol presentations of emotional facial expressions
9:193–202 modulate amygdala activity without explicit
40. Peper M, Irle E (1997) The decoding of emo- knowledge. J Neurosci 18:411–418
tional concepts in patients with focal cerebral 55. Pessoa L, Japee S, Sturman D, Ungerleider
lesions. Brain Cogn 34:360–387 LG (2006) Target visibility and visual aware-
41. Ekman P (1994) Strong evidence for uni- ness modulate amygdala responses to fearful
versals in facial expressions. Psychol Bull faces. Cereb Cortex 16:366–375
115:268–287 56. McIntosh AR, Gonzalez-Lima F (1994)
42. Russell JA (1994) Is there universal recog- Network interactions among limbic cor-
nition of emotion from facial expressions? tices, basal forebrain, and cerebellum dif-
Psychol Bull 115:102–141 ferentiate a tone conditioned as a Pavlovian
43. Stemmler G (1998) Emotionen. In: F excitor or inhibitor: fluorodeoxyglucose map-
Rösler (Ed) Enzyklopädie der Psychologie: ping and covariance structural modeling.
Themenbereich C Theorie und Forschung, Serie J Neurophysiol 72:1717–1733
1 Biologische Psychologie, Band 5 Ergebnisse 57. Büchel C, Friston KJ (1997) Modulation of
und Anwendungen der Psychophysiologie. connectivity in visual pathways by attention:
Göttingen: Hogrefe, pp 95–163 cortical interactions evaluated with structural
44. Scherer KR (1984) On the nature and equation modelling. Magn Reson Imaging
function of emotion: a component process 15:763–770
approach. In: Scherer KR, Ekman P (eds) 58. Büchel C, Friston KJ (2000) Assessing interac-
Approaches to emotion. Erlbaum, Hillsdale, tions among neuronal systems using functional
NJ, pp 203–317 neuroimaging. Neural Netw 13:871–882
45. Lacey JI (1967) Somatic response patterning 59. Friston KJ, Harrison L, Penny W (2003)
and stress: some revisions of activation theory. Dynamic causal modelling. NeuroImage
In: Appley MH, Trumbull R (eds) Psychological 19:1273–1302
stress: issues in research. Appleton-Century 60. Vuilleumier P, Richardson MP, Armony JL,
Crofts, New York, NY, pp 14–42 Driver J, Dolan RJ (2004) Distant influences
46. Peper M (2000) Awareness of emotions: of amygdala lesion on visual cortical activa-
a neuropsychological perspective. Adv tion during emotional face processing. Nat
Consciousness Stud 16:245–270 Neurosci 7:1271–1278
490 Simon Robinson et al.
61. Friston KJ (1994) Functional and effective 75. Canli T, Sivers H, Whitfield SL, Gotlib IH,
connectivity in neuroimaging: a synthesis. Gabrieli JDE (2002) Amygdala response to
Hum Brain Mapp 2:56–78 happy faces as a function of extraversion.
62. Ochsner KN, Ray RD, Cooper JC et al (2004) Science 296:2191
For better or for worse: neural systems support- 76. Schienle A, Schafer A, Stark R, Walter B, Vaitl
ing the cognitive down- and up-regulation of D (2005) Relationship between disgust sensi-
negative emotion. Neuroimage 23:483–499 tivity, trait anxiety and brain activity during
63. Fahrenberg J, Peper M (2000) disgust induction. Neuropsychobiology
Psychophysiologie. In: Sturm W, Herrmann 51:86–92
M, Wallesch CW (eds) Lehrbuch der neuro- 77. Canli T, Amin Z (2002) Neuroimaging of
psychologie, chapter 1. 10. Swets and emotion and personality: scientific evidence
Zeitlinger, Amsterdam, pp 154–68 and ethical considerations. Brain Cogn
64. Passingham RE, Stephan KE, Kötter R 50:414–431
(2002) The anatomical basis of functional 78. Small SL, Nusbaum HC (2004) On the neu-
localization in the cortex. Nat Rev Neurosci robiological investigation of language under-
3:606–616 standing in context. Brain Lang 89:300–311
65. Büchel C, Dolan RJ, Armony JL, Friston KJ 79. Fahrenberg J, Myrtek M (eds) (2001)
(1999) Amygdala-hippocampal involvement in Progress in ambulatory assessment. Hogrefe
human aversive trace conditioning revealed and Huber, Seattle, WA
through event-related functional magnetic res- 80. Adolphs R (2002) Neural systems for recog-
onance imaging. J Neurosci 19:10869–10876 nizing emotion. Curr Opin Neurobiol
66. Foerster F (1995) On the problems of initial- 12:169–177
value-dependencies and measurement of 81. Deichmann R, Josephs O, Hutton C, Corfield
change. J Psychophysiol 9:324–341 DR, Turner R (2002) Compensation of
67. Peper M, Herpers M, Spreer J, Hennig J, susceptibility-induced BOLD sensitivity losses
Zentner J (2006) Functional neuroimaging in echo-planar fMRI imaging. NeuroImage
studies of emotional learning and autonomic 15:120–135
reactions. J Physiol Paris 99:342–354 82. Yablonskiy DA (1998) Quantitation of intrin-
68. Cardinal RN, Parkinson JA, Hall J, Everitt BJ sic magnetic susceptibility-related effects in a
(2002) Emotion and motivation: the role of tissue matrix. Phantom study. Magn Reson
the amygdala, ventral striatum, and prefrontal Med 39:417–428
cortex. Neurosci Biobehav Rev 26:321–352 83. Lipschutz B, Friston KJ, Ashburner J, Turner
69. O’Doherty J, Dayan P, Schultz J, Deichmann R, Price CJ (2001) Assessing study-specific
R, Friston K, Dolan RJ (2004) Dissociable regional variations in fMRI signal.
roles of ventral and dorsal striatum in instru- Neuroimage 13:392–398
mental conditioning. Science 304:452–454 84. Toga AW, Thompson PM (2001) Maps of the
70. Grandjean D, Sander D, Pourtois G et al brain. Anat Rec 265:37–53
(2005) The voices of wrath: brain responses 85. Hutton C, Bork A, Josephs O, Deichmann R,
to angry prosody in meaningless speech. Nat Ashburner J, Turner R (2002) Image distor-
Neurosci 8:145–146 tion correction in fMRI: a quantitative evalu-
71. Critchley H, Daly E, Phillips M et al (2000) ation. NeuroImage 16:217–240
Explicit and implicit neural mechanisms for 86. Raj D, Paley DP, Anderson AW, Kennan RP,
processing of social information from facial Gore JC (2000) A model for susceptibility
expressions: a functional Magnetic Resonance artefacts from respiration in functional echo-
Imaging study. Hum Brain Mapp 9:93–105 planar magnetic resonance imaging. Phys
72. Hariri S, Bookheiner SY, Mazziotta JC Med Biol 45:3809–3820
(2000) Modulating emotional responses: 87. Windischberger C, Langenberger H, Sycha T
effects of a neocortical network on the limbic et al (2002) On the origin of respiratory arti-
system. Neuroreport 11:43–48 facts in BOLD-EPI of the human brain.
73. Furmark T, Fischer H, Wik G, Larsson M, Magn Reson Imaging 20:575–582
Fredrikson M (1997) The amygdala and indi- 88. Wise RG, Ide K, Poulin MJ, Tracey I (2004)
vidual differences in human fear conditioning. Resting fluctuations in arterial carbon dioxide
Neuroreport 8:3957–3960 induce significant low frequency variations in
74. Canli T, Zhao Z, Desmond JE, Kang E, Gross BOLD signal. Neuroimage 21:1652–1664
J, Gabrieli JDE (2001) An fMRI study of per- 89. Dagli MS, Ingeholm JE, Haxby JV (1999)
sonality influences on brain reactivity to emo- Localization of cardiac-induced signal change
tional stimuli. Behav Neurosci 115:33–42 in fMRI. Neuroimage 9:407–415
fMRI of Emotion 491
90. Critchley HD, Rotshtein P, Nagai Y, 103. Beckmann CF, De Luca M, Devlin JT, Smith
O’Doherty J, Mathias CJ, Dolan RJ (2005) SM (2005) Investigations into resting-state
Activity in the human brain predicting differ- connectivity using independent component
ential heart rate responses to emotional facial analysis. Philos Trans R Soc Lond B Biol Sci
expressions. Neuroimage 24:751–762 360:1001–1013
91. Frysinger RC, Harper RM (1989) Cardiac 104. Damoiseaux JS, Rombouts SA, Barkhof F
and respiratory correlations with unit dis- et al (2006) Consistent resting-state networks
charge in human amygdala and hippocampus. across healthy subjects. Proc Natl Acad Sci U
Electroencephalogr Clin Neurophysiol S A 103:13848–13853
72:463–470 105. De Luca M, Beckmann CF, De Stefano N,
92. Shmueli K, van Gelderen P, de Zwart JA et al Matthews PM, Smith SM (2006) fMRI rest-
(2007) Low-frequency fluctuations in the ing state networks define distinct modes of
cardiac rate as a source of variance in the long-distance interactions in the human brain.
resting-state fMRI BOLD signal. Neuroimage Neuroimage 29:1359–1367
38:306–320 106. Robinson S, Soldati N, Basso G et al (2008)
93. Cohen MA, Taylor JA (2002) Short-term car- A resting state network in the basal ganglia.
diovascular oscillations in man: measuring Proc Intl Soc Magn Res Med 16:746
and modelling the physiologies. J Physiol 107. Beckmann CF, Smith SM (2005) Tensorial
542:669–683 extensions of independent component analy-
94. Birn RM, Diamond JB, Smith MA, Bandettini sis for multisubject FMRI analysis.
PA (2006) Separating respiratory-variation- Neuroimage 25:294–311
related fluctuations from neuronal-activity- 108. Raichle M, MacLeod A, Snyder A, Powers W,
related fluctuations in fMRI. Neuroimage Gusnard D, Shulman G (2001) A default
31:1536–1548 mode of brain function. Proc Natl Acad Sci U
95. Edward V, Windischberger C, Cunnington R S A 98:676–682
et al (2000) Quantification of fMRI artifact 109. Shulman G, Fiez J, Corbetta M et al (1997)
reduction by a novel plaster cast head holder. Common blood flow changes across visual
Hum Brain Mapp 11:207–213 tasks: II. Decreases in cerebral cortex. J Cogn
96. Hajnal J, Myers R, Oatridge A, Schwieso J, Neurosci 9:648–663
Young I, Bydder G (1994) Artifacts due to 110. Merboldt KD, Fransson P, Bruhn H, Frahm
stimulus correlated motion in functional J (2001) Functional MRI of the human
imaging of the brain. Magn Reson Med amygdala? Neuroimage 14:253–257
31:283–291 111. Triantafyllou C, Hoge RD, Krueger G et al
97. Field A, Yen Y, Burdette J, Elster A (2000) (2005) Comparison of physiological noise at
False cerebral activation on BOLD functional 1.5 T, 3 T and 7 T and optimization of fMRI
MR images: study of low-amplitude motion acquisition parameters. Neuroimage 26:
weakly correlated to stimulus. AJNR Am 243–250
J Neuroradiol 21:1388–1396 112. Krasnow B, Tamm L, Greicius MD et al (2003)
98. Robinson S, Moser E (2004) Positive results Comparison of fMRI activation at 3 and 1.5 T
in amygdala fMRI: Emotion or head motion? during perceptual, cognitive, and affective pro-
NeuroImage 22:S47, WE 294 cessing. Neuroimage 18:813–826
99. Biswal B, Yetkin FZ, Haughton VM, Hyde JS 113. Dickerson BC, Wright CI, Miller S et al
(1995) Functional connectivity in the motor (2006) Ultrahigh-field differentiation of
cortex of resting human brain using echo- medial temporal lobe function: sub-millimeter
planar MRI. Magn Reson Med 34:537–541 fMRI of amygdala and hippocampal activa-
100. Lowe MJ, Mock BJ, Sorenson JA (1998) tion at 7 Tesla. NeuroImage 31:S154
Functional connectivity in single and mul- 114. Merboldt KD, Finsterbusch J, Frahm J (2000)
tislice echoplanar imaging using resting-state Reducing inhomogeneity artifacts in func-
fluctuations. Neuroimage 7:119–132 tional MRI of human brain activation-thin
101. Fox MD, Snyder AZ, Zacks JM, Raichle ME sections vs gradient compensation. J Magn
(2006) Coherent spontaneous activity accounts Reson 145:184–191
for trial-to-trial variability in human evoked 115. Bellgowan PS, Bandettini PA, van Gelderen
brain responses. Nat Neurosci 9:23–25 P, Martin A, Bodurka J (2006) Improved
102. Fox MD, Snyder AZ, Vincent JL, Raichle ME BOLD detection in the medial temporal
(2007) Intrinsic fluctuations within cortical region using parallel imaging and voxel
systems account for intertrial variability in volume reduction. Neuroimage 29:
human behavior. Neuron 56:171–184 1244–1251
492 Simon Robinson et al.
116. Triantafyllou C, Hoge RD, Wald LL (2006) 129. Haacke E, Brown R, Thompson M,
Effect of spatial smoothing on physiological Venkatesan R (1999) Magnetic resonance
noise in high-resolution fMRI. Neuroimage imaging: physical principles and sequence
32:551–557 design. Wiley-Liss, New York, NY
117. Robinson S, Windischberger C, Rauscher A, 130. Scouten A, Papademetris X, Constable RT
Moser E (2004) Optimized 3 T EPI of the (2006) Spatial resolution, signal-to-noise
amygdalae. NeuroImage 22:203–210 ratio, and smoothing in multi-subject func-
118. Frahm J, Merboldt KD, Hänicke W (1988) tional MRI studies. Neuroimage
Direct FLASH MR imaging of magnetic field 30:787–793
inhomogeneities by gradient compensation. 131. Parrish T, Gitelman D, LaBar K, Mesulam M
Magn Reson Med 6:474–480 (2000) Impact of signal-to-noise on func-
119. Deichmann R, Gottfried JA, Hutton C, tional MRI. Magn Reson Med 44:925–932
Turner R (2003) Optimized EPI for fMRI 132. LaBar K, Gitelman D, Mesulam M, Parrish T
studies of the orbitofrontal cortex. (2001) Impact of signal-to-noise on func-
Neuroimage 19:430–441 tional MRI of the human amygdala.
120. Cho Z, Ro Y (1992) Reduction of suscepti- Neuroreport 12:3461–3464
bility artifact in gradient-echo imaging. Magn 133. Robinson S, Hoheisel B, Windischberger C,
Reson Med 23:193–200 Habel U, Lanzenberger R, Moser E (2005)
121. Stenger VA, Boada FE, Noll DC (2000) FMRI of the emotions, towards an improved
Three-dimensional tailored RF pulses for the understanding of amygdala function. Curr
reduction of susceptibility artifacts in T(*) Med Imaging Rev 1:115–129
(2)-weighted functional MRI. Magn Reson 134. Morawetz C, Holz P, Lange C et al (2008)
Med 44:525–531 Improved functional mapping of the human
122. Yip CY, Fessler JA, Noll DC (2006) Advanced amygdala using a standard functional magnetic
three-dimensional tailored RF pulse for signal resonance imaging sequence with simple modi-
recovery in T2*-weighted functional mag- fications. Magn Reson Imaging 26:45–53
netic resonance imaging. Magn Reson Med 135. Stocker T, Kellermann T, Schneider F et al
56:1050–1059 (2006) Dependence of amygdala activation
123. De Panfilis C, Schwarzbauer C (2005) Positive on echo time: results from olfactory fMRI
or negative blips? The effect of phase encod- experiments. Neuroimage 30:151–159
ing scheme on susceptibility-induced signal 136. Gorno-Tempini M, Hutton C, Josephs O,
losses in EPI. Neuroimage 25:112–121 Deichmann R, Price C, Turner R (2002)
124. Ojemann JG, Akbudak E, Snyder AZ, Echo time dependence of BOLD contrast and
McKinstry RC, Raichle ME, Conturo TE susceptibility artifacts. NeuroImage
(1997) Anatomic localization and quantitative 15:136–142
analysis of gradient refocused echo-planar fMRI 137. Sodickson DK, Manning WJ (1997)
susceptibility artifacts. Neuroimage 6:156–167 Simultaneous acquisition of spatial harmonics
125. Chen N, Dickey CC, Guttman CRG, Panych (SMASH): fast imaging with radiofrequency
LP (2003) Selection of voxel size and slice ori- coil arrays. Magn Reson Med 38:591–603
entation for fMRI in the presence of suscepti- 138. Pruessmann K, Weiger M, Scheidegger M,
bility field gradients: application to imaging of Boesiger P (1999) SENSE: sensitivity encod-
the amygdala. NeuroImage 19:817–825 ing for fast MRI. Magn Reson Med
126. Weiskopf N, Hutton C, Josephs O, Deichmann 42:952–962
R (2006) Optimal EPI parameters for reduc- 139. Lutcke H, Merboldt KD, Frahm J (2006)
tion of susceptibility-induced BOLD sensitiv- The cost of parallel imaging in functional
ity losses: a whole-brain analysis at 3 T and 1.5 MRI of the human brain. Magn Reson
T. Neuroimage 33:493–504 Imaging 24:1–5
127. Young IR, Cox IJ, Bryant DJ, Bydder GM 140. Schmidt CF, Degonda N, Luechinger R,
(1988) The benefits of increasing spatial reso- Henke K, Boesiger P (2005) Sensitivity-
lution as a means of reducing artifacts due to encoded (SENSE) echo planar fMRI at 3 T in
field inhomogeneities. Magn Reson Imaging the medial temporal lobe. NeuroImage
6:585–590 25:625–641
128. Hyde S, Biswal B, Jesmanowicz A (2001) 141. Fürsatz M, Windischberger C, Karlsson KÆ,
High-resolution fMRI using multislice partial Moser E (2008) Successful fMRI of the
k-space GR-EPI with cubic voxels. Magn hypothalamus at 3T. Proc Intl Soc Magn
Reson Med 46:114–125 Reson Med 16:2501
fMRI of Emotion 493
142. Fürsatz M, Windischberger C, Karlsson KÆ, 156. Witoszynskyj S, Rauscher A, Reichenbach JR,
Mayr W, Moser E. Valence-dependent modu- Barth M (2007) ΠUN (Πhase UNwrapping)
lation of hypothalamic activity. Neuroimage validation of a 2D region-growing phase
(in press) unwrapping program. Proc Intl Soc Magn
143. Dowell NG, Tofts PS (2007) Fast, accurate, Reson Med 15:3436
and precise mapping of the RF field in vivo 157. Windischberger C, Robinson S, Rauscher A,
using the 180 degrees signal null. Magn Barth M, Moser E (2004) Robust field map
Reson Med 58:622–630 generation using a triple-echo acquisition.
144. Speck O, Hennig J (1998) Functional imag- J Magn Reson Imaging 20:730
ing by I0- and T2*-parameter mapping using 158. Roemer PB, Edelstein WA, Hayes CE, Souza
multi-image EPI. Magn Reson Med SP, Mueller OM (1990) The NMR phased
40:243–248 array. Magn Reson Med 16:192–225
145. Posse S, Wiese S, Gembris D et al (1999) 159. Bernstein MA, Grgic M, Brosnan TJ, Pelc NJ
Enhancement of BOLD-contrast sensitivity (1994) Reconstructions of phase contrast,
by single-shot multi-echo functional MR phased array multicoil data. Magn Reson Med
imaging. Magn Reson Med 42:87–97 32:330–334
146. Posse S, Shen Z, Kiselev V, Kemna LJ (2003) 160. Hammond KE, Lupo JM, Xu D et al (2008)
Single-shot T(2)* mapping with 3D compen- Development of a robust method for generat-
sation of local susceptibility gradients in mul- ing 7.0 T multichannel phase images of the
tiple regions. Neuroimage 18:390–400 brain with application to normal volunteers
147. Posse S, Holten D, Gao K, Rick J, Speck O and patients with neurological diseases.
(2006) Evaluation of interleaved XYZ- Neuroimage 39:1682–1692
shimming with multi-echo EPI in prefrontal 161. Robinson S, Jovicich J (2008) EPI distortion
cortex and amygdala at 4 Tesla. NeuroImage corrections at 4 T: Multi-channel field map-
31:S154 ping and a comparison with the point-spread
148. Weiskopf N, Klose U, Birbaumer N, Mathiak function method. Proc Intl Soc Magn Reson
K (2005) Single-shot compensation of image Med 16:3031
distortions and BOLD contrast optimization 162. Robson MD, Gore JC, Constable RT (1997)
using multi-echo EPI for real-time Measurement of the point spread function in
fMRI. Neuroimage 24:1068–1079 MRI using constant time imaging. Magn
149. Glover GH, Law CS (2001) Spiral-in/out Reson Med 38:733–740
BOLD fMRI for increased SNR and reduced 163. Zeng H, Constable RT (2002) Image distor-
susceptibility artifacts. Magn Reson Med tion correction in EPI: Comparison of field
46:515–522 mapping with point spread function mapping.
150. Guo H, Song AW (2003) Single-shot spiral Magn Reson Med 48:137–146
image acquisition with embedded z-shimming 164. Zaitsev M, Hennig J, Speck O (2004) Point
for susceptibility signal recovery. J Magn spread function mapping with parallel imag-
Reson Imaging 18:389–395 ing techniques and high acceleration factors:
151. Truong TK, Song AW (2008) Single-shot fast, robust, and flexible method for echo-
dual-z-shimmed sensitivity-encoded spiral- planar imaging distortion correction. Magn
in/out imaging for functional MRI with
Reson Med 52:1156–1166
reduced susceptibility artifacts. Magn Reson 165. Hu X, Kim SG (1994) Reduction of signal
Med 59:221–227 fluctuation in functional MRI using navigator
152. Li Z, Wu G, Zhao X, Luo F, Li SJ (2002) echoes. Magn Reson Med 31:495–503
Multiecho segmented EPI with z-shimmed 166. Bruder H, Fischer H, Reinfelder HE,
background gradient compensation Schmitt F (1992) Image reconstruction for
(MESBAC) pulse sequence for fMRI. Magn echo planar imaging with nonequidistant
Reson Med 48:312–321 k-space sampling. Magn Reson Med
153. Weisskoff RM, Davis TL (1992) Correcting 23:311–323
gross distortion on echo planar images. Paper 167. Barry RL, Klassen LM, Williams JM, Menon RS
presented at the SMRM, Berlin (2008) Hybrid two-dimensional navigator cor-
154. Jezzard P, Balaban RS (1995) Correction for rection: a new technique to suppress respiratory-
geometric distortion in echo planar images from induced physiological noise in multi-shot
B0 field variations. Magn Reson Med 34:65–73 echo-planar functional MRI. Neuroimage
155. Jenkinson M (2003) Fast, automated, 39:1142–1150
N-dimensional phase-unwrapping algorithm. 168. Lund TE, Madsen KH, Sidaros K, Luo WL,
Magn Reson Med 49:193–197 Nichols TE (2006) Non-white noise in fMRI:
494 Simon Robinson et al.
does modelling have an impact? Neuroimage 181. Preibisch C, Raab P, Neumann K et al (2003)
29:54–66 Event-related fMRI for the suppression of
169. Glover GH, Li TQ, Ress D (2000) Image-based speech-associated artifacts in stuttering.
method for retrospective correction of physio- Neuroimage 19:1076–1084
logical motion effects in fMRI: RETROICOR. 182. Birn RM, Bandettini PA, Cox RW, Shaker R
Magn Reson Med 44:162–167 (1999) Event-related fMRI of tasks involving
170. Windischberger C, Friedreich S, Hoheisel B, brief motion. Hum Brain Mapp 7:106–114
Moser E (2004) The importance of correct- 183. Phelps EA, O’Connor KJ, Cunningham WA
ing for physiological artifacts for functional et al (2000) Performance on indirect mea-
MRI in deep brain structures. NeuroImage sures of race evaluation predicts amygdala
22:S28 activation. J Cogn Neurosci 12:729–738
171. Josephs O, Howseman A, Friston K, Turner 184. Winston JS, Strange BA, O’Doherty J, Dolan
R (1997) Physiological noise modelling for RJ (2002) Automatic and intentional brain
multi-slice EPI fMRI using SPM. Proc Intl responses during evaluation of trustworthi-
Soc Magn Reson Med 5:1682 ness of faces. Nat Neurosci 5:277–283
172. Weissenbacher A, Windischberger C, 185. Singer T, Kiebel SJ, Winston JS, Dolan RJ,
Lanzenberger R, Moser E (2008) Efficient Frith CD (2004) Brain responses to the acquired
correction for artificial signal fluctuations in moral status of faces. Neuron 41:653–662
resting-state fMRI-data. Proc Intl Soc Magn 186. Campbell DT, Stanley JC (1966)
Reson Med 16:2467 Experimental and quasi-experimental designs
173. Beckmann CF, Smith SM (2004) Probabilistic for research. Houghton Mifflin, Boston
independent component analysis for func- 187. Poldrack RA, Wagner AD (2004) What can
tional magnetic resonance imaging. IEEE neuroimaging tell us about the mind? Insights
Trans Med Imaging 23:137–152 from prefrontal cortex. Curr Dir Psychol Sci
174. Calhoun V, Adali T, Stevens M, Kiehl K, 13:177–181
Pekar J (2005) Semi-blind ICA of fMRI: a 188. Logothetis NK, Pauls J, Augath M, Trinath
method for utilizing hypothesis-derived time T, Oeltermann A (2001) Neurophysiological
courses in a spatial ICA analysis. Neuroimage investigation of the basis of the fMRI signal.
25:527–538 Nature 412:150–157
175. Thomas CG, Harshman RA, Menon RS 189. Janz C, Heinrich SP, Kornmayer J, Bach M,
(2002) Noise reduction in BOLD-based Hennig J (2001) Coupling of neural activity
fMRI using component analysis. Neuroimage and BOLD fMRI response: new insights by
17:1521–1537 combination of fMRI and VEP experiments
176. Kochiyama T, Morita T, Okada T, Yonekura Y, in transition from single events to continu-
Matsumura M, Sadato N (2005) Removing the ous stimulation. Magn Reson Med 46:
effects of task-related motion using independent- 482–486
component analysis. Neuroimage 25:802–814 190. Baas D, Aleman A, Kahn RS (2004)
177. Perlbarg V, Bellec P, Anton JL, Pelegrini-Issac Lateralization of amygdala activation: a sys-
M, Doyon J, Benali H (2007) CORSICA: tematic review of functional neuroimaging
correction of structured noise in fMRI by studies. Brain Res Brain Res Rev 4:96–103
automatic identification of ICA components. 191. Robinson S, Pripfl J, Bauer H, Moser M
Magn Reson Imaging 25:35–46 (2005) Empirical evidence for the minimum
178. Tohka J, Foerde K, Aron AR, Tom SM, Toga voxel size required for reliable 3 T fMRI of
AW, Poldrack RA (2008) Automatic indepen- the amygdala. NeuroImage 26:S795
dent component labeling for artifact removal 192. Habel U, Windischberger C, Derntl B et al
in fMRI. Neuroimage 39:1227–1245 (2007) Amygdala activation and facial expres-
179. Bullmore ET, Brammer MJ, Rabe-Hesketh S sions: explicit emotion discrimination versus
et al (1999) Methods for diagnosis and treat- implicit emotion processing.
ment of stimulus-correlated motion in generic Neuropsychologia 45:2369–2377
brain activation studies using fMRI. Hum 193. Sack AT, Linden DE (2003) Combining transcra-
Brain Mapp 7:38–48 nial magnetic stimulation and functional imaging
180. Morgan VL, Dawant BM, Li Y, Pickens DR in cognitive brain research: possibilities and limita-
(2007) Comparison of fMRI statistical soft- tions. Brain Res Brain Res Rev 43:41–56
ware packages and strategies for analysis of 194. Thayer JF, Brosschot JF (2005)
images containing random and stimulus- Psychosomatics and psychopathology: look-
correlated motion. Comput Med Imaging ing up and down from the brain.
Graph 31:436–446 Psychoneuroendocrinology 30:1050–1058
Chapter 16
fMRI of Pain
Emma G. Duerden, Roberta Messina, Maria A. Rocca,
Massimo Filippi, and Gary H. Duncan
Abstract
Pain was first considered to be a hard-wired system in which noxious input was passively transmitted along
sensory channels to the brain. However, today it is generally accepted that the experience of pain is not
simply driven by noxious stimulus characteristics, but that the brain is the structure where the subjective
perception of pain emerges and is critically linked with other cognitive processes.
The field of pain research has progressed immensely due to the advancement of brain imaging tech-
niques. The initial goal of this research was to expand our understanding of the cerebral mechanisms
underlying the perception of pain; more recently the research objectives have shifted toward chronic
pain—understanding its origins, developing methods for its diagnosis, and exploring potential avenues for
its treatment. While several different neuroimaging approaches have certain advantages for the study of
pain, fMRI has ultimately become the most widely utilized imaging technique over the past decade because
of its noninvasive nature, high-temporal and spatial resolution, and general availability; thus, the following
chapter will focus on fMRI and the special aspects of this technique that are particular to pain research.
1 Introduction
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_16, © Springer Science+Business Media New York 2016
495
496 Emma G. Duerden et al.
2.1 Nociceptive For cortical nociceptive processing related to cutaneous heat stimuli,
BOLD Signal the hemodynamic response function (HRF) peaks slightly later and
fMRI of Pain 497
2.2 BOLD fMRI A newly developing field in pain fMRI is spinal cord imaging, which
of Spinal Nociceptive is crucial for a better understanding of central nervous system (CNS)
Signals pain processing. The spinal cord and brainstem receive input from
the periphery before relaying this information on to the cortex. These
subcortical regions are involved in the modulation of nociceptive
input and the potentially abnormal processing of that input that may
lead to chronic pain syndromes. Therefore, knowledge concerning
the peripheral mechanisms of nociceptive processing is crucial to
understanding a number of pathological pain conditions resulting
from nerve injury or inflammation. These factors contribute to the
generation and maintenance of two key components of chronic pain,
namely hyperalgesia and allodynia. Hyperalgesia is the phenomenon
where an exaggerated response occurs after exposure to a noxious
stimulus. Allodynia is an exaggerated response toward nonpainful
498 Emma G. Duerden et al.
3.1 Pain Assessment A key issue in functional imaging of the cortical nociceptive signal
is to ensure that the stimuli delivered to the subjects are perceived
as noxious. Pain thresholds are commonly determined during a
separate session prior to the scan. This procedure also serves to
familiarize participants with the stimuli and reduce anxiety, thereby
minimizing anxiety-related fluctuations in cardiovascular activity
[16]. Stimuli utilized for the scanning session are frequently tai-
lored to each individual’s pain threshold; conversely, all subjects
can be administered the same level of noxious stimulation, which
has been determined to evoke the perception of pain in all subjects.
A corollary to the appropriate choice of noxious stimuli is the con-
firmation that predetermined levels of stimulation are actually per-
ceived as painful, within the scanning environment. A number of
contextual factors can alter the perception of stimuli that were
originally considered painful during a pre-scanning test, including
the temperature of the scanning suite, the position of the body in
the scanner, and distractions of noise, possible feelings of claustro-
phobia, and other conditions specific to the scanning paradigm.
fMRI of Pain 499
Before describing how fMRI measures the cortical and spinal noci-
ceptive signal, it is important to understand how this signal is
transferred to the cortex. In the periphery, a painful stimulus
applied to the body is transmitted to the CNS through nociceptors
[31]. Myelinated A-delta fibers transmit sharp pricking pain [32],
while unmyelinated C-fibers transmit slow burning pain, often
referred to as second pain [33]. The cell bodies of A-delta and
C-fibers are located in the dorsal root ganglia, receiving afferent
input from the periphery and then sending the information into
the spinal cord to terminate in the dorsal horn [34, 35]. Axons
from the second-order dorsal horn neurons rise through several
ascending pathways that transmit nociceptive information to the
thalamus, reticular formation, and cortex [8, 36]. Pain and tem-
perature information applied to the face is relayed through cranial
nerves to the spinal nucleus V terminating in the thalamus via the
trigeminothalamic tract, which is then relayed to the cortex.
fMRI of Pain 501
4.1 Supraspinal During the past 3–5 years, neuroimaging studies have extensively
Processing investigated the neural basis of pain perception, thus showing that
of Nociceptive Stimuli nociceptive stimuli commonly elicit activity within a very wide array
of subcortical and cortical brain structures [44, 45]. Regions most
frequently activated by painful stimuli include primary somatosen-
sory cortex (SI), SII, ACC, the insula, the PFC, and the thalamus.
Regions responsible for pain processing are categorized along
two functional lines—the first being the sensory-discriminative (lat-
eral pain system) component involved in the perception of temporal,
intensity, and localization aspects of pain processing, and the second,
the affective-motivational (medial) component associated with the
emotional aspects of pain [46, 47]. Dissociations between the two
systems are made through subjective reports on pain scales. After
exposure to noxious stimuli, subjects are asked to quantify separately
how intense and how unpleasant is the perceived pain. Regions
implicated in the lateral pain system include SI, SII, posterior insula,
and lateral thalamus, while the medial pain system consists of the
medial thalamic nuclei, the ACC, and the PFC. Much of what is
known regarding the two components in pain processing was ini-
tially explored through single-unit recordings in nonhuman pri-
mates and lesion studies in humans. However, the more recent
ability to study these functional components noninvasively in
humans using fMRI and other brain mapping techniques has allowed
pain researchers to advance rapidly in their understanding of the role
of these cortical regions in pain processing and how they interact.
4.1.1 Primary SI is located in the postcentral gyrus, is composed of four areas (areas
Somatosensory Cortex 3a, 3b, 1, and 2) [48], and is involved in the processing of both
tactile and noxious stimuli [49]. It was long debated whether SI was
necessary to perceive pain. Early studies of patients with brain lesions
suggested that deficits in nociceptive processing were rather com-
mon following lesions to the thalamus, but were very rare when
damage was restricted to the area believed to incorporate SI [2].
Likewise, later studies, using electrical stimulation of the human cor-
tex during awake brain surgery, reported that direct stimulation of
SI rarely evoked any perception of pain in patients [1].
The advent of imaging technology allowed a more global
exploration of the role of SI and other cortical regions involved in
pain processing, and these studies could be conducted in healthy
volunteers, rather than in patients with brain injuries that might
alter normal function. The first of these studies involved PET and
demonstrated that noxious stimuli applied to the hands were
fMRI of Pain 503
4.1.2 Secondary SII is also considered to be an important region for processing the
Somatosensory Cortex (SII) sensory-discriminative component of pain. SII is located in the
parietal operculum in the dorsal bank of the lateral sulcus. Like SI,
this region receives projections from the ventroposterior lateral
nucleus (VPL) of the thalamus, but its major nociceptive input
comes directly from the ventroposterior inferior (VPI) nucleus
[52]. Studies of patients with lesions that include SII have demon-
strated deficits in the perception of pain intensity [67, 68]; how-
ever, lesions comprised additional cortical regions that may work in
concert with SII to process this piece of information. In addition
to these clinical findings, converging evidence from a number of
studies supports the notion that SII possesses a functional capacity
to discriminate between different intensities of noxious stimuli pre-
sented to the contralateral side of the body. Evidence from PET
provides a role for this region in intensity processing in that sub-
jects’ ratings of pain intensity in response to thermal heat pain have
been shown to be highly correlated with activation of SII [69].
Additionally, an fMRI study by Maihofner et al. [70] found
increased activation in SII in response to painful mechanical stimuli
compared to thermal heat pain. In turn, ratings of subjective inten-
sity were correlated with the intensity of mechanical pain. However,
dissociative processing was noted in this region as ratings of
unpleasantness were not found to correlate with SII activation.
Contrary to these findings, evidence from fMRI suggests this
region may be involved in some emotional aspects of pain process-
ing. For example, Gracely et al. [71] found that fibromyalgia
patients who scored higher on a pain catastrophizing questionnaire
showed increased activation in both the ACC and SII in response
to noxious stimuli. Catastrophizing (and in turn anxiety about
painful stimuli) is inherently linked with pain perception, where
the individual’s emotional state augments neural processing of
these stimuli. In line with these findings are data that show
increased activity in SII during the anticipation of painful stimuli,
indicative of an enhanced emotional response [66, 72].
4.1.3 Insular Cortex The insula is extensively connected to other brain regions such as
the prefrontal cortex, cingulum, amygdala, SI, SII, and also tha-
lamic nuclei (VPI, the centromedian-parafasicular, the medial dor-
sal [MD], and the ventral medial posterior [Vmpo] nuclei). It may
fMRI of Pain 505
4.1.4 Anterior Cingulate The ACC plays a prominent role in pain processing. This region
Cortex (ACC) receives thalamo-cortical input from nociceptive neurons in the
thalamus and contains nociceptive-specific neurons responsive to
noxious stimuli [87]. Additionally, the ACC is implicated in medi-
ating antinociceptive responses as it contains high numbers of opi-
ate receptors [88, 89].
Historically, the ACC was considered key to affective process-
ing, as it was classified along with the retrosplenial cortex, hippo-
campus, amygdala, and several basal forebrain structures as part of
506 Emma G. Duerden et al.
Fig. 1 Pain (a) and temperature (d) processing brain regions: brain areas with significantly increased activation
during noxious than innocuous stimulation and during warm stimulation are coded red, while regions coded
blue show significantly increased activation during innocuous than noxious stimulation and during cold stimu-
lation. Insular clusters (seed clusters) with pain- (b) and temperature- (e) specific activity divided in aINS
(green and yellow) and pINS (red and blue), that were used for the insular functional connectivity analysis: both
aINS and pINS were functionally connected to a large brain network, which predominantly includes areas
involved in nociception and thermoception (SI, SII, cingulate gyrus, PFC, and parietal association cortices).
Comparison of pain- (c) and temperature- (f) specific functional connectivity of the two insular areas (areas
with significantly stronger functional connectivity to aINS than to pINS are coded red–yellow, while areas with
significantly stronger functional connectivity to pINS than to aINS are coded blue–green): the aINS was more
strongly connected to PFC and to ACC than was pINS; pINS meanwhile was more strongly connected to SI and
to the primary motor cortex. From [79] with permission
4.1.5 Prefrontal Cortex Regions of the PFC have been implicated in both pain processing
(PFC) and pain modulation. PFC activation seen in brain imaging studies
of pain is believed to reflect attention toward the stimuli [69, 101],
but it has also been shown to be directly involved in modulating
responses to painful stimuli. Recently, Lobanov et al. [102] dem-
onstrated that attention to both spatial and intensity feature of the
noxious stimulus was associated with activation of fronto-parietal
areas, including the PFC.
Functional imaging studies have shown that activity in the
PFC reduces the pain magnitude or hyperalgesia, suggesting that
the PFC can regulate the amount of pain an individual perceives.
Activity in the PFC is also associated with episodes of emotional
detachment, when the “suffering” element of pain is absent.
Negative emotional responses can heighten the experience of pain,
and the ventrolateral PFC seems to regulate these responses via
interactions with the nucleus accumbens and the amygdala [103].
508 Emma G. Duerden et al.
4.1.6 Amygdala The amygdala, buried beneath the uncus and located at the tail of
the caudate nucleus, is a key limbic structure involved in the pro-
cessing of emotional stimuli. The amygdala is suited for such pro-
cessing as it is the sole subcortical structure to receive projections
from every sensory area.
Functional neuroimaging studies utilizing various types of
aversive stimuli including pain, habitually report amygdala activa-
tion [105]. Studies using fMRI have demonstrated that amygdala
activation is associated with extremely unpleasant noxious stimuli,
suggesting an involvement of this region in processing the affective
component of pain [106, 107]. Other evidence from fMRI has
implicated the amygdala in processing uncertainty associated with
painful stimuli [108].
Fig. 2 (a) Correlation analysis between the BOLD response in the PAG and pain threshold, recorded in seconds,
during the cold pressor test: the pain threshold directly correlated with BOLD activation in the PAG (cluster level
corrected threshold p < 0.05, Pearson’s r = 0.63). (b) Correlation analysis between the BOLD response in the
PAG and pain intensity ratings, as assessed by the numerical rating scale, during the cold pressor test: the pain
rating inversely correlated with BOLD activation in the PAG (cluster level corrected threshold p < 0.05, Pearson’s
r = 0.45). From [113] with permission
4.1.8 Motor Cortices A number of other cortical and subcortical regions are commonly
activated during fMRI studies of pain including many regions
involved in motor processing. Motor regions include the primary
motor cortex, premotor cortex, supplementary motor area,
cerebellum, and basal ganglia. Frequently, these regions are con-
comitantly activated along with those involved with affective and
sensory aspects of pain processing [117].
The perception of a painful stimulus involves an orienting
response and subsequent retraction of the body part being tar-
geted. Activation of motor areas during functional neuroimaging
studies is believed to reflect motor preparatory responses. However,
several of these areas, such as the nuclei associated with the basal
ganglia, are directly responsive to noxious stimuli [118, 119].
Using fMRI, a reliable somatotopic organization has been shown
in the putamen [120, 121] in response to noxious stimuli, which
indicates that this region may be involved in sensory-discriminative
processing of pain.
4.2 Spinal Cord To date, only a few reports have assessed the feasibility of studying
Processing nociception using fMRI of the spinal cord [123]. One study by
of Nociceptive Stimuli Brooks et al. [14] examined the spinal nociceptive signal at 1.5 T
in response to noxious heat pain stimuli. Using a tailored, high-
resolution scanning protocol and postprocessing techniques for
controlling physiological noise, they demonstrated reliable pain-
related activation in the ipsilateral dorsal horn.
A recent connectivity analysis revealed functional coupling
between the spinal cord dorsal horn and typical ascending thalamo-
cortical pain pathways. More importantly, the spinal cord was also
functionally connected with brain regions involved in descending
pain modulation, such as the PAG. A positive correlation between
the individual strength of connectivity within this descending pain
modulatory pathway and the behavioral pain ratings was found,
thus pointing to the functional relevance of this system during the
processing of physiological nociceptor pain [124]. In a similar
study, BOLD fMRI response in the spinal cord was correlated with
individual pain ratings, further supporting the contribution of
spinal cord activity to the perception of pain [125].
fMRI of Pain 511
5.1 Pain Modulation What is clear from several studies is that nociceptive information
processing, and consequent pain perception, is subject to signifi-
cant pro- and anti-nociceptive modulations that can be influenced
dramatically by cognitive, emotional, and contextual factors [126].
Pain modulation can occur through both endogenous mecha-
nisms and as a result of exogenously administered agents. One final
common pathway for analgesic mechanisms is believed to be through
the release of endogeneous opioids [127] acting on sites in the
brainstem and midbrain that block the nociceptive signal through
their descending pathways; the final effects of this descending mod-
ulation are exerted either on the spinal cord and/or at the site of
peripheral nerves that transmit the nociceptive stimuli. Additionally,
recent research has implicated endocannibinoids in pain modula-
tion, which may act on similar descending pathways [128].
fMRI is a useful tool for examining cerebral mechanisms of pain
modulation, whereby subjects experience either analgesia or hyperal-
gesia—a decrease or increase in perceived pain, respectively. First, the
anatomical resolution of fMRI is sufficient to localize some of the
small brain regions involved in pain modulation, such as the RVM or
PAG [112, 129], and the temporal resolution allows an assessment of
the time course of activations within those regions. fMRI is also well
suited to study procedures that evoke changes in pain perception
since it accommodates the use of parametric data, whereby experi-
mental parameters such as pain ratings (intensity, expectation,
unpleasantness) can be correlated with brain activations and thus
used to characterize cortical structures according to their response
profile to various experimental parameters. As a corollary of increased
temporal resolution, a major advantage of using fMRI to study pain
modulation is the possibility of utilizing event-related designs
whereby the time course of brain activations over different phases of
the modulation period can be studied—the anticipation of the nox-
ious stimulus, the onset of pain perception, changes in pain percep-
tion over time, and post-stimulus ratings. Anticipation of the painful
stimulus is a crucial phase of the pain modulation process, since at
this time point neural mechanisms act on descending modulatory
systems to diminish or enhance the response to the stimulus [130].
fMRI has been widely applied to study modulatory processes
triggered either through endogenous mechanisms utilizing cogni-
tive strategies, such as attention [77, 131], hypnosis [132–134],
placebo and nocebo effects [104, 135], or through exogenous
agents, such as pharmacological [126, 136–138] and non pharma-
cological interventions [139].
Several lines of evidence strengthen the notion that pain mod-
ulation occurs via an integrated “frontal to brainstem to spinal
cord” system. Disruption of the descending pain modulatory
system may represent a point of vulnerability for the development
and maintenance of chronic pain [126].
512 Emma G. Duerden et al.
5.2 Pain Empathy Inherent to processing the emotional component of pain is the
ability to understand the emotional reactions of other people who
are experiencing pain—i.e., pain empathy [140]. This rapidly
growing field of empathy research is directed toward studying the
mental representation of pain—both that which is perceived to be
experienced by others, as well as that which is perceived as one’s
own. Several different types of experimental stimuli implicating
other people in pain have been used in these fMRI paradigms,
including photographic images [141–145], or short animations
[146] of body parts in potentially tissue-damaging situations, view-
ing the faces of actors evoking facial expressions of pain [147], or
subjects actually receiving painful stimuli [148], or those of chronic
pain patients [149], or being cued that a loved one in the room
was receiving painful stimuli [150].
A common finding from these studies is that the processing of
pain in others recruits brain regions involved in affective processing—
namely the ACC and insula. In a recent meta-analysis, Lamm et al.
[151] compiled brain activation coordinates from 32 studies that
had investigated empathy for pain using fMRI. Authors identified a
core network, consisting of bilateral anterior insular cortex and
medial/anterior cingulate cortex, that was associated with empathy
for pain. Activation in these areas overlapped with activation during
directly experienced pain, thus linking their involvement to repre-
senting global feeling states and the guidance of adaptive behavior
for both self- and other-related experiences. Moreover, the analysis
demonstrates that—depending on the type of experimental para-
digm—this core network was coactivated with distinct brain regions:
viewing pictures of body parts in painful situations recruited areas
underpinning action understanding (inferior parietal/ventral pre-
motor cortices) to a stronger extent; eliciting empathy by means of
abstract visual information about the other’s affective state more
strongly engaged areas associated with inferring and representing
mental states of self and others (precuneus, ventral medial prefrontal
cortex, superior temporal cortex, and temporo-parietal junction).
Several transcranial magnetic stimulation studies reported modu-
lation of sensory-discriminative regions associated with pain empathy,
thus suggesting a somatotopic specificity in the perceived pain of oth-
ers. Although early functional imaging studies suggested that somato-
sensory areas contribute very little to the neural response when seeing
others’ pain or empathizing with it, Morrison et al. have recently dem-
onstrated that just viewing others’ painful actions biases participants
to report tactile stimulation even when none occurred [145]. Such
discrepancies might have resulted from differences in experimental
paradigms, since it was shown that only the picture-based paradigms
activated somatosensory areas during empathy for pain [151].
fMRI has provided considerable insight into the neural mecha-
nisms of processing pain in others, and suggests a number of inter-
esting clinical implications. Since pain is a sensory and emotional
fMRI of Pain 513
9 Conclusions
References
1. Penfield W, Boldrey E (1937) Somatic motor to painful stimuli during sphygmomanom-
and sensory representation in the cerebral etry. Int J Psychophysiol 33(3):253–257
cortex of man as studied by electrical stimula- 17. Mobascher A et al (2010) Brain activation
tion. Brain 60(4):389–443 patterns underlying fast habituation to painful
2. Head H, Holmes G (1911) Sensory dis- laser stimuli. Int J Psychophysiol 75(1):16–24
turbances from cerebral lesions. Brain 18. Becerra LR et al (1999) Human brain acti-
34(2–3):102–254 vation under controlled thermal stimulation
3. Talbot J et al (1991) Multiple representations and habituation to noxious heat: an fMRI
of pain in human cerebral cortex. Science study. Magn Reson Med 41(5):1044–1057
251(4999):1355–1358 19. Price DD et al (1994) A comparison of pain
4. Jones AKP et al (1991) Cortical and subcor- measurement characteristics of mechanical
tical localization of response to pain in man visual analogue and simple numerical rating
using positron emission tomography. Proc R scales. Pain 56(2):217–226
Soc B Biol Sci 244(1309):39–44 20. Rainville P et al (2004) Rapid deterioration
5. Apkarian AV et al (1992) Persistent pain of pain sensory-discriminative information in
inhibits contralateral somatosensory cor- short-term memory. Pain 110(3):605–615
tical activity in humans. Neurosci Lett 21. Charron J, Rainville P, Marchand S (2006)
140(2):141–147 Direct comparison of placebo effects on
6. Davis KD et al (1995) fMRI of human clinical and experimental pain. Clin J Pain
somatosensory and cingulate cortex dur- 22(2):204–211
ing painful electrical nerve stimulation. 22. Price DD et al (1999) An analysis of factors
Neuroreport 7(1):321–325 that contribute to the magnitude of placebo
7. Flor H (2000) The functional organization analgesia in an experimental paradigm. Pain
of the brain in chronic pain. Prog Brain Res 83(2):147–156
129:313–322 23. Apkarian AV et al (1999) Differentiating corti-
8. Tracey I, Mantyh PW (2007) The cerebral cal areas related to pain perception from stim-
signature for pain perception and its modula- ulus identification: temporal analysis of fMRI
tion. Neuron 55(3):377–391 activity. J Neurophysiol 81(6):2956–2963
9. Wager TD et al (2013) An fMRI-based neu- 24. Porro CA et al (2004) Percept-related activ-
rologic signature of physical pain. N Engl ity in the human somatosensory system: func-
J Med 368(15):1388–1397 tional magnetic resonance imaging studies.
10. deCharms RC et al (2005) Control over Magn Reson Imaging 22(10):1539–1548
brain activation and pain learned by using 25. Andrew D, Greenspan JD (1999) Peripheral
real-time functional MRI. Proc Natl Acad Sci coding of tonic mechanical cutaneous pain:
102(51):18626–18631 comparison of nociceptor activity in rat
11. Rance M et al (2014) Real time fMRI feed- and human psychophysics. J Neurophysiol
back of the anterior cingulate and posterior 82(5):2641–2648
insular cortex in the processing of pain. Hum 26. Adriaensen H et al (1984) Nociceptor dis-
Brain Mapp 35(12):5784–5798 charges and sensations due to prolonged nox-
12. Chen JI et al (2002) Differentiating noxious- ious mechanical stimulation--a paradox. Hum
and innocuous-related activation of human Neurobiol 3(1):53–58
somatosensory cortices using temporal analy- 27. Gallez A et al (2005) Attenuation of sensory
sis of fMRI. J Neurophysiol 88(1):464–474 and affective responses to heat pain: evidence
13. Iramina K et al (1999) Effects of stimulus for contralateral mechanisms. J Neurophysiol
intensity on fMRI and MEG in somatosen- 94(5):3509–3515
sory cortex using electrical stimulation. IEEE 28. Bingel U et al (2007) Habituation to painful
Trans Magn 35(5):4106–4108 stimulation involves the antinociceptive sys-
14. Brooks J, Tracey I (2005) From nociception tem. Pain 131(1):21–30
to pain perception: imaging the spinal and 29. Valeriani M et al (2003) Reduced habituation
supraspinal pathways. J Anat 207(1):19–33 to experimental pain in migraine patients:
15. Mackey S et al (2006) FMRI evidence of nox- a CO2 laser evoked potential study. Pain
ious thermal stimuli encoding in the human 105(1):57–64
spinal cord. J Pain 7(4):S25 30. Nickel FT et al (2013) Brain correlates of
16. Rollnik JD, Schmitz N, Kugler J (1999) short-term habituation to repetitive electrical
Anxiety moderates cardiovascular responses noxious stimulation. Eur J Pain 18(1):56–66
fMRI of Pain 517
31. Willis WD Jr (1985) The pain system. The DR (ed) The skin senses. Charles C. Thomas
neural basis of nociceptive transmission in the Publishers, Springfield, IL, pp 423–443
mammalian nervous system. Pain Headache 47. Tracey I (2008) Imaging pain. Br J Anaesth
8:1–346 101(1):32–39
32. Adriaensen H et al (1983) Response properties 48. Kaas J et al (1979) Multiple represen-
of thin myelinated (A-delta) fibers in human tations of the body within the primary
skin nerves. J Neurophysiol 49(1):111–122 somatosensory cortex of primates. Science
33. Ochoa J, Torebjörk E (1989) Sensations 204(4392):521–523
evoked by intraneural microstimulation of 49. Kenshalo DR Jr, Isensee O (1983) Responses
C nociceptor fibres in human skin nerves. of primate SI cortical neurons to noxious
J Physiol 415(1):583–599 stimuli. J Neurophysiol 50(6):1479–1496
34. Cervero F, Iggo A (1980) The substantia 50. Casey KL et al (1996) Comparison of human
gelatinosa of the spinal cord: a critical review. cerebral activation pattern during cutane-
Brain 103(4):717–772 ous warmth, heat pain, and deep cold pain.
35. Wilson P, Kitchener PD (1996) Plasticity J Neurophysiol 76(1):571–581
of cutaneous primary afferent projections 51. Gelnar PA et al (1998) Fingertip representa-
to the spinal dorsal horn. Prog Neurobiol tion in the human somatosensory cortex: an
48(2):105–129 fMRI study. Neuroimage 7(4):261–283
36. Craig AD et al (1994) A thalamic nucleus 52. Liang M, Mouraux A, Iannetti GD (2011)
specific for pain and temperature sensation. Parallel processing of nociceptive and non-
Nature 372(6508):770–773 nociceptive somatosensory information in the
37. Legrain V et al (2011) The pain matrix human primary and secondary somatosensory
reloaded: a salience detection system for the cortices: evidence from dynamic causal mod-
body. Prog Neurobiol 93(1):111–124 eling of functional magnetic resonance imag-
38. Bromm B, Treede RD (1984) Nerve fibre ing data. J Neurosci 31(24):8976–8985
discharges, cerebral potentials and sensa- 53. Cheng JC et al (2015) Individual dif-
tions induced by CO2 laser stimulation. Hum ferences in temporal summation of pain
Neurobiol 3(1):33–40 reflect pronociceptive and antinociceptive
39. Carmon A, Dotan Y, Sarne Y (1978) brain structure and function. J Neurosci
Correlation of subjective pain experience with 35(26):9689–9700
cerebral evoked responses to noxious thermal 54. Derbyshire GSW, Jones PAK (1998) Cerebral
stimulations. Exp Brain Res 33(3–4):445–453 responses to a continual tonic pain stimulus
40. Iannetti GD et al (2004) Aδ nociceptor measured using positron emission tomogra-
response to laser stimuli: selective effect phy. Pain 76(1):127–135
of stimulus duration on skin temperature, 55. Disbrow E et al (1998) Somatosensory cor-
brain potentials and pain perception. Clin tex: a comparison of the response to noxious
Neurophysiol 115(11):2629–2637 thermal, mechanical, and electrical stimuli
41. Spiegel J, Hansen C, Treede RD (2000) using functional magnetic resonance imaging.
Clinical evaluation criteria for the assess- Hum Brain Mapp 6(3):150–159
ment of impaired pain sensitivity by thulium- 56. Vierck CJ et al (2013) Role of primary
laser evoked potentials. Clin Neurophysiol somatosensory cortex in the coding of pain.
111(4):725–735 Pain 154(3):334–344
42. Leandri M et al (2006) Measurement of skin 57. Tommerdahl M et al (1996) Anterior pari-
temperature after infrared laser stimulation. etal cortical response to tactile and skin-
Neurophysiol Clin 36(4):207–218 heating stimuli applied to the same skin site.
43. Helmchen C et al (2008) Common neural J Neurophysiol 75(6):2662–2670
systems for contact heat and laser pain stimu- 58. Bushnell MC et al (1999) Pain perception: is
lation reveal higher-level pain processing. there a role for primary somatosensory cor-
Hum Brain Mapp 29(9):1080–1091 tex? Proc Natl Acad Sci 96(14):7705–7709
44. Iannetti GD, Mouraux A (2010) From the 59. Seminowicz DA, Mikulis DJ, Davis KD
neuromatrix to the pain matrix (and back). (2004) Cognitive modulation of pain-related
Exp Brain Res 205(1):1–12 brain responses depends on behavioral strat-
45. Apkarian AV et al (2005) Human brain mech- egy. Pain 112(1):48–58
anisms of pain perception and regulation in 60. Oshiro Y et al (2007) Brain mechanisms
health and disease. Eur J Pain 9(4):463–484 supporting spatial discrimination of pain.
46. Melzack R, Casey KL (1968) Sensory, moti- J Neurosci 27(13):3388–3394
vational and central control determinants of 61. Andersson JLR et al (1997) Somatotopic
pain: a new conceptual model. In: Kenshalo organization along the central sulcus, for
518 Emma G. Duerden et al.
91. Pillay PK, Hassenbusch SJ (1992) Bilateral flow: an fMRI study. Neuropsychobiology
MRI-guided stereotactic cingulotomy for 43(3):175–185
intractable pain. Stereotact Funct Neurosurg 107. Berna C et al (2010) Induction of depressed
59(1–4):33–38 mood disrupts emotion regulation neurocir-
92. Gybels JM, Sweet WH (1989) Neurosurgical cuitry and enhances pain unpleasantness. Biol
treatment of persistent pain. Physiological Psychiatry 67(11):1083–1090
and pathological mechanisms of human pain. 108. Bornhovd K et al (2002) Painful stimuli evoke
Pain Headache 11:1–402 different stimulus–response functions in the
93. Vogt BA et al (1995) Human cingulate cor- amygdala, prefrontal, insula and somatosen-
tex: surface features, flat maps, and cytoarchi- sory cortex: a single-trial fMRI study. Brain
tecture. J Comp Neurol 359(3):490–506 125(6):1326–1336
94. Devinsky O, Morrell MJ, Vogt BA (1995) 109. Schulte LH, Sprenger C, May A (2015)
Contributions of anterior cingulate cortex to Physiological brainstem mechanisms of tri-
behaviour. Brain 118(1):279–306 geminal nociception: an fMRI study at
95. Vogt BA (2005) Pain and emotion interac- 3T. Neuroimage 124(Pt A):518–525
tions in subregions of the cingulate gyrus. 110. Mason P (2005) Deconstructing endog-
Nat Rev Neurosci 6(7):533–544 enous pain modulations. J Neurophysiol
96. Davis KD et al (1997) Functional MRI of 94(3):1659–1663
pain- and attention-related activations in 111. Fields HL, Heinricher MM (1985) Anatomy
the human cingulate cortex. J Neurophysiol and physiology of a nociceptive modulatory
77(6):3370–3380 system. Philos Trans R Soc Lond B Biol Sci
97. Rainville P et al (1997) Pain affect encoded in 308(1136):361–374
human anterior cingulate but not somatosen- 112. Fields HL (2000) Pain modulation: expecta-
sory cortex. Science 277(5328):968–971 tion, opioid analgesia and virtual pain. Prog
98. Wilcox CE et al (2015) The subjective expe- Brain Res 122:245–253
rience of pain: an FMRI study of percept- 113. La Cesa S et al (2014) fMRI pain activation in
related models and functional connectivity. the periaqueductal gray in healthy volunteers
Pain Med 16(11):2121–33 during the cold pressor test. Magn Reson
99. Arienzo D et al (2006) Somatotopy of ante- Imaging 32(3):236–240
rior cingulate cortex (ACC) and supplemen- 114. Dunckley P et al (2005) A comparison
tary motor area (SMA) for electric stimulation of visceral and somatic pain processing
of the median and tibial nerves: an fMRI in the human brainstem using functional
study. Neuroimage 33(2):700–705 magnetic resonance imaging. J Neurosci
100. Kroger IL, Menz MM, May A (2015) 25(32):7333–7341
Dissociating the neural mechanisms of 115. Tracey I, Iannetti GD (2006) Brainstem
pain consistency and pain intensity in the functional imaging in humans. Suppl Clin
trigemino-nociceptive system. Cephalalgia Neurophysiol 58:52–67
[published online before print October 22, 116. Guimaraes AR et al (1998) Imaging subcorti-
2015, doi:10.1177/0333102415612765] cal auditory activity in humans. Hum Brain
101. Casey KL (1999) Forebrain mechanisms of Mapp 6(1):33–41
nociception and pain: analysis through imag- 117. Farina S et al (2003) Pain-related modula-
ing. Proc Natl Acad Sci 96(14):7668–7674 tion of the human motor cortex. Neurol Res
102. Lobanov OV et al (2013) Frontoparietal 25(2):130–142
mechanisms supporting attention to loca- 118. Chudler EH, Dong WK (1995) The role of
tion and intensity of painful stimuli. Pain the basal ganglia in nociception and pain.
154(9):1758–1768 Pain 60(1):3–38
103. Tracey I (2011) Can neuroimaging studies 119. Borsook D et al (2010) A key role of the basal
identify pain endophenotypes in humans? Nat ganglia in pain and analgesia--insights gained
Rev Neurol 7(3):173–181 through human functional imaging. Mol Pain
104. Wager TD et al (2004) Placebo-induced 6:27
changes in FMRI in the anticipation and expe- 120. Tomycz ND, Friedlander RM (2011) The
rience of pain. Science 303(5661):1162–1167 experience of pain and the putamen: a
105. Zald DH (2003) The human amygdala and new link found with functional MRI and
the emotional evaluation of sensory stimuli. diffusion tensor imaging. Neurosurgery
Brain Res Rev 41(1):88–123 69(4):N12–N13
106. Schneider F et al (2001) Subjective ratings of 121. Bingel U et al (2004) Somatotopic repre-
pain correlate with subcortical-limbic blood sentation of nociceptive information in the
520 Emma G. Duerden et al.
putamen: an event-related fMRI study. Cereb 136. Maihöfner C et al (2007) Brain imaging
Cortex 14(12):1340–1345 of analgesic and antihyperalgesic effects of
122. Loggia ML et al (2015) Evidence for brain cyclooxygenase inhibition in an experimental
glial activation in chronic pain patients. Brain human pain model: a functional MRI study.
138(Pt 3):604–615 Eur J Neurosci 26(5):1344–1356
123. Cahill CM, Stroman PW (2011) Mapping of 137. Wise RG et al (2007) The anxiolytic effects
neural activity produced by thermal pain in the of midazolam during anticipation to pain
healthy human spinal cord and brain stem: a revealed using fMRI. Magn Reson Imaging
functional magnetic resonance imaging study. 25(6):801–810
Magn Reson Imaging 29(3):342–352 138. Sanders D et al (2015) Pharmacologic modu-
124. Sprenger C, Finsterbusch J, Buchel C (2015) lation of hand pain in osteoarthritis: a double-
Spinal cord-midbrain functional connectivity blind placebo-controlled functional magnetic
is related to perceived pain intensity: a com- resonance imaging study using naproxen.
bined spino-cortical FMRI study. J Neurosci Arthritis Rheumatol 67(3):741–751
35(10):4248–4257 139. Li K et al (2015) The effects of acupuncture
125. Khan HS, Stroman PW (2015) Inter- treatment on the right frontoparietal network
individual differences in pain processing in migraine without aura patients. J Headache
investigated by functional magnetic reso- Pain 16:518
nance imaging of the brainstem and spinal 140. Thompson E (2001) Empathy and conscious-
cord. Neuroscience 307:231–241 ness. J Conscious Stud 8(5–7):1–32
126. Bingel U, Tracey I (2008) Imaging CNS 141. Jackson PL et al (2006) Empathy examined
modulation of pain in humans. Physiology through the neural mechanisms involved
(Bethesda) 23:371–380 in imagining how I feel versus how you feel
127. Levine JD et al (1978) The narcotic antago- pain. Neuropsychologia 44(5):752–761
nist naloxone enhances clinical pain. Nature 142. Jackson PL, Meltzoff AN, Decety J (2005)
272(5656):826–827 How do we perceive the pain of others? A
128. Hohmann AG, Suplita RL (2006) window into the neural processes involved in
Endocannabinoid mechanisms of pain modu- empathy. Neuroimage 24(3):771–779
lation. AAPS J 8(4):E693–E708 143. Lamm C et al (2007) What are you feeling?
129. Tracey I et al (2002) Imaging attentional Using functional magnetic resonance imaging
modulation of pain in the periaqueductal gray to assess the modulation of sensory and affec-
in humans. J Neurosci 22(7):2748–2752 tive responses during empathy for pain. PLoS
130. Porro CA (2003) Functional imaging and One 2(12):e1292
pain: behavior, perception, and modulation. 144. Moriguchi Y et al (2006) Empathy and judg-
Neuroscientist 9(5):354–369 ing other’s pain: an fMRI study of alexi-
131. Bantick SJ et al (2002) Imaging how atten- thymia. Cereb Cortex 17(9):2223–2234
tion modulates pain in humans using func- 145. Morrison I et al (2013) “Feeling” others’
tional MRI. Brain 125(2):310–319 painful actions: the sensorimotor integration
132. Roder CH et al (2007) Pain response in of pain and action information. Hum Brain
depersonalization: a functional imaging study Mapp 34(8):1982–1998
using hypnosis in healthy subjects. Psychother 146. Morrison I, Peelen MV, Downing PE (2007)
Psychosom 76(2):115–121 The sight of others’ pain modulates motor
133. Schulz-Stübner S et al (2004) Clinical hypno- processing in human cingulate cortex. Cereb
sis modulates functional magnetic resonance Cortex 17(9):2214–2222
imaging signal intensities and pain percep- 147. Simon D et al (2006) Brain responses to
tion in a thermal stimulation paradigm. Reg dynamic facial expressions of pain. Pain
Anesth Pain Med 29(6):549–556 126(1):309–318
134. Nakata H, Sakamoto K, Kakigi R (2014) 148. Botvinick M et al (2005) Viewing facial expressions
Meditation reduces pain-related neural activ- of pain engages cortical areas involved in the direct
ity in the anterior cingulate cortex, insula, sec- experience of pain. Neuroimage 25(1):312–319
ondary somatosensory cortex, and thalamus. 149. Saarela MV et al (2007) The compassionate
Front Psychol 5:1489 brain: humans detect intensity of pain from
135. Tracey I (2010) Getting the pain you another’s face. Cereb Cortex 17(1):230–237
expect: mechanisms of placebo, nocebo 150. Singer T et al (2004) Empathy for pain
and reappraisal effects in humans. Nat Med involves the affective but not sensory compo-
16(11):1277–1283 nents of pain. Science 303(5661):1157–1162
fMRI of Pain 521
151. Lamm C, Decety J, Singer T (2011) Meta- time functional magnetic resonance imaging
analytic evidence for common and distinct (fMRI): methodology and exemplary data.
neural networks associated with directly Neuroimage 19(3):577–586
experienced pain and empathy for pain. 162. Posse S et al (2003) Real-time fMRI of tem-
Neuroimage 54(3):2492–2502 porolimbic regions detects amygdala activa-
152. Apkarian AV et al (2004) Chronic back pain tion during single-trial self-induced sadness.
is associated with decreased prefrontal and Neuroimage 18(3):760–768
thalamic gray matter density. J Neurosci 163. Weiskopf N (2012) Real-time fMRI and its
24(46):10410–10415 application to neurofeedback. Neuroimage
153. Schmidt-Wilcke T et al (2006) Affective com- 62(2):682–692
ponents and intensity of pain correlate with 164. Guan M et al (2015) Self-regulation of brain
structural differences in gray matter in chronic activity in patients with postherpetic neural-
back pain patients. Pain 125(1):89–97 gia: a double-blind randomized study using
154. Jensen KB et al (2013) Overlapping struc- real-time FMRI neurofeedback. PLoS One
tural and functional brain changes in patients 10(4):e0123675
with long-term exposure to fibromyalgia 165. Ramsey NF et al (1996) Functional map-
pain. Arthritis Rheum 65(12):3293–3303 ping of human sensorimotor cortex with 3D
155. Vernon DJ (2005) Can neurofeedback training BOLD fMRI correlates highly with H215O
enhance performance? An evaluation of the evi- PET rCBF. J Cereb Blood Flow Metab
dence with implications for future research. Appl 16(5):755–759
Psychophysiol Biofeedback 30(4):347–364 166. Detre JA et al (1992) Perfusion imaging.
156. Tao JX et al (2005) Intracranial EEG sub- Magn Reson Med 23(1):37–45
strates of scalp EEG interictal spikes. Epilepsia 167. Owen DG et al (2008) Quantification of
46(5):669–676 pain-induced changes in cerebral blood flow
157. Lantz G et al (2001) Localization of distrib- by perfusion MRI. Pain 136(1):85–96
uted sources and comparison with functional 168. Maleki N et al (2013) Pain response measured
MRI. Epileptic Disord, Special Issue:45–58. with arterial spin labeling. NMR Biomed
158. Stern JM (2006) Simultaneous electroen- 26(6):664–673
cephalography and functional magnetic reso- 169. Wang J et al (2004) Reduced susceptibility
nance imaging applied to epilepsy. Epilepsy effects in perfusion fMRI with single-shot
Behav 8(4):683–692 spin-echo EPI acquisitions at 1.5 tesla. Magn
159. Cox RW, Jesmanowicz A, Hyde JS (1995) Reson Imaging 22(1):1–7
Real-time functional magnetic resonance 170. Devlin JT et al (2000) Susceptibility-induced
imaging. Magn Reson Med 33(2):230–236 loss of signal: comparing PET and fMRI on a
160. Yoo S-S, Jolesz FA (2002) Functional MRI semantic task. Neuroimage 11(6):589–600
for neurofeedback: feasibility study on a hand 171. Ojemann JG et al (1997) Anatomic local-
motor task. Neuroreport 13(11):1377–1381 ization and quantitative analysis of gradient
161. Weiskopf N et al (2003) Physiological self- refocused echo-planar fMRI susceptibility
regulation of regional brain activity using real- artifacts. Neuroimage 6(3):156–167
Chapter 17
Abstract
The extensive application of fMRI to the assessment of the human sensorimotor system has disclosed a
complexity that is largely beyond our original understanding. From the available data, it is accepted
that this system consists of a large, and somewhat yet unknown, number of cortical and subcortical
areas, with a precise location and a specialized function. In particular, a large number of regions in the
frontal and parietal lobes contribute to different aspects of motor act performance. It is also evident
that the properties and potentialities of this network still need to be fully elucidated by further research.
Defining how the human sensorimotor system works is of outmost importance for understanding its
dysfunction in case of diseases and also to develop potential therapeutic strategies capable to enhance
its functional plasticity and reserve.
1 Introduction
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_17, © Springer Science+Business Media New York 2016
523
524 Massimo Filippi et al.
2 Sensorimotor Paradigms
Fig. 1 Comparison of mean activation in old vs. young healthy subjects during the
performance of wrist extension/flexion and index finger abduction/adduction
with the left and right upper limb, respectively. Areas more significantly activated
in old subjects are coded in red spectrum, while areas more significantly acti-
vated in young subjects are coded in blue spectrum. Activations have been over-
laid on a standard T1 brain image in neurological view. For each motor task, the
contralateral primary sensorimotor cortex and the premotor cortex had signifi-
cantly greater activation in the young group and caudal supplementary motor
area had significantly greater activation in the old group. Ipsilateral sensorimotor
cortex was more significantly activated in the old group for index finger motor
tasks of both hands (From ref. [35], with permission)
In addition, this has allowed to define the role that the different
components of the network have during the performance of a
motor act.
3.1 The Primary Anatomically, the primary SMC is the cortex lying within the ante-
Sensorimotor Cortex rior and posterior banks of the central sulcus [1]. In line with clini-
cal and electrical stimulation studies, fMRI studies confirmed the
somatotopic organization of the primary SMC of the left
fMRI of the Sensorimotor System 527
3.2 The Another important component of the motor network is the SMA,
Supplementary Motor which is the cortex lying above the cingulate sulcus and anteriorly
Area within the paracentral lobule [1]. The SMA contributes to the
preparation, coordination, temporal course, and execution of
movements [50–52]. Studies of healthy individuals suggest that
the movement-related activity of the primary SMC might be
mediated by the extensive input it receives from the SMA [53],
which might act as facilitator or suppressor according to the task
conditions [54], and that the SMA recruitment might increase by
increasing task difficulty and complexity [16, 21, 24]. The extent
of SMA activation has been inversely related to the amount of
training an individual has gained with that specific task [50, 52,
55]. Inter- and intrahemispheric connections between the primary
SMC, the premotor cortex, and the SMA are likely to be mediated
primarily by the SMA [56]. In addition, strong bilateral connec-
tions exist between bilateral SMA and the basal ganglia, thus
528 Massimo Filippi et al.
3.3 The Frontal The frontal cortex contains many areas contributing to the motor
Cortex network [66, 67]. In addition to the primary SMC, these areas
include the ventral premotor areas (including the inferior frontal
gyrus [IFG]), the dorsal premotor cortex (sometimes divided into
a caudal and a rostral part) (PMd), and a set of motor areas on the
medial wall of the hemispheres, such as the SMA and the cingulate
motor area (CMA). The premotor areas in the frontal lobe influ-
ence motor output through connections with the primary SMC
and direct projections to the spinal cord [68]. All previous premo-
tor areas contain corticospinal neurons that give a substantial con-
tribution to corticospinal projections, which have a high degree of
topographic organization [39, 69].
The role of the left inferior frontal lobe (ventral premotor cor-
tex/Broca’s area) in motor sequence control is well documented
by several studies [70–74]. Activation of Broca’s area has been
reported in various functional imaging studies based on finger
movements [71, 72], movement imagination, and motor learning
[70]. This area is supposed to receive rich sensory information
originating from the parietal lobe (including the SII) and to use it
for action [75]. In addition, modulation of this area’s activity by
task complexity has been clearly documented [73]. Studies in
humans have shown that this region is important for encoding
hand/object interactions [3, 75] and mediating motor response
inhibition via connectivity with the preSMA [74].
The PMd has an important role in motor preparation, selec-
tion, and initiation of voluntary actions [76–78]. Imaging and
TMS experiments suggest that the PMd cortex of the left hemi-
sphere is dominant in right-handed people [79]. This area is recip-
rocally connected with the ipsilateral and contralateral primary
SMC, as well as with the parietal cortex and the contralateral PMd
fMRI of the Sensorimotor System 529
3.5 The Basal The basal ganglia have extensive connections to the motor and
Ganglia, Insula, somatosensory cortices and are involved in motor programming,
and Thalamus execution, and control [104, 105]. In particular, basal ganglia activ-
ity has been associated with motor program selection and suppres-
sion at early stages of motor planning, as well as with control of
movement simulation [57, 106]. In addition, they are implicated in
the formation of motor skills and are part of subsystems whose
activity has been associated with timing of motor acts [107–110].
The thalamus [104, 111] and the insula [112] have extensive
connections with the motor and somatosensory cortices and are
involved in motor execution [104, 112]. Interestingly, the thala-
mus is an important relay station of the complex re-entrant cir-
cuitry that links the motor and the prefrontal cortices to the basal
ganglia, which is part of the feedback loops of the limbic system
able to modulate the cortical motor output [57, 113].
The insular cortex has been shown to play a role in crossmodal
transfer of information [114]. In addition, the insular cortex,
which has connections with numerous cortical and subcortical
motor regions, is involved in the synchronization of movement
kinematic [115] and skeletomotor body orientation [116].
3.6 The Cerebellum The cerebellum integrates sensory information and motor pro-
grams to coordinate fine movements. Functionally, the cerebellum
is organized in modules arranged in the medio-lateral direction,
being the medial part responsible for control of posture and the
lateral regions for coordination and movements. Anatomically, the
cerebellum is divided along the rostro-caudal axis in the anterior
lobe, which contains a somatotopic representation of movement of
532 Massimo Filippi et al.
the ipsilateral side and contributes to motor control [117], and the
posterior lobe, which is thought to be related to motor imagery
[75] and motor learning [118, 119]. The posterior lobe of the
cerebellum has projections from and to regions of the parietal cor-
tex, involved in the processing of sensory information [120–122],
which is then used to correct movements. However, recent studies
revealed functionally distinct areas within and across cerebellar lob-
ules, thus demonstrating a functional parcellation that is indepen-
dent of anatomical lobular divisions [122].
Several imaging studies have reported a cerebellar recruitment
associated to timing of rhythmic movements [123]. Some studies
also described increased cerebellar activation corresponding to
increase in movement frequency and velocity [50, 110, 124, 125].
The cerebellum has also been involved in the “automatization”
(improvement of motor performance) of learned skills, establish-
ment of movement strategies, and consolidation of such a motor
knowledge [110, 126, 127]. Further evidence supporting the role
of the cerebellum in motor learning is based on data from patients
with focal cerebellar lesions, who have shown impairment in learn-
ing new motor skills [126, 128, 129], imaging studies that high-
lighted its contribution to motor recovery after SCI [62] and other
studies of motor learning in healthy individuals, who showed
prominent cerebellar recruitment [130, 131].
Fig. 3 Activation patterns in the basal ganglia and cerebellum during acquisition of motor skills. (a Upper)
Activation maps obtained in the putamen superimposed on a coronal T1-weighted image. There was a pro-
gressive activation decrease in the dorsal part of the putamen (arrows) and an increase in a more ventrolateral
area (arrowheads) bilaterally, which persisted after 4 weeks of training. (a Lower) Percentage signal
increase ± SEM averaged across all subjects for each run of the trained sequence confirmed the activation
decrease in the dorsal putamen and increase in the ventral putamen. (b Top) Activation maps obtained in the
substantia nigra (SN) and subthalamic nucleus (STN) superimposed on EPI images. During session 1, STN
activation was observed during the first run of T-sequence (T1). After 4 weeks of training, these areas were no
more activated during the T-sequence. There was no significant signal change in the SN across runs. (b
Bottom) Signal-to-time curves ± SEM in the STN averaged across all subjects and epochs confirm the activa-
tion decrease. (c Left) Activation maps obtained in the cerebellum during the T-sequence (T1 on day 1 and T5
on day 28). Activation in the lateral cerebellar hemispheres, the left dentate nucleus (DN), and the pons
decreased with training. (c Right) Percentage signal increase ± SEM averaged across all subjects for each run
of the trained sequence in the left and right DN. In the right DN, activation increased transiently during T2
(10 min of practice) and returned to pretraining values (From ref. [141], with permission)
the IFG, the adjacent premotor cortex [156], and the rostral part of
the inferior parietal lobule [157]. The MNS is connected with the
superior temporal sulcus (STS) that provides a higher-order visual
description of the observed action [152, 158]. Mirror neurons are
likely to be multimodal, as they respond to both the visual observation
of an action as well as the sound associated with specific actions [159].
A recent fMRI study showed that regions supporting visuomotor
integration and MNS abilities are multimodal convergent zones of the
visual and motor streams (Fig. 4). In addition, the MNS seems to be
a privilege position for incorporating and integrating basic sensory-
motor information into higher-order cognitive centers [160].
In humans, mirror neurons are part of a system serving the
imitation of actions and speech generation. Therefore, the MNS
fMRI of the Sensorimotor System 535
Stepwise Convergence of
Visual and Motor Cortices
Visual
MNS: Mirror Neuron System
Motor
vPM/I: ventral Premotor / Insula
Major Visuo-Motor Integration aOP4: anterior Operculum Parietale-4
Other Visuo-Motor Integration OPI: Operculum Parietale-I
Functional Connections
SPL: Superior Parietal Lobe
(no white matter tracts)
Fig. 4 Diagram showing the convergence of the visual (green nodes) and motor
(red nodes) systems into the multimodal integration network (blue nodes) or mir-
ror neuron system. Visual cortex streams converge into a common destiny in the
brain network. The main motor functional stream connects motor areas with the
same network as the visual system does. Therefore, motor and visual functional-
related streams meet in a multimodal integration network. Other regions such as
premotor, dorsolateral prefrontal, anterior lateral occipital, or even in situ primary
motor areas may be relevant for visuomotor integration as well, although they
engage fewer convergent functional pathways than the multimodal integration
core (light blue nodes) (From ref. [160] with permission)
6 Conclusions
Fig. 5 Brain activity patterns between healthy participants who were trained on simple motor tasks with their
right hand with (mirror training group—MG) and without a mirror (control training group—CG). The fMRI analy-
sis revealed activation changes within the right dorsal premotor cortex (dPMC), left inferior parietal lobule
(vPMC), left ventral premotor cortex (IPL) and left primary sensorimotor cortex (SMC) in the MG in comparison
with the CG (schema 1). The functional connectivity (FC) analysis revealed an increased FC between previous
regions and the left supplementary motor area (SMA) in the MG over the CG (schema 2). The dynamic causal
modeling, which estimates and makes inferences about the coupling among brain areas, revealed that the
right dPMC and left vPMC interacted with the left SMA, which in turn had access to the left SMC (and the left
IPL interacted with the left vPMC) (schema 3) (From ref. [167], with permission)
538 Massimo Filippi et al.
References
1. Fink GR et al (1997) Multiple nonpri- 15. van der Meulen M et al (2014) The influence
mary motor areas in the human cortex. of individual motor imagery ability on cere-
J Neurophysiol 77(4):2164–2174 bral recruitment during gait imagery. Hum
2. Ciccarelli O et al (2005) Identifying brain Brain Mapp 35(2):455–470
regions for integrative sensorimotor process- 16. Wexler BE et al (1997) An fMRI study of
ing with ankle movements. Exp Brain Res the human cortical motor system response to
166(1):31–42 increasing functional demands. Magn Reson
3. Nowak DA, Glasauer S, Hermsdorfer Imaging 15(4):385–396
J (2013) Force control in object manipula- 17. Deiber MP et al (1999) Mesial motor areas
tion—a model for the study of sensorimotor in self-initiated versus externally triggered
control strategies. Neurosci Biobehav Rev movements examined with fMRI: effect of
37(8):1578–1586 movement type and rate. J Neurophysiol
4. Debaere F et al (2001) Brain areas involved 81(6):3065–3077
in interlimb coordination: a distributed net- 18. VanMeter JW et al (1995) Parametric analy-
work. Neuroimage 14(5):947–958 sis of functional neuroimages: application
5. Rocca MA et al (2007) Influence of body to a variable-rate motor task. Neuroimage
segment position during in-phase and anti- 2(4):273–283
phase hand and foot movements: a kinematic 19. Dettmers C et al (1995) Relation between
and functional MRI study. Hum Brain Mapp cerebral activity and force in the motor areas of
28(3):218–227 the human brain. J Neurophysiol 74:802–815
6. Reddy H et al (2001) Altered cortical acti- 20. Keisker B et al (2009) Differential force scal-
vation with finger movement after peripheral ing of fine-graded power grip force in the
denervation: comparison of active and passive sensorimotor network. Hum Brain Mapp
tasks. Exp Brain Res 138(4):484–491 30(8):2453–2465
7. Reddy H et al (2002) Functional brain reor- 21. Neely KA et al (2013) Segregated and over-
ganization for hand movement in patients lapping neural circuits exist for the produc-
with multiple sclerosis: defining distinct tion of static and dynamic precision grip
effects of injury and disability. Brain 125(Pt force. Hum Brain Mapp 34(3):698–712
12):2646–2657 22. Schlaug G, Knorr U, Seitz R (1994) Inter-
8. Petsas N et al (2013) Evidence of impaired subject variability of cerebral activations in
brain activity balance after passive sensorimo- acquiring a motor skill: a study with posi-
tor stimulation in multiple sclerosis. PLoS tron emission tomography. Exp Brain Res
One 8(6):e65315 98(3):523–534
9. Jaeger L et al (2014) Brain activation associ- 23. Karni A et al (1995) Functional MRI evidence
ated with active and passive lower limb step- for adult motor cortex plasticity during motor
ping. Front Hum Neurosci 8:828 skill learning. Nature 377(6545):155–158
10. Logothetis NK et al (2001) Neurophysiological 24. Rao SM et al (1993) Functional magnetic
investigation of the basis of the fMRI signal. resonance imaging of complex human move-
Nature 412(6843):150–157 ments. Neurology 43(11):2311–2318
11. Mehta JP et al (2012) The effect of move- 25. Annett M (1973) Handedness in families.
ment rate and complexity on functional Ann Hum Genet 37(1):93–105
magnetic resonance signal change during 26. Kim SG et al (1993) Functional magnetic
pedaling. Motor Control 16(2):158–175 resonance imaging of motor cortex: hemi-
12. Francis S et al (2009) fMRI analysis of active, spheric asymmetry and handedness. Science
passive and electrically stimulated ankle dorsi- 261(5121):615–617
flexion. Neuroimage 44(2):469–479 27. Singh LN et al (1998) Comparison of ipsilat-
13. Decety J et al (1994) Mapping motor repre- eral activation between right and left handers:
sentations with positron emission tomogra- a functional MR imaging study. Neuroreport
phy. Nature 371(6498):600–602 9(8):1861–1866
14. Porro CA et al (1996) Primary motor and 28. Solodkin A et al (2001) Lateralization of
sensory cortex activation during motor per- motor circuits and handedness during finger
formance and motor imagery: a functional movements. Eur J Neurol 8(5):425–434
magnetic resonance imaging study. J Neurosci 29. Verstynen T et al (2005) Ipsilateral motor
16(23):7688–7698 cortex activity during unimanual hand
fMRI of the Sensorimotor System 539
among female subjects. Hum Brain Mapp of sequential movements. J Cogn Neurosci
34(5):1194–1207 12:56–77
58. Brinkman C (1981) Lesions in supplemen- 73. Haslinger B et al (2002) The role of lateral
tary motor area interfere with a monkey’s premotor-cerebellar-parietal circuits in motor
performance of a bimanual coordination task. sequence control: a parametric fMRI study.
Neurosci Lett 27(3):267–270 Brain Res Cogn Brain Res 13(2):159–168
59. Brinkman C (1984) Supplementary motor 74. Duann JR et al (2009) Functional connec-
area of the monkey’s cerebral cortex: short- tivity delineates distinct roles of the inferior
and long-term deficits after unilateral ablation frontal cortex and presupplementary motor
and the effects of subsequent callosal section. area in stop signal inhibition. J Neurosci
J Neurosci 4(4):918–929 29(32):10171–10179
60. Akkal D, Dum RP, Strick PL (2007) 75. Grafton ST et al (1996) Localization of grasp
Supplementary motor area and presupple- representations in humans by positron emission
mentary motor area: targets of basal gan- tomography. 2. Observation compared with
glia and cerebellar output. J Neurosci imagination. Exp Brain Res 112(1):103–111
27(40):10659–10673 76. Scott SH, Sergio LE, Kalaska JF (1997)
61. Martino AM, Strick PL (1987) Corticospinal Reaching movements with similar hand paths
projections originate from the arcuate premo- but different arm orientations. II. Activity
tor area. Brain Res 404(1–2):307–312 of individual cells in dorsal premotor cor-
62. Hou JM et al (2014) Alterations of resting- tex and parietal area 5. J Neurophysiol
state regional and network-level neu- 78(5):2413–2426
ral function after acute spinal cord injury. 77. Grafton ST, Fagg AH, Arbib MA (1998)
Neuroscience 277:446–454 Dorsal premotor cortex and conditional
63. Humberstone M et al (1997) Functional movement selection: a PET functional map-
magnetic resonance imaging of single motor ping study. J Neurophysiol 79(2):1092–1097
events reveals human presupplementary 78. Bestmann S et al (2008) Dorsal premotor
motor area. Ann Neurol 42(4):632–637 cortex exerts state-dependent causal influ-
64. Zhang S, Ide JS, Li CS (2012) Resting-state ences on activity in contralateral primary
functional connectivity of the medial superior motor and dorsal premotor cortex. Cereb
frontal cortex. Cereb Cortex 22(1):99–111 Cortex 18(6):1281–1291
65. Weilke F et al (2001) Time-resolved fMRI 79. Schluter ND et al (2001) Cerebral dominance
of activation patterns in M1 and SMA for action in the human brain: the selection of
during complex voluntary movement. actions. Neuropsychologia 39(2):105–113
J Neurophysiol 85(5):1858–1863 80. Schluter ND et al (1998) Temporary inter-
66. Picard N, Strick PL (1996) Motor areas of ference in human lateral premotor cortex
the medial wall: a review of their location suggests dominance for the selection of
and functional activation. Cereb Cortex movements. A study using transcranial mag-
6(3):342–353 netic stimulation. Brain 121(Pt 5):785–799
67. Rizzolatti G, Luppino G (2001) The cortical 81. Moisa M et al (2012) Uncovering a
motor system. Neuron 31(6):889–901 context-specific connectional fingerprint of
68. Dum RP, Strick PL (1991) The origin of cor- human dorsal premotor cortex. J Neurosci
ticospinal projections from the premotor areas 32(21):7244–7252
in the frontal lobe. J Neurosci 11(3):667–689 82. Marconi B et al (2003) Callosal connec-
69. Dum RP, Strick PL (2002) Motor areas in tions of dorso-lateral premotor cortex. Eur
the frontal lobe of the primate. Physiol Behav J Neurosci 18(4):775–788
77(4–5):677–682 83. O’Shea J et al (2007) Functionally specific
70. Stephan KM et al (1995) Functional anat- reorganization in human premotor cortex.
omy of the mental representation of upper Neuron 54(3):479–490
extremity movements in healthy subjects. 84. Cunnington R et al (2006) The selection
J Neurophysiol 73:373–386 of intended actions and the observation of
71. Binkofski F et al (1999) A fronto-parietal others’ actions: a time-resolved fMRI study.
circuit for object manipulation in man: evi- Neuroimage 29(4):1294–1302
dence from an fMRI-study. Eur J Neurosci 85. Deiber MP et al (1991) Cortical areas and
11(9):3276–3286 the selection of movement: a study with posi-
72. Harrington DL et al (2000) Specialized tron emission tomography. Exp Brain Res
neural systems underlying representations 84(2):393–402
fMRI of the Sensorimotor System 541
86. Devinsky O, Morrell MJ, Vogt BA (1995) 101. Culham JC, Kanwisher NG (2001)
Contributions of anterior cingulate cortex to Neuroimaging of cognitive functions in
behaviour. Brain 118:279–306 human parietal cortex. Curr Opin Neurobiol
87. Hoffstaedter F et al (2014) The role of ante- 11(2):157–163
rior midcingulate cortex in cognitive motor 102. Monaco S et al (2015) Neural correlates of
control: evidence from functional connectivity object size and object location during grasp-
analyses. Hum Brain Mapp 35(6):2741–2753 ing actions. Eur J Neurosci 41(4):454–465
88. Paus T et al (1993) Role of the human anterior 103. Barany DA et al (2014) Feature interactions
cingulate cortex in the control of oculomo- enable decoding of sensorimotor transforma-
tor, manual, and speech responses: a positron tions for goal-directed movement. J Neurosci
emission tomography study. J Neurophysiol 34(20):6860–6873
70(2):453–469 104. Parent A, Hazrati LN (1995) Functional
89. Vogt BA, Finch DM, Olson CR (1992) anatomy of the basal ganglia. I. The cortico-
Functional heterogeneity in cingulate cortex: basal ganglia-thalamo-cortical loop. Brain Res
the anterior executive and posterior evaluative Brain Res Rev 20(1):91–9127
regions. Cereb Cortex 2(6):435–443 105. Choi EY, Yeo BT, Buckner RL (2012) The
90. Botvinick M et al (1999) Conflict monitoring organization of the human striatum esti-
versus selection-for-action in anterior cingu- mated by intrinsic functional connectivity.
late cortex. Nature 402(6758):179–181 J Neurophysiol 108(8):2242–2263
91. Carter CS et al (1998) Anterior cingulate cor- 106. Kessler K et al (2006) Investigating the human
tex, error detection, and the online monitoring mirror neuron system by means of cortical syn-
of performance. Science 280(5364):747–749 chronization during the imitation of biological
92. Wenderoth N et al (2005) The role of anterior movements. Neuroimage 33(1):227–238
cingulate cortex and precuneus in the coor- 107. Harrington DL, Haaland KY, Knight
dination of motor behaviour. Eur J Neurosci RT (1998) Cortical networks underlying
22(1):235–246 mechanisms of time perception. J Neurosci
93. Rizzolatti G, Fogassi L, Gallese V (1997) 18(3):1085–1095
Parietal cortex: from sight to action. Curr 108. Ivry RB, Keele SW, Diener HC (1988)
Opin Neurobiol 7(4):562–567 Dissociation of the lateral and medial cer-
94. Karhu J, Tesche CD (1999) Simultaneous ebellum in movement timing and movement
early processing of sensory input in human pri- execution. Exp Brain Res 73(1):167–180
mary (SI) and secondary (SII) somatosensory 109. Jantzen KJ, Steinberg FL, Kelso JAS (2004)
cortices. J Neurophysiol 81(5):2017–2025 Brain networks underlying human timing
95. Hamalainen H, Hiltunen J, Titievskaja I behavior are influenced by prior context. Proc
(2000) fMRI activations of SI and SII cortices Natl Acad Sci U S A 101(17):6815–6820
during tactile stimulation depend on atten- 110. Walz AD et al (2014) Changes in corti-
tion. Neuroreport 11(8):1673–1676 cal, cerebellar and basal ganglia repre-
96. Huttunen J et al (1996) Significance of sentation after comprehensive long term
the second somatosensory cortex in senso- unilateral hand motor training. Behav Brain
rimotor integration: enhancement of sen- Res 278C:393–403
sory responses during finger movements. 111. Brooks DJ (1995) The role of the basal gan-
Neuroreport 7(5):1009–1012 glia in motor control: contributions from
97. Shergill SS et al (2013) Modulation of somato- PET. J Neurol Sci 128(1):1–13
sensory processing by action. Neuroimage 112. Mesulam MM (1998) From sensation to cog-
70:356–362 nition. Brain 121(Pt 6):1013–1052
98. Mima T et al (1998) Attention modulates 113. Chaudhuri A, Behan PO (2000) Fatigue and
both primary and second somatosensory corti- basal ganglia. J Neurol Sci 179(S 1–2):34–42
cal activities in humans: a magnetoencephalo- 114. Hadjikhani N, Roland PE (1998) Cross-
graphic study. J Neurophysiol 80(4):2215–2221 modal transfer of information between the
99. Dobkin BH (2003) Functional MRI: a poten- tactile and the visual representations in the
tial physiologic indicator for stroke rehabilita- human brain: a positron emission tomo-
tion interventions. Stroke 34(5):e23–e28 graphic study. J Neurosci 18(3):1072–1084
100. Del Gratta C et al (2002) Topographic orga- 115. Mosier K, Bereznaya I (2001) Parallel cortical
nization of the human primary and second- networks for volitional control of swallowing
ary somatosensory cortices: comparison in humans. Exp Brain Res 140(3):280–289
of fMRI and MEG findings. Neuroimage 116. Taylor KS, Seminowicz DA, Davis KD (2009)
17(3):1373–1383 Two systems of resting state connectiv-
542 Massimo Filippi et al.
ity between the insula and cingulate cortex. tron emission tomography. Rev Neurol
Hum Brain Mapp 30(9):2731–2745 (Paris) 149:647–653
117. Nitschke MF et al (1996) Somatotopic motor 131. Jenkins IH et al (1994) Motor sequence
representation in the human anterior cerebel- learning: a study with positron emission
lum. A high-resolution functional MRI study. tomography. J Neurosci 14(6):3775–3790
Brain 119(Pt 3):1023–1029 132. Krings T et al (2000) Cortical activation pat-
118. Sakai K et al (1998) Separate cerebel- terns during complex motor tasks in piano
lar areas for motor control. Neuroreport players and control subjects. A functional
9(10):2359–2363 magnetic resonance imaging study. Neurosci
119. Kim JJ, Thompson RF (1997) Cerebellar Lett 278(3):189–193
circuits and synaptic mechanisms involved 133. Pearce AJ et al (2000) Functional reorgan-
in classical eyeblink conditioning. Trends isation of the corticomotor projection to the
Neurosci 20(4):177–181 hand in skilled racquet players. Exp Brain Res
120. Ehrsson HH, Kuhtz-Buschbeck JP, Forssberg 130(2):238–243
H (2002) Brain regions controlling nonsyn- 134. Milton J et al (2007) The mind of expert
ergistic versus synergistic movement of the motor performance is cool and focused.
digits: a functional magnetic resonance imag- Neuroimage 35(2):804–813
ing study. J Neurosci 22(12):5074–5080 135. Bishop DT et al (2013) Neural bases for
121. Allen GI, Tsukahara N (1974) anticipation skill in soccer: an FMRI study.
Cerebrocerebellar communication systems. J Sport Exerc Psychol 35(1):98–109
Physiol Rev 54(4):957–951006 136. Dirnberger G et al (2004) Habituation in
122. Kipping JA et al (2013) Overlapping and a simple repetitive motor task: a study with
parallel cerebello-cerebral networks contrib- movement-related cortical potentials. Clin
uting to sensorimotor control: an intrinsic Neurophysiol 115(2):378–384
functional connectivity study. Neuroimage 137. Loubinoux I et al (2001) Within-session and
83:837–848 between-session reproducibility of cerebral
123. Ramnani N, Passingham RE (2001) Changes sensorimotor activation: a test–retest effect
in the human brain during rhythm learning. evidenced with functional magnetic reso-
J Cogn Neurosci 13(7):952–966 nance imaging. J Cereb Blood Flow Metab
124. Jancke L, Shah NJ, Peters M (2000) Cortical 21(5):592–607
activations in primary and secondary motor 138. Tracy JI et al (2001) A comparison of ‘Early’
areas for complex bimanual movements in and ‘Late’ stage brain activation during brief
professional pianists. Brain Res Cogn Brain practice of a simple motor task. Brain Res
Res 10(1–2):177–183 Cogn Brain Res 10(3):303–316
125. Wenzel U et al (2014) Functional and struc- 139. Morgen K et al (2004) Kinematic specificity of
tural correlates of motor speed in the cerebel- cortical reorganization associated with motor
lar anterior lobe. PLoS One 9(5):e96871 training. Neuroimage 21(3):1182–1187
126. Doyon J et al (1998) Role of the stria- 140. Kruger B et al (2014) Parietal and premo-
tum, cerebellum and frontal lobes in the tor cortices: activation reflects imitation
automatization of a repeated visuomotor accuracy during observation, delayed imita-
sequence of movements. Neuropsychologia tion and concurrent imitation. Neuroimage
36(7):625–641 100:39–50
127. Jueptner M, Weiller C (1998) A review of 141. Lehericy S et al (2005) Distinct basal ganglia
differences between basal ganglia and cer- territories are engaged in early and advanced
ebellar control of movements as revealed motor sequence learning. Proc Natl Acad Sci
by functional imaging studies. Brain 121(Pt U S A 102(35):12566–12571
8):1437–1449 142. Kim YT et al (2011) Neural correlates related
128. Sanes JN, Dimitrov B, Hallett M (1990) to action observation in expert archers. Behav
Motor learning in patients with cerebellar Brain Res 223(2):342–347
dysfunction. Brain 113(Pt 1):103–120 143. Baeck JS et al (2012) Brain activation patterns
129. Bracha V et al (2000) The human cerebel- of motor imagery reflect plastic changes asso-
lum and associative learning: dissociation ciated with intensive shooting training. Behav
between the acquisition, retention and extinc- Brain Res 234(1):26–32
tion of conditioned eyeblinks. Brain Res 144. Balser N et al (2014) Prediction of human
860(1–2):87–94 actions: expertise and task-related effects on
130. Jenkins IH, Frackowiak RS (1993) Functional neural activation of the action observation
studies of the human cerebellum with posi- network. Hum Brain Mapp 35(8):4016–4034
fMRI of the Sensorimotor System 543
145. Fadiga L et al (1995) Motor facilitation dur- 158. Iacoboni M (2005) Neural mechanisms of imi-
ing action observation: a magnetic stimula- tation. Curr Opin Neurobiol 15(6):632–637
tion study. J Neurophysiol 73(6):2608–2611 159. Kohler E et al (2002) Hearing sounds, under-
146. Hari R et al (1998) Activation of human pri- standing actions: action representation in
mary motor cortex during action observation: mirror neurons. Science 297:846–848
a neuromagnetic study. Proc Natl Acad Sci U 160. Sepulcre J (2014) Integration of visual and
S A 95(25):15061–15065 motor functional streams in the human brain.
147. Grezes J et al (2003) Activations related to Neurosci Lett 567:68–73
“mirror” and “canonical” neurones in the 161. Rizzolatti G, Arbib MA (1998) Language
human brain: an fMRI study. Neuroimage within our grasp. Trends Neurosci
18(4):928–937 21(5):188–194
148. Rizzolatti G et al (1996) Localization of 162. Oztop E, Kawato M, Arbib MA (2013)
grasp representations in humans by PET: 1. Mirror neurons: functions, mechanisms and
Observation versus execution. Exp Brain Res models. Neurosci Lett 540:43–55
111(2):246–252 163. Buccino G et al (2001) Action observa-
149. Mengotti P, Corradi-Dell’acqua C, Rumiati tion activates premotor and parietal areas in
RI (2012) Imitation components in the a somatotopic manner: an fMRI study. Eur
human brain: an fMRI study. Neuroimage J Neurosci 13(2):400–404
59(2):1622–1630 164. Filippi M et al (2013) The “vegetarian brain”:
150. Buccino G et al (2004) Neural circuits under- chatting with monkeys and pigs? Brain Struct
lying imitation learning of hand actions: Funct 218(5):1211–1227
an event-related fMRI study. Neuron 165. Rocca MA et al (2008) The mirror-neuron
42(2):323–334 system and handedness: a “right” world?
151. Rizzolatti G, Craighero L (2004) The Hum Brain Mapp 29(11):1243–1254
mirror-neuron system. Annu Rev Neurosci 166. Buccino G (2014) Action observation
27:169–192 treatment: a novel tool in neurorehabilita-
152. Rizzolatti G, Sinigaglia C (2010) The func- tion. Philos Trans R Soc Lond B Biol Sci
tional role of the parieto-frontal mirror cir- 369(1644):20130185
cuit: interpretations and misinterpretations. 167. Hamzei F et al (2012) Functional plastic-
Nat Rev Neurosci 11(4):264–274 ity induced by mirror training: the mirror as
153. Johnson SH et al (2002) Selective activation of a the element connecting both hands to one
parietofrontal circuit during implicitly imagined hemisphere. Neurorehabil Neural Repair
prehension. Neuroimage 17(4):1693–1704 26(5):484–496
154. Leslie KR, Johnson-Frey SH, Grafton ST 168. Aziz-Zadeh L et al (2006) Lateralization of
(2004) Functional imaging of face and hand the human mirror neuron system. J Neurosci
imitation: towards a motor theory of empa- 26(11):2964–2970
thy. Neuroimage 21(2):601–607 169. Cheng Y-W et al (2006) Gender differ-
155. Filippi M et al (2010) The brain functional ences in the human mirror system: a mag-
networks associated to human and animal netoencephalography study. Neuroreport
suffering differ among omnivores, vegetarians 17(11):1115–1119
and vegans. PLoS One 5(5):e10847 170. Theoret H et al (2005) Impaired motor facili-
156. Cerri G et al (2015) The mirror neuron tation during action observation in individu-
system and the strange case of Broca’s area. als with autism spectrum disorder. Curr Biol
Hum Brain Mapp 36(3):1010–1027 15(3):R84–R85
157. Rizzolatti G, Fogassi L, Gallese V (2001) 171. Dapretto M et al (2006) Understanding emo-
Neurophysiological mechanisms underlying tions in others: mirror neuron dysfunction in
the understanding and imitation of action. children with autism spectrum disorders. Nat
Nat Rev Neurosci 2(9):661–670 Neurosci 9(1):28–30
Chapter 18
Abstract
The human visual system consists of a large, yet unknown number of cortical areas. We summarize the efforts
which have led to the identification of 19 retinotopic areas in human occipital cortex, using the macaque
visual cortex as a guide. In this process retinotopic mapping has proven far superior to the study of functional
properties. Macaques and humans share early areas (V1, V2, and V3), a motion-sensitive middle temporal
(MT/V5) cluster as well as six other areas. The remaining human occipital areas either result from reorgani-
zation of a group of monkey areas or seem to be specifically human. Several regions sensitive to motion and
even higher-order motion have been described in parietal cortex, the retinotopic organization of which is still
under debate. On the other hand, both dorsal and ventral regions are sensitive to shape, which is most pro-
nounced in the lateral occipital complex (LOC) extending into the fusiform gyrus. The anterior part of this
complex is flanked by specialized regions devoted to processing faces and bodies and represents “visual
objects” rather than image properties. Its exact organization requires further investigation.
Key words Vision, Retinotopy, Cortical area, Visual field, Motion, 2D and 3D Shape, Actions
1 Introduction
The human visual system is located in the occipital lobe and extends
rostrally into the parietal and temporal lobes. It is estimated to
encompass 30 % of human cortex [1]. Functional imaging gives us
direct access to the function of this important part of human cortex.
One way to study this system is to consider a number of perceptual
or visual cognitive functions and to localize their neural correlates.
An alternative is to consider the visual system as an anatomically orga-
nized collection of cortical areas and subcortical centers that process
retinal information and transform it into messages appropriate for
processing in the nonvisual cerebral regions to which the visual sys-
tem projects. A critical aim in visual neuroscience is to define the
different cortical areas that make up the human visual system. In
other species, such as the nonhuman primates, cortical areas are
defined by the combination of four criteria: (1) cyto- and myelo-
architectonics, (2) anatomical connections with other (known) areas,
(3) topographic organization, i.e., retinotopic organization, and
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_18, © Springer Science+Business Media New York 2016
545
546 Guy A. Orban and Stefania Ferri
2 Methodological Issues
2.2 Tasks One of the main challenges in brain imaging is investigating the
link between neural activity and human behavior. Recent studies
using parametric stimulus manipulation employ detection or dis-
crimination tasks [17–19] rather than passive viewing of the
548 Guy A. Orban and Stefania Ferri
2.3 Control of Eye Control of fixation is mandatory in motion response studies, reti-
Movements notopic mapping experiments, and in spatial attention studies.
Although in the past it was acceptable to show that the subjects
fixated well based on off-line measurements, standards have
evolved. In addition, precise eye movement records, provided by
infrared corneal reflection methods, allow one to remove the effect
of residual eye movements that occur despite fixation. In general,
in all visual experiments, control of fixation will ensure that the
part of visual field stimulated is known and will remove eye move-
ments as a source of unwanted and uncontrolled activations.
2.4 fMRI Designs The conventional fMRI approach for identifying cortical areas
and Paradigms involved in different processes and cognitive tasks entails a
subtraction of activations between different stimulus types that are
presented in blocked or event-related designs.
One of the limitations of these fMRI paradigms is that they
average across neural populations that may respond homogeneously
across stimulus properties or may be differentially tuned to different
stimulus attributes. Thus, in most cases, it is impossible to infer the
properties of the underlying imaged neural populations. fMRI
adaptation (or repetition suppression) paradigms [22–27] have
recently been employed to study the properties of neuronal popula-
tions beyond the limited spatial resolution of fMRI. These para-
digms capitalize on the reduction of neural responses for stimuli
Functional Imaging of the Human Visual System 549
that have been presented for prolonged time or repeatedly [28, 29].
A change in a specific stimulus dimension that elicits increased
responses (i.e., rebound of activity) identifies neural populations
that are tuned to the modified stimulus attributes. fMRI adaptation
paradigms have been used in both monkey and human fMRI stud-
ies as a sensitive tool that allows us to investigate: (a) the sensitivity
of the neural populations to stimulus properties, and (b) the invari-
ance of their responses within the imaged voxels. Adaptation across
a change between two stimuli suggests a common neural represen-
tation invariant to that change, while recovery from adaptation sug-
gests neural representations sensitive to specific stimulus properties.
For example, recent imaging studies tested whether fMRI measure-
ments can reveal neural populations in early visual areas sensitive to
elementary visual features, e.g., orientation, color, and direction of
motion [30–34]. Consider the case of motion direction: after pro-
longed exposure to the adapting motion direction, observers were
tested with the same stimulus in the same or in an orthogonal
motion direction. Decreased fMRI responses were observed in MT
when the test stimuli were at the same motion direction as the
adapting stimulus. However, recovery from this adaptation effect
was observed for stimuli presented at an orthogonal direction.
These studies suggest that the neural populations in human MT are
sensitive to direction of motion [31, 34]. Using the same procedure
in the monkey, Nelissen et al. [2] indeed observed adaptation in
MT/V5 but also in other motion-sensitive regions, such as the
medial superior temporal (MST) region. Similarly, recent studies
have shown stronger adaptation in hMT/V5+ for coherently than
transparently moving plaid stimuli. These findings provide evidence
that fMRI adaptation responses are linked to the activity of pattern-
motion rather than component-motion cells in MT/MST [32].
Thus, these studies suggest that the fMRI signal can reveal neural
selectivity consistent with the selectivity established by neurophysi-
ological methods. However, recent studies comparing fMRI adap-
tation and neurophysiology in monkeys call for cautious
interpretation of the relationship between fMRI adaptation effects
and neural selectivity or invariance at higher levels in the system
[10]. In particular, fMRI adaptation in a given cortical area may be
the result of adaptation at earlier or later stages of processing that is
propagated along the visual areas. Hence in higher-order areas
receiving from multiple inputs fMRI adaptation might reflect adap-
tation of one of the inputs, while recordings show that local neuro-
nal responses driven by the other inputs are not adapted.
Interestingly, novel MVPA methods [11, 35, 36] provide an alter-
native approach for investigating neural selectivity based on fMRI sig-
nals. Unlike conventional univariate analysis, MVPA takes advantage of
the information across multiple voxels in a cortical area and allows us
to characterize neural representations of features that are encoded at a
higher spatial resolution in the brain than the typical resolution of
550 Guy A. Orban and Stefania Ferri
fMRI. These classification analyses have been used successfully for the
decoding of elementary visual features (e.g., orientation [13, 37],
motion direction [38], and object categories [39–42]). The weakness
of the MVPA approach is its dependence on the clustering of neurons
with similar properties. This is also the case for a third technique which
is has been proposed to infer neuronal selectivity from fMRI measure-
ments: measuring the tuning of individual voxels [43]. Just as MVPA,
tuning of voxels is prone to false negative results, as the grouping of
neurons for higher-order selectivity is frequently unknown. In contrast
adaptation fMRI is prone to false positives as inputs may adapt and not
the local neuronal activity. For all these methods greater caution is
required at higher level in the cortex.
2.5 Whole Brain The statistical evaluation of activation differences between stimulus
Versus Region and tasks is typically conducted by comparing responses for each voxel
of Interest Analyses using the general linear model. Analysis of activation patterns across
the whole brain (whole brain analysis) reveals clusters of activations in
different anatomical regions that show significant differences in their
functional processing. This approach has allowed researchers to iden-
tify and localize cortical regions with different functions and evaluate
their involvement in various cognitive tasks. In contrast, region of
interest (ROI) analysis focuses on specific cortical areas identified ana-
tomically or functionally following standard mapping procedures
(e.g., retinotopic mapping). The advantage of this approach is that it
allows us to zoom in on specific cortical regions and investigate their
neural computations using parametric stimulus manipulations. Such
manipulations result in fine stimulus variations and differences in
behavioral performance. Identifying fMRI activations that reflect
these fine differences in neural processing may require the high signal-
to-noise ratio that is possible when scanning and analyzing smaller
regions of cortex. However, ROI analyses are limited in two respects:
(a) the ROI may be outside the volume scanned or analyzed, (b) the
voxels of interest (i.e., voxels that show differential activations across
conditions) may cover a smaller cortical volume than the ROI; as a
result, the differential activations may be averaged out within the
ROI. Taken together, whole brain and ROI analyses can be used as
complementary tools for studying the functional roles of cortical
regions. Whole brain analyses search the entire brain for regions
involved in the analysis of a given stimulus or a cognitive task, while
ROI methods are more appropriate for finer investigation of the neu-
ral processing in these cortical regions [44, 45].
3 Retinotopic Organization
3.1 Early Visual Initially, positron emission tomography (PET) studies have concen-
Areas (V1, V2, V3) trated on the retinotopy of V1 [46], which is a large area of known
localization in the calcarine sulcus. With the advent of fMRI,
Functional Imaging of the Human Visual System 551
Fig. 1 The sweeping stimulus retinotopy paradigm. Two stimuli are used to measure the retinotopic maps in
the cortex. Expanding ring stimuli map eccentricity, and rotating wedge stimuli map polar angle. The phase of
the best-fitting sinusoid for each voxel indicates the position in the visual field that produces maximal activa-
tion for that voxel. Thus, these pseudocolor phase maps are used to visualize the retinotopic maps. Data area
is shown for the left hemisphere (medial view) of one subject. Because of the heavy folding of human cortex,
these retinotopic maps are best seen on flattened hemispheres (from Dougherty et al. [50])
552 Guy A. Orban and Stefania Ferri
Fig. 2 The 18 retinotopic areas defined in the polar angle (a) and eccentricity (b) maps by Georgieva et al. [54], and
Kolster et al. [55]; right hemisphere of subject 1. Stars: central visual field, purple: eccentricity ridge, white dotted
lines: horizontal meridian, black full and dashed lines: lower and upper vertical meridian (from Abdollahi et al. [56])
Fig. 4 Comparison of retinotopic layout of monkey (a) and human (b) visual cortex. Adapted with permission
from Vanduffel et al. [58]. Stars indicate central representations
554 Guy A. Orban and Stefania Ferri
Fig. 5 (a) Outlines (black lines) of retinotopic MPMs superimposed on myelin density(blue to red color) maps
of left hemisphere (L) of 196 subjects. Color code: myelin content in percentiles of the normalized T1w/T2w
distribution. (b) outlines (white) of retinotopic MPM superimposed on the cytoarchitectonic MPM of left (L)
hemisphere. Black lines: 50 % contours of PAs of hOc1, hOc2, hOc5 from Fischl et al. [64]. Inset color code of
cytoarchitectonic areas (from Abdollahi et al. [56])
Functional Imaging of the Human Visual System 555
3.2 Two Middle- V5 or the Middle Temporal (MT) area in humans was initially
Level Areas: Human localized in the ascending branch of the inferior temporal sulcus
MT/V5 and V3A (ITS) [66, 67]. This identification was supported by the fMRI
study of Tootell et al. [68], showing that this region of human cor-
tex has properties, such as luminance and color contrast sensitivity,
similar to those of macaque MT/V5. Subsequently this region has
been referred to as human MT/V5+ [52] to indicate that probably
it corresponds not just to MT/V5 of the macaque but also to sev-
eral of its satellites. It has proven difficult to demonstrate a retino-
topic organization in this region. Huk et al. [69] have suggested
that the MT/V5 complex in humans contains a posterior retino-
topic part, considered the homolog of MT/V5, and an anterior
part driven by ipsilateral stimuli [70], considered the homolog of
MST. One of the drawbacks of this parcellation was the absence of
an homolog of the fundus of the superior temporal (FST) area. Also
the retinotopic organization of what was believed to be MT in
humans [69, 71] seems opposite to that of macaque MT in which
the lower visual field projects in the dorsal part of MT [72, 73]. The
breakthrough occurred when refining the sweeping technique
proved that MT and its satellites could be mapped in the macaque
(Fig. 4) [74]. Applying the same strategy to humans yielded a MT
cluster organized exactly as in the monkey and including four reti-
notopic areas, considered homologues of MT, MSTv, FST, and V4t
(Figs. 2 and 3). The critical point was to identify the central visual
field representation in the eccentricity maps, as it corresponds to the
center of the cluster from which the four areas radiate. This center
is distinct from the central confluence (Fig. 2) and separated from
it by a representation of the periphery, the so-called peripheral edge
(purple in Fig. 2), which was initially noted by Tootell and cowork-
ers [75]. It is noteworthy that in both species the cluster does not
included MSTd involved in optic flow processing [76]. There is at
present little consensus on the criteria to define the human counter-
part of this MST component [69, 77].
In humans, V3A has a similar retinotopic organization as in
macaque: it is defined by a hemifield representation in which the
representations of the two quadrants, separated by the HM, are
neighbors and occupies the banks of the transverse sulcus [78]. The
posterior quadrant is the lower quadrant, separated from that of
V3d by a lower VM. In contrast to macaque V3A, hV3A is motion
sensitive [14, 78, 79]. In the initial mapping study [68] the central
representation of V3A was considered to be fused with that of V1–
V2–V3. Subsequent studies [80–82] have shown that the central
representation is separated from and located more dorsal than that
of the V1–3 confluence, as it generally is in monkeys (5/8 hemi-
spheres in [73]). It has also been noted in humans that this foveal
projection, which V3A shares with V3B (see below), can vary con-
siderably in clarity, being well defined in about half (13/30) hemi-
spheres [83]. In fact the retinotopic organization of dorsal occipital
556 Guy A. Orban and Stefania Ferri
3.3 The Fate of V4 In their 1995 study, Sereno et al. [51] reported an upper quadrant
in Human representation anterior to V3v, that they labeled V4v as it occupied
Visual Cortex the same position as ventral V4 in macaque. Many studies have
replicated that finding of a lower quadrant in front of V3v, but it
has proven difficult to identify a corresponding dorsal V4 quadrant
in front of dorsal V3 [75]. One possible explanation was that stan-
dard mapping technique locating meridians did not apply. Indeed,
in the macaque the horizontal meridian, which represents the ante-
rior border of ventral V4, forms the boundary of dorsal V4 only
over a short distance, as it curves to join the HM splitting MT/V5
into two halves [73, 85]. Hence, we [3] and others [75] have sug-
gested that the region between V3/V3A and hMT/V5+, which we
refer to as LOS [20], is the homolog of macaque dorsal V4. Indeed
it is located in a position similar to that of dorsal V4 and has func-
tional properties relatively similar to those of macaque dorsal V4,
for example, is sensitive to 3D shape from motion (Fig. 7), to 2D
shape [20], and kinetic boundaries [87, 88].
Yet, subsequent mapping studies concentrating on the central
6° of the visual field have suggested that the two halves of macaque
V4 have become separated in humans and are each integrated into
a separate representation of the contralateral hemifield. Brewer
et al. [89] have shown that a lower quadrant was located in front
of the upper quadrant initially labeled V4v, with the eccentricity
running at right angle to the polar variations. They proposed that
this hemifield, located in front of V3v (Figs. 2 and 3) should be
considered human V4. They went on to describe two additional
maps located in front of hV4: ventral occipital (VO)1 and VO2,
each supposedly containing a hemifield representation.
Interestingly, the two face areas, the fusiform and occipital face
areas are located just lateral to hV4 and VO2, respectively. In the
same vein, Larsson and Heeger [83] have described a complete
hemifield representation in front of V3d, which they refer to as
lateral occipital (LO)1. The posterior half of this region is a lower
quadrant that was initially described by Smith et al. [90] as
V3B. Thus, the posterior parts of hV4 and LO1 apparently seem
more responsive, explaining why they were discovered first. Just as
is the case ventrally, a second hemifield representation has been
described in front of LO1: LO2, of which the anterior border is
close to hMT/V5+. The LO1–2/hV4 scheme led to the sugges-
tion that beyond V1–3 the monkey occipital cortex was not an
adequate model for human cortex [81], prompting some [91] to
attempt to rescue the monkey model by suggesting that human V4
was similar to that of the monkey.
Functional Imaging of the Human Visual System 557
3.4 Dorsal Occipital Human V3A has been suggested to share its central representation
and Intraparietal Areas with an area referred to as V3B, located in front of V3A and dor-
sally from the LO1/LO2 pair [83, 96]. V3B occupied in this
scheme a position initially referred to as V7 [97]. V7 is now instead
described as an area rostro-dorsal to a complex of dorsal occipital
areas, the V3A complex, which includes four hemifields organized
pairwise (Figs. 3 and 4). The lower pair, V3A/V3B shares its
peripheral representation (P-cluster) like hV4 and VO1, while the
upper pair V3C/V3D shares a central representation (C-cluster).
Area V7 instead is a parietal area corresponding to IPS0 of Swisher
et al. [98] and seems to correspond to the ventral intraparietal sul-
cus (VIPS) motion-sensitive region [79, 82, 99], located in the
most ventral part of the occipital part of human intraparietal sulcus
(IPS) [100]. In fact V7 is part of another C-cluster sharing its cen-
ter with V7A [101], corresponding to IPS1 and likely the homo-
logue of the pair CIP1–2 described in the monkey [102, 103].
In the human parieto-occipital sulcus (POS) Pitzalis et al. [104]
have described human V6, which borders the dorsal parts of V2 and
V3, representing large eccentricities in the lower visual field (Fig. 3b),
and seems to be homologous in both species. It represents the con-
tralateral hemifield, but with an emphasis on the periphery of the
visual field rather than the center. Pitzalis et al. [105] described
lower-field only representation in the opposite bank of the POS
(Fig. 3b), which they labeled human V6A. This area shows strong
pointing responses, unlike V6, and likely belongs to parietal cortex.
Finally, several attempts have been made to parcel visual regions
in human IPS. Using standard retinotopic mapping, Swisher et al.
[98] described four retinotopic maps, labeled IPS1–4, separated by
VM representations. Konen and Kastner [106] added IPS5 and
SPL1, relying again only on polar angle maps. Responses to stan-
dard retinotopic stimuli are weak in this region, and within ante-
rior parts of IPS moving stimuli are more appropriate to map
558 Guy A. Orban and Stefania Ferri
3.5 Conclusions Human occipital cortex is now almost completely mapped and
includes 19 areas: early areas V1–V3, middle areas LO1–2, hV4,
ventral areas VO1–2, dorsal areas V3A–D and hV6, plus the
occipito-temporal MT and PIT clusters. The competing scheme
using only polar angle maps to define areas [111] only lists 12
occipital retinotopic areas.
Most (13/19) areas are similar to those in the monkey (Fig. 4),
if we admit the proposal of Orban et al. [112] that TFO1–2 located
ventrally to V4/PITv in the monkey are the homologues of VO1–
2. The main inter-species differences are the reorganization of V4/
V4A/OTd into LO1–2/hV4, perhaps related to the separation of
the PITs from the central confluence [55], and the emergence of
areas V3B–D. These latter areas seem to have no counterpart in
the monkey in which V3A neighbors CIP1–2, and may relate to
the expansion of IPL in humans giving rise to the occipital part of
IPS. It is noteworthy that clear homologies are present both at
early and high-order level in the occipital cortex, refuting the idea
that the human visual system divergence more and more from its
monkey counterpart as one ascends into the hierarchy. Also homol-
ogous areas may differ in functional properties, e.g. V3A is motion
sensitive in humans and not in monkeys.
In human occipital cortex all areas beyond V1–3 have a hemi-
field organization, while in macaque hemifield representations
seemed for a long time the exception and split representations, with
separate dorsal and ventral quadrants, the rule. Indeed most initially
known areas (V1–4) had split organization with MT/V5 and V3A
being the exceptions. With most areas mapped, only 5/16 areas
have a split representation in the monkey (Fig. 4), still a larger pro-
portion than in humans (3/19). What is the benefit of the hemi-
field arrangement? As noticed earlier the dorsal region between
V3/V3A and hMT/V5+, in macaque as well as in human, has some
particular functional characteristics, such as 3D shape from motion
sensitivity. The advantage of the human arrangement is that this
sensitivity applies to the whole visual field, while in macaque it
applies only to the lower field. This might be an evolutionary
Functional Imaging of the Human Visual System 559
4 Motion-Sensitive Regions
4.1 Low-Level The two most prominent motion-sensitive regions in human visual
Motion Regions cortex are human MT/V5+ and V3A (see earlier). They display the
highest z scores in a contrast between moving and static random
dots. Their activation remains significant at low stimulus contrasts
typical of the magnocellular stream [78]. In the occipital cortex
motion responses have also been noted in lingual gyrus, probably
corresponding to ventral V2, V3, and in parts of LOS [20, 79,
114, 115]. This activation pattern depends heavily on the size of
the stimuli. With large stimuli, lower-order motion additionally
recruits hV6 [116].
In the early studies it was noted that some parietal regions
were also responsive to motion in a contrast between moving and
static random dots. Sunaert et al. [79] described four motion-
sensitive regions in the IPS. The ventral IPS (VIPS) region is
located at the bottom of the IPS near hV3A. This region, we
believe corresponds to V7 (see above). The parieto-occipital IPS
(POIPS) region is located dorsally with respect to VIPS, at the
junction of the parieto-occipital sulcus and IPS, in the vicinity of
hV6. Not surprisingly, it represents mainly the peripheral visual
field [82] (Fig. 6). The dorsal IPS medial and anterior (DIPSM
and DIPSA) regions are located in the horizontal part of IPS, and
both represent mainly the central visual field [82] (Fig. 6). They
are considered the homolog of anterior part of lateral intraparietal
(LIP) region (DIPSM) and posterior part of anterior intraparietal
(AIP) region (DIPSA), and indeed DIPSA is located just behind
the region referred to as human homolog of AIP based on activa-
tion by grasping actions [117]. All these regions are also activated
by 3D shape from motion [100], which just as motion itself has a
much more extensive representation in human IPS than in macaque
IPS (Fig. 7) [86, 118]. We have speculated that this might in part
be due to the more extensive tool use in humans than in monkeys,
and using a tool indeed activates DIPSM and DIPSA [119]. These
different parietal regions may be engaged in different visuomotor
Fig. 6 Human motion-sensitive regions: distinction between central and peripheral visual field. (a) Stimulus
configuration in experiment 1: the randomly textured pattern (RTP) was positioned either centrally or 5° into
left and right visual field (red dot indicates fixation point). (b and c) Statistical parametric maps (SPMs) showing
voxels significant (yellow: p < 0.0001 uncorrected for multiple comparisons, corresponding to a false discovery
rate of less than 5 % false positives; red: p < 0.001 uncorrected) in the group random-effects analysis (experi-
ment 1, n = 16) for the subtraction moving minus stationary conditions for the centrally (b) and peripherally
(right visual field) (c) positioned stimulus, rendered on the posterior and superior views of the standard human
brain. Further statistical testing revealed that the interaction between type of stimulus (motion, stationary) and
location (center, periphery) was significant (random effects analysis) in DIPSA (Z = 3.12, p < 0.001 uncorrected
and Z = 3.58, p < 0.001 uncorrected for right and left, respectively), DIPSM (Z = 3.58, p < 0.001 uncorrected
and Z = 4.35, p < 0.0001 uncorrected for right and left, respectively) and weakly in POIPS (Z = 2.69, p < 0.01
uncorrected and Z = 2.24, p < 0.01 uncorrected for right and left, respectively). (d) Overlap of voxels (p < 0.001
uncorrected; yellow) in the group random-effects analysis for the subtraction moving minus stationary condi-
tions for the centrally (red) and peripherally (right and left visual field; green) positioned stimulus (experiment
1), rendered on the posterior and superior views of the standard human brain. (e) Stimulus configuration in
experiment 2: RTP was positioned centrally or at 5° eccentricity on upper or lower vertical or horizontal merid-
ian. (f–i) SPMs showing voxels significant (p < 0.05 corrected) in experiment 2 (n = 3) for the subtraction mov-
ing minus stationary conditions for the stimuli positioned in the central visual field (f, red), peripherally left and
right on the horizontal meridian (g, green), and on the lower (h, blue) and upper vertical meridian (i, white),
rendered on the superior view of the standard human brain (posterior part). (j) SPM showing voxels that are
active only in the central condition (obtained by exclusive masking of the subtraction in (f) with those in (g–i)).
Functional Imaging of the Human Visual System 561
Fig. 7 Visual cortical regions sensitive to 3D shape from motion in human and macaque. Statistical parametric
maps (SPMs) for the subtraction viewing of 3D rotating lines minus viewing of 2D translating lines (p < 0.05,
corrected) of a single human (a) and monkey (M4) (b) subject projected on the posterior part of the flattened
right hemisphere. White stippled and solid lines: vertical and horizontal meridian projections (from separate
retinotopic mapping experiments); black stippled lines: motion-responsive regions from separate motion local-
izing tests; purple stippled lines: region of interspecies difference encompassing V3 and intraparietal sulcus.
PCS post-central sulcus, IPS intraparietal sulcus, LaS lateral sulcus, POS parieto-occipital sulcus, CAS calca-
rine sulcus, STS superior temporal sulcus, ITS inferior temporal sulcus, CoS collateral sulcus, IOS inferior
occipital sulcus, OTS occipito-temporal sulcus, PMTS posterior middle temporal sulcus, AMTS anterior middle
temporal sulcus (modified from Vanduffel et al. [86])
Fig. 6 (continued) The opposite procedure, subtractions (g–i) masked by that in (f), yielded no active voxels.
R right, L left, VF visual field. White and yellow numbers in (a) and (e) indicate eccentricity and diameter (diam-
eter), respectively. Numbers in (b–d) correspond to the activation sites listed: 1 and 8: hV3A; 2 and 9: lingual
gyrus; 3, 10, and 11: hMT/V5+; 4: LOS; 5 and 12: VIPS; 13: POIPS; 6: DIPSM; and 7: DIPSA [82]
562 Guy A. Orban and Stefania Ferri
4.2 The Kinetic Using kinetic gratings, that is, stimuli in which random dots
Occipital (KO) Region move in opposite directions in alternate stripes, and comparing
them to luminance gratings or uniform motion, our group [87,
126, 127] discovered a region located between V3/V3A and
hMT/V5+ that appeared selective for kinetic boundaries and
that we referred to as the kinetic occipital (KO) region. Recent
work by Zeki et al. [128] has proposed that KO responds to
boundaries defined by other cues (e.g., colors). These findings
do not dispute the responsiveness of KO to kinetic gratings as
several groups have observed these responses [83, 129].
Although they have been presented differently, these findings are
in fact consistent with our PET [127] and fMRI studies [87]
showing responses in KO for both kinetic and luminance grat-
ings, suggesting that KO responds to contours of different
nature, not just kinetic contours. However, it is important to
emphasize that in contrast with responses in hMT/V5+ and
other motion-sensitive regions, KO is selective for kinetic con-
tours as opposed to uniform motion. Thus, we meant selectivity
in the motion domain, not in the domain of cues defining con-
tours, when we stated [87] that KO is selective for kinetic
boundaries. In the Van Oostende et al. study [87] we observed
overlap of the KO region with response to the LO localizer.
Indeed, Larsson and Heeger [83] in their study identifying
LO1/2 showed that the maximal response to kinetic gratings
compared to transparent motion, the contrast most sharply
defining KO [87], was strongest in LO1 and V3A/B. The coor-
dinates of LO1 [83] are very similar to those of KO (±31, −91,
0, and −32, −92, 0 [87]), supporting the identifying LO1 as the
core region of KO. Thus KO is another functionally defined
region that is incorporated into retinotopic regions, as those
become known, the human motion area [66], or hMT/V5+,
being the primary example, and EBA [130] another one [92].
4.3 High-Level All these motion-sensitive regions are low-level motion regions in
Motion Area the sense that they are driven by motion of light over the retina.
Claeys et al. [99] provided evidence for an attention-based motion-
sensitive region in the inferior parietal lobule (IPL). This region
has activated equiluminant color gratings in which one of the col-
ors is more salient than the other, a paradigm tapping third-order
motion [131, 132]. In addition this region has a bilateral represen-
tation of the visual field, while all other motion-sensitive areas have
mainly a contralateral representation.
Functional Imaging of the Human Visual System 563
5 Shape-Sensitive Regions
Fig. 8 (a) Shape-sensitive regions in human occipital, temporal, and parietal cortex adapted from Sawamura et al.
[136]: yellow: voxels significant (p < 0.05; corrected) in the subtraction 32-objects minus identical condition; black
and blue lines: borders of shape-sensitive regions [i.e., voxels significant in the subtractions intact vs. scrambled
images] obtained by Sawamura et al. [136] and Denys et al. [20] respectively. Numbers: local maxima listed in
black. (b) Face, place, and body patches in human occipital and occipito-temporal cortex: activations are projected
onto flattened right hemisphere of fsaverage atlas. Faces (red)- and body (dark blue) selective regions: approximate
probabilistic data of Engell and McCarthy [137]; Dynamic facial expressions (ocre): real data from Zhu et al. [138];
ATFP: approximate data from (Rajimehr et al. [139]; aSTS: approximate data from (Pitcher et al. [140], showing
Talairach coordinates of individual activations (local maxima). Scene patches (light blue) from Nasr et al. [141].
Biological motion sensitivity (green shape main effect, white kinematics main effect, white interaction): approxi-
mate data from Jastorff and Orban [142]. Regions sensitive for primate vocalizations (black outlines) and intelligible
speech (dashed yellow outlines): real data from Joly et al. [143]. Retinotopy: approximate probabilistic data from
PALS-B12 atlas; VIPS, POIPS, DIPSM, DIPSA, and phAIP (green dashed outlines): are approximate locations from
Jastorff et al. [144],. White star is central representation in V6
564 Guy A. Orban and Stefania Ferri
8 Conclusions
References
1. Van Essen DC (2004) Organization of visual content as revealed by t1- and t2-weighted
areas in macaque and human cerebral cortex. MRI. J Neurosci 31:11597–11616
In: Chalupa LM, Werner JS (eds) The visual 5. Dougherty RF et al (2005) Occipital-callosal
neurosciences, vol 1. MIT Press, Cambridge, pathways in children validation and atlas devel-
MA, pp 507–521 opment. Ann N Y Acad Sci 1064:98–112
2. Nelissen K, Vanduffel W, Orban GA (2006) 6. Schmahmann JD et al (2007) Association
Charting the lower superior temporal region, a fibre pathways of the brain: parallel observa-
new motion-sensitive region in monkey supe- tions from diffusion spectrum imaging and
rior temporal sulcus. J Neurosci 26:5929–5947 autoradiography. Brain 130:630–653
3. Orban GA, Van Essen D, Vanduffel W (2004) 7. Van Essen DC, Jbabdi S, Sotiropoulos SN,
Comparative mapping of higher visual areas Chen C, Dikranian K, Coalson T, Harwell
in monkeys and humans. Trends Cogn Sci J, Behrens TE, Glasser MT (2013) Mapping
8:315–324 connections in humans and non-human pri-
4. Glasser MF, Van Essen DC (2011) Mapping mates: aspirations and challenges for diffu-
human cortical areas in vivo based on myelin sion imaging. In: Johansen-Berg H, Behrens
Functional Imaging of the Human Visual System 567
patterns of fMRI activity in human visual cor- 52. DeYoe EA et al (1996) Mapping striate and
tex. Neuroimage 19:261–270 extrastriate visual areas in human cerebral cor-
37. Kamitani Y, Tong F (2005) Decoding the tex. Proc Natl Acad Sci U S A 93:2382–2386
visual and subjective contents of the human 53. Engel SA, Glover GH, Wandell BA (1997)
brain. Nat Neurosci 8:679–685 Retinotopic organization in human visual
38. Kamitani Y, Tong F (2006) Decoding seen cortex and the spatial precision of functional
and attended motion directions from activ- MRI. Cereb Cortex 7:181–192
ity in the human visual cortex. Curr Biol 54. Georgieva S, Peeters R, Kolster H, Todd JT,
16:1096–1102 Orban GA (2009) The processing of three-
39. Williams MA, Dang S, Kanwisher NG (2007) dimensional shape from disparity in the
Only some spatial patterns of fMRI response human brain. J Neurosci 29:727–742
are read out in task performance. Nat 55. Kolster H, Peeters R, Orban GA (2010) The
Neurosci 10:685–686 retinotopic organization of the human middle
40. O’Toole AJ, Jiang F, Abdi H, Haxby JV temporal area MT/V5 and its cortical neigh-
(2005) Partially distributed representations of bors. J Neurosci 30:9801–9820
objects and faces in ventral temporal cortex. 56. Abdollahi RO et al (2014) Correspondences
J Cogn Neurosci 17:580–590 between retinotopic areas and myelin maps in
41. Hanson SJ, Matsuka T, Haxby JV (2004) human visual cortex. Neuroimage 99:509–524
Combinatorial codes in ventral temporal lobe 57. Robinson EC et al (2013) Multimodal surface
for object recognition: Haxby (2001) revisited: matching: fast and generalisable cortical reg-
is there a “face” area? Neuroimage 23:156–166 istration using discrete optimisation. In: Lect
42. Haxby JV et al (2001) Distributed and Notes Comput Sci (including Subser. Lect
overlapping representations of faces and Notes Artif Intell Lect Notes Bioinformatics)
objects in ventral temporal cortex. Science 7917 LNCS. pp 475–486
293:2425–2430 58. Vanduffel W, Zhu Q, Orban GA (2014)
43. Serences JT, Saproo S, Scolari M, Ho T, Monkey cortex through fMRI glasses.
Muftuler LT (2009) Estimating the influence Neuron 83:533–550
of attention on population codes in human 59. Lyon DC, Kaas JH (2002) Evidence for a
visual cortex using voxel-based tuning func- modified V3 with dorsal and ventral halves in
tions. Neuroimage 44:223–231 macaque monkeys. Neuron 33:453–461
44. Friston KJ, Rotshtein P, Geng JJ, Sterzer P, 60. Rosa MGP, Tweedale R (2005) Brain maps,
Henson RN (2006) A critique of functional great and small: lessons from comparative
localisers. Neuroimage 30:1077–1087 studies of primate visual cortical organiza-
45. Saxe R, Brett M, Kanwisher N (2006) Divide tion. Philos Trans R Soc Lond B Biol Sci
and conquer: a defense of functional localiz- 360:665–691
ers. Neuroimage 30:1088–1096 61. Duncan RO, Boynton GM (2003) Cortical
46. Fox PT et al (1986) Mapping human visual magnification within human primary visual
cortex with positron emission tomography. cortex correlates with acuity thresholds.
Nature 323:806–809 Neuron 38:659–671
47. Schneider W, Noll DC, Cohen JD (1993) 62. Adams DL, Sincich LC, Horton JC (2007)
Functional topographic mapping of the corti- Complete pattern of ocular dominance
cal ribbon in human vision with conventional columns in human primary visual cortex.
MRI scanners. Nature 365:150–153 J Neurosci 27:10391–10403
48. Shipp S, Watson JD, Frackowiak RS, Zeki S 63. Van Essen DC, Newsome WT, Maunsell JHR
(1995) Retinotopic maps in human prestriate (1984) The visual field representation in stri-
visual cortex: the demarcation of areas V2 and ate cortex of the macaque monkey: asymme-
V3. Neuroimage 2:125–132 tries, anisotropies, and individual variability.
49. Engel SA et al (1994) fMRI of human visual Vision Res 24:429–448
cortex. Nature 369:525 64. Fischl B et al (2008) Cortical folding pat-
50. Dougherty RF et al (2003) Visual field represen- terns and predicting cytoarchitecture. Cereb
tations and locations of visual areas V1/2/3 in Cortex 18:1973–1980
human visual cortex. J Vis 3:586–598 65. Hagmann P et al (2008) Mapping the struc-
51. Sereno MI et al (1995) Borders of multiple tural core of human cerebral cortex. PLoS
visual areas in humans revealed by func- Biol 6:1479–1493
tional magnetic resonance imaging. Science 66. Zeki S et al (1991) A direct demonstration of
268:889–893 functional specialization in human visual cor-
tex. J Neurosci 11:641–649
Functional Imaging of the Human Visual System 569
67. Watson JD et al (1993) Area V5 of the human 82. Orban GA et al (2006) Mapping the parietal
brain: evidence from a combined study using cortex of human and non-human primates.
positron emission tomography and magnetic Neuropsychologia 44:2647–2667
resonance imaging. Cereb Cortex 3:79–94 83. Larsson J, Heeger DJ (2006) Two retinotopic
68. Tootell RB et al (1995) Functional analy- visual areas in human lateral occipital cortex.
sis of human MT and related visual cortical J Neurosci 26:13128–13142
areas using magnetic resonance imaging. 84. Lyon DC, Kaas JH (2002) Evidence from V1
J Neurosci 15:3215–3230 connections for both dorsal and ventral sub-
69. Huk AC, Dougherty RF, Heeger DJ (2002) divisions of V3 in three species of new world
Retinotopy and functional subdivision of monkeys. J Comp Neurol 449:281–297
human areas MT and MST. J Neurosci 85. Gattass R, Sousa AP, Gross CG (1988)
22:7195–7205 Visuotopic organization and extent of V3 and
70. Dukelow SP et al (2001) Distinguishing sub- V4 of the macaque. J Neurosci 8:1831–1845
regions of the human MT+ complex using 86. Vanduffel W et al (2002) Extracting 3D from
visual fields and pursuit eye movements. motion: differences in human and monkey
J Neurophysiol 86:1991–2000 intraparietal cortex. Science 298:413–415
71. Smith AT, Wall MB, Williams AL, Singh KD 87. Van Oostende S, Sunaert S, Van Hecke P,
(2006) Sensitivity to optic flow in human Marchal G, Orban GA (1997) The kinetic
cortical areas MT and MST. Eur J Neurosci occipital (KO) region in man: an fMRI study.
23:561–569 Cereb Cortex 7:690–701
72. Van Essen DC, Maunsell JH, Bixby JL 88. Nelissen K, Vanduffel W, Sunaert S, Janssen
(1981) The middle temporal visual area in P, Tootell RB, Orban GA (2000) Processing
the macaque: myeloarchitecture, connec- of kinetic boundaries investigated using fMRI
tions, functional properties and topographic and double-label deoxyglucose technique in
organization. J Comp Neurol 199:293–326 awake monkeys. Soc Neurosci Abstr 26:1584
73. Fize D et al (2003) The retinotopic organi- 89. Brewer AA, Liu J, Wade AR, Wandell BA
zation of primate dorsal V4 and surround- (2005) Visual field maps and stimulus selec-
ing areas: a functional magnetic resonance tivity in human ventral occipital cortex. Nat
imaging study in awake monkeys. J Neurosci Neurosci 8:1102–1109
23:7395–7406 90. Smith AT, Greenlee MW, Singh KD, Kraemer
74. Kolster H et al (2009) Visual field map clus- FM, Hennig J (1998) The processing of first-
ters in macaque extrastriate visual cortex. and second-order motion in human visual cor-
J Neurosci 29:7031–7039 tex assessed by functional magnetic resonance
75. Tootell RB, Hadjikhani N (2001) Where imaging (fMRI). J Neurosci 18:3816–3830
is “dorsal V4” in human visual cortex? 91. Hansen KA, Kay KN, Gallant JL (2007)
Retinotopic, topographic and functional evi- Topographic organization in and near human
dence. Cereb Cortex 11:298–311 visual area V4. J Neurosci 27:11896–11911
76. Tanaka K et al (1986) Analysis of local and 92. Ferri S, Kolster H, Jastorff J, Orban GA
wide-field movements in the superior tem- (2013) The overlap of the EBA and the MT/
poral visual areas of the macaque monkey. V5 cluster. Neuroimage 66:412–425
J Neurosci 6:134–144 93. Boussaoud D, Desimone R, Ungerleider LG
77. Morrone MC et al (2000) A cortical area that (1991) Visual topography of area TEO in the
responds specifically to optic flow, revealed by macaque. J Comp Neurol 306:554–575
fMRI. Nat Neurosci 3:1322–1328 94. Janssens T, Zhu Q, Popivanov ID, Vanduffel
78. Tootell R et al (1997) Functional analysis of W (2014) Probabilistic and single-subject ret-
V3A and related areas in human visual cortex. inotopic maps reveal the topographic organi-
J Neurosci 17:7060–7078 zation of face patches in the macaque cortex.
79. Sunaert S, Van Hecke P, Marchal G, Orban J Neurosci 34:10156–10167
GA (1999) Motion-responsive regions of the 95. Kolster H, Janssens T, Orban GA, Vanduffel
human brain. Exp Brain Res 127:355–370 W (2014) The retinotopic organization of
80. Press WA, Brewer AA, Dougherty RF, Wade macaque occipitotemporal cortex anterior to
AR, Wandell BA (2001) Visual areas and spa- V4 and caudoventral to the middle temporal
tial summation in human visual cortex. Vision (MT) cluster. J Neurosci 34:10168–10191
Res 41:1321–1332 96. Wandell BA, Brewer AA, Dougherty RF
81. Wandell BA, Dumoulin SO, Brewer AA (2005) Visual field map clusters in human
(2007) Visual field maps in human cortex. cortex. Philos Trans R Soc Lond B Biol Sci
Neuron 56:366–383 360:693–707
570 Guy A. Orban and Stefania Ferri
97. Tootell RBH, Tsao D, Vanduffel W (2003) raphy in human cortex. Cereb Cortex
Neuroimaging weighs in: humans meet 25(10):3911–3931
macaques in “primate” visual cortex. 112. Orban GA, Jastorff J (2014) Functional map-
J Neurosci 23:3981–3989 ping of motion regions in human and non-
98. Swisher JD, Halko MA, Merabet LB, human primates. In: Chalupa LM, Werner
McMains SA, Somers DC (2007) Visual JS (eds) The new visual neuroscience, vol 1.
topography of human intraparietal sulcus. MIT Press, Cambridge, MA, pp 777–791
J Neurosci 27:5326–5337 113. Arcaro MJ, McMains SA, Singer BD,
99. Claeys KG, Lindsey DT, De Schutter E, Kastner S (2009) Retinotopic organization
Orban GA (2003) A higher order motion of human ventral visual cortex. J Neurosci
region in human inferior parietal lobule: evi- 29:10638–10652
dence from fMRI. Neuron 40:631–642 114. Sunaert S, Van Hecke P, Marchal G, Orban
100. Orban GA, Sunaert S, Todd JT, Van Hecke GA (2000) Attention to speed of motion,
P, Marchal G (1999) Human cortical regions speed discrimination, and task difficulty: an
involved in extracting depth from motion. fMRI study. Neuroimage 11:612–623
Neuron 24:929–940 115. Rees G, Friston K, Koch C (2000) A direct
101. Kolster H, Peeters R, Orban GA (2011) Ten quantitative relationship between the func-
retinotopically organized areas in human tional properties of human and macaque V5.
paritetal cortex. Soc Neurosci Abstr 851.10 Nat Neurosci 3:716–723
102. Arcaro MJ, Pinsk MA, Li X, Kastner S 116. Pitzalis S et al (2010) Human V6: the medial
(2011) Visuotopic organization of macaque motion area. Cereb Cortex 20:411–424
posterior parietal cortex: a functional mag- 117. Binkofski F et al (1998) Human anterior
netic resonance imaging study. J Neurosci intraparietal area subserves prehension: a
31:2064–2078 combined lesion and functional MRI activa-
103. Orban GA, Zhu Q, Vanduffel W (2014) The tion study. Neurology 50:1253–1259
transition in the ventral stream from feature 118. Orban GA et al (2003) Similarities and dif-
to real-world entity representations. Front ferences in motion processing between the
Psychol 5:695 human and macaque brain: evidence from
104. Pitzalis S et al (2006) Wide-field retino- fMRI. Neuropsychologia 41:1757–1768
topy defines human cortical visual area v6. 119. Stout D, Chaminade T (2007) The evo-
J Neurosci 26:7962–7973 lutionary neuroscience of tool making.
105. Pitzalis S et al (2013) The human homo- Neuropsychologia 45:1091–1100
logue of macaque area V6A. Neuroimage 120. Peuskens H et al (2001) Human brain regions
82:517–530 involved in heading estimation. J Neurosci
106. Konen CS, Kastner S (2008) Two hierar- 21:2451–2461
chically organized neural systems for object 121. Gori M et al (2012) Long integration time
information in human visual cortex. Nat for accelerating and decelerating visual, tac-
Neurosci 11:224–231 tile and visuo-tactile stimuli. Multisens Res
107. Silver MA, Ress D, Heeger DJ (2005) 26:53–68
Topographic maps of visual spatial attention 122. Braddick OJ, O’Brien JMD, Wattam-Bell
in human parietal cortex. J Neurophysiol J, Atkinson J, Turner R (2000) Form and
94:1358–1371 motion coherence activate independent, but
108. Sereno MI, Pitzalis S, Martinez A (2001) not dorsal/ventral segregated, networks in
Mapping of contralateral space in retino- the human brain. Curr Biol 10:731–734
topic coordinates by a parietal cortical area in 123. Eickhoff SB, Grefkes C, Zilles K, Fink GR
humans. Science 294:1350–1354 (2007) The somatotopic organization of
109. Schluppeck D, Curtis CE, Glimcher PW, cytoarchitectonic areas on the human parietal
Heeger DJ (2006) Sustained activity in operculum. Cereb Cortex 17:1800–1811
topographic areas of human posterior pari- 124. Grüsser OJ, Pause M, Schreiter U (1990)
etal cortex during memory-guided saccades. Vestibular neurones in the parieto-insular
J Neurosci 26:5098–5108 cortex of monkeys (Macaca fascicularis):
110. Schluppeck D, Glimcher P, Heeger DJ visual and neck receptor responses. J Physiol
(2005) Topographic organization for delayed 430:559–583
saccades in human posterior parietal cortex. 125. Grüsser O-J, Guldin WO, Mirring S, Salah-
J Neurophysiol 94:1372–1384 Eldin A (1994) Comparative physiological
111. Wang L, Mruczek REB, Arcaro MJ, Kastner and anatomical studies of the primate vestibu-
S (2015) Probabilistic maps of visual topog- lar cortex. In: Albowitz B, Albus K, Kuhnt
Functional Imaging of the Human Visual System 571
U, Nothdurft H-C, Wahle P (eds) Structural ity for dynamic versus static information in
and functional organization of the neocortex. face-selective cortical regions. Neuroimage
Proceedings of a Symposium in the Memory 56:2356–2363
of Otto D. Creutzfeldt, May 1993, Exp Brain 141. Nasr S et al (2011) Scene-selective cortical
Res Series 24. pp 358–371 regions in human and nonhuman primates.
126. Orban GA et al (1995) A motion area in J Neurosci 31:13771–13785
human visual cortex. Proc Natl Acad Sci U S 142. Jastorff J, Orban GA (2009) Human func-
A 92:993–997 tional magnetic resonance imaging reveals
127. Dupont P et al (1997) The kinetic occipital region separation and integration of shape and
in human visual cortex. Cereb Cortex 7:283–292 motion cues in biological motion processing.
128. Zeki S, Perry RJ, Bartels A (2003) The pro- J Neurosci 29:7315–7329
cessing of kinetic contours in the brain. Cereb 143. Joly O et al (2012) Processing of vocaliza-
Cortex 13:189–202 tions in humans and monkeys: a comparative
129. Tyler CW, Likova LT, Kontsevich LL, Wade AR fMRI study. Neuroimage 62:1376–1389
(2006) The specificity of cortical region KO to 144. Jastorff J, Begliomini C, Fabbri-Destro M,
depth structure. Neuroimage 30:228–238 Rizzolatti G, Orban GA (2010) Coding
130. Downing PE, Jiang Y, Shuman M, Kanwisher observed motor acts: different organizational
N (2001) A cortical area selective for visual principles in the parietal and premotor cortex
processing of the human body. Science of humans. J Neurophysiol 104:128–140
293:2470–2473 145. Felleman DJ, Van Essen DC (1991)
131. Lu ZL, Sperling G (1995) Attention-generated Distributed hierarchical processing in the pri-
apparent motion. Nature 377:237–239 mate cerebral cortex. Cereb Cortex 1:1–47
132. Lu ZL, Lesmes LA, Sperling G (1999) 146. Tanaka K, Saito H, Fukada Y, Moriya M
The mechanism of isoluminant chromatic (1991) Coding visual images of objects in the
motion perception. Proc Natl Acad Sci U S A inferotemporal cortex of the macaque mon-
96:8289–8294 key. J Neurophysiol 66:170–189
133. Ungerleider LG, Mishkin M (1982) Two 147. Grill-Spector K, Malach R (2004) The human
cortical visual systems. Anal Vis Behav visual cortex. Annu Rev Neurosci 27:649–677
549:549–586 148. Quiroga RQ, Reddy L, Kreiman G, Koch C,
134. Malach R et al (1995) Object-related activ- Fried I (2005) Invariant visual representation
ity revealed by functional magnetic resonance by single neurons in the human brain. Nature
imaging in human occipital cortex. Proc Natl 435:1102–1107
Acad Sci U S A 92:8135–8139 149. Reddy L, Kanwisher N (2006) Coding of
135. Kanwisher N, Chun MM, McDermott J, visual objects in the ventral stream. Curr Opin
Ledden PJ (1996) Functional imaging of human Neurobiol 16:408–414
visual recognition. Cogn Brain Res 5:55–67 150. Privman E et al (2007) Enhanced category
136. Sawamura H, Georgieva S, Vogels R, tuning revealed by intracranial electroenceph-
Vanduffel W, Orban GA (2005) Using func- alograms in high-order human visual areas.
tional magnetic resonance imaging to assess J Neurosci 27:6234–6242
adaptation and size invariance of shape pro- 151. Altmann CF, Bülthoff HH, Kourtzi Z (2003)
cessing by humans and monkeys. J Neurosci Perceptual organization of local elements into
25:4294–4306 global shapes in the human visual cortex.
137. Engell AD, McCarthy G (2013) Probabilistic Curr Biol 13:342–349
atlases for face and biological motion percep- 152. Kourtzi Z, Tolias AS, Altmann CF, Augath
tion: an analysis of their reliability and over- M, Logothetis NK (2003) Integration of
lap. Neuroimage 74:140–151 local features into global shapes: monkey and
138. Zhu Q et al (2012) Dissimilar processing human fMRI studies. Neuron 37:333–346
of emotional facial expressions in human 153. Kourtzi Z, Kanwisher N (2001) Human
and monkey temporal cortex. Neuroimage lateral occipital complex representation of
66C:402–411 perceived object shape by the human lateral
139. Rajimehr R, Young JC, Tootell RBH (2009) occipital complex. Science 293:1506–1509
An anterior temporal face patch in human 154. Downing PE, Chan AW-Y, Peelen MV, Dodds
cortex, predicted by macaque maps. Proc CM, Kanwisher N (2006) Domain specificity
Natl Acad Sci U S A 106:1995–2000 in visual cortex. Cereb Cortex 16:1453–1461
140. Pitcher D, Dilks DD, Saxe RR, Triantafyllou 155. Freedman DJ, Riesenhuber M, Poggio T,
C, Kanwisher N (2011) Differential selectiv- Miller EK (2003) A comparison of primate
572 Guy A. Orban and Stefania Ferri
prefrontal and inferior temporal cortices human object-related visual areas. Hum Brain
during visual categorization. J Neurosci Mapp 15:67–79
23:5235–5246 171. Orban GA (2007) Three-dimensional shape:
156. Vogels R (1999) Categorization of com- cortical mechanisms of shape extraction. In:
plex visual images by rhesus monkeys. Masland RH, Albright T (eds) Handbook of
Part 2: Single-cell study. Eur J Neurosci the senses, vol 5, Vision. Elsevier, Amsterdam
11:1239–1255 172. Durand JB et al (2007) Anterior regions of
157. Levy I, Hasson U, Avidan G, Hendler T, Malach monkey parietal cortex process visual 3D
R (2001) Center-periphery organization of shape. Neuron 55:493–505
human object areas. Nat Neurosci 4:533–539 173. James TW et al (2002) Haptic study of three-
158. Hasson U, Levy I, Behrmann M, Hendler T, dimensional objects activates extrastriate visual
Malach R (2002) Eccentricity bias as an orga- areas. Neuropsychologia 40:1706–1714
nizing principle for human high-order object 174. Kourtzi Z, Erb M, Grodd W, Bülthoff HH
areas. Neuron 34:479–490 (2003) Representation of the perceived 3-D
159. Rolls ET (2000) Functions of the primate object shape in the human lateral occipital
temporal lobe cortical visual areas in invariant complex. Cereb Cortex 13:911–920
visual object and face recognition. Neuron 175. Murray SO, Olshausen BA, Woods DL
27:205–218 (2003) Processing shape, motion and three-
160. Freiwald WA, Tsao DY (2010) Functional dimensional shape-from-motion in the
compartmentalization and viewpoint gener- human cortex. Cereb Cortex 13:508–516
alization within the macaque face-processing 176. Sereno ME, Trinath T, Augath M, Logothetis
system. Science 330:845–851 NK (2002) Three-dimensional shape represen-
161. Cumming BG, DeAngelis GC (2001) The tation in monkey cortex. Neuron 33:635–652
physiology of stereopsis. Annu Rev Neurosci 177. Shikata E et al (2001) Surface orienta-
24:203–238 tion discrimination activates caudal and
162. Parker AJ (2007) Binocular depth perception and anterior intraparietal sulcus in humans: an
the cerebral cortex. Nat Rev Neurosci 8:379–391 event-related fMRI study. J Neurophysiol
163. Neri P, Bridge H, Heeger DJ (2004) 85:1309–1314
Stereoscopic processing of absolute and 178. Taira M, Nose I, Inoue K, Tsutsui K (2001)
relative disparity in human visual cortex. Cortical areas related to attention to 3D sur-
J Neurophysiol 92:1880–1891 face structures based on shading: an fMRI
164. Orban GA, Janssen P, Vogels R (2006) study. Neuroimage 14:959–966
Extracting 3D structure from disparity. 179. Welchman AE, Deubelius A, Conrad V,
Trends Neurosci 29:466–473 Bülthoff HH, Kourtzi Z (2005) 3D shape per-
165. Gulyas B, Roland PE (1994) Processing and ception from combined depth cues in human
analysis of form, colour and binocular dis- visual cortex. Nat Neurosci 8:820–827
parity in the human brain: functional anat- 180. Perrett DI et al (1985) Visual analysis of body
omy by positron emission tomography. Eur movements by neurones in the temporal cor-
J Neurosci 6:1811–1828 tex of the macaque monkey: a preliminary
166. Mendola JD, Dale AM, Fischl B, Liu AK, report. Behav Brain Res 16:153–170
Tootell RB (1999) The representation of illu- 181. Abdollahi RO, Jastorff J, Orban GA (2013)
sory and real contours in human cortical visual Common and segregated processing of
areas revealed by functional magnetic reso- observed actions in human SPL. Cereb
nance imaging. J Neurosci 19:8560–8572 Cortex 23:2734–2753
167. Backus BT, Fleet DJ, Parker AJ, Heeger 182. Jastorff J, Popivanov ID, Vogels R, Vanduffel
DJ (2001) Human cortical activity corre- W, Orban GA (2012) Integration of shape and
lates with stereoscopic depth perception. motion cues in biological motion processing
J Neurophysiol 86:2054–2068 in the monkey STS. Neuroimage 60:911–921
168. Tsao DY et al (2003) Stereopsis activates V3A 183. Grossman E et al (2000) Brain areas involved
and caudal intraparietal areas in macaques and in perception of biological motion. J Cogn
humans. Neuron 39:555–568 Neurosci 12:711–720
169. Brouwer GJ, van Ee R, Schwarzbach J (2005) 184. Glasser MF, Robinson EC, Coalson TS, Smith
Activation in visual cortex correlates with the SM, Jenkinson M, Hacker CS, Laumann TO,
awareness of stereoscopic depth. J Neurosci Van Essen DC (2014) Partial correlation
25:10403–10413 functional connectivity gradients for cortical
170. Gilaie-Dotan S, Ullman S, Kushnir T, Malach parcellation: methods and multi-modal com-
R (2002) Shape-selective stereo processing in parisons SFN abstract WCC 147B
Chapter 19
Abstract
Over the years, blood oxygen level-dependent (BOLD) fMRI has made important contributions to the under-
standing of central auditory processing in humans. Although there are significant technical challenges to over-
come in the case of auditory fMRI, the unique methodological advantage of fMRI as an indicator of population
neural activity lies in its spatial precision. It can be used to examine the neural basis of auditory representation
at a number of spatial scales, from the micro-anatomical scale of population assemblies to the macro-anatomical
scale of cortico-cortical circuits. The spatial resolution of fMRI is maximized in the case of mapping individual
brain activity, and here it has been possible to demonstrate known organizational features of the auditory system
that have hitherto been possible only using invasive electrophysiological recording methods. Frequency coding
in the primary auditory cortex is one such example that we shall discuss in this chapter. Of course, noninvasive
procedures for neuroscience are the ultimate aim and as the field moves towards this goal by recording in awake,
behaving animals so human neuroimaging techniques will be increasingly relied upon to provide an interpretive
link between animal neurophysiology at the multi-unit level and the operation of larger neuronal assemblies, as
well as the mechanisms of auditory perception itself. For example, the neural effects of intentional behavior on
stimulus-driven coding have been explored both in animals, using electrophysiological techniques, and in
humans, using fMRI. While the feature-specific effects of selective attention are well established in the visual
cortex, the effect of auditory attention in the auditory cortex has generally been examined at a very coarse spatial
scale. Ongoing research in our laboratory has started to address this question and here we present preliminary
evidence for frequency-specific effects of attentional enhancement in the human auditory cortex. We end with
a brief discussion of several future directions for auditory fMRI research.
Key words Technical challenges, Frequency coding, Selective attention, Perceptual representation,
Task specificity
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_19, © Springer Science+Business Media New York 2016
573
574 Deborah Ann Hall and Aspasia Eleni Paltoglou
functional brain activation relies on two factors: first that local neu-
ral activity is a metabolically demanding process that is closely asso-
ciated with a local increase in the supply of oxygenated blood to
those active parts of the brain, and second that the different para-
magnetic properties of oxygenated and deoxygenated blood pro-
duce measurable effects on the MR signal. The functional signal
detected during fMRI is known as the blood oxygen level-
dependent (BOLD) response. Essentially, the functional image
represents the spatial distribution of blood oxygenation levels in
the brain, and the small fluctuations in these levels over time are
correlated with the stimulus input or cognitive task.
MR scanners operate using three different types of electromag-
netic fields: a very high static field generated by a superconducting
magnet, time-varying gradient magnetic fields, and pulsed RF fields.
The latter two fields are much weaker than the first, but all pose a
number of unique and considerable technical challenges for conduct-
ing auditory fMRI research within this hostile environment. In the
first place, the static and time-varying magnetic fields preclude the use
of many types of electronic sound presentation equipment, as well as
preventing the safe scanning of patients who are wearing listening
devices such as hearing aids or implants. Additionally, the high levels
of scanner noise generated by the flexing of the gradient coils in the
static magnetic field can potentially cause hearing difficulties. The
scanner noise masks the perception of the acoustic stimuli presented
to the subject in the scanner making it difficult to calibrate audible
hearing levels and adding to the difficulty of the listening task. And
finally, the scanner noise not only activates parts of the auditory brain,
but also interacts with the patterns of activity evoked by experimental
stimuli. Auditory fMRI poses a number of other challenges, not
related to the hostile environment of the MR scanner, but related
instead to the nature of the neural coding in the auditory cortex. The
response of auditory cortical neurons to a particular class of sound is
determined not only by the acoustic features of that sound, but also
by its presentation context. For example, neurons respond strongly to
the onset of sound events and thereafter tend to show rapid adapta-
tion to that sound in terms of a reduction in their firing rate. Thus, the
result of any particular auditory fMRI experiment will depend not
only on the physical attributes of a stimulus, but also on the way in
which the stimuli are presented. In this first section, we shall take each
one of these issues in turn, introducing the problems in more detail as
well as proposing some solutions.
1.1 Use of Electronic The ideal requirement is a sound presentation system that produces
Equipment for Sound a range of sound levels [up to 100-dB sound pressure level (SPL)],
Presentation in the MR with low distortion, a flat frequency response, and a smooth and
Scanner predictable phase response. The first commercially available solu-
tion utilized loudspeakers, placed away from the high static mag-
1.1.1 Problems
netic field, from which the sound was delivered through plastic
tubes inserted into the ear canal (Fig. 1a) through a protective ear
fMRI of the Central Auditory System 575
Fig. 1 MR compatible headsets for sound delivery and noise reduction: (a) tube phones system with foam ear
inserts, (b) circum-aural ear defenders, plus foam ear plugs for passive noise reduction, (c) MRC IHR sound
presentation headset combining commercially available electrostatic transducers in an industry standard ear
defender, and (d) modified MRC IHR headset for sound presentation and for active noise cancellation (ANC),
including an optical error microphone positioned underneath the ear defender
1.1.2 Solutions Despite the restriction on the materials that can be used in a
scanner, a number of different MR-compatible active head-
phone driving units have been produced. An ingenious system
has been developed and marketed by one auditory neuroimag-
ing research group (MR confon GmbH, Magdeburg, Germany,
www.mr-confon.de). This system incorporates a unique, elec-
trodynamic driver that uses the scanner’s static magnetic field in
place of the permanent magnets that are found in conventional
headphones and loudspeakers. It produces a wide frequency
range (less than 200 Hz–35 kHz) with a flat frequency response
(±6 dB). Another company manufactures and supplies high-
quality products for MRI, with a special focus on the fast-grow-
ing field of functional imaging (NordicNeuroLab AS, Bergen,
Norway, www.nordicneurolab.com/). Their audio system uses
electrostatic transducers to ensure high performance.
Electrostatic headphones generate sound using a conductive
diaphragm placed next to a fixed conducting panel. A high volt-
age polarizes the fixed panel and the audio signal passing
through the diaphragm rapidly switches between a positive and
a negative signal, attracting or repelling it to the fixed panel and
thus vibrating the air. Their technical specification claims a flat
frequency response from 8 Hz to 35 kHz. The signal is trans-
ferred from the audio source to the headphones in the RF
screened scanner room using either filters through a filter panel
or fiber-optic cable through the waveguide.
Here at the MRC Institute of Hearing Research, we became
engaged in auditory fMRI research well before such commercial sys-
tems were widely available and so, for our own purposes, we developed
an MR-compatible headset (Fig. 1c) based on commercially available
electrostatic headphones, modified to remove or replace their ferro-
magnetic components, and combined with standard industrial ear
defenders to provide good acoustic isolation [1]. Our custom-built
system delivers a flat frequency response (±10 dB) across the frequency
range 50 Hz–10 kHz and has an output level capability up to 120-dB
SPL. Again, the digital audio source, electronics, and power supply
that drive the system are housed outside the RF screened scanner room
to avoid electromagnetic interference with MR scanning, and all elec-
trical signals passing into the screened scanner room are RF filtered.
fMRI of the Central Auditory System 577
1.2 Risk to Patients No ferromagnetic components can be placed in the scanner bore as
Who Are Wearing they would experience a strong attraction by the static magnetic
Listening Devices field and potentially cause damage not only to the scanner and the
in the MR Scanner listening device, but also to the patient. Induced currents in the
electronics, caused directly by the time-varying gradient magnetic
1.2.1 Problems fields or the RF pulses, are an additional hazard to the electronic
devices themselves, while some materials can also absorb the RF
energy causing local tissue heating and even burns if in contact with
soft tissue. For these reasons, there are restrictions on scanning peo-
ple who have electronic listening devices. These include hearing
aids, cochlear implants, and brainstem implants. Hearing aids
amplify sound for people who have moderate to profound hearing
loss. The aid is battery-operated and worn in or around the ear.
Hearing aids are available in different shapes, sizes, and types, but
they all work in a similar way. They all have a built-in microphone
that picks up sound from the environment. These sounds are pro-
cessed electronically and made louder, either by analogue circuits or
digitally, and the resulting signals are passed to a receiver in the
hearing aid where they are converted back into audible sounds. In
contrast, cochlear and brainstem implants are both small, complex
electronic devices that can help to provide a sense of sound to peo-
ple who are profoundly deaf or severely hard-of-hearing. Cochlear
implants bypass damaged portions of the inner ear (the cochlea)
and directly stimulate the auditory nerve, while auditory brainstem
implants bypass the vestibulocochlear nerve in cases when it is dam-
aged by tumors or surgery and directly stimulate the lower part of
the auditory brain (the cochlear nucleus). In general, both types of
implant consist of an external portion that sits behind the ear and a
second portion that is surgically placed under the skin. They con-
tain a microphone, a sound processor (which converts sounds
picked up by the microphone into an electrical code), a transmitter
and receiver/stimulator (which receive signals from the processor
and convert them into electric impulses), and finally an electrode
array (which is a set of electrodes that collect the impulses from the
stimulator and stimulate groups of auditory neurons). Coded infor-
mation from the sound processor is delivered across the skin via
electromagnetic induction to the implanted receiver/stimulator,
which is surgically placed on a bone behind the ear.
1.2.2 Solutions Official approval for the manufacture of implant devices requires
rigorous testing for susceptibility to electromagnetic fields, radi-
ated electromagnetic fields, and electrical safety testing (including
susceptibility to electrical discharge). However, such tests are con-
ducted under normal conditions, not in the magnetic fields of an
MR scanner. Some implant designs have been proven to be MR
compatible [2–5], but they are not routinely supplied in clinical
practice. Standard listening devices do not meet MR compatibility
criteria and, for the patient, risks include movement of the device
578 Deborah Ann Hall and Aspasia Eleni Paltoglou
and localized heating of brain tissue, whereas, for the device, the
electronic components may be damaged. Magnetic Resonance
Safety Testing Services (MRSTS) is a highly experienced testing
company that conducts comprehensive evaluations of implants,
devices, objects, and materials in the MR environment (MRSTS,
Los Angeles, CA, www.magneticresonancesafetytesting.com/).
Testing includes approved assessment of magnetic field interac-
tions, heating, induced electrical currents, and artifacts. A database
of the devices and results of implant testing is accessible to the
interested reader (www.mrisafety.com/). However, auditory
devices have generally been tested only at low magnetic fields (up
to 1.5 T) because most clinical MR systems operate at this field
strength. Since research systems typically operate at 3.0 T (for
improved BOLD signal-to-noise ratio, BOLD SNR) it may be nec-
essary for individual research teams to ensure the safety of their
patients. For example, here at the MRC Institute of Hearing
Research, we have recently assessed the risks of movement and
localized tissue heating for two middle ear piston devices [6]. For
the safety reasons discussed in this subsection, listeners who nor-
mally wear hearing aids could be scanned without their aid but, to
compensate, have been presented with sounds amplified to an
audible level. Given that implanted devices cannot be removed
without surgical intervention, clinical imaging research of implan-
tees has generally used other brain imaging methods, namely posi-
tron emission tomography [7].
1.3 Intense MR The scanning sequence used to measure the BOLD fMRI signal
Scanner Noise and Its requires rapid on and off switching of electrical currents through
Effects on Hearing the three gradient coils of wire in order to create time-varying mag-
netic fields that are required for selecting and encoding the three-
1.3.1 Problems
dimensional image volume (in the x, y, and z planes). This rapid
switching in the static magnetic field induces bending and buckling
of the gradient coils during MRI. As a result, the gradient coils act
like a moving coil loudspeaker to produce a compression wave in
the air, which is heard as acoustic noise during the image acquisi-
tion. Scanner noise increases nonlinearly with static magnetic field
strength, such that ramping from 0.5 to 2 T could account for a rise
in sound level of as much as 11-dB SPL [8]. A brain scan is com-
posed of a set of two-dimensional “slices” through the brain.
Gradient switching is required for each slice acquisition and so an
intense scanner “ping” occurs each time a brain slice is collected.
Each ping lasts about 50 ms and so during fMRI, each scan is audi-
ble as a rapid sequence of such “pings” (see inset in Fig. 2 for an
example of the amplitude envelope of the scanner noise).
The dominant components of the noise spectrum are com-
posed of a peak of sound energy at the gradient switching frequency
plus its higher harmonics. Most of the energy lies below 3 kHz.
Secondary acoustic noise can be produced if the vibration of the
fMRI of the Central Auditory System 579
−18
−30
−54
−66
−78
uncancelled
−90
uncancelled
Fig. 2 Typical frequency spectrum of the scanner noise generated during blood
oxygen level-dependent (BOLD) fMRI. This example was measured in the bore of
a Philips Intera 3.0 Tesla scanner. The black line (uncancelled) indicates the
acoustic energy of the noise recorded under normal scanning conditions. The
gray line (canceled) indicates the residual acoustic energy at the ear when the
active noise cancellation (ANC) system is operative. The inset (upper right) shows
an example of the amplitude envelope of the scanner noise for a brain scan
consisting of 16 slices corresponding to a sequence of 16 intense “pings”
coils and the core on which they are wound conducts through the
core supports to the rest of the scanner structure. These secondary
noise characteristics depend more on the mechanical resonances of
the coil assemblies than on the type of imaging sequence and they
tend to be the dominant contributor to the bandwidth and the
spectral envelope of the noise. In this example of the frequency
spectrum captured from a BOLD fMRI scanning sequence that was
run on a Philips 3 Tesla Intera (Fig. 2), the spectrum has a peak
component at 600 Hz with several other prominent pseudo-har-
monics at 300, 1080, and 1720 Hz. The sound level measured in
the bore of the scanner is typically 99-dB SPL [98 dB(A) using an
A-weighting], measured using the maximum “fast” root-mean-
square (RMS) time constant (125 ms). Clearly, exposure to such an
intense sound levels without protection is likely to cause a tempo-
rary threshold shift in hearing and tinnitus, and it could be perma-
nently damaging over a prolonged dosage [9].
1.3.2 Solutions The simplest way to treat the intense noise is to use ear protection in
the form of ear defenders and/or ear plugs (shown in Fig. 1b). Foam
ear plugs can compromise the acoustic quality of the experimental
sounds delivered to the subject and so ear defenders are preferable.
580 Deborah Ann Hall and Aspasia Eleni Paltoglou
Fig. 3 Acoustic waveforms of the scanner noise measured with and without a lining of acoustic damping foam
in the bore of the scanner. Our data demonstrate that the foam reduces the sound pressure level (SPL) at the
position of the subject’s head and in scanner room by a significant margin (about 8 dB). The segment of scan-
ner noise that is illustrated here has a duration of approximately 1 s
1.4 The Effect Not only is the intense scanner noise a risk for hearing, but it also
of Scanner Noise masks the perception of the acoustic stimuli presented to the sub-
on Stimulus Audibility ject. The exact specification of the acoustic signal-to-scanner-noise
ratio (acoustic SNR) in fMRI studies using auditory stimuli is a
1.4.1 Problems
potentially complicated matter. Nevertheless, we have sought to
establish the relative difference between the stimulus level and the
scanner noise level at the ear, by measuring these signals using a
reference microphone placed inside the cup of the ear defender
while participants perform a signal detection in noise task.
Detection thresholds for a narrow band noise centered at the peak
frequency of the scanner noise (600 Hz) are elevated when the
target coincides with the scanner noise. We have demonstrated an
average 11-dB shift in the 71 % detection threshold for the 600-Hz
target when we modulate the perceived level of the scanner noise
using active noise cancelation (ANC) methods (see later).
This evidence suggests that even with hearing protection,
whenever the scanner noise coincides with the presented sound
stimulus it produces changes in task performance and probably
also increases the attentional demands of the listening task. The
frequency range of the scanner acoustic noise is crucial for speech
intelligibility, and speech experiments can be particularly compro-
mised by a noisy environment ([15]; for review, see [16]). A recent
study has quantified the effect of acoustic SNR using four listening
tasks: pitch discrimination of complex tones, same/different judg-
ments of minimal-pair nonsense syllables, lexical decision, and
judgement of sentence plausibility [17]. Across these tasks, perfor-
mance was assessed in silence (acoustic SNR = infinity) and in a
background of MR scanner noise at the three acoustic SNR levels
582 Deborah Ann Hall and Aspasia Eleni Paltoglou
Fig. 4 Mean performance in a simulated scanning environment across four acoustic signal-to-noise ratios
[17]. The top panel plots the proportion of correct responses on the individual tasks, while the bottom panel
shows the overall mean performance (SNR signal-to-noise ratio, dB decibels)
1.4.2 Solutions The aggregate noise dosage can be reduced by acquiring either a
single or at least very few brain slices, but at the expense of only a
partial view of brain activity [18]. For whole brain fMRI, other
strategies are required.
One novel method that has been developed and evaluated at
our Institute combines optical microphone technology with an
active noise controller for significant attenuation of ambient noise
received at the ears [19]. The canceller is based upon a variation of
the single channel feed-forward filtered-x adaptive controller and
uses a digital signal processor to achieve the noise reduction in real
time. The canceler minimizes the noise pressure level at a specific
control point in space that is defined by the position of the error
microphone, positioned underneath the circum-aural ear defender
of the headset (see Fig. 1d). In 2001, we published a psychophysical
assessment of the system using a prototype system built in the labo-
ratory that utilized a loudspeaker as the noise generator [19]. This
system produced 10–20 dB of subjective noise reduction between
250 Hz and 1 kHz and smaller amounts at higher frequencies.
More recently, we have obtained psychophysical threshold data in a
Philips 3 Tesla scanner confirming that the same level of cancella-
tion is achieved in the real scanner environment (Fig. 5; [20]).
Again, the subjective impression of the scanner noise is the volume
of gentle drumming when the sound system is operating in its can-
celed mode. Thus, it is possible to achieve a high level of noise
attenuation by combining both passive and active methods.
A much more common strategy for reducing the masking influ-
ence of the concomitant scanner noise combines a passive method of
ear protection with an experimental protocol that carefully controls
35
30
25
uncancelled
20
cancelled
15
10
5
1 2 3 4 5
Participant
1.5 The Effect of To increase the BOLD SNR, it is necessary to acquire a large num-
Scanner Noise on ber of scans in each condition in an fMRI experiment. Typically, an
Sound-Related experimenter would collect many hundreds of brain scans in a sin-
Activation in the Brain gle session, with the time in between each scan chosen to be as
short as the scanner hardware and software will permit. Remember
1.5.1 Problems
that, for fMRI, an intense “ping” is generated for each slice of the
scan and so of course this means that the participant can easily be
subjected to several thousand repeated “pings” of noise during the
experiment. Not only does this scanner noise acoustically mask the
presented sound stimuli, but the elevated baseline of sound-evoked
activation due to the ambient scanner noise also makes the experi-
mentally induced auditory activation more difficult to detect statis-
tically. Much of the work examining the influence of acoustic
scanner noise has been directed toward its capacity to interfere
with the study of audition or speech perception by producing acti-
vation of various brain regions, especially the auditory cortex [22–
25]. Several studies highlight the reduced activation signal (i.e. the
difference between stimulation and baseline conditions) in the
auditory cortex when the amount of prior scanner noise is increased,
demonstrating that the scanner noise effectively masks the detec-
tion of auditory activation [22, 26, 27]. In another example, taken
from one of the early fMRI experiments conducted at the MRC
Institute of Hearing Research, we used a specially tailored scanning
protocol to measure the amplitude and the time course of the
BOLD response to a high-quality recording of a single burst of
scanner noise presented to participants over headphones [24]. Our
results revealed a reliable transient increase in the BOLD signal
across a large part of the auditory cortex. As in many other brain
regions, the evoked response to this single brief stimulus event was
smoothed and delayed in time. It rose to a peak by 4–5 s after
stimulus onset and decayed by 5–8 s after stimulus offset [24]. Its
amplitude reached about 1.5 % of the overall signal change, which
is considerable considering that stimulus-related activation usually
accounts for a BOLD signal change of approximately 2–5 %.
Figure 6 illustrates the canonical BOLD response to a noise onset.
In many fMRI experimental paradigms, regions of stimulus-
evoked activation are detected by comparing the BOLD scans
acquired during one sound condition with the BOLD scans acquired
during another condition, which could be either a condition in
which a different type of sound was presented or no sound (known
as a baseline “silent” condition) was presented. Activation is defined
as those parts of the brain that demonstrate a statistically significant
fMRI of the Central Auditory System 585
3.5
1.5
0.5
–0.5
–1
–1.5
0 5 10 15 20 25 30
Peristimulus time (seconds)
1.5.2 Solutions A number of different scanning protocols have been used to mini-
mize the effect of the scanner acoustic noise on the measured pat-
terns of auditory cortical activation. In this section, we will describe
two of these, but before we do, we need to consider some impor-
tant details about the time course of the BOLD response to the
scanner noise and introduce some new terms.
During an fMRI experiment, the BOLD response to the scan-
ner noise spans two different temporal scales. First, the “ping”
generated by the acquisition of one slice early in the scan may
induce a BOLD response in a slice, which is acquired later in the
same scan if that later scan is positioned over the auditory cortex.
We shall call this inter-slice interference. Inter-slice interference is
maximally reduced when all slices in the scan are acquired in rapid
586 Deborah Ann Hall and Aspasia Eleni Paltoglou
a Sparse sampling
true scans
EPI readout
inter-scan interval = 10 s
Fig. 7 Two scanning protocols that have been used to minimize the effect of the
scanner acoustic noise on the measured patterns of auditory cortical activation.
See text for further explanation (s seconds, EPI echo-planar imaging, RF
radiofrequency)
succession and the total duration of the scan is not more than 2 s
[26]. A common term for the scanning protocol that uses a mini-
mum inter-slice interval is a clustered-acquisition sequence.
Edmister et al. [28] found that the clustered-acquisition sequence
provides an advantageous auditory BOLD SNR compared with a
conventional scanning protocol. The second form of interference
is called inter-scan interference. This occurs when the scanner noise
evokes an auditory BOLD response that extends across time to
subsequent scans, predominantly when the interval between scans
is as short as the MR system will permit. Reducing the inter-scan
interference can easily be achieved by extending the period between
scans (the inter-scan interval). By separately manipulating the tim-
ing between slices and between scans, we can reduce the inter-slice
and inter-scan interference independently of one another. When
the clustered-acquisition sequence is combined with a long (e.g.
10 s) inter-scan interval, the activation associated with the experi-
mental sound can be separated from the activation associated with
the scanner sound (Fig. 7a). Furthermore, because the scanner
sound is temporally offset, it does not produce acoustical masking
and does not distract the listener. This scanning protocol is com-
monly known as sparse sampling [21]. Sparse sampling is often the
scanning protocol of choice for identifying auditory cortical evoked
responses in the absence of scanner noise (see e.g. [29–33]).
However, it requires a scanning session that is longer than that of
conventional “continuous” protocols in order to acquire the same
amount of imaging data, and participants can be intolerant of long
fMRI of the Central Auditory System 587
1.6 The Effect of The acoustic environment is typically composed of one or more
Stimulus Context: sound sources that change over time. Over the years, both psycho-
Neural Adaptation to physical and electrophysiological studies have amply demonstrated
Sounds that stimulus context strongly influences the perception and neural
coding of individual sounds, especially in the context of stream seg-
1.6.1 Problems regation and grouping [35–37]. A simple example of the influence
of stimulus context is forward masking, which occurs when the
presence of one sound increases the detection threshold for the
subsequent sound. The perceptual effects of forward masking are
strongest when the spectral content of the first sound is similar to
the second sound, when there is no delay between the two sounds,
and when the masker duration is long [38]. Forward inhibition
typically lasts from 70 to 200 ms. This type of suppression has not
only been demonstrated in anesthetized preparations, but also in
awake primates. In the latter case, suppression was seen to extend
up to 1 s in time [39]. As well as tone–tone interactions, neural fir-
ing rate is sensitive to stimulus duration. Neurons respond strongly
to the onset of a sound and their response decays thereafter. Many
illustrative examples can be found in the literature, especially in
cases where longer duration sounds are presented (e.g. 750–
1500 ms in the case of Bartlett and Wang [38], see their Fig. 4).
By transporting these well-established paradigms into a neuro-
imaging experiment, researchers are beginning to address the con-
text dependency of neural coding in humans. One way in which
the effect of sound context on the auditory BOLD fMRI signal has
been examined is in terms of different repetition rates [19, 40].
This is conceptually analogous to the presentation rate manipula-
tions of the forward masking studies described earlier, but goes
beyond the simple case of two-tone interactions. In the fMRI stud-
ies, stimuli were long trains of noise bursts presented at different
rates. The slowest rate was 2 Hz and the fastest rate was 35 Hz,
with intermediate rates being 10 and 20 Hz. Noise bursts at each
588 Deborah Ann Hall and Aspasia Eleni Paltoglou
0.3
0.2
BOLD response in right posterior auditory cortex
0.1
0.0
−0.1
−0.2
−0.3
−0.4
0.2
0.1
0.0
−0.1
−0.2
−0.3
−0.4
0 8 16 24 32 40 48
Peristimulus time (secs)
Fig. 8 Adjusted blood oxygen level-dependent (BOLD) response (measured in arbitrary units) across the 32-s
stimulus epoch shaded in gray (a) for a sound from a fixed source and (b) for a sound from a rotating source.
Adjusted values are combined for all six participants and the trend line is indicated using a polynomial sixth
order function. The response for both stimulus types is plotted using the same voxel location in the planum
temporale region of the right auditory cortex (coordinates x 63, y −30, z 15 mm). The position of this voxel is
shown in the inserted panel. The activation illustrated in this insert represents the subtraction of the fixed sound
location from the rotating sound conditions (p < 0.001)
2.1 The Within the inner ear, an incoming sound is separated into its indi-
Representation of vidual frequency components by the way in which the energy at
Frequency in the different frequencies travels along the cochlear partition [45].
Auditory Cortex High-frequency tones maximally stimulate those nerve fibers near
the base of the cochlea while low-frequency tones are best coded
fMRI of the Central Auditory System 591
a b
A
6 L M
1a
FTTS
P
1b HG
2
3
8 4
HS
High-Frequency sensitive area
c d
Fig. 9 (a) Sagittal view of the brain with the oblique white line denoting the approximate location and orienta-
tion of the schematic view shown in panel (b) along the supratemporal plane. (b) Schematic representation of
the most consistently found high (red) and low (blue) frequency-sensitive areas across the human auditory
cortex reported by Talavage et al. [50, 51]. The primary area is shown in white and the nonprimary areas are
shown by dotted shading. Panels (c) and (d) illustrate the high- (red) and low- (blue) frequency sensitive areas
across the left auditory cortex of one participant (unpublished data). Two planes in the superior-inferior dimen-
sion are shown (z = 5 mm and z = 0 mm above the CA-CP line). A anterior, P posterior, M medial, L lateral, HG
Heschl’s gyrus, HS Heschl’s sulcus, FTTS first transverse temporal sulcus, STP supratemporal plane
2.2 The Influence of We live in a complex sound environment in which many different
Selective Attention on overlapping auditory sources contribute to the incoming acoustical
Frequency signal. Our brains have a limited processing capacity and so one of
Representations in the most important functions of neural coding is to separate out
Human Auditory these competing sources of information. One way to achieve this is
Cortex by filtering out the uninformative signals (the “ground”) and
attending to the signal of interest (the “figure”). Competition
between incoming signals can be resolved by a bottom-up, stimulus-
driven process (such as a highly salient stimulus that evokes an
involuntary orienting response), or it can be resolved by a top-
down, goal-directed process (such as selective attention). Selective
attention provides a modulatory influence that enables a listener to
focus on the figure and to filter out or attenuate the ground [53].
Visual scientists have shown that attention can be directed to
the features of the figure (feature-based attention, for a review see
[54]) or to the entire figure (object-based attention, for a review see
[55]). Given that so little is known about the mechanisms by which
auditory objects are coded [56], we shall focus on those studies of
auditory feature-based attention. A sound can be defined according
to many different feature dimensions including frequency spec-
trum, temporal envelope, periodicity, spatial location, sound level,
and duration. The experimenter can instruct listeners to attend to
any feature dimension in order to investigate the effect of selective
attention on the neural coding of that feature. Different listening
conditions have been used for comparison with the “attend” condi-
tion. The least controlled of these is a passive listening condition in
which participants are not given any explicit task instructions [30,
57, 58]. Even if there are cases where a task is required, but the
cognitive demand of that task is low, participants are able to divide
their attention across both relevant and irrelevant stimulus dimen-
sions (see [59] for a review on attentional load). Again, this leads to
an uncontrolled experimental situation. For greater control, some
studies have employed a visual distractor task to compete for atten-
tional resources and pull selective attention away from the auditory
modality [60, 61]. However, there is some evidence that the mere
presence of a visual stimulus exerts a significant influence on audi-
tory cortical responses [62, 63] and hence modulation related to
selective attention might interact with that related to the presence
of visual stimuli in a rather complex manner. This can make com-
parison between the results from bimodal studies [60, 61] and uni-
modal auditory studies [32, 64] somewhat problematic.
One paradigm that has been commonly used to examine
feature-based attention manipulates two different feature dimen-
sions independently within the same experimental session and lis-
teners are required to make a discrimination judgement to one
feature or the other. Studies have compared attention to spatial
features such as location, motion, and ear of presentation with
attention to nonspatial features such as pitch and phonemes [60,
fMRI of the Central Auditory System 595
2.295
2.285
2.280
2.275
Just listen Attend Attend
BF off BF
Fig. 10 Response to the three listening conditions: just listen, attend to the best frequency (BF) tones, and
attend OFF BF. The data shown are for those stimuli in which BF tones formed the majority (80 %) of the tones
in the sound sequence, combining responses across areas 1–3. The error bars denote the 95 % confidence
intervals
3 Future Directions
3 3
Intensity above normal Equivalent loudness
hearing threshold percept
(% signal change)
2 2
BOLD activation
1 1
0 0
Normal hearing
Hearing impaired
−1 −1
0 20 40 60 80 0 20 40 60 80
intensity (dB) loudness (dB)
degree of sensitivity to pitch salience was found at all sites, but the
preference appeared greater in the higher centers than in the
cochlear nucleus [76]. Thus, the evidence supports the notion of
an increasing responsiveness to percept attributes of sound
throughout the ascending auditory system, culminating in the
nonprimary auditory cortex. These findings are consistent with the
hierarchical processing of sound attributes.
Encoding the perceptual properties of a sound is integral to
identifying the object properties of that sound source. The nonpri-
mary auditory cortex probably plays a key role in this process
because it has widespread cortical projections to frontal and parietal
brain regions and is therefore ideally suited to access distinct higher
level cortical mechanisms for sound identification and localization.
Recent trends in auditory neuroscience are increasingly concerned
with auditory coding beyond the conventional limits of the audi-
tory cortex (the superior temporal gyrus in humans), particularly
with respect to the hierarchical organization of sensory coding via
dorsal and ventral auditory processing routes. At the top of this
hierarchy stands the brain’s representation of an auditory “object.”
The concept of an auditory object still remains controversial [56].
Although it is clear that the brain needs to code information about
the invariant properties of a sound source, research in this field is
considerably underdeveloped. Future directions are likely to begin
to address critical issues such as the definition of an auditory object,
whether the concept is informative for auditory perception, and
optimal paradigms for studying object coding.
3.2 Cortical Listeners interact with complex auditory environments that, at any
Activation Also one time point, contain multiple auditory objects located at
Reflects Behaviourally dynamically varying spatial locations. One of the primary chal-
Relevant Coding lenges for the auditory system is to analyze this external environ-
ment in order to inform goal-directed behavior. In Sect. 2.2 we
introduced some of the neurophysiological evidence for the impor-
tance of the attentional focus of the task in determining the pattern
of auditory cortical activity [69]. Here, we consider the contribu-
tion of human auditory fMRI research to this question. In
particular, we present the interesting findings of one group who
have started to address how the auditory cortex responds to the
context and the procedural and cognitive demands of the listening
task (see [77] for a review).
In that review, Scheich and colleagues report a series of research
studies in which they suggest that the function of different audi-
tory cortical areas is not determined so much by stimulus features
(such as timbre, pitch, motion, etc.), but rather by the task that is
performed. For example, one study reported the results of two
fMRI experiments in which the same frequency-modulated stimuli
were presented under different task conditions [78]. Top-down
influences strongly affected the strength of the auditory response.
602 Deborah Ann Hall and Aspasia Eleni Paltoglou
References
1. Palmer AR, Bullock DC, Chambers JD (1998) 10. Ravicz ME, Melcher JR (2001) Isolating the
A high-output, high-quality sound system for auditory system from acoustic noise during func-
use in auditory fMRI. Neuroimage 7:S357 tional magnetic resonance imaging: examination
2. Chou CK, McDougall JA, Chan KW (1995) of noise conduction through the ear canal, head,
Absence of radiofrequency heating from audi- and body. J Acoust Soc Am 109(1):216–231
tory implants during magnetic-resonance 11. Price DL, De Wilde JP, Papadaki AM, Curran
imaging. Bioelectromagnetics 16(5):307–316 JS, Kitney RI (2001) Investigation of acoustic
3. Heller JW, Brackmann DE, Tucci DL, noise on 15 MRI scanners from 0.2 T to 3 T. J
Nyenhuis JA, Chou CK (1996) Evaluation Magn Reson Imaging 13(2):288–293
of MRI compatibility of the modified nucleus 12. Hedeen RA, Edelstein WA (1997) Characterization
multichannel auditory brainstem and cochlear and prediction of gradient acoustic noise in MR
implants. Am J Otol 17(5):724–729 imagers. Magn Reson Med 37(1):7–10
4. Shellock FG, Morisoli S, Kanal E (1993) MR 13. Hennel F, Girard F, Loenneker T (1999)
procedures and biomedical implants, mate- “Silent” MRI with soft gradient pulses. Magn
rials, and devices—1993 update. Radiology Reson Med 42:6–10
189(2):587–599 14. Brechmann A, Baumgart F, Scheich H (2002)
5. Weber BP, Neuburger J, Battmer RD, Lenarz Sound-level-dependent representation of fre-
T (1997) Magnetless cochlear implant: rel- quency modulations in human auditory cortex: a
evance of adult experience for children. Am low-noise fMRI study. J Neurophysiol 87:423–433
J Otol 18(6):S50–S51 15. Sumby WH, Pollack I (1954) Visual contribu-
6. Wild DC, Head K, Hall DA (2006) Safe tion to speech intelligibility in noise. J Acoust
magnetic resonance scanning of patients with Soc Am 26(2):212–215
metallic middle ear implants. Clin Otolaryngol 16. Assmann P, Summerfield Q (2004) Perception
31(6):508–510 of speech under adverse conditions. In:
7. Giraud AL, Truy E, Frackowiak R (2001) Greenberg S, Ainsworth WA, Popper AN, Fay
Imaging plasticity in cochlear implant patients. RR (eds) Speech processing in the auditory sys-
Audiol Neurootol 6(6):381–393 tem. Springer, New York, pp 231–308
8. Moelker A, Wielopolski PA, Pattynama PM 17. Healy EW, Moser DC, Morrow-Odom KL,
(2003) Relationship between magnetic field Hall DA, Fridriksson J (2007) Speech percep-
strength and magnetic-resonance-related tion in MRI scanner noise by persons with
acoustic noise levels. MAGMA 16:52–55 aphasia. J Speech Lang Hear Res 50:323–334
9. Foster JR, Hall DA, Summerfield AQ, Palmer 18. Harms MP, Melcher JR (2002) Sound repetition
AR, Bowtell RW (2000) Sound-level measure- rate in the human auditory pathway: representa-
ments and calculations of safe noise dosage during tions in the waveshape and amplitude of fMRI
fMRI at 3T. J Magn Reson Imaging 12:157–163 activation. J Neurophysiol 88:1433–1450
fMRI of the Central Auditory System 603
19. Chambers JD, Akeroyd MA, Summerfield absence of background scanner noise. J Acoust
AQ, Palmer AR (2001) Active control of the Soc Am 109(4):1559–1570
volume acquisition noise in functional mag- 32. Hart HC, Palmer AR, Hall DA (2004) Different
netic resonance imaging: method and psy- areas of human non-primary auditory cortex are
choacoustical evaluation. J Acoust Soc Am activated by sounds with spatial and nonspatial
110(6):3041–3054 properties. Hum Brain Mapp 21:178–190
20. Hall DA, Chambers J, Foster J, Akeroyd MA, 33. Langers DRM, Backes WH, Van Dijk P (2007)
Coxon R, Palmer AR (2009) Acoustic, psycho- Representation of lateralization and tonotopy
physical, and neuroimaging measurements of in primary versus secondary human auditory
the effectiveness of active cancellation during cortex. Neuroimage 34:264–273
auditory functional magnetic resonance imag- 34. Schwarzbauer C, Davis MH, Rodd JM,
ing. J Acoust Soc Am 125(1):347–359 Johnsrude I (2006) Interleaved silent steady
21. Hall DA, Haggard MP, Akeroyd MA, Palmer state (ISSS) imaging: a new sparse imaging
AR, Summerfield AQ, Elliott MR, Gurney EM, method applied to auditory fMRI. Neuroimage
Bowtell RW (1999) ‘Sparse’ temporal sampling 29(3):774–782
in auditory fMRI. Hum Brain Mapp 7:213–223 35. Bregman AS (1990) Auditory scene analysis:
22. Bandettini PA, Jesmanowicz A, Van Kylen J, the perceptual organisation of sound. MIT,
Birn RM, Hyde JS (1998) Functional MRI of Cambridge, MA
brain activation induced by scanner acoustic 36. Fishman YI, Arezzo JC, Steinschneider M
noise. Magn Reson Med 39:410–416 (2004) Auditory stream segregation in mon-
23. Bilecen D, Scheffler K, Schmid N, Tschopp K, key auditory cortex: effects of frequency sepa-
Seelig J (1998) Tonotopic organization of the ration, presentation rate, and tone duration.
human auditory cortex as detected by BOLD- J Acoust Soc Am 116(3):1656–1670
FMRI. Hear Res 126:19–27 37. Fishman YI, Reser DH, Arezzo JC, Steinschneider
24. Hall DA, Summerfield AQ, Gonçalves MS, M (2001) Neural correlates of auditory stream
Foster JR, Palmer AR, Bowtell RW (2000) segregation in primary auditory cortex of the
Time-course of the auditory BOLD response to awake monkey. Hear Res 151:167–187
scanner noise. Magn Reson Med 43:601–606 38. Brosch M, Schreiner CE (1997) Time course of
25. Shah NJ, Jäncke L, Grosse-Ruyken M-L, forward masking tuning curves in cat primary
Müller-Gärtner HW (1999) Influence of auditory cortex. J Neurophysiol 77:923–943
acoustic masking noise in fMRI of the auditory 39. Bartlett EL, Wang X (2005) Long-lasting
cortex during phonetic discrimination. J Magn modulation by stimulus context in primate
Reson Imaging 9(1):19–25 auditory cortex. J Neurophysiol 94:83–104
26. Talavage TM, Edmister WB, Ledden PJ, 40. Harms MP, Guinan JJ, Sigalovsky IS, Melcher
Weisskoff RM (1999) Quantitative assessment JR (2005) Short-term sound temporal enve-
of auditory cortex responses induced by imager lope characteristics determine multisecond time
acoustic noise. Hum Brain Mapp 7(2):79–88 patterns of activity in human auditory cortex as
27. Elliott MR, Bowtell RW, Morris PG (1999) shown by fMRI. J Neurophysiol 93:210–222
The effect of scanner sound in visual, motor, 41. Palmer AR, Hall DA, Sumner C, Barrett DJK,
and auditory functional MRI. Magn Reson Jones S, Nakamoto K, Moore DR (2007)
Med 41(6):1230–1235 Some investigations into non-passive listening.
28. Edmister WB, Talavage TM, Ledden PJ, Hear Res 229:148–157
Weisskoff RM (1999) Improved auditory cor- 42. Hall DA, Edmondson-Jones M, Fridriksson
tex imaging using clustered volume acquisi- J (2006) Periodicity and frequency coding
tions. Hum Brain Mapp 7:89–97 in human auditory cortex. Eur J Neurosci
29. Formisano E, Kim DS, Di Salle F, van de 24:3601–3610
Moortele PF, Ugurbil K, Goebel R (2003) 43. Hall DA, Johnsrude IS, Haggard MP, Palmer
Mirror-symmetric tonotopic maps in human pri- AR, Akeroyd MA, Summerfield AQ (2002)
mary auditory cortex. Neuron 40(4):859–869 Spectral and temporal processing in human
30. Hall DA, Haggard MP, Akeroyd MA, auditory cortex. Cereb Cortex 12:140–149
Summerfield AQ, Palmer AR, Elliott MR, 44. Hart HC, Hall DA, Palmer AR (2003) The
Bowtell RW (2000) Modulation and task sound-level-dependent growth in the extent of
effects in auditory processing measured using fMRI activation in Heschl’s gyrus is different
fMRI. Hum Brain Mapp 10(3):107–119 for low- and high-frequency tones. Hear Res
31. Hall DA, Haggard MP, Summerfield AQ, 179(1–2):104–112
Akeroyd MA, Palmer AR, Bowtell RW (2001) 45. Von Békésy G (1947) The variations of phase
Functional magnetic resonance imaging mea- along the basilar membrane with sinusoidal
surements of sound-level encoding in the vibrations. J Acoust Soc Am 19:452–460
604 Deborah Ann Hall and Aspasia Eleni Paltoglou
46. Kosaki H, Hashikawa T, He J, Jones EG (1997) 62. Kayser C, Petkov CI, Augath M, Logothetis
Tonotopic organization of auditory cortical fields NK (2007) Functional imaging reveals visual
delineated by parvalbumin immunoreactivity in modulation of specific fields in auditory cortex.
macaque monkeys. J Comp Neurol 386:304–316 J Neurosci 27(8):1824–1835
47. Merzenich MM, Brugge JF (1973) 63. Lehmann C, Herdener M, Esposito F, Hubl
Representation of the cochlear partition on the D, di Salle F, Scheffler K, Bach DR, Federspiel
superior temporal plane of the macaque mon- A, Kretz R, Dierks T, Seifritz E (2006)
key. Brain Res 50:275–296 Differential patterns of multisensory interac-
48. Petkov CL, Kayser C, Augath M, Logothetis tions in core and belt areas of human auditory
NK (2006) Functional imaging reveals numer- cortex. Neuroimage 31(1):294–300
ous fields in the monkey auditory cortex. PLoS 64. Ahveninen J, Jaaskelainen IP, Raij T, Bonmassar
Biol 4(7):213–226 G, Devore S, Hamalainen M, Levanen S, Lin
49. Hall DA, Hart HC, Johnsrude IS (2003) F-H, Sams M, Shinn-Cunningham BG, Witzel
Relationships between human auditory corti- T, Belliveau JW (2006) Task-modulated “what”
cal structure and function. Audiol Neurootol and “where” pathways in human auditory cortex.
8(1):1–18 Proc Natl Acad Sci U S A 103(39):14608–14613
50. Schönwiesner M, Von Cramon DY, Rubsamen 65. Lewald J, Meister IG, Weidemann J, Topper R
R (2002) Is it tonotopy after all? Neuroimage (2004) Involvement of the superior temporal
17:1144–1161 cortex and the occipital cortex in spatial hearing:
51. Talavage TM, Ledden PJ, Benson RR, Rosen evidence from repetitive transcranial magnetic
BR, Melcher JR (2000) Frequency-dependent stimulation. J Cogn Neurosci 16(5):828–838
responses exhibited by multiple regions in 66. Degerman A, Rinne T, Pekkola J, Autti T,
human auditory cortex. Hear Res 150:225–244 Jaaskelainen IP, Sams M, Alho K (2007)
52. Talavage TM, Sereno MI, Melcher JR, Ledden Human brain activity associated with audio-
PJ, Rosen BR, Dale AM (2004) Tonotopic visual perception and attention. Neuroimage
organization in human auditory cortex 34(4):1683–1691
revealed by progressions of frequency sensitiv- 67. Greenberg GZ, Larkin WD (1968) Frequency-
ity. J Neurophysiol 91:1282–1296 response characteristic of auditory observers
53. Kastner S, Ungerleider LG (2000) Mechanisms detecting signals of a single frequency in noise:
of visual attention in the human cortex. Annu the probe-signal method. J Acoust Soc Am
Rev Neurosci 23(1):315–341 44(6):1513–1523
54. Maunsell JHR, Treue S (2006) Feature-based 68. Schlauch RS, Hafter ER (1991) Listening
attention in visual cortex. Trends Neurosci bandwidths and frequency uncertainty in
29(6):317–322 pure-tone signal detection. J Acoust Soc Am
90(3):1332–1339
55. Scholl BJ (2001) Objects and attention: the
state of the art. Cognition 80(1–2):1–46 69. Fritz JB, Elhilali M, David SV, Shamma SA
(2007) Does attention play a role in dynamic
56. Griffiths TD, Warren JD (2004) What receptive field adaptation to changing acoustic
is an auditory object? Nat Rev Neurosci salience in A1? Hear Res 229:186–203
5(11):887–892
70. Shackleton TM, Carlyon RP (1994) The role
57. Johnson JA, Zatorre RJ (2005) Attention to of resolved and unresolved harmonics in pitch
simultaneous unrelated auditory and visual perception and frequency modulation discrimi-
events: behavioral and neural correlates. Cereb nation. J Acoust Soc Am 95:3529–3540
Cortex 15(10):1609–1620
71. Penagos H, Melcher JR, Oxenham AJ (2004)
58. Johnson JA, Zatorre RJ (2006) Neural sub- A neural representation of pitch salience in
strates for dividing and focusing attention nonprimary human auditory cortex revealed
between simultaneous auditory and visual with functional magnetic resonance imaging.
events. Neuroimage 31(4):1673–1681 J Neurosci 24(30):6810–6815
59. Lavie N (2005) Distracted and confused? 72. Griffiths TD, Büchel C, Frackowiak RSJ,
Selective attention under load. Trends Cogn Patterson RD (1998) Analysis of temporal
Sci 9(2):75–82 structure in sound by the human brain. Nat
60. Degerman A, Rinne T, Salmi J, Salonen O, Neurosci 1:422–427
Alho K (2006) Selective attention to sound 73. Blauert J, Lindemann W (1986) Spatial map-
location or pitch studied with fMRI. Brain Res ping of intracranial auditory events for various
1077(1):123–134 degrees of interaural coherence. J Acoust Soc
61. Petkov CI, Kang X, Alho K, Bertrand O, Yund Am 79(3):806–813
EW, Woods DL (2004) Attentional modula- 74. Budd TW, Hall DA, Goncalves MS,
tion of human auditory cortex. Nat Neurosci Akeroyd MA, Foster JR, Palmer AR, Head
7(6):658–663 K, Summerfield AQ (2003) Binaural
fMRI of the Central Auditory System 605
Abstract
The variable effectiveness of reparative and recovery mechanisms following tissue damage is among the
factors that might contribute to explain, at least partially, the paucity of the correlation between clinical
and magnetic resonance imaging (MRI) findings in patients with white matter disorders. Among the
mechanisms of recovery, brain plasticity is likely to be one of the most important with several possible dif-
ferent substrates (including increased axonal expression of sodium channels, synaptic changes, increased
recruitment of parallel existing pathways or “latent” connections, and reorganization of distant sites). The
application of fMRI has shown that plastic cortical changes do occur after white matter injury of different
etiology, that such changes are related to the extent of white matter damage, and that they can contribute
in limiting the clinical consequences of brain damage. Conversely, the failure or exhaustion of the adaptive
properties of the cerebral cortex might be among the factors responsible for the accumulation of “fixed”
neurological deficits in patients with white matter disorders.
Key words Multiple sclerosis, Functional magnetic resonance imaging, White matter, Adaptation,
Maladaptation, Myelitis, Vasculitides
1 Introduction
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_20, © Springer Science+Business Media New York 2016
609
610 Massimo Filippi and Maria A. Rocca
2 fMRI in MS
2.1 General The main problem in the interpretation of fMRI studies in dis-
Considerations eased people is that the observed changes might be biased by
differences in task performance between patients and controls.
Clearly, this is a major issue in MS, which typically causes impair-
ment of various functional systems. Therefore, despite providing
several important pieces of information, the value of the earliest
fMRI studies of patients with MS [7–13] has to be weight against
this background. For this reason, more recent fMRI studies in
MS have been based on larger and more selected patients’ groups
than the seminal studies. These studies have investigated the
brain patterns of cortical activations during the performance of a
number of motor, visual, and cognitive tasks in patients with all
the major clinical phenotypes of the disease. Another appealing
strategy which has been introduced for the study of functional
network rewiring in clinically impaired patients is based on the
assessment of functional abnormalities at rest in the main brain
functional networks (resting state networks). One of the most
solid conclusions that can be drawn from fMRI studies of MS is
that cortical reorganization does occur in patients affected by
this condition. The correlation between various measures of
structural MS damage and the extent of cortical activations also
suggests an adaptive role of such cortical changes in contributing
to clinical recovery and maintaining a normal level of function-
ing in patients with MS, despite the presence of irreversible
axonal/neuronal loss.
Application of fMRI to Multiple Sclerosis and Other White Matter Disorders 611
2.2 Visual System The method usually applied to investigate the visual system con-
sists of the application of a 8 Hz photic stimulation to one or both
eyes [8, 12–18]. A study of the visual system [12] in patients who
had recovered from a single episode of acute unilateral optic neu-
ritis demonstrated that these patients, relative to healthy volun-
teers, had an extensive activation of the visual network, including
the claustrum, lateral temporal and posterior parietal cortices, and
thalamus, in addition to the primary visual cortex, when the clini-
cally affected eye was studied. When the unaffected eye was stimu-
lated, only activations of the visual cortex and the right insula/
claustrum were observed. A strong correlation was found in these
patients between the volume of the extra-occipital activation and
the latency of the visual evoked potential (VEP) P100, suggesting
that the functional reorganization of the cortex might represent an
adaptive response to a persistently abnormal visual input. The
results of this preliminary study have been confirmed and extended
by subsequent studies [14, 15, 17]. Using fMRI and VEP to mon-
itor the functional recovery after an acute unilateral optic neuritis,
Russ et al. [15] found a strong relationship between fMRI and
VEP latencies, suggesting that fMRI might contribute to the
assessment of the temporal evolution of the visual deficits during
recovery. Levin et al. [17] showed reduced activation of the pri-
mary visual cortex and increased activation of the lateral occipital
complex (LOC) in eight subjects who recovered clinically from an
episode of optic neuritis, but who still had prolonged VEP laten-
cies in comparison with healthy controls.
Structural MRI, electrophysiology, and fMRI have been com-
bined in another study to investigate why ON patients exhibit a
wide variation in severity of acute visual loss. Optic nerve lesion
length and VEP amplitude were associated with visual loss. Bilateral
activation in the extra-striate occipital cortex correlated directly
with vision, after adjusting for optic nerve lesion length, VEP
amplitude, and demographic characteristics [19] (Fig. 1).
These data suggest that acute visual loss is associated with the
extent of inflammation and conduction block in the optic nerve,
but not with pathology in the optic radiations or occipital cortex.
The association of better vision with greater fMRI responses, after
accounting for factors which reduce afferent input, suggested a
role for adaptive neuroplasticity within the association cortex of
the dorsal stream of higher visual processing [19]. In a 1-year
follow-up study, Toosy et al. [14], using a novel technique that
modeled the fMRI response and optic nerve structure together
with clinical function, demonstrated a potential adaptive role of
cortical reorganization within the extra-striate visual areas. An
increased optic nerve gadolinium-enhanced lesion length at base-
line was associated with a reduced functional activation within the
visual cortex and poorer vision. At 3 months, more severe optic
nerve damage was associated with an increased fMRI response in
612 Massimo Filippi and Maria A. Rocca
2
2 1
1
BOLD_response
0
0
-1 -2
0 .5 1 1.5 2
logmar_aff1
Fig. 1 Statistical parametric maps showing group correlations between functional magnetic resonance imag-
ing (fMRI) response and visual acuity, after correcting for age, gender, side affected, gadolinium-enhanced
lesion length, fast spin-echo lesion length, and visual evoked potential (VEP) amplitude. A correlation is seen
in the region of the cuneus bilaterally, where better visual acuity is associated with a greater fMRI response.
The graph plots logMAR visual acuity against the mean corrected fMRI response (approximate percentage
blood oxygenated level dependent signal change), at the peak voxel. The statistical parametric maps are thres-
holded at cluster level p < 0.05 (corrected), and the scale bar indicates the voxel level t-scores (from ref. [19])
2.3 Motor System The investigation of the motor system in patients with MS has
mainly focused on the analysis of the performance of simple motor
tasks with the dominant right upper limbs [9–11, 21–40]. Such
tasks were either self-paced or paced by a metronome. A few stud-
ies assessed the performance of simple motor tasks with the domi-
nant right lower limbs [23, 27, 33], while even fewer studies have
investigated the performance of more complex tasks, including
phasic movements of dominant hand and foot [27, 33], object
manipulation [41], and visuomotor integration tasks [42].
An altered brain pattern of movement-associated cortical acti-
vations, characterized by an increased recruitment of the contralat-
eral primary sensorimotor cortex (SMC) during the performance of
simple tasks [23, 27] and by the recruitment of additional “classi-
cal” and “higher-order” sensorimotor areas during the performance
of more complex tasks [27], has been demonstrated in patients with
clinically isolated syndrome (CIS) suggestive of MS. The clinical
and conventional MRI follow-up of these patients has shown that,
at disease onset, CIS patients with a subsequent evolution to clini-
cally definite MS tend to recruit a more widespread sensorimotor
network than those without short-term disease evolution [36].
These findings suggest that in CIS patients the extent of early corti-
cal reorganization might be a factor associated with a different clini-
cal evolution. This would support the notion that, whereas increased
recruitment of a widespread sensorimotor network contributes to
614 Massimo Filippi and Maria A. Rocca
Fig. 2 Comparisons of patients at presentation with clinically isolated syndrome (CIS) suggestive of MS and
patients with relapsing-remitting (RR) MS and no disability during a simple, right-hand, motor task. Patients
with CIS showed an increased activation of the contralateral primary sensorimotor cortex when compared with
patients with RRMS and no disability (top row). Patients with RRMS and no disability had a more significant
activation of the supplementary motor area, bilaterally, when compared with patients with a CIS (bottom row).
Images are color-coded for activation and arrows show t cut-off values. Activations were superimposed on a
high-resolution T1-weighted scan obtained from one healthy individual and normalized into a standard statisti-
cal parametric mapping space (neurological convention) (from ref. [28])
2.4 Cognition Several fMRI studies have suggested that functional cortical
changes might have an adaptive role also in limiting MS-related
cognitive impairment [52–68]. Therefore, brain plasticity might,
in part, explain the weak relationship found in MS between neuro-
psychological deficits and conventional MRI measures of disease
burden [69].
Several cognitive domains have been investigated in MS
patients with fMRI. Working memory has been the most exten-
sively studied by means of the Paced Auditory Serial Addition Test
(PASAT) or the Paced Visual Serial Addition Task (PVSAT) [52–
55, 60, 64, 70] (which also involve sustained attention, information
processing speed, and simple calculation), the n-back task [59,
61–63, 65, 71], or a task adapted from the Sternberg paradigm
[57]. Additional cognitive domains including attention [58], epi-
sodic memory [72], planning [68], and emotional processing [73]
have also been interrogated.
In patients at presentation with CIS suggestive of MS, an
altered pattern of cortical activations has been described during the
performance of the PASAT [55, 56, 70], confirming the presence
of cortical reorganization at the earliest clinical stage of the disease.
Staffen et al. [52] found that, during the performance of the
PVSAT, MS patients with intact task performance had an increased
activation of several regions located in the frontal and parietal
lobes, bilaterally, compared with healthy volunteers, suggesting the
Application of fMRI to Multiple Sclerosis and Other White Matter Disorders 617
Fig. 3 Brain patterns of cortical activations on a rendered brain during the execution of the Paced Auditory
Serial Addition Task (PASAT) in (a) 22 healthy controls and in (b) 22 patients with MS. (b1, b2) Rendered images
for patients with MS subgrouped according to their performance at the PASAT during fMRI showing significant
activated foci in (b1) for 12 patients whose performance was similar to that of healthy controls and in (b2) the
activations found in the 10 patients who exhibited lower scores (from ref. [60])
Fig. 4 Analysis of functional connectivity during the performance of the n-back task with different levels of
difficulty in healthy individuals and patients with MS. The most significant correlations between activation in
regions involved in processing increasing task demand are indicated in (a). In (b) are those connections more
significant for controls (p < 0.05). The image in (c) shows connections that were more significant in patients
than controls (p < 0.05). C cingulate, SF superior medial frontal, RF right dorsolateral prefrontal, LF left dorso-
lateral prefrontal, RP right parietal, LP left parietal (from ref. [65])
Fig. 5 Areas showing increased activations in patients with BMS in comparison with healthy controls during
the analysis of the Stroop facilitation condition (random effect interaction analysis, ANOVA, p < 0.05 corrected
for multiple comparisons). BMS patients had increased activations of several areas located in the frontal and
parietal lobes, bilaterally, including the anterior cingulate cortex, the superior frontal sulcus, the inferior frontal
gyrus, the precuneus, the secondary sensorimotor cortex, the bilateral visual cortex, and the cerebellum, bilat-
erally. Note that the color-encoded activations have been superimposed on a rendered brain and normalized
into standard SPM space (neurological convention) (see ref. [67])
2.6 Functional Structural damage of white matter pathways that connect functional
Cortical relevant areas for a given task has been shown to modify the observed
Reorganization brain patterns of cortical activations in patients with MS. Damage to
and Regional Damage the corticospinal tract [30, 31] (Fig. 6) as well as damage to the
in MS corpus callosum (CC) [75, 76] has been related to a more bilateral
movement-associated brain pattern of cortical activations.
The role of the CC in interhemispheric connectivity and in elic-
iting functional cortical changes has been underpinned by a study by
Lowe et al. [32], who showed, by measuring low-frequency BOLD
fluctuations, a reduced functional connectivity between the right
and the left hemisphere primary motor cortices in MS patients.
The recent development of diffusion-based tractography
methods that allow to define with precision the pathways connect-
ing different CNS structures and their application to patients with
MS resulted in an improvement of the correlation between struc-
tural and functional abnormalities. Several studies combined
measures of abnormal functional connectivity with DT MR mea-
sures of damage within selected white matter fiber bundles in
patients with RRMS [40], BMS [67] and PPMS [77]. In patients
with RRMS and no clinical disability [40], measures of abnormal
connectivity inside the motor network were correlated with
Application of fMRI to Multiple Sclerosis and Other White Matter Disorders 621
2.7 Adaptive Role Although the actual role of cortical reorganization on the clinical
of Functional Cortical manifestations of MS remains to be established, there are several
Reorganization in MS pieces of evidence which suggest that cortical adaptive changes are
likely to contribute in limiting the clinical consequences of MS-related
structural damage. In nondisabled patients with RRMS [22], an
increased activation of several motor regions, mainly located in the
contralateral cerebral hemisphere, has been seen during the perfor-
mance of a simple motor task. The correlations found in this study
[22] between the extent of fMRI activations and several MT and DT
MRI metrics of structural brain damage suggested that an increased
recruitment of movement-associated cortical network contributes to
limiting the functional impact of MS-related damage.
The notion that an increased recruitment of areas that are usu-
ally activated by healthy individuals when performing different/
more complex motor tasks might be one of the mechanisms play-
ing a role in MS recovery/maintenance of function has been high-
lighted by the results of two experiments [39, 41]. The first showed
that MS patients, during the performance of a simple motor task,
activate some regions that are part of a fronto-parietal circuit,
whose recruitment occurs typically in healthy subjects during
object manipulation (Fig. 7) [41]. The second, which assessed the
fMRI patterns of activation during the performance of a simple
motor task and of a task aimed at investigating the mirror-neuron
system, demonstrated activations of regions that are part of the
mirror-neuron system in patients with MS during the performance
of the simple motor task [39].
The compensatory role of cortical reorganization has also been
demonstrated by studies investigating the cognitive domains,
which showed increased recruitment of several cortico-subcortical
areas in cognitively preserved MS patients [52–58]. In patients
complaining of fatigue, when compared with matched nonfatigued
MS patients [78], a reduced activation of a complex movement-
associated cortical/subcortical network, including the cerebellum,
the thalamus, and regions in the frontal lobes, has been shown.
The correlation found in these patients between the reduction of
thalamic activity and the clinical severity of fatigue indicates that a
“pseudoreduction” of brain functional recruitment might be asso-
ciated with the appearance of MS symptomatology. Additional
work has shown that the pattern of movement-associated cortical
activations in MS is determined by both the extent of brain injury
and disability and that these changes are distinct [26, 33].
2.8 Maladaptive Role The results of several studies suggest that an increased cortical
of Functional Cortical recruitment might not always be beneficial for patients with MS. As
Reorganization already mentioned, disease progression and accrual of disability has
been observed in patients with SPMS, despite the widespread acti-
vations of regions in the frontal and parietal lobes during the per-
formance of simple motor tasks [24, 28]. fMRI studies of the
Application of fMRI to Multiple Sclerosis and Other White Matter Disorders 623
Fig. 7 Comparison of simple vs. complex task with the dominant right hands in healthy subjects (top row; a–c)
and patients with MS (bottom row; d–g) (paired t test for each group, corrected p value <0.05). The ipsilateral
anterior lobe of the cerebellum (a and d), bilateral insula/basal-ganglia (b, e, f) and contralateral primary sen-
sorimotor cortex and supplementary motor area (c and g) were identified in both groups. Compared with
healthy subjects, MS patients also had a significant activation of the contralateral inferior frontal gyrus and
bilateral secondary sensorimotor cortex (e). Note that the activations are color-coded according to their t val-
ues. Images are in neurological convention (see ref. [41])
motor system [21, 33, 44] of patients with PPMS suggested a lack
of “classical” adaptive mechanisms as a potential additional factor
contributing to the accumulation of disability. In these patients
during the performance of different motor tasks with the nonim-
paired dominant limbs, a recruitment of a widespread movement-
associated cortical network usually considered to function in
motor, sensory, and multimodal integration processing (i.e., the
frontal and temporal lobes, and the insula) was detected [21, 33,
44]. The absence of a concomitant recruitment of the “classical”
motor areas, including the primary SMC, the SMA, the infrapari-
etal sulcus, and the SII, was interpreted as a failure of part of the
adaptive capacity of the cerebral cortex in this severely disabling
phenotype of the disease [21, 33]. The notion that multimodal
integration areas might have a critical role in PPMS patients has
been strengthened by another study which showed increased
624 Massimo Filippi and Maria A. Rocca
2.9 Use of fMRI Dynamic functional changes have been described in an MS patient
to Assess Longitudinal following an acute relapse [10]. These results have been confirmed
Changes of Cortical and extended by another study which assessed the early cortical
Reorganization changes following acute motor relapses secondary to pseudotu-
moral lesions in 12 MS patients and the evolution over time of
cortical reorganization in a subgroup of these patients [38]. Short-
term cortical changes were mainly characterized by the recruitment
of pathways in the unaffected hemisphere. A recovery of function
of the primary SMC of the affected hemisphere was found in
patients with clinical improvement, while in patients without clini-
cal recovery, there was a persistent recruitment of the primary
SMC of the unaffected hemisphere, suggesting that the restoration
of function of motor areas of the affected hemisphere might be a
critical factor for a favorable recovery (Fig. 9).
A longitudinal (time interval of 15–26 months) fMRI study of
the motor system has been conducted in a group of patients with
early RRMS [93]. Patients exhibited greater bilateral activations
than controls in both fMRI studies. Although no significant
differences between the two fMRI scans were observed in controls,
a reduction of the functional activity of the ipsilateral SMC and the
contralateral cerebellum was seen in patients at follow-up.
Moreover, activation changes in ipsilateral motor areas correlated
inversely with age, extent, and progression of T1 lesion load, and
occurrence of a new relapse, suggesting that younger patients with
less structural brain damage and a favorable clinical course demon-
strate brain plasticity that follows a more lateralized pattern of
brain activations [93].
Longitudinal modifications of cognitive networks’ recruitment
and their impact on patients’ cognitive status have been marginally
explored. A 1-year longitudinal study in patients with early MS
found an association between increased levels of activation in the
right dorsolateral prefrontal cortex during a cognitive task and
improved working memory and processing speed performance
[94]. A 20 month longitudinal study [95] showed that worsening
of SDMT performance in RRMS patients is correlated with
increased activity of the left inferior parietal lobule over time, prob-
ably reflecting a maladaptive mechanism.
Application of fMRI to Multiple Sclerosis and Other White Matter Disorders 627
Fig. 9 Longitudinal evolution of cortical activations in the primary sensorimotor cortex (SMC), bilaterally, during
task performance with impaired hand compared with unimpaired hand in one patient with good clinical recov-
ery during follow-up (a and b) and in one patient with poor/absent clinical recovery (c and d). Scans obtained
during left hand motor task have been flipped to keep the left hemisphere contralateral to movement. At
baseline, both patients showed an increased activation of the primary SMC of the unaffected (ipsilateral)
hemisphere (a and c). During follow-up, the patient with good clinical recovery showed an increased recruit-
ment of the primary SMC of the affected hemisphere (b), while the patient with poor/absent clinical recovery
continued to show an increased recruitment of the primary SMC of the unaffected hemisphere (d). Note that
the activations are color-coded according to their t values (see ref. [38])
2.10 fMRI to Monitor The potential of fMRI in a multicentre setting has been explored by
Treatment a few studies of the motor [96, 97] and cognitive [92] networks.
Only a few fMRI studies have been performed to monitor the
effect of treatments in MS [58, 98, 99]. Different patterns of
brain response to lower dose acute and higher dose chronic
administration of rivastigmine have been demonstrated in MS
patients using cognitive tasks [58, 99]. Rivastigmine was shown to
enhance the prefrontal function and alter the functional connec-
tivity associated with cognition. In MS patients, increased activa-
tion in the ipsilateral primary SMC and SMA has been observed
after a single dose of 3,4-diaminopyridine (a potassium channel
628 Massimo Filippi and Maria A. Rocca
4.1 Isolated Spinal Studies of patients with spinal cord injury of different etiology (i.e.,
Cord Injury traumatic and/or demyelinating) with no or only partial clinical
recovery have shown movement-associated cortical changes, consist-
ing of an abnormal location of the activated areas and in a more wide-
spread recruitment of motor areas, mainly located in the hemisphere
contralateral to the limb used to perform the task [106–108].
In patients with isolated myelitis of probable demyelinating
origin and normal function in the investigated limbs, an abnormal
pattern of movement-associated cortical activation has been
described [34, 109, 110] and has been related to the degree of
daily hand use [109], the severity of cervical cord damage [34,
110], and the level of spinal cord involvement [110].
Application of fMRI to Multiple Sclerosis and Other White Matter Disorders 629
a b e 1
5
3 0.5
LSMC response
1
0
c d
0
-0.5
0 5 10 15
Dual-echo lesion load (ml)
Fig. 10 Relative cortical activations in patients with neuropsychiatric systemic lupus erythematosus during a
simple motor task with the right hand in comparison with healthy volunteers (color-coded t values). (a)
Contralateral primary sensorimotor cortex. (b) Contralateral putamen, contralateral middle frontal gyrus (MFG),
bilateral MT/V5 complex, contralateral middle occipital gyrus (MOG). (c) Contralateral putamen, ipsilateral
inferior frontal gyrus, bilateral MT/V5 complex, ipsilateral MOG. (d) Contralateral dentate nucleus. The relative
activation of the contralateral primary sensorimotor cortex was significantly correlated with brain dual-echo
lesion load (e) (p < 0.001, r = 0.79) (from ref. [111])
630 Massimo Filippi and Maria A. Rocca
Fig. 11 Source localization and time of onset of the blood oxygenation level-dependent (BOLD) signal changes
during induced and spontaneous visual aura attack in patients with migraine: (a) data are represented on
inflated cortical surface shown from a posterior-medial view; (b) a fully flattened view of the cortical surface
of the involved region. Cortical regions showing the first BOLD perturbations are coded in red (according to the
color-coded scale representing variation in time) and locations showing the BOLD perturbations at progres-
sively later times are coded by green and blue (according to the color-coded scale representing variation in
time). The aura-related changes appeared first in extrastriate cortex (area V3A) and then progressed contigu-
ously and slowly over the cortex following the same retinotopic progression of visual disturbance. From [112]
with permission
5 Conclusions
References
1. Filippi M, Rocca MA (2004) Magnetization ing optic neuritis recovery. Neurosci Lett
transfer magnetic resonance imaging in 330(3):255–259
the assessment of neurological diseases. 15. Russ MO et al (2002) Functional magnetic
J Neuroimaging 14(4):303–313 resonance imaging in acute unilateral optic
2. Filippi M, Rocca MA, Comi G (2003) The neuritis. J Neuroimaging 12(4):339–350
use of quantitative magnetic-resonance-based 16. Toosy AT et al (2005) Adaptive cortical plas-
techniques to monitor the evolution of mul- ticity in higher visual areas after acute optic
tiple sclerosis. Lancet Neurol 2(6):337–346 neuritis. Ann Neurol 57(5):622–633
3. Hesselink JR (2006) Differential diagnostic 17. Levin N et al (2006) Normal and abnor-
approach to MR imaging of white matter dis- mal fMRI activation patterns in the visual
eases. Top Magn Reson Imaging 17:243–263 cortex after recovery from optic neuritis.
4. Rocca MA, Filippi M (2006) Functional MRI Neuroimage 33(4):1161–1168
to study brain plasticity in clinical neurology. 18. Korsholm K et al (2007) Recovery from optic
Neurol Sci 27(Suppl 1):S24–S26 neuritis: an ROI-based analysis of LGN and
5. Rocca MA, Filippi M (2007) Functional visual cortical areas. Brain 130(Pt 5):1244–1253
MRI in multiple sclerosis. J Neuroimaging 19. Jenkins T et al (2010) Dissecting structure-
17(Suppl 1):36S–41S function interactions in acute optic neuritis to
6. Waxman SG (1998) Demyelinating diseases— investigate neuroplasticity. Hum Brain Mapp
new pathological insights, new therapeutic 31(2):276–286
targets. N Engl J Med 338(5):323–325 20. Jenkins TM et al (2010) Neuroplasticity pre-
7. Clanet M, Berry I, Boulanouar K (1997) dicts outcome of optic neuritis independent
Functional imaging in multiple sclerosis. Int of tissue damage. Ann Neurol 67(1):99–113
MS J 4:26–32 21. Filippi M et al (2002) Correlations
8. Rombouts SA et al (1998) Visual activa- between structural CNS damage and func-
tion patterns in patients with optic neu- tional MRI changes in primary progressive
ritis: an fMRI pilot study. Neurology MS. Neuroimage 15(3):537–546
50(6):1896–1899 22. Rocca MA et al (2002) Adaptive functional
9. Lee M et al (2000) The motor cortex shows changes in the cerebral cortex of patients with
adaptive functional changes to brain injury from nondisabling multiple sclerosis correlate with
multiple sclerosis. Ann Neurol 47(5):606–613 the extent of brain structural damage. Ann
10. Reddy H et al (2000) Relating axonal injury Neurol 51(3):330–339
to functional recovery in MS. Neurology 23. Rocca MA et al (2003) Evidence for axonal
54(1):236–239 pathology and adaptive cortical reorganiza-
11. Reddy H et al (2000) Evidence for adaptive tion in patients at presentation with clinically
functional changes in the cerebral cortex with isolated syndromes suggestive of multiple
axonal injury from multiple sclerosis. Brain sclerosis. Neuroimage 18(4):847–855
123(Pt 11):2314–2320 24. Rocca MA et al (2003) A functional mag-
12. Werring DJ et al (2000) Recovery from optic netic resonance imaging study of patients
neuritis is associated with a change in the distri- with secondary progressive multiple sclerosis.
bution of cerebral response to visual stimulation: Neuroimage 19(4):1770–1777
a functional magnetic resonance imaging study. 25. Rocca MA et al (2003) Functional cortical
J Neurol Neurosurg Psychiatry 68(4):441–449 changes in patients with multiple sclerosis and
13. Langkilde AR et al (2002) Functional MRI of nonspecific findings on conventional mag-
the visual cortex and visual testing in patients netic resonance imaging scans of the brain.
with previous optic neuritis. Eur J Neurol Neuroimage 19(3):826–836
9(3):277–286 26. Reddy H et al (2002) Functional brain reorgani-
14. Toosy AT et al (2002) Functional magnetic zation for hand movement in patients with mul-
resonance imaging of the cortical response tiple sclerosis: defining distinct effects of injury
to photic stimulation in humans follow- and disability. Brain 125(Pt 12):2646–2657
634 Massimo Filippi and Maria A. Rocca
27. Filippi M et al (2004) Simple and complex 41. Filippi M et al (2004) A functional MRI
movement-associated functional MRI changes study of cortical activations associated
in patients at presentation with clinically iso- with object manipulation in patients with
lated syndromes suggestive of multiple sclero- MS. Neuroimage 21(3):1147–1154
sis. Hum Brain Mapp 21(2):108–117 42. Cerasa A et al (2006) Adaptive cortical
28. Rocca MA et al (2005) Cortical adaptation in changes and the functional correlates of visuo-
patients with MS: a cross-sectional functional motor integration in relapsing-remitting mul-
MRI study of disease phenotypes. Lancet tiple sclerosis. Brain Res Bull 69(6):597–605
Neurol 4(10):618–626 43. Calautti C, Baron J-C (2003) Functional
29. Pantano P et al (2002) Cortical motor reor- neuroimaging studies of motor recov-
ganization after a single clinical attack of mul- ery after stroke in adults: a review. Stroke
tiple sclerosis. Brain 125(Pt 7):1607–1615 34(6):1553–1566
30. Pantano P et al (2002) Contribution of 44. Ciccarelli O et al (2006) Functional response
corticospinal tract damage to cortical to active and passive ankle movements with
motor reorganization after a single clini- clinical correlations in patients with pri-
cal attack of multiple sclerosis. Neuroimage mary progressive multiple sclerosis. J Neurol
17(4):1837–1843 253(7):882–891
31. Rocca MA et al (2004) Pyramidal tract lesions 45. Rocca MA et al (2010) Preserved brain adap-
and movement-associated cortical recruitment in tive properties in patients with benign mul-
patients with MS. Neuroimage 23(1):141–147 tiple sclerosis. Neurology 74(2):142–149
32. Lowe MJ et al (2002) Multiple sclerosis: 46. Petsas N et al (2013) Evidence of impaired
low-frequency temporal blood oxygen level- brain activity balance after passive sensorimo-
dependent fluctuations indicate reduced func- tor stimulation in multiple sclerosis. PLoS
tional connectivity initial results. Radiology One 8(6):e65315
224(1):184–192 47. Agosta F et al (2008) Tactile-associated
33. Rocca MA et al (2002) Evidence for wide- recruitment of the cervical cord is altered in
spread movement-associated functional MRI patients with multiple sclerosis. Neuroimage
changes in patients with PPMS. Neurology 39(4):1542–1548
58(6):866–872 48. Agosta F et al (2008) Evidence for enhanced
34. Rocca MA et al (2003) Cord damage elic- functional activity of cervical cord in relaps-
its brain functional reorganization after ing multiple sclerosis. Magn Reson Med
a single episode of myelitis. Neurology 59(5):1035–1042
61(8):1078–1085 49. Valsasina P et al (2010) Cervical cord func-
35. Rocca MA et al (2004) A functional MRI tional MRI changes in relapse-onset MS
study of movement-associated cortical patients. J Neurol Neurosurg Psychiatry
changes in patients with Devic’s neuromyeli- 81(4):405–408
tis optica. Neuroimage 21(3):1061–1068 50. Valsasina P et al (2012) Cervical cord fMRI
36. Rocca MA et al (2005) A widespread pat- abnormalities differ between the progressive
tern of cortical activations in patients at forms of multiple sclerosis. Hum Brain Mapp
presentation with clinically isolated symp- 33(9):2072–2080
toms is associated with evolution to definite 51. Rocca MA et al (2012) Abnormal cervical
multiple sclerosis. AJNR Am J Neuroradiol cord function contributes to fatigue in mul-
26(5):1136–1139 tiple sclerosis. Mult Scler 18(11):1552–1559
37. Rocca MA et al (2007) fMRI changes in 52. Staffen W et al (2002) Cognitive function
relapsing-remitting multiple sclerosis patients and fMRI in patients with multiple sclerosis:
complaining of fatigue after IFNbeta-1a evidence for compensatory cortical activa-
injection. Hum Brain Mapp 28(5):373–382 tion during an attention task. Brain 125(Pt
38. Mezzapesa DM et al (2008) Functional corti- 6):1275–1282
cal changes of the sensorimotor network are 53. Au Duong MV et al (2005) Altered func-
associated with clinical recovery in multiple tional connectivity related to white matter
sclerosis. Hum Brain Mapp 29(5):562–573 changes inside the working memory network
39. Rocca MA et al (2008) The “mirror-neu- at the very early stage of MS. J Cereb Blood
ron system” in MS: a 3 tesla fMRI study. Flow Metab 25(10):1245–1253
Neurology 70(4):255–262 54. Au Duong MV et al (2005) Modulation of
40. Rocca MA et al (2007) Altered functional and effective connectivity inside the working mem-
structural connectivities in patients with MS: ory network in patients at the earliest stage of
a 3-T study. Neurology 69(23):2136–2145 multiple sclerosis. Neuroimage 24(2):533–538
Application of fMRI to Multiple Sclerosis and Other White Matter Disorders 635
55. Audoin B et al (2003) Compensatory cortical moderately advanced multiple sclerosis. Mult
activation observed by fMRI during a cog- Scler 10(5):549–555
nitive task at the earliest stage of MS. Hum 69. Comi G et al (2001) Clinical and MRI assess-
Brain Mapp 20(2):51–58 ment of brain damage in MS. Neurol Sci
56. Audoin B et al (2005) Magnetic resonance 22(Suppl 2):123–127
study of the influence of tissue damage and 70. Forn C et al (2012) Functional magnetic res-
cortical reorganization on PASAT perfor- onance imaging correlates of cognitive per-
mance at the earliest stage of multiple sclero- formance in patients with a clinically isolated
sis. Hum Brain Mapp 24(3):216–228 syndrome suggestive of multiple sclerosis at
57. Hillary FG et al (2003) An investigation of presentation: an activation and connectivity
working memory rehearsal in multiple scle- study. Mult Scler 18(2):153–163
rosis using fMRI. J Clin Exp Neuropsychol 71. Cerasa A et al (2010) The effects of BDNF
25(7):965–978 Val66Met polymorphism on brain function in
58. Parry AM et al (2003) Potentially adaptive controls and patients with multiple sclerosis:
functional changes in cognitive processing for an imaging genetic study. Behav Brain Res
patients with multiple sclerosis and their acute 207(2):377–386
modulation by rivastigmine. Brain 126(Pt 72. Bobholz JA et al (2006) fMRI study of epi-
12):2750–2760 sodic memory in relapsing-remitting MS: cor-
59. Penner IK et al (2003) Analysis of impairment relation with T2 lesion volume. Neurology
related functional architecture in MS patients 67(9):1640–1645
during performance of different attention 73. Jehna M et al (2011) Cognitively preserved
tasks. J Neurol 250(4):461–472 MS patients demonstrate functional differ-
60. Mainero C et al (2004) fMRI evidence ences in processing neutral and emotional
of brain reorganization during attention faces. Brain Imaging Behav 5(4):241–251
and memory tasks in multiple sclerosis. 74. Rocca MA et al (2012) Differential cerebel-
Neuroimage 21(3):858–867 lar functional interactions during an interfer-
61. Sweet LH et al (2004) Functional magnetic ence task across multiple sclerosis phenotypes.
resonance imaging of working memory among Radiology 265(3):864–873
multiple sclerosis patients. J Neuroimaging 75. Lenzi D et al (2007) Effect of corpus callosum
14(2):150–157 damage on ipsilateral motor activation in patients
62. Sweet LH et al (2006) Functional magnetic with multiple sclerosis: a functional and anatom-
resonance imaging response to increased ical study. Hum Brain Mapp 28(7):636–644
verbal working memory demands among 76. Manson SC et al (2006) Loss of interhemi-
patients with multiple sclerosis. Hum Brain spheric inhibition in patients with multiple
Mapp 27(1):28–36 sclerosis is related to corpus callosum atrophy.
63. Wishart HA et al (2004) Brain activation Exp Brain Res 174(4):728–733
patterns associated with working mem- 77. Ceccarelli A et al (2010) Structural and func-
ory in relapsing-remitting MS. Neurology tional magnetic resonance imaging correlates
62(2):234–238 of motor network dysfunction in primary
64. Chiaravalloti N et al (2005) Cerebral activa- progressive multiple sclerosis. Eur J Neurosci
tion patterns during working memory per- 31(7):1273–1280
formance in multiple sclerosis using FMRI. J 78. Filippi M et al (2002) Functional magnetic
Clin Exp Neuropsychol 27(1):33–54 resonance imaging correlates of fatigue in mul-
65. Cader S et al (2006) Reduced brain functional tiple sclerosis. Neuroimage 15(3):559–567
reserve and altered functional connectivity in 79. Raichle ME, Snyder AZ (2007) A default
patients with multiple sclerosis. Brain 129(Pt mode of brain function: a brief history of
2):527–537 an evolving idea. Neuroimage 37(4):1083–
66. Li Y et al (2004) Differential cerebellar activa- 1090, discussion 1097–1099
tion on functional magnetic resonance imag- 80. Roosendaal SD et al (2010) Resting state net-
ing during working memory performance in works change in clinically isolated syndrome.
persons with multiple sclerosis. Arch Phys Brain 133(Pt 6):1612–1621
Med Rehabil 85(4):635–639 81. Rocca MA et al (2010) Default-mode network
67. Rocca MA et al (2009) Structural and func- dysfunction and cognitive impairment in pro-
tional MRI correlates of Stroop control in gressive MS. Neurology 74(16):1252–1259
benign MS. Hum Brain Mapp 30(1):276–290 82. Bonavita S et al (2011) Distributed changes
68. Lazeron RHC et al (2004) An fMRI study of in default-mode resting-state connectivity in
planning-related brain activity in patients with multiple sclerosis. Mult Scler 17(4):411–422
636 Massimo Filippi and Maria A. Rocca
Abstract
Stroke is a major cause of long-term disability worldwide. One of the key factors underpinning recovery of
function is reorganization of surviving neural networks. Noninvasive techniques such as fMRI allow this
reorganization to be studied in humans. However, the design of experiments involving patients with impair-
ment requires careful consideration and is often constrained. Difficulty with some tasks can lead to a number
of performance confounds, and so tasks and task parameters that avoid or minimize this should be selected.
Furthermore, when studying patients with cerebrovascular disease, it is important to consider the possibility
that the blood oxygen level-dependent signal may be altered and affect interpretation of results. Despite these
potential problems, careful experimental design can provide real insights into system-level reorganization
after stroke and how it is related to functional recovery. Currently, results suggest that functionally relevant
reorganization does occur in cerebral networks in human stroke patients. For example, it is apparent that
initial attempts to move a paretic limb following stroke are associated with widespread activity within the
distributed motor system in both cerebral hemispheres. This reliance on nonprimary motor output pathways
is unlikely to support full recovery, but improved efficiency of the surviving networks is associated with
behavioral gains. This reorganization can only occur in structurally and functionally intact brain regions.
Understanding the dynamic process of system-level reorganization will allow greater understanding of the
mechanisms of recovery and potentially improve our ability to deliver effective restorative therapy.
Key words fMRI, Stroke, Blood oxygen level-dependent, Motor cortex, Premotor cortex, Plasticity,
Rehabilitation
1 Introduction
Studying patients who have suffered from stroke with functional brain
imaging is difficult for a variety of reasons. The motivation behind
such studies is a desire to understand and subsequently improve the
process of functional recovery. Stroke and other forms of neurological
damage account for nearly half of all severely disabled adults [1–3].
Longitudinal studies of recovery suggest that only 50 % of stroke sur-
vivors with significant initial upper limb paresis recover useful function
of the limb [4]. Furthermore, those with poor recovery of arm func-
tion have dramatically impaired quality of life and sense of well-being
[5, 6]. It is clear that effective treatment of motor impairment after
stroke is critically important to many people.
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_21, © Springer Science+Business Media New York 2016
639
640 Nick S. Ward
The results from any functional imaging study are only as reliable
as the care with which the experiment is constructed and executed,
but studies involving patients who have had a stroke raise some
specific issues. The selection of patients, choice of experimental
paradigm, within scanner monitoring of performance, and the
approaches to data analysis all require careful consideration.
3.1 Subject Selection In general, stroke patients are a heterogeneous group differing in
several important ways, not least the site and size of infarct, patency
of the vascular system, age, comorbidities, and concurrent medica-
tion. The criteria for patient selection will to an extent depend on
the experimental question. It is unlikely that averaging the results
from a wide variety of patient types will prove useful because of this
variability, but there are two other ways of approaching the experi-
mental design. First, it may be desirable to use a group of patients
highly selected on the basis of lesion location, for example. Results
from this type of controlled study are powerful, but do not gener-
alize outside the subgroup selected. Alternatively, it might be more
useful to study a group of patients who vary in a specific factor of
interest (e.g., outcome), to explore the relationship between this
factor and task-related brain activity. Results from studies using
this approach can be generalized more easily.
3.2 Performance The choice of experimental task is critical and is dependent on the
Confounds experimental question. For example, a study of the relationship
between brain activation and outcome after stroke will by necessity
involve patients with different performance abilities. Similarly, a lon-
gitudinal study will require that patients are studied at different stages
of recovery. For an active motor or language task this can result in the
problem of performance confounds, because the ability to perform
the task is not the same across patients or sessions. A change in exper-
imental task performance can have significant effects upon the pat-
tern of brain activation. In other words, comparison across patients
or time points is made difficult if the patients are performing the task
differently. Thus, each patient must perform the same task during the
fMRI experiment, so that a meaningful comparison can be made
across subjects or scanning sessions. Maintaining a consistent task is
therefore of great importance, but in stroke recovery studies equality
of task may be interpreted in a number of ways. In particular, a task
may be consistent across patients with different abilities in terms of
absolute or relative parameters.
fMRI in Cerebrovascular Disorders 645
3.3 Task Frequency The rate of task performance is also something that has implications
for both data analysis and interpretation of results. Consider once
again a simple motor task such as finger tapping or hand squeezing.
The rate of performance of a repetitive task will influence how
effortful the task is, in the same way as the target force. Most experi-
ments are conducted in a “block design”; that is to say a period of
activity (usually for 16–30 s) followed by a period of rest. If subjects
are asked to perform at the same rate (even if the target force is
scaled according to each subject’s own performance abilities), for
example, finger tapping at 1 Hz for 20 s, then differences across
subjects/sessions could be due to differences in perceived effort,
just as with equal absolute target forces. Some investigators have
varied the rate at which subjects are asked to perform a task to try
to control for effort exerted. However, comparing blocks with dif-
ferent numbers of “events” within them is problematic because the
BOLD signal summates depending on how many events there are.
The BOLD response needs approximately 10 s and longer to return
to baseline, but “events” are usually more frequent than this (e.g.,
1 Hz finger tapping). It is usually assumed that there is a summa-
tion of the overlapping BOLD responses, which is largely (but not
entirely) linear [35, 36]. The basis function (boxcar design) in the
general linear model will have the same “height,” and so more fre-
quent events will result in a larger parameter estimate, for the same
amount of event-related activity. In fact, it is the quantity of events,
646 Nick S. Ward
3.4 Task Complexity Investigators are often tempted to use more complex tasks when
studying patients with impairment in the hope that this will maxi-
mize differences between patients and control subjects. This is
sometimes done in the hope of exploring a more ecologically valid
task, i.e., one which is relevant to function in the real world.
However, it is never possible to study the neural correlates of a task
that a subject cannot themselves perform. By introducing more
complexity into the task, patients with significant impairment are
more likely to adopt new operational strategies toward these
experimental tasks in an attempt to adapt to their impairment.
These differences in strategy could therefore account for differ-
ences between subject groups. Although of clinical interest, differ-
ences in strategy across a group represent a potential experimental
confound if they are unexpected and not measured. One approach
is therefore to use a simple task that minimizes difference in strate-
gic approach to the task so that valid comparisons can then be
made across subjects/sessions.
3.5 Task Monitoring Once a paradigm has been selected it is important that task perfor-
mance is monitored during the experiment. Intersubject variability
may be greater after stroke and new sources of variability can arise,
such as mirror or associated movements. To take account of this,
some investigators record behavior during a prescan rehearsal,
whilst others incorporate the increasingly available instrumentation
that is compatible with the MRI setting. Prescan rehearsal provides
some idea of whether a task can be performed correctly, or whether
mirror movements are present, for example. However, in-scanner
recordings allow this information to be incorporated into image
analysis as a covariate, and thus improving statistical power by
accounting for correlated variance in the measured scan signal.
The experimental approach is therefore dictated by the experi-
mental question. Not all investigators will have the same question,
fMRI in Cerebrovascular Disorders 647
but the issues discussed earlier need to be considered in all cases. For
most questions, this approach is entirely appropriate and standard-
ization of experimental paradigms, patient selection, and method of
analysis across experiments is not required. In the case of experimen-
tal questions that require a multicenter approach that is technically
feasible, standardization of such factors would be required.
4.1 Residual Early studies of motor system organization after stroke compared
Functional brain activation during movement in well-recovered patients and
Architecture After normal controls. Early group studies of stroke patients with sub-
Stroke: The Story cortical lesions described greater activation within a number of
So Far? motor-related cortical regions compared with controls during a
finger tapping task [38–42]. It was suggested that nonprimary (or
secondary) cortical motor regions were thus responsible for recov-
ery of motor function in these patients. Strick [43] had proposed
this as a potential mechanism of restoration of function, some years
before based on an understanding of the organization of the corti-
cal motor system in primates. Normal distal motor function is facil-
itated largely through the corticospinal pathway, from the cortical
motor system to the spinal cord motor neurons. The majority of
corticospinal fibers originate in the M1, but there are contribu-
tions from other cortical regions [44]. In primates, the M1, arcu-
ate (or lateral) premotor cortex (PM), and supplementary motor
area (SMA) are each part of parallel, independent motor networks
with (1) separate projections to spinal cord motoneuron and (2)
interactions at the level of the cortex [43]. There is some similarity
between the corticospinal projections from the hand regions of
M1, PM, and SMA. Thus, it seemed feasible that a number of
motor networks acting in parallel could generate an output to the
spinal cord necessary for movement, and that damage in one of
these networks could be at least partially compensated for by activ-
ity in another [45, 46]. Subsequently, many studies have demon-
strated that the performance of a simple motor task with the
affected limb is associated with greater bilateral brain activation in
a number of cortical motor-related areas compared with healthy
volunteers, including dorsal PM (PMd) and ventral PM (PMv),
SMA, and cingulate motor areas (CMA) [23, 38–42, 47–52].
A critical question is whether these differences are related to
recovery. As discussed previously in this chapter, this question
requires that the group of patients examined have a wide variety of
outcomes, or else longitudinal studies should be performed. In the
first such cross-sectional study, a group of chronic stroke patients
with infarcts sparing M1 were scanned during a hand grip with
visual feedback task using fMRI [30]. The target forces used were
always a proportion of each subject’s own maximum grip force, so
648 Nick S. Ward
Fig. 1 Brain regions in which there is a negative correlation between corticospinal system integrity (as assessed
with transcranial magnetic stimulation) and task-related signal change during hand grip with the affected
hand. Increasing task-related activity is seen in a number of secondary motor areas including premotor regions
and supplementary motor area as damage to the corticospinal system increases. The affected hand was on
the left side. Results are displayed on a “glass brain” shown from the right side (top left image), from behind
(top right image), and from above (bottom left image). Voxels are significant at P < 0.001 (uncorrected), and
clusters are significant at P < 0.05 (corrected) (Reproduced from ref. [31], Oxford University Press.)
ipsilesional M1
Fig. 2 Brain regions in which the blood oxygen level-dependent (BOLD) signal varies linearly with force exerted
during hand grip change as a function of corticospinal tract (CST) integrity (as assessed with transcranial
magnetic stimulation). The affected hand was on the left side. In the left panel, increasing force leads to
greater modulation of BOLD signal in ipsilesional M1 in patients with less damage to CST. In the right panel,
increasing force leads to greater modulation of BOLD signal in contralesional dorsal premotor cortex, contral-
esional cerebellum, contralesional, and ipsilesional ventral premotor cortex. This demonstrates that brain
regions involved in force modulation shift away from primary motor cortex to premotor regions with increasing
CST damage. Results are overlaid onto the average T1-weighted structural scan obtained from all stroke
patients in the study (Adapted from ref. [32], Blackwell Publishing.)
5 Conclusions
References
1. Hoffman C, Rice D, Sung HY (1996) Persons being one year after stroke. Clin Rehabil
with chronic conditions. Their prevalence and 11(2):139–145
costs. JAMA 276(18):1473–1479 7. Stroke Unit Trialists’ Collaboration (2000)
2. Office of Population Censuses and Surveys Organised inpatient (stroke unit) care for
(1988) OPCS surveys of disability in Great stroke (Cochrane Review). The Cochrane
Britain. I. The prevalence of disability among Library, Issue 2. Oxford: Update Software
adults. HMSO, London 8. Ward NS, Cohen LG (2004) Mechanisms
3. Wade DT, Hewer RL (1987) Epidemiology of underlying recovery of motor function after
some neurological diseases with special refer- stroke. Arch Neurol 61(12):1844–1848
ence to work load on the NHS. Int Rehabil 9. The Academy of Medical Sciences (2004)
Med 8(3):129–137 Restoring neurological function: putting the
4. Wade DT (1989) Measuring arm impairment neurosciences to work in neurorehabilitation.
and disability after stroke. Int Disabil Stud Academy of Medical Sciences, London
11(2):89–92 10. Loubinoux I, Carel C, Pariente J, Dechaumont
5. Nichols-Larsen DS, Clark PC, Zeringue A, S, Albucher JF, Marque P et al (2003)
Greenspan A, Blanton S (2005) Factors influ- Correlation between cerebral reorganization
encing stroke survivors’ quality of life during and motor recovery after subcortical infarcts.
subacute recovery. Stroke 36(7):1480–1484 Neuroimage 20(4):2166–2180
6. Wyller TB, Sveen U, Sodring KM, Pettersen 11. Tombari D, Loubinoux I, Pariente J, Gerdelat A,
AM, Bautz-Holter E (1997) Subjective well- Albucher JF, Tardy J et al (2004) A longitudinal
fMRI in Cerebrovascular Disorders 653
fMRI study: in recovering and then in clinically disease on the BOLD signal in
stable sub-cortical stroke patients. Neuroimage fMRI. Neuroimage 20(2):1393–1399
23(3):827–839 26. Rossini PM, Altamura C, Ferretti A, Vernieri F,
12. Ward NS, Brown MM, Thompson AJ, Zappasodi F, Caulo M et al (2004) Does cere-
Frackowiak RS (2006) Longitudinal changes in brovascular disease affect the coupling between
cerebral response to proprioceptive input in neuronal activity and local haemodynamics?
individual patients after stroke: an FMRI study. Brain 127(Pt 1):99–110
Neurorehabil Neural Repair 20(3):398–405 27. Rother J, Knab R, Hamzei F, Fiehler J,
13. Lee A, Kannan V, Hillis AE (2006) The contri- Reichenbach JR, Buchel C et al (2002)
bution of neuroimaging to the study of language Negative dip in BOLD fMRI is caused by
and aphasia. Neuropsychol Rev 16(4):171–183 blood flow-oxygen consumption uncoupling
14. Price CJ, Crinion J (2005) The latest on func- in humans. Neuroimage 15(1):98–102
tional imaging studies of aphasic stroke. Curr 28. Krainik A, Hund-Georgiadis M, Zysset S, von
Opin Neurol 18(4):429–434 Cramon DY (2005) Regional impairment of
15. Wise RJ (2003) Language systems in normal cerebrovascular reactivity and BOLD signal in
and aphasic human subjects: functional imag- adults after stroke. Stroke 36(6):1146–1152
ing studies and inferences from animal studies. 29. Murata Y, Sakatani K, Hoshino T, Fujiwara N,
Br Med Bull 65:95–119 Kano T, Nakamura S et al (2006) Effects of
16. Buxton RB (2002) An introduction to func- cerebral ischemia on evoked cerebral blood
tional magnetic resonance imaging: principles oxygenation responses and BOLD contrast
and techniques. Cambridge University Press, functional MRI in stroke patients. Stroke
Cambridge 37(10):2514–2520
17. Magistretti PJ, Pellerin L (1999) Cellular 30. Ward NS, Brown MM, Thompson AJ,
mechanisms of brain energy metabolism and Frackowiak RS (2003) Neural correlates of
their relevance to functional brain imaging. outcome after stroke: a cross-sectional fMRI
Philos Trans R Soc Lond B Biol Sci study. Brain 126(Pt 6):1430–1448
354(1387):1155–1163 31. Ward NS, Newton JM, Swayne OB, Lee L,
18. Magistretti PJ, Pellerin L, Rothman DL, Thompson AJ, Greenwood RJ et al (2006)
Shulman RG (1999) Energy on demand. Motor system activation after subcortical stroke
Science 283(5401):496–497 depends on corticospinal system integrity.
19. Iadecola C (2004) Neurovascular regulation in Brain 129(Pt 3):809–819
the normal brain and in Alzheimer’s disease. 32. Ward NS, Newton JM, Swayne OB, Lee L,
Nat Rev Neurosci 5(5):347–360 Frackowiak RS, Thompson AJ et al (2007) The
20. Attwell D, Iadecola C (2002) The neural basis relationship between brain activity and peak
of functional brain imaging signals. Trends grip force is modulated by corticospinal system
Neurosci 25(12):621–625 integrity after subcortical stroke. Eur J Neurosci
25(6):1865–1873
21. Friston KJ, Josephs O, Rees G, Turner R
(1998) Nonlinear event-related responses in 33. D’Esposito M, Zarahn E, Aguirre GK, Rypma
fMRI. Magn Reson Med 39(1):41–52 B (1999) The effect of normal aging on the
coupling of neural activity to the bold hemody-
22. Newton J, Sunderland A, Butterworth SE, namic response. Neuroimage 10(1):6–14
Peters AM, Peck KK, Gowland PA (2002) A
pilot study of event-related functional mag- 34. Ward NS, Swayne OB, Newton JM (2008)
netic resonance imaging of monitored wrist Age-dependent changes in the neural corre-
movements in patients with partial recovery. lates of force modulation: an fMRI study.
Stroke 33(12):2881–2887 Neurobiol Aging 29(9):1434–1446
23. Pineiro R, Pendlebury S, Johansen-Berg H, 35. Pollmann S, Dove A, Yves von Cramon D,
Matthews PM (2001) Functional MRI detects Wiggins CJ (2000) Event-related fMRI: com-
posterior shifts in primary sensorimotor cortex parison of conditions with varying BOLD
activation after stroke: evidence of local adap- overlap. Hum Brain Mapp 9(1):26–37
tive reorganization? Stroke 32(5):1134–1139 36. Wager TD, Vazquez A, Hernandez L, Noll DC
24. Carusone LM, Srinivasan J, Gitelman DR, (2005) Accounting for nonlinear BOLD
Mesulam MM, Parrish TB (2002) Hemodynamic effects in fMRI: parameter estimates and a
response changes in cerebrovascular disease: model for prediction in rapid event-related
implications for functional MR imaging. AJNR studies. Neuroimage 25(1):206–218
Am J Neuroradiol 23(7):1222–1228 37. Kim JA, Eliassen JC, Sanes JN (2005) Movement
25. Hamzei F, Knab R, Weiller C, Rother J (2003) quantity and frequency coding in human motor
The influence of extra- and intracranial artery areas. J Neurophysiol 94(4):2504–2511
654 Nick S. Ward
38. Cao Y, D’Olhaberriague L, Vikingstad EM, ment and focusing of brain activation. Stroke
Levine SR, Welch KM (1998) Pilot study of 33(6):1610–1617
functional MRI to assess cerebral activation of 51. Johansen-Berg H, Rushworth MF, Bogdanovic
motor function after poststroke hemiparesis. MD, Kischka U, Wimalaratna S, Matthews PM
Stroke 29(1):112–122 (2002) The role of ipsilateral premotor cortex
39. Chollet F, DiPiero V, Wise RJ, Brooks DJ, in hand movement after stroke. Proc Natl Acad
Dolan RJ, Frackowiak RS (1991) The func- Sci U S A 99(22):14518–14523
tional anatomy of motor recovery after stroke 52. Seitz RJ, Hoflich P, Binkofski F, Tellmann
in humans: a study with positron emission L, Herzog H, Freund HJ (1998) Role of
tomography. Ann Neurol 29(1):63–71 the premotor cortex in recovery from mid-
40. Cramer SC, Nelles G, Benson RR, Kaplan JD, dle cerebral artery infarction. Arch Neurol
Parker RA, Kwong KK et al (1997) A func- 55(8):1081–1088
tional MRI study of subjects recovered from 53. Ward NS, Brown MM, Thompson AJ,
hemiparetic stroke. Stroke 28(12):2518–2527 Frackowiak RS (2004) The influence of time
41. Weiller C, Chollet F, Friston KJ, Wise RJ, after stroke on brain activations during a motor
Frackowiak RS (1992) Functional reorga- task. Ann Neurol 55(6):829–834
nization of the brain in recovery from stria- 54. Dancause N, Barbay S, Frost SB, Plautz EJ,
tocapsular infarction in man. Ann Neurol Stowe AM, Friel KM et al (2006) Ipsilateral
31(5):463–472 connections of the ventral premotor cor-
42. Weiller C, Ramsay SC, Wise RJ, Friston KJ, tex in a new world primate. J Comp Neurol
Frackowiak RS (1993) Individual patterns 495(4):374–390
of functional reorganization in the human 55. Dancause N, Barbay S, Frost SB, Mahnken JD,
cerebral cortex after capsular infarction. Ann Nudo RJ (2007) Interhemispheric connections
Neurol 33(2):181–189 of the ventral premotor cortex in a new world
43. Strick PL (1988) Anatomical organization of primate. J Comp Neurol 505(6):701–715
multiple motor areas in the frontal lobe: impli- 56. Dum RP, Strick PL (1991) The origin of cor-
cations for recovery of function. Adv Neurol ticospinal projections from the premotor areas
47:293–312 in the frontal lobe. J Neurosci 11(3):667–689
44. Porter R, Lemon RN (1993) Corticospinal 57. He SQ, Dum RP, Strick PL (1993) Topographic
function and voluntary movement. Oxford organization of corticospinal projections
University Press, Oxford, UK from the frontal lobe: motor areas on the
45. Dum RP, Strick PL (1996) Spinal cord termina- lateral surface of the hemisphere. J Neurosci
tions of the medial wall motor areas in macaque 13(3):952–980
monkeys. J Neurosci 16(20):6513–6525 58. He SQ, Dum RP, Strick PL (1995) Topographic
46. Rouiller EM, Moret V, Tanne J, Boussaoud organization of corticospinal projections from
D (1996) Evidence for direct connections the frontal lobe: motor areas on the medial
between the hand region of the supplemen- surface of the hemisphere. J Neurosci 15(5 Pt
tary motor area and cervical motoneurons 1):3284–3306
in the macaque monkey. Eur J Neurosci 59. Boudrias MH, Belhaj-Saif A, Park MC, Cheney
8(5):1055–1059 PD (2006) Contrasting properties of motor
47. Calautti C, Leroy F, Guincestre JY, Baron JC output from the supplementary motor area
(2001) Dynamics of motor network overac- and primary motor cortex in rhesus macaques.
tivation after striatocapsular stroke: a longi- Cereb Cortex 16(5):632–638
tudinal PET study using a fixed-performance 60. Maier MA, Armand J, Kirkwood PA, Yang
paradigm. Stroke 32(11):2534–2542 HW, Davis JN, Lemon RN (2002) Differences
48. Calautti C, Leroy F, Guincestre JY, Baron JC in the corticospinal projection from primary
(2003) Displacement of primary sensorimotor motor cortex and supplementary motor area
cortex activation after subcortical stroke: a lon- to macaque upper limb motoneurons: an ana-
gitudinal PET study with clinical correlation. tomical and electrophysiological study. Cereb
Neuroimage 19(4):1650–1654 Cortex 12(3):281–296
49. Cramer SC, Shah R, Juranek J, Crafton KR, 61. Baker SN, Zaaimi B, Fisher KM, Edgley SA,
Le V (2006) Activity in the peri-infarct rim Soteropoulos DS (2015) Pathways medi-
in relation to recovery from stroke. Stroke ating functional recovery. Prog Brain Res
37(1):111–115 218:389–412
50. Feydy A, Carlier R, Roby-Brami A, Bussel B, 62. Dettmers C, Fink GR, Lemon RN, Stephan
Cazalis F, Pierot L et al (2002) Longitudinal KM, Passingham RE, Silbersweig D et al
study of motor recovery after stroke: recruit- (1995) Relation between cerebral activity and
fMRI in Cerebrovascular Disorders 655
force in the motor areas of the human brain. nectivity after subcortical stroke assessed with
J Neurophysiol 74(2):802–815 functional magnetic resonance imaging. Ann
63. Thickbroom GW, Phillips BA, Morris I, Byrnes Neurol 63(2):236–246
ML, Sacco P, Mastaglia FL (1999) Differences 71. Bestmann S, Swayne O, Blankenburg F, Ruff
in functional magnetic resonance imaging of CC, Teo J, Weiskopf N et al (2010) The role of
sensorimotor cortex during static and dynamic contralesional dorsal premotor cortex after
finger flexion. Exp Brain Res 126(3):431–438 stroke as studied with concurrent TMS-fMRI. J
64. Ward NS, Frackowiak RS (2003) Age-related Neurosci 30(36):11926–11937
changes in the neural correlates of motor per- 72. Zemke AC, Heagerty PJ, Lee C, Cramer SC
formance. Brain 126(Pt 4):873–888 (2003) Motor cortex organization after stroke
65. Fridman EA, Hanakawa T, Chung M, is related to side of stroke and level of recovery.
Hummel F, Leiguarda RC, Cohen LG (2004) Stroke 34(5):e23–e28
Reorganization of the human ipsilesional 73. Crafton KR, Mark AN, Cramer SC (2003)
premotor cortex after stroke. Brain 127(Pt Improved understanding of cortical injury by
4):747–758 incorporating measures of functional anatomy.
66. Lotze M, Markert J, Sauseng P, Hoppe J, Brain 126(Pt 7):1650–1659
Plewnia C, Gerloff C (2006) The role of mul- 74. Marshall RS, Perera GM, Lazar RM, Krakauer
tiple contralesional motor areas for complex JW, Constantine RC, DeLaPaz RL (2000)
hand movements after internal capsular lesion. Evolution of cortical activation during recov-
J Neurosci 26(22):6096–6102 ery from corticospinal tract infarction. Stroke
67. Grefkes C, Fink GR (2014) Connectivity-based 31(3):656–661
approaches in stroke and recovery of function. 75. Small SL, Hlustik P, Noll DC, Genovese C,
Lancet Neurol 13(2):206–216 Solodkin A (2002) Cerebellar hemispheric
68. Carter AR, Patel KR, Astafiev SV, Snyder AZ, activation ipsilateral to the paretic hand corre-
Rengachary J, Strube MJ et al (2012) Upstream lates with functional recovery after stroke.
dysfunction of somatomotor functional con- Brain 125(Pt 7):1544–1557
nectivity after corticospinal damage in stroke. 76. Ward NS, Brown MM, Thompson AJ,
Neurorehabil Neural Repair 26(1):7–19 Frackowiak RS (2003) Neural correlates of
69. Carter AR, Astafiev SV, Lang CE, Connor LT, motor recovery after stroke: a longitudinal
Rengachary J, Strube MJ et al (2010) Resting fMRI study. Brain 126(Pt 11):2476–2496
interhemispheric functional magnetic reso- 77. Kleim JA, Chan S, Pringle E, Schallert K,
nance imaging connectivity predicts perfor- Procaccio V, Jimenez R et al (2006) BDNF val-
mance after stroke. Ann Neurol 67(3):365–375 66met polymorphism is associated with modi-
70. Grefkes C, Nowak DA, Eickhoff SB, Dafotakis fied experience-dependent plasticity in human
M, Küst J, Karbe H et al (2008) Cortical con- motor cortex. Nat Neurosci 9(6):735–737
Chapter 22
Abstract
Functional neuroimaging has become an important tool for clinical research, with the potentiality to
provide information on psychiatric disease pathology and treatment response. We review functional mag-
netic resonance imaging (fMRI) research findings for five psychiatric disorders: schizophrenia, major
depressive disorder, bipolar disorder, obsessive-compulsive disorder, and posttraumatic stress disorder.
Brain functional abnormalities and possible underlying mechanisms for disease symptoms are discussed,
with a focus on future clinical implications for fMRI in psychiatric disease.
Key words fMRI, Blood oxygen level dependent, Psychiatric disorders, Schizophrenia, Major depres-
sive disorder, Bipolar disorder, Obsessive-compulsive disorder, Posttraumatic stress disorder
1 Introduction
1.1 Overview of fMRI Functional magnetic resonance imaging (fMRI) is a unique, noninva-
sive method of measuring neural activation through changes in oxida-
tion and regional blood flow. An important clinical research tool that
has been used more and more frequently in recent years, fMRI is able
to indirectly detect brain activity in the working brain, allowing for the
assessment of psychiatric disease physiology and treatment effects.
fMRI does not involve exposure to radioactive tracers, thus allowing
patients and subjects to undergo multiple scans over a short period of
time, if necessary. Most fMRI studies involve the measurement of sig-
nal arising from hydrogen nuclei [1, 2]. Common types of fMRI used
in psychiatric neuroimaging include blood oxygen level dependent
(BOLD) and arterial spin labeling (ASL).
Instead of incorporating a radioactive tracer as in positron
emission tomography (PET) or single photon emission computed
tomography (SPECT), fMRI makes use of the unique properties
of hemoglobin (BOLD and BOLD contrast methods) or the water
molecules of flowing blood (ASL) to produce images of neural
activation. Most fMRI studies today are BOLD studies that make
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_22, © Springer Science+Business Media New York 2016
657
658 Erin L. Habecker et al.
1.3 ASL ASL differs from BOLD in that it depends on T1 mechanisms and
the magnetic labeling of water molecules to generate images. Water
molecules in flowing blood are tagged through the saturation or
inversion of the longitudinal component of the MR signal [12];
these molecules then diffuse from capillaries into brain tissue where
fMRI in Psychiatric Disorders 659
they alter the magnetization of the local tissue [1]. As blood flow
into the imaging slice increases, there is a more significant differ-
ence between the magnetized condition and the control condition,
during which the magnetization of arterial blood is fully relaxed
[1]. Control and tagged images are then taken, and the difference
between them is proportional to the CBF. ASL can be used to
measure global CBF changes dynamically; its use of water mole-
cules as an endogenous blood flow tracer means that the images
generated by ASL are not susceptible to neurovascular changes
that are not related to neuronal activation. ASL has another advan-
tage over BOLD in that effects of frequency drifts tend to be mini-
mized in ASL, making this method more suitable for longer
duration scans [12]. However, BOLD acquisitions tend to have
greater temporal resolution, greater maximum number of slices,
and appear to be more sensitive to parametric manipulations of
task demands [1, 13, 14]. BOLD maps also usually have larger
activation areas than ASL maps [15, 16], which could either be
due to the decreased sensitivity or improved signal localization
inherent in ASL [13]. There are three classes of ASL methods:
pulsed ASL, continuous ASL, and velocity selective ASL [1]. A
discussion of the relative methodologies and merits of the three
techniques is beyond the scope of this chapter.
2 fMRI in Psychiatry
2.1 Clinical A wide range of neuropsychiatric disorders have now been investi-
Disorders gated using fMRI techniques and protocols. This review will
explore the paradigms employed, imaging results, and future
research opportunities in five mental disorders: schizophrenia,
major depression, bipolar disorder (BD), obsessive-compulsive dis-
order (OCD), and posttraumatic stress disorder (PTSD). These
particular disorders were selected due to the fact that they repre-
sent a subset of psychotic, mood, and anxiety disorders; have a
significant prevalence in the general population (0.4–14 % depend-
ing on age and gender of the sample); are popular candidates for
fMRI research; and have each been the subject of research on diag-
nosis, disease progression, and treatment using imaging. Table 1
summarizes the nature, range, and prevalence of the selected disor-
ders in the general population.
Modern imaging techniques have been crucial to the delinea-
tion of the brain structures and functions that are negatively
impacted in psychiatric disorders such as the ones reviewed here.
Traditionally, such disorders have been characterized primarily via
clinical psychiatric evaluation of abnormal symptoms, and treat-
ments consist of a trial-and-error strategy combined with patient
self-selection of treatment options or option combinations [17].
The use of fMRI to evaluate the underlying cognitive disturbances
Table 1
660
Schizophrenia Psychotic Delusions, hallucinations (often Subtypes are defined by the 0.5–1.5 % among Complete remission Antipsychotic
disorder auditory), disorganized speech, predominant symptom at adults uncommon. Some medications,
grossly disorganized or catatonic the time of evaluation individuals display including
behavior, negative symptoms (paranoid, disorganized, exacerbations and clozapine,
(affective flattening, alogia, catatonic, undifferentiated, remissions of risperidone,
aviolotion) residual). However, symptoms while olanzapine,
Erin L. Habecker et al.
3 Psychotic Disorders
Schizophrenia
Manoach 9 schizophrenic subjects (8 under stable dose of Working memory: Sternberg Schizophrenic patients exhibited deficiencies in working
et al. [20] antipsychotic medication, 1 unmedicated) and 9 Item Recognition memory, despite similar activity in the DLPFC compared to
healthy controls Paradigm adapted to healthy controls. Compared with controls, during the task,
include monetary reward schizophrenic patients showed activation in the basal ganglia
for correct responses and thalamus
Rubia et al. 6 male patients with schizophrenia (all under Inhibitory control: “stop” Schizophrenic patients exhibited reduced left prefrontal
[21] atypical antipsychotic medication) and 7 and “go/no-go” tasks activation compared to healthy controls
matched healthy controls
Hempel 10 partially remitted schizophrenic patients (all Facial affect discrimination Compared with controls, schizophrenic patients showed a
et al. [22] under atypical antipsychotic medication) and 10 and labeling significantly decreased activation in the ACC during the
healthy controls discrimination task, and decreased activity in the amygdala-
hippocampal complex bilaterally during labeling
Hofer et al. 10 male outpatients with schizophrenia (all under Episodic encoding/ Compared with controls, patients with schizophrenia
[23] atypical antipsychotic medication) and 10 male recognition of words demonstrated similar cognitive performance in word
healthy controls recognition, but decreased activation in the bilateral DLPFC
and lateral temporal cortices during both tasks
Kubicki 9 male chronic schizophrenic patients (unspecified Semantic encoding Schizophrenic patients had decreased activation of the left
et al. [24] medication) and 9 control subjects inferior prefrontal cortex and increased activation of the left
superior temporal gyrus compared to healthy controls
Habel et al. 13 male patients with schizophrenia (8 under Positive and negative mood Schizophrenic patients and brothers of schizophrenic patients
[25] typical and 4 under atypical antipsychotic evocation showed reduced amygdala activity compared to healthy
medication, and 1 unmedicated), 13 of their non controls
affected brothers (asymptomatic, unmedicated),
and 26 unrelated matched healthy controls
fMRI in Psychiatric Disorders
Ragland 14 patients with schizophrenia (2 under typical and 9 Word encoding/recognition Compared with controls, schizophrenic patients demonstrated
et al. [26] under atypical antipsychotic medication, and 3 reduced activation of the prefrontal cortex and increased
under both typical/atypical medication) and 15 activation of the parahippocampal gyri during encoding
healthy controls
663
(continued)
Table 2
664
(continued)
Takahashi 15 schizophrenic patients (11 under atypical Affective pictures Compared with controls, patients with schizophrenia had
et al. [27] antipsychotic medication; 4 unmedicated) and decreased activity in the right amygdala and medial
15 healthy volunteers prefrontal cortex
Williams 27 schizophrenic patients and 22 healthy controls Facial expressions of fear Compared with controls, schizophrenic patients had enhanced
et al. [28] arousal responses coupled with reduction in amygdala and
medial prefrontal activity
Erin L. Habecker et al.
Honey et al. 12 healthy volunteers were administered with Episodic memory task Subjects on ketamine showed left frontal activation during
[29] 100 ng/ml plasma ketamine or placebo semantic processing at encoding and in these subjects
successful encoding was supplemented by additional
nonverbal processing
Morey et al. 52 subjects: 10 ultra-high risk for schizophrenia, Visual oddball task Compared with controls, early and chronic schizophrenic
[30] 15 with early schizophrenia, 11 with chronic patients showed lower activation associated with the target
schizophrenia and 16 healthy controls stimuli in the ACC, inferior frontal gyrus, and the medial
frontal gyrus. Although the ultra-high-risk group did not
reach the significance when compared with controls, they
showed a trend toward the early group
Yurgelun- 12 schizophrenic patients under 8 weeks of Word fluency task Compared with patients receiving placebo, patients receiving
Todd d-cycloserine treatment or placebo d-cycloserine showed a significant increased activation,
et al. [31] associated with a reduction in negative symptoms, in the
temporal lobe
Juckel et al. 20 schizophrenic patients (10 under typical and 10 Incentive monetary delay task Schizophrenic patients showed reduced ventral striatal
[32] under atypical antipsychotic medication) and 10 activation during the presentation of reward-indicating cues
age-matched male healthy controls as compared to healthy controls. Decreased activation of the
left ventral striatum was inversely correlated with the severity
of symptoms
Vink et al. 21 schizophrenic patients (all under stable atypical Inhibitory control: stop cues Compared with controls, schizophrenic patients and
[33] antipsychotic medication), 15 unaffected unaffected siblings did not activate the striatum when
siblings, and 36 matched healthy controls responding to motor cues
ACC anterior cingulate cortex, DLPFC dorsolateral prefrontal cortex
fMRI in Psychiatric Disorders 665
3.1.1 fMRI Features In the last decades, several studies have attempted to define the
of Subjects at Risk prodromal traits predictive of future conversion to schizophrenia in
to Develop Schizophrenia at-risk subjects, especially after demonstration that early interven-
tion improves disease prognosis [37]. The eventual “risk” to
develop schizophrenia is defined in terms of genetic aspects, such as
a positive family history for schizophrenia (e.g., individuals who
have an affected first-degree family member-FHR), or in terms of
clinical (prodromal) features. Among individuals with a clinical
high risk (CHR) to develop schizophrenia, subjects ‘at risk mental
state’ (ARMS) [38] are those having clinically defined sub-psychotic
symptoms and subjects at ultra-high risk (UHR) are defined by the
presence of prevalent positive clinical symptoms [38, 39].
Although altered activation of frontal and prefrontal cortices
has been amply reported in people with increased risk of psychosis,
at present it is still not clear if this neurofunctional alteration
increases in line with the level of psychosis risk. On this purpose, a
study observed a relationship between the level of working mem-
ory task-related deactivation in the mPFC and precuneus and the
level of psychosis risk, with deactivation weakest in the UHR
group, at an intermediate level in the FHR group, and greatest in
healthy controls [40]. In another study, while controls showed a
negative association between age and frontal functional activation
during verbal working memory, clinical high risk youth who con-
verted to psychosis showed the opposite [41], likely reflecting an
emerging hyperactivity in frontal regions for compensative pur-
poses [41]. During an executive task, a significant reduction in the
topological centrality of the ACC in ARMS subjects which later
converted into a psychotic disorder suggested this as a potential
biomarker for the transition to psychosis [42].
In subjects at risk, functional alterations at the frontal regions
are likely to subtend not only their cognitive but also their emo-
tional features. In FHR subjects, it has been reported a reduced
coupling between amygdala and prefrontal cortex during facial
expression processing [43]. Adolescent FHR subjects have shown
reduced ACC activation during emotional processing [44] and
specific hyperactivation in the right superior frontal gyrus and right
precentral gyrus during fearful face presentations [45]. Patients
fMRI in Psychiatric Disorders 667
4 Mood Disorders
Fu et al. [19] 19 medication free, acutely depressed Facial expressions of sadness: Compared with controls, depressed subjects had decreased activation in
subjects and 19 matched healthy low, medium, and high regions of the left brain: hippocampus and parahippocampal gyrus,
volunteers intensity. Subjects asked to amygdala, insula, caudate nucleus, thalamus, dorsal cingulate gyrus,
identify sex of face inferior parietal cortex. They also showed negative relationship
between increased differential response to variable affective intensity
and activation in the rostral prefrontal cortex
Surguladze 16 individuals with major depressive Facial expressions: happy and sad Healthy individuals displayed increased activation in bilateral fusiform
et al. [60] disorder and 14 healthy controls gyri and right putamen in response to increasing happiness. Depressed
subjects demonstrated activation in left putamen, left
parahippocampal gyrus/amygdala, and right fusiform gyrus in
response to increasing sadness
Del-Ben 12 healthy male subjects, under 7.5 mg IV Go/No-go, Loss/No-loss, Citalopram enhanced activations in the right BA47 during Go/No-go
et al. [61] citalopram or placebo, single blind covert (averse) face emotion task but attenuated BA47 response to aversive faces. Citalopram
crossover design recognition attenuated the right amygdala response to aversive faces and the
BA11 response during Loss/No-loss task
Wagner et al. 16 patients with unipolar depression (free Adapted version of Stroop task No differences in reaction time and accuracy between groups.
[62] of psychotropic medication for a week at Compared with controls, patients showed increased activation in the
the time of the fMRI); 16 matched rostral ACC and left DLPFC
healthy controls
Siegle et al. 27 unmedicated, unipolar depressive Executive control task: digit Compared with controls, depressed subjects displayed sustained
[59] subjects (free of antidepressant sorting; emotional information amygdala reactivity on emotional tasks and decreased DLPFC activity
medication for 2 weeks before testing) processing: personal relevance on the digit-sorting task
and 25 never-depressed healthy controls rating of words
Vollm et al. 45 healthy male subjects under Go/No-go, Reward/No-reward, The Mirtazapine treatment was associated with enhanced activation of the
[63] Mirtazapine or placebo, double blind, Loss/No-loss right orbitofrontal cortex during Go/No-go and Reward/No-reward
placebo controlled and of the bilateral parietal cortex during Reward/No-reward
Walter et al. 12 partially remitted, medicated inpatients Working memory: delayed match Compared with controls, depressed patients were slower and less
[64] with major depressive disorder; 17 to sample accurate in task. They showed increased activation in the left DLPFC
healthy control during highest cognitive load and in the VMPFC during control
condition
Bipolar disorder
Yurgelun- 14 BD subjects (12 under mood stabilizers; Faces: happy and fearful affect Compared with controls, BD subjects had reduced DLPFC activation
Todd et al. 13 under atypical and 1 under typical recognition paradigm and increased amygdala response to fearful facial affect
[65] antipsychotic medication) and 10 healthy
controls
Blumberg 36 BD subjects: 11 elevated mood state, Color-word Stroop task Compared with controls, elevated mood group showed small signal
et al. [66] 10 depressed mood state, 15 euthymic increase in right frontal cortex; and depressed mood group showed
mood state (13 unmedicated; 11 under large signal increase in the left frontal cortex. Regardless of the mood
lithium; 11 under anticonvulsants; 13 state, BD subjects showed increased activation in rostral region of left
under antidepressants; 3 under atypical VPFC
and 2 typical antipsychotic medication);
20 matched healthy controls
Adler et al. 12 euthymic BD patients (8 medicated Working Memory: Two-back BD patients performed more poorly than healthy controls. Compared
[67] with mood stabilizers and/or task, zero-back control/ with controls, BD patients had increased activation in fronto-polar
antipsychotic) and 10 healthy controls attention task prefrontal cortex, temporal cortex, basal ganglia, thalamus, and
posterior parietal cortex
Chang et al. 12 young (12–18 years old) male BD Working Memory: two-back task Compared with controls, BD subjects showed increased activations in
[68] subjects off medication for 24 h; 10 bilateral ACC, left putamen, left thalamus, left DLPFC, and right
age-matched healthy controls inferior frontal gyrus
Gruber et al. 14 BD patients on stable pharmacotherapy Stroop test Compared with controls, BD patients had reduced activations in the
[69] regimen (11 under mood stabilizers; 7 right subdivision of the ACC and increased activation in DLPFC
under antipsychotic medication; 4 under
antidepressants; 3 unmedicated) 10
healthy controls
Lawrence 12 euthymic BD patients (5 under SSRIs Faces: fear, happiness, and Compared with controls, BD patients had increased activations in the
et al. [70] medication; 5 under atypical sadness ventral striatal, thalamic, hippocampal, and ventral prefrontal cortical
antipsychotics; 9 under mood stabilizers) areas in response to intense fear, mild happiness, and mild sadness
fMRI in Psychiatric Disorders
ACC anterior cingulate cortex, BD bipolar disorder, DLPFC dorsolateral prefrontal cortex, VPFC ventral prefrontal cortex
fMRI in Psychiatric Disorders 673
4.2.1 fMRI Features Given that the strongest risk factor for developing mania is a posi-
of Subjects at Risk tive family history (FH+) for BD [88], several studies have
to Develop Bipolar Disorder attempted to characterize the fMRI features in non symptomatic
subjects at risk using cognitive and emotional processing tasks. In
a motor inhibition task, compared to healthy controls and BD
patients, asymptomatic youths with a first-degree BD relative
exhibited increased activation of the putamen during unsuccessful
inhibition [89]. Compared with their low-risk peers, children
without disorders born from parents with BD showed aberrant
prefrontal neural responses to reward, aberrant connectivities
among reward-related regions, and neural correlates in mesolimbic
regions to novelty seeking and impulsive traits [90]. During a WM
task, compared to controls, both BD patients and non symptom-
atic FH+ subjects exhibited failure to suppress emotional arousal
and functional activity of the anterior insular and frontopolar cor-
tices [91]. While processing facial expressions, relative to HC, both
BD patients and FH+ subjects rated anger faces as less hostile and
they showed decreased modulation in the amygdala and inferior
frontal gyrus during anger face presentation [92]. Youth at
increased genetic risk for BD demonstrated reduced brain signal of
the left inferior frontal gyrus when inhibiting responses to fearful
face stimuli, compared with subjects from control families [93].
5 Anxiety Disorders
Obsessive-compulsive disorder
van den Heuvel 22 unmedicated OCD patients and Tower of London Compared with healthy controls, OCD patients showed decreased frontal-
et al. [97] 22 healthy controls striatal responsiveness, mainly in dorsolateral prefrontal cortex and caudate
nucleus; and increased involvement of anterior cingulate, ventrolateral
prefrontal, and parahippocampal cortices
Remijnse et al. 20 unmedicated OCD patients and Reversal learning task Compared with healthy controls, patients with OCD showed reduced number of
[98] 27 healthy controls correct responses but similar responses to receipt of punishment and
demonstrated normal affective switching. During the task patients showed
reduced activations in the right medial and lateral OFC and in right caudate
nucleus. Patients recruited the left posterior OFC, bilateral insular cortex,
bilateral dorsolateral, and bilateral anterior prefrontal cortex to a lesser extent
than control subjects
Roth et al. [99] 12 adults with OCD (6 under SSRIs, Response inhibition: During response inhibition, healthy controls demonstrated right-hemisphere
and 6 had not taken any Go/No-go activation while the patient group showed a more diffuse and bilateral
psychotropic medication for at least pattern of activation. The OCD group had less activation than control group
6 weeks before scanning) and 14 during response inhibition in several right-hemisphere regions. Severe OCD
healthy control subjects symptoms were positively correlated with thalamic and posterior cortical
activations and inversely correlated with right OFC and anterior cingulate
gyri activations
Posttraumatic stress disorder
Rauch et al. 8 Vietnam combat veterans with Masked-fearful vs. Subjects with PTSD had an increased amygdala response to the masked-fearful
[100] PTSD and 8 Vietnam combat masked-happy faces faces as compared to control group and to the PTSD group’s responses to
veterans free from PTSD the masked-happy faces
Lanius et al. 9 traumatized subjects with PTSD, 9 Script-driven symptom Compared with controls, PTSD subjects showed reduced activation of the
[101] traumatized subjects without PTSD provocation thalamus, anterior cingulate gyrus, and medial frontal gyrus
Shin et al. [102] 8 Vietnam veterans with PTSD, 8 Emotional counting Compared with controls, PTSD group exhibited diminished response in rostral
fMRI in Psychiatric Disorders
Hendler et al. 21 male veterans, 10 with PTSD (1 Parametric factorial Compared with veterans without PTSD, PTSD group showed increased
[103] unmedicated; 7 under design with combat activation in amygdala in response to all images and increased activation in
antidepressants; 1 under slides and visual cortex of PTSD group in response to combat content
antipsychotics; 1 under mood noncombat slides
stabilizers; 6 under
benzodiazepams,, and 11 (all
unmedicated) without
Lanius et al. 10 traumatized subjects with PTSD, Script-driven symptom Compared with controls, PTSD subjects had less activation of the thalamus and
[104] 10 traumatized subjects without provocation the anterior cingulate gyrus
PTSD
Driessen et al. 12 traumatized female patients with Autobiographical Subgroup without PTSD: predominant bilateral activation of OFC and Broca’s
[105] BPD, 6 with PTSD and 6 without recall of traumatic area. Subjects with PTSD: predominant activation of right anterior temporal
PTSD vs. negative but lobes, mesiotemporal areas, amygdala, posterior cingulate gyrus, occipital
nontraumatic events areas, and cerebellum
Protopopescu 11 patients with assault-related PTSD, Trauma and Compared with controls, PTSD patients had increased initial amygdala
et al. [106] 21 healthy controls nontrauma-related response to trauma-related emotional words and did not become habituated
emotional words to negative stimuli
Shin et al. [107] 13 traumatized subjects with PTSD, Emotional faces Compared with traumatized subjects without PTSD, the PTSD group
13 traumatized subjects without exhibited increased amygdala response and diminished medial prefrontal
PTSD cortex response to fearful facial expressions and decreased ability to habituate
right amygdala response to fearful faces
BPD borderline personality disorder, OCD obsessive-compulsive disorder, OFC orbitofrontal cortex, PTSD posttraumatic stress disorder
fMRI in Psychiatric Disorders 681
5.1.1 fMRI Features Genetic epidemiological studies have revealed that OCD has a sig-
of Subjects at Risk nificant familial aggregation [116]. The aggregate risk in first-
to Develop Obsessive- degree relative of probands with OCD has been estimated at
Compulsive Disorder approximately 8–23 % [116]. As relatives of patients are at a signifi-
cantly higher risk of developing OCD symptoms than the general
population, young relatives at risk represent a valuable group to
examine potential neurobiological precursors of the disorder, how-
ever up to date few fMRI studies have been carried on in this popu-
lation. One of these studies observed that a WM increased task
682 Erin L. Habecker et al.
5.2.1 fMRI Features Only one study so far investigated the fMRI features related to the
of Subjects at Risk risk to develop PTSD [143]. Authors assessed the fMRI differ-
to Develop Posttraumatic ences between combat-exposed veterans with PTSD and their
Stress Disorder identical combat-unexposed co-twins vs. combat-exposed veterans
without PTSD and their identical combat-unexposed co-twins.
During a Multi-Source Interference Task, combat-exposed veter-
ans with PTSD and their unexposed co-twins had significantly
greater activation in the dorsal ACC and tended to have larger
response time difference scores, as compared to combat-exposed
veterans without PTSD and their co-twins [143]. This cerebral
activation in the unexposed twins was positively correlated with
their combat exposed co-twins’ PTSD symptom severity [143]
leading authors to conclude that hyperresponsivity in the dorsal
ACC appears to be a familial risk factor for the development of
PTSD following psychological trauma [143].
Stimulation
Function paradigms Observed activation differences in affected group Disorder
Emotional Evocation of happy ↓ Anterior cingulate Schizophrenia
processing and sad mood
↓ Amygdala–hippocampal complex (Hempel et al. [22])
(autobiographical
↓ Amygdala (Habel et al. [24], Williams et al. [28])
scripts, positive,
↓ Right amygdala
and negative
↓ mPFC (Takahashi et al. [27])
words, etc.),
↑ Left amygdala Major depressive disorder
affective pictures,
↑ Ventral striatum
facial expressions
↑ Left hippocampus, insula, caudate nucleus, thalamus, dorsal cingulate gyrus, inferior
parietal cortex (Fu et al. [19])
↑ Amygdala (Siegle et al. [58, 59])
↓ DLPFC (Yurgulen-Todd et al. [65]) Bipolar disorder
↑ Amygdala (Yurgulen-Todd et al. [65], Malhi et al. [72], Blumberg et al. [66],
Pavuluri et al. [76])
↑ Thalamus (Lawrence et al. [70], Malhi et al. [71, 72])
↑ Ventral striatum, ventral PFC
↑ Hippocampus (Lawrence et al. [70])
↑ Caudate nucleus (Malhi et al. [71])
↑ Hypothalamus, medial globus pallidus (Malhi et al. [72])
↓ Rostral anterior cingulate (Blumberg et al. [66])
↓ Right rostral ventrolateral PFC (Pavuluri et al. [76])
↑ Right pregenual anterior cingulate, paralimbic cortex (Pavuluri et al. [76])
↑ Amygdala (Rauch et al. [100], Hendler et al. [103], Driessen et al. [105], Posttraumatic stress disorder
Protopopescu et al. [106], Shin et al. [107])
↓ Thalamus, anterior cingulate gyrus (Lanius et al. [104])
↑ Right anterior temporal lobes, mesiotemporal areas, posterior cingulate gyrus,
occipital areas, cerebellum (Driessen et al. [105])
fMRI in Psychiatric Disorders
(continued)
Stimulation
Function paradigms Observed activation differences in affected group Disorder
Working Word encoding/ ↑ Basal ganglia Schizophrenia
memory/ recognition,
↑ Thalamus (Manoach et al. [20])
attention delayed match to
↑ Parahippocampus (Ragland et al. [26])
sample, Sternberg
↓ Prefrontal cortex (Ragland et al. [26], Morey et al. [30])
test, N-back,
Erin L. Habecker et al.
ACC anterior cingulate cortex, DLPFC dorsolateral prefrontal cortex, VPFC ventral prefrontal cortex
fMRI in Psychiatric Disorders
689
690 Erin L. Habecker et al.
References
1. Brown GG, Perthen JE, Liu TT, Buxton RB in PaCO2 on cerebral blood volume, blood
(2007) A primer on functional magnetic reso- flow, and vascular mean transit time. Stroke
nance imaging. Neuropsychol Rev 17(2): 5(5):630–639
107–125 11. Reiman EM, Raichle ME, Robins E, Butler
2. Francati V, Vermetten E, Bremner JD (2007) FK, Herscovitch P, Fox P, Perlmutter J
Functional neuroimaging studies in posttrau- (1986) The application of positron emission
matic stress disorder: review of current methods tomography to the study of panic disorder.
and findings. Depress Anxiety 24(3):202–218 Am J Psychiatry 143(4):469–477
3. Giardino ND, Friedman SD, Dager SR 12. Aguirre GK, Detre JA, Wang J (2005)
(2007) Anxiety, respiration, and cerebral Perfusion fMRI for functional neuroimaging.
blood flow: implications for functional brain Int Rev Neurobiol 66:213–236
imaging. Compr Psychiatry 48(2):103–112 13. Liu TT, Brown GG (2007) Measurement
4. Yurgelun-Todd DA, Renshaw PF, Femia LA of cerebral perfusion with arterial spin label-
(2006) Applications of fMRI to psychiatry. ing: Part 1. Methods. J Int Neuropsychol
In: Faro SH, Mohamed FB (eds) Functional Soc 13(3):517–525. doi:10.1017/
MRI: basic principles and clinical applica- S1355617707070646
tions. Springer, New York, pp 183–220 14. Rao SM, Salmeron BJ, Durgerian S, Janowiak
5. Lai S, Hopkins AL, Haacke EM, Li D, JA, Fischer M, Risinger RC, Conant LL,
Wasserman BA, Buckley P, Friedman L, Stein EA (2000) Effects of methylpheni-
Meltzer H, Hedera P, Friedland R (1993) date on functional MRI blood-oxygen-
Identification of vascular structures as a major level-dependent contrast. Am J Psychiatry
source of signal contrast in high resolution 157(10):1697–1699
2D and 3D functional activation imaging of 15. Mildner T, Zysset S, Trampel R, Driesel W,
the motor cortex at 1.5T: preliminary results. Moller HE (2005) Towards quantifica-
Magn Reson Med 30(3):387–392 tion of blood-flow changes during cognitive
6. Saad ZS, Ropella KM, DeYoe EA, Bandettini task activation using perfusion-based fMRI.
PA (2003) The spatial extent of the BOLD Neuroimage 27(4):919–926
response. Neuroimage 19(1):132–144 16. Tjandra T, Brooks JCW, Figueiredo P, Wise R,
7. Ide K, Eliasziw M, Poulin MJ (2003) Matthews PM, Tracey I (2005) Quantitative
Relationship between middle cerebral artery assessment of the reproducibility of functional
blood velocity and end-tidal PCO2 in the activation measured with BOLD and MR
hypocapnic-hypercapnic range in humans. perfusion imaging: implications for clinical
J Appl Physiol (1985) 95(1):129–137 trial design. Neuroimage 27(2):393–401
8. Poulin MJ, Liang PJ, Robbins PA (1996) 17. Mayberg HS (2003) Modulating dysfunc-
Dynamics of the cerebral blood flow tional limbic-cortical circuits in depression:
response to step changes in end-tidal PCO2 towards development of brain-based algo-
and PO2 in humans. J Appl Physiol (1985) rithms for diagnosis and optimised treatment.
81(3):1084–1095 Br Med Bull 65:193–207
9. Rostrup E, Knudsen GM, Law I, Holm S, 18. Deckersbach T, Dougherty DD, Rauch SL
Larsson HBW, Paulson OB (2005) The rela- (2006) Functional imaging of mood and anx-
tionship between cerebral blood flow and vol- iety disorders. J Neuroimaging 16(1):1–10
ume in humans. Neuroimage 24(1):1–11 19. Fu CHY, Williams SCR, Cleare AJ, Brammer
10. Grubb RL, Raichle ME, Eichling JO, Ter- MJ, Walsh ND, Kim J, Andrew CM, Pich EM,
Pogossian MM (1974) The effects of changes Williams PM, Reed LJ, Mitterschiffthaler MT,
fMRI in Psychiatric Disorders 691
Suckling J, Bullmore ET (2004) Attenuation 29. Honey GD, Honey RAE, O’Loughlin C,
of the neural response to sad faces in major Sharar SR, Kumaran D, Suckling J, Menon
depression by antidepressant treatment: a DK, Sleator C, Bullmore ET, Fletcher PC
prospective, event-related functional mag- (2005) Ketamine disrupts frontal and hippo-
netic resonance imaging study. Arch Gen campal contribution to encoding and retrieval
Psychiatry 61(9):877–889 of episodic memory: an fMRI study. Cereb
20. Manoach DS, Gollub RL, Benson ES, Cortex 15(6):749–759
Searl MM, Goff DC, Halpern E, Saper CB, 30. Morey RA, Inan S, Mitchell TV, Perkins DO,
Rauch SL (2000) Schizophrenic subjects Lieberman JA, Belger A (2005) Imaging fron-
show aberrant fMRI activation of dorso- tostriatal function in ultra-high-risk, early, and
lateral prefrontal cortex and basal ganglia chronic schizophrenia during executive pro-
during working memory performance. Biol cessing. Arch Gen Psychiatry 62(3):254–262
Psychiatry 48(2):99–9109 31. Yurgelun-Todd DA, Coyle JT, Gruber SA,
21. Rubia K, Russell T, Bullmore ET, Soni W, Renshaw PF, Silveri MM, Amico E, Cohen
Brammer MJ, Simmons A, Taylor E, Andrew B, Goff DC (2005) Functional magnetic
C, Giampietro V, Sharma T (2001) An fMRI resonance imaging studies of schizophrenic
study of reduced left prefrontal activation in patients during word production: effects of
schizophrenia during normal inhibitory func- D-cycloserine. Psychiatry Res 138(1):23–31
tion. Schizophr Res 52(1–2):47–55 32. Juckel G, Schlagenhauf F, Koslowski M,
22. Hempel A, Hempel E, Schonknecht P, Filonov D, Wustenberg T, Villringer A,
Stippich C, Schroder J (2003) Impairment Knutson B, Kienast T, Gallinat J, Wrase J,
in basal limbic function in schizophrenia Heinz A (2006) Dysfunction of ventral stria-
during affect recognition. Psychiatry Res tal reward prediction in schizophrenic patients
122(2):115–124 treated with typical, not atypical, neuroleptics.
23. Hofer A, Weiss EM, Golaszewski SM, Psychopharmacology (Berl) 187(2):222–228
Siedentopf CM, Brinkhoff C, Kremser C, 33. Vink M, Ramsey NF, Raemaekers M, Kahn
Felber S, Fleischhacker WW (2003) An FMRI RS (2006) Striatal dysfunction in schizophre-
study of episodic encoding and recognition of nia and unaffected relatives. Biol Psychiatry
words in patients with schizophrenia in remis- 60(1):32–39
sion. Am J Psychiatry 160(5):911–918 34. Braver TS, Barch DM, Kelley WM, Buckner
24. Kubicki M, McCarley RW, Nestor PG, RL, Cohen NJ, Miezin FM, Snyder AZ,
Huh T, Kikinis R, Shenton ME, Wible CG Ollinger JM, Akbudak E, Conturo TE,
(2003) An fMRI study of semantic process- Petersen SE (2001) Direct comparison of
ing in men with schizophrenia. Neuroimage prefrontal cortex regions engaged by working
20(4):1923–1933 and long-term memory tasks. Neuroimage
25. Habel U, Klein M, Shah NJ, Toni I, Zilles 14(1 Pt 1):48–59
K, Falkai P, Schneider F (2004) Genetic load 35. Cohen NJ, Ryan J, Hunt C, Romine L,
on amygdala hypofunction during sadness Wszalek T, Nash C (1999) Hippocampal
in nonaffected brothers of schizophrenia system and declarative (relational) memory:
patients. Am J Psychiatry 161(10):1806–1813 summarizing the data from functional neuro-
26. Ragland JD, Gur RC, Valdez J, Turetsky BI, imaging studies. Hippocampus 9(1):83–98
Elliott M, Kohler C, Siegel S, Kanes S, Gur 36. Nyberg L, Marklund P, Persson J, Cabeza
RE (2004) Event-related fMRI of fronto- R, Forkstam C, Petersson KM, Ingvar M
temporal activity during word encoding and (2003) Common prefrontal activations dur-
recognition in schizophrenia. Am J Psychiatry ing working memory, episodic memory, and
161(6):1004–1015 semantic memory. Neuropsychologia 41(3):
27. Takahashi H, Koeda M, Oda K, Matsuda T, 371–377
Matsushima E, Matsuura M, Asai K, Okubo 37. Klosterkotter J, Schultze-Lutter F, Bechdolf
Y (2004) An fMRI study of differential neural A, Ruhrmann S (2011) Prediction and pre-
response to affective pictures in schizophre- vention of schizophrenia: what has been
nia. Neuroimage 22(3):1247–1254 achieved and where to go next? World
28. Williams LM, Das P, Harris AWF, Liddell BB, Psychiatry 10(3):165–174
Brammer MJ, Olivieri G, Skerrett D, Phillips 38. Yung AR, Phillips LJ, Yuen HP, McGorry PD
ML, David AS, Peduto A, Gordon E (2004) (2004) Risk factors for psychosis in an ultra
Dysregulation of arousal and amygdala- high-risk group: psychopathology and clinical
prefrontal systems in paranoid schizophrenia. features. Schizophr Res 67(2–3):131–142.
Am J Psychiatry 161(3):480–489 doi:10.1016/S0920-9964(03)00192-0
692 Erin L. Habecker et al.
39. Yung AR, Phillips LJ, Yuen HP, Francey 48. Rausch F, Mier D, Eifler S, Esslinger C,
SM, McFarlane CA, Hallgren M, McGorry Schilling C, Schirmbeck F, Englisch S, Meyer-
PD (2003) Psychosis prediction: 12-month Lindenberg A, Kirsch P, Zink M (2014)
follow up of a high-risk (“prodromal”) Reduced activation in ventral striatum and
group. Schizophr Res 60(1):21–32, ventral tegmental area during probabilistic
S0920996402001676 [pii] decision-making in schizophrenia. Schizophr
40. Falkenberg I, Chaddock C, Murray RM, Res 156(2–3):143–149. doi:10.1016/j.
McDonald C, Modinos G, Bramon E, Walshe schres.2014.04.020
M, Broome M, McGuire P, Allen P (2015) 49. Chung YS, Kang DH, Shin NY, Yoo SY,
Failure to deactivate medial prefrontal cor- Kwon JS (2008) Deficit of theory of mind
tex in people at high risk for psychosis. Eur in individuals at ultra-high-risk for schizo-
Psychiatry 30(5):633–640. doi:10.1016/j. phrenia. Schizophr Res 99(1–3):111–118.
eurpsy.2015.03.003 doi:10.1016/j.schres.2007.11.012
41. Karlsgodt KH, van Erp TG, Bearden CE, 50. Marjoram D, Job DE, Whalley HC,
Cannon TD (2014) Altered relationships Gountouna VE, McIntosh AM, Simonotto E,
between age and functional brain activation in Cunningham-Owens D, Johnstone EC, Lawrie
adolescents at clinical high risk for psychosis. S (2006) A visual joke fMRI investigation into
Psychiatry Res 221(1):21–29. doi:10.1016/j. Theory of Mind and enhanced risk of schizo-
pscychresns.2013.08.004 phrenia. Neuroimage 31(4):1850–1858.
42. Lord LD, Allen P, Expert P, Howes O, doi:10.1016/j.neuroimage.2006.02.011
Broome M, Lambiotte R, Fusar-Poli P, 51. Drevets WC (2000) Neuroimaging studies of
Valli I, McGuire P, Turkheimer FE (2012) mood disorders. Biol Psychiatry 48(8):813–829
Functional brain networks before the onset 52. Mayberg HS (1997) Limbic-cortical dys-
of psychosis: a prospective fMRI study with regulation: a proposed model of depression.
graph theoretical analysis. NeuroImage Clin J Neuropsychiatry Clin Neurosci 9(3):471–481
1(1):91–98. doi:10.1016/j.nicl.2012.09.008 53. Mayberg HS, Liotti M, Brannan SK,
43. Pulkkinen J, Nikkinen J, Kiviniemi V, Maki McGinnis S, Mahurin RK, Jerabek PA, Silva
P, Miettunen J, Koivukangas J, Mukkala JA, Tekell JL, Martin CC, Lancaster JL, Fox
S, Nordstrom T, Barnett JH, Jones PB, PT (1999) Reciprocal limbic-cortical func-
Moilanen I, Murray GK, Veijola J (2015) tion and negative mood: converging PET
Functional mapping of dynamic happy and findings in depression and normal sadness.
fearful facial expressions in young adults with Am J Psychiatry 156(5):675–682
familial risk for psychosis—Oulu brain and 54. Baxter LR, Schwartz JM, Phelps ME,
mind study. Schizophr Res 164(1–3):242– Mazziotta JC, Guze BH, Selin CE, Gerner
249. doi:10.1016/j.schres.2015.01.039 RH, Sumida RM (1989) Reduction of pre-
44. Hart SJ, Bizzell J, McMahon MA, Gu H, frontal cortex glucose metabolism common to
Perkins DO, Belger A (2013) Altered fronto- three types of depression. Arch Gen Psychiatry
limbic activity in children and adolescents 46(3):243–250
with familial high risk for schizophrenia. 55. Bench CJ, Friston KJ, Brown RG, Scott
Psychiatry Res 212(1):19–27. doi:10.1016/j. LC, Frackowiak RS, Dolan RJ (1992) The
pscychresns.2012.12.003 anatomy of melancholia—focal abnormalities
45. Li HJ, Chan RC, Gong QY, Liu Y, Liu SM, Shum of cerebral blood flow in major depression.
D, Ma ZL (2012) Facial emotion processing in Psychol Med 22(3):607–615
patients with schizophrenia and their non-psy- 56. Drevets WC, Price JL, Simpson JR, Todd
chotic siblings: a functional magnetic resonance RD, Reich T, Vannier M, Raichle ME
imaging study. Schizophr Res 134(2–3):143– (1997) Subgenual prefrontal cortex
150. doi:10.1016/j.schres.2011.10.019 abnormalities in mood disorders. Nature
46. Barbour T, Pruitt P, Diwadkar VA (2012) 386(6627):824–827
fMRI responses to emotional faces in chil- 57. Liotti M, Mayberg HS, Brannan SK, McGinnis
dren and adolescents at genetic risk for psy- S, Jerabek P, Fox PT (2000) Differential lim-
chiatric illness share some of the features of bic–cortical correlates of sadness and anxiety
depression. J Affect Disord 136(3):276–285. in healthy subjects: implications for affective
doi:10.1016/j.jad.2011.11.036 disorders. Biol Psychiatry 48(1):30–42
47. Habel U, Chechko N, Pauly K, Koch K, 58. Siegle GJ, Steinhauer SR, Thase ME, Stenger
Backes V, Seiferth N, Shah NJ, Stocker T, VA, Carter CS (2002) Can’t shake that
Schneider F, Kellermann T (2010) Neural feeling: event-related fMRI assessment of sus-
correlates of emotion recognition in schizo- tained amygdala activity in response to emo-
phrenia. Schizophr Res 122(1–3):113–123. tional information in depressed individuals.
doi:10.1016/j.schres.2010.06.009 Biol Psychiatry 51(9):693–707
fMRI in Psychiatric Disorders 693
59. Siegle GJ, Thompson W, Carter CS, 69. Gruber SA, Rogowska J, Yurgelun-Todd DA
Steinhauer SR, Thase ME (2007) Increased (2004) Decreased activation of the anterior
amygdala and decreased dorsolateral pre- cingulate in bipolar patients: an fMRI study.
frontal BOLD responses in unipolar depres- J Affect Disord 82(2):191–201
sion: related and independent features. Biol 70. Lawrence NS, Williams AM, Surguladze
Psychiatry 61(2):198–209 S, Giampietro V, Brammer MJ, Andrew C,
60. Surguladze SA, Young AW, Senior C, Brebion Frangou S, Ecker C, Phillips ML (2004)
G, Travis MJ, Phillips ML (2004) Recognition Subcortical and ventral prefrontal cortical
accuracy and response bias to happy and sad neural responses to facial expressions distin-
facial expressions in patients with major depres- guish patients with bipolar disorder and major
sion. Neuropsychology 18(2):212–218 depression. Biol Psychiatry 55(6):578–587
61. Del-Ben CM, Deakin JF, McKie S, Delvai NA, 71. Malhi GS, Lagopoulos J, Sachdev P, Mitchell
Williams SR, Elliott R, Dolan M, Anderson IM PB, Ivanovski B, Parker GB (2004) Cognitive
(2005) The effect of citalopram pretreatment generation of affect in hypomania: an
on neuronal responses to neuropsychological fMRI study. Bipolar Disord 6(4):271–285.
tasks in normal volunteers: an FMRI study. doi:10.1111/j.1399-5618.2004.00123.x
Neuropsychopharmacology 30(9):1724– 72. Malhi GS, Lagopoulos J, Ward PB, Kumari
1734. doi:10.1038/sj.npp.1300728 V, Mitchell PB, Parker GB, Ivanovski B,
62. Wagner G, Sinsel E, Sobanski T, Kohler S, Sachdev P (2004) Cognitive generation of
Marinou V, Mentzel H-J, Sauer H, Schlosser affect in bipolar depression: an fMRI study.
RGM (2006) Cortical inefficiency in patients Eur J Neurosci 19(3):741–754
with unipolar depression: an event-related 73. Monks PJ, Thompson JM, Bullmore ET,
FMRI study with the Stroop task. Biol Suckling J, Brammer MJ, Williams SCR,
Psychiatry 59(10):958–965 Simmons A, Giles N, Lloyd AJ, Harrison CL,
63. Vollm B, Richardson P, McKie S, Elliott Seal M, Murray RM, Ferrier IN, Young AH,
R, Deakin JFW, Anderson IM (2006) Curtis VA (2004) A functional MRI study
Serotonergic modulation of neuronal of working memory task in euthymic bipolar
responses to behavioural inhibition and rein- disorder: evidence for task-specific dysfunc-
forcing stimuli: an fMRI study in healthy vol- tion. Bipolar Disord 6(6):550–564
unteers. Eur J Neurosci 23(2):552–560 74. Frangou S, Raymont V, Bettany D (2002)
64. Walter H, Wolf RC, Spitzer M, Vasic N (2007) The Maudsley bipolar disorder project. A
Increased left prefrontal activation in patients survey of psychotropic prescribing patterns in
with unipolar depression: an event-related, bipolar I disorder. Bipolar Disord 4(6):
parametric, performance-controlled fMRI 378–385
study. J Affect Disord 101(1–3):175–185 75. Strakowski SM, Delbello MP, Adler CM
65. Yurgelun-Todd DA, Gruber SA, Kanayama (2005) The functional neuroanatomy of
G, Killgore WD, Baird AA, Young AD (2000) bipolar disorder: a review of neuroimaging
fMRI during affect discrimination in bipolar findings. Mol Psychiatry 10(1):105–116
affective disorder. Bipolar Disord 2(3 Pt 2): 76. Pavuluri MN, O’Connor MM, Harral E,
237–248 Sweeney JA (2007) Affective neural cir-
66. Blumberg HP, Donegan NH, Sanislow CA, cuitry during facial emotion processing in
Collins S, Lacadie C, Skudlarski P, Gueorguieva pediatric bipolar disorder. Biol Psychiatry
R, Fulbright RK, McGlashan TH, Gore JC, 62(2):158–167
Krystal JH (2005) Preliminary evidence for 77. Lagopoulos J, Ivanovski B, Malhi GS (2007)
medication effects on functional abnormali- An event-related functional MRI study of
ties in the amygdala and anterior cingulate in working memory in euthymic bipolar disor-
bipolar disorder. Psychopharmacology (Berl) der. J Psychiatry Neurosci 32(3):174–184
183(3):308–313 78. Sheline YI, Barch DM, Donnelly JM,
67. Adler CM, Holland SK, Schmithorst V, Ollinger JM, Snyder AZ, Mintun MA (2001)
Tuchfarber MJ, Strakowski SM (2004) Increased amygdala response to masked emo-
Changes in neuronal activation in patients with tional faces in depressed subjects resolves with
bipolar disorder during performance of a work- antidepressant treatment: an fMRI study. Biol
ing memory task. Bipolar Disord 6(6):540–549 Psychiatry 50(9):651–658
68. Chang K, Adleman NE, Dienes K, Simeonova 79. Chen C-H, Ridler K, Suckling J, Williams S,
DI, Menon V, Reiss A (2004) Anomalous Fu CHY, Merlo-Pich E, Bullmore E (2007)
prefrontal-subcortical activation in familial Brain imaging correlates of depressive symp-
pediatric bipolar disorder: a functional mag- tom severity and predictors of symptom
netic resonance imaging investigation. Arch improvement after antidepressant treatment.
Gen Psychiatry 61(8):781–792 Biol Psychiatry 62(5):407–414
694 Erin L. Habecker et al.
80. Casement MD, Guyer AE, Hipwell AE, motor inhibition in children at risk for bipolar
McAloon RL, Hoffmann AM, Keenan KE, disorder. Prog Neuropsychopharmacol Biol
Forbes EE (2014) Girls’ challenging social Psychiatry 38(2):127–133. doi:10.1016/j.
experiences in early adolescence predict pnpbp.2012.02.014
neural response to rewards and depressive 90. Singh MK, Kelley RG, Howe ME, Reiss
symptoms. Dev Cogn Neurosci 8:18–27. AL, Gotlib IH, Chang KD (2014) Reward
doi:10.1016/j.dcn.2013.12.003 processing in healthy offspring of parents
81. Sharp C, Kim S, Herman L, Pane H, Reuter with bipolar disorder. JAMA Psychiatry
T, Strathearn L (2014) Major depression in 71(10):1148–1156. doi:10.1001/
mothers predicts reduced ventral striatum jamapsychiatry.2014.1031
activation in adolescent female offspring with 91. Thermenos HW, Goldstein JM, Milanovic
and without depression. J Abnorm Psychol SM, Whitfield-Gabrieli S, Makris N,
123(2):298–309. doi:10.1037/a0036191 Laviolette P, Koch JK, Faraone SV, Tsuang
82. Whalley HC, Sussmann JE, Romaniuk L, MT, Buka SL, Seidman LJ (2010) An fMRI
Stewart T, Papmeyer M, Sprooten E, Hackett study of working memory in persons with
S, Hall J, Lawrie SM, McIntosh AM (2013) bipolar disorder or at genetic risk for bipolar
Prediction of depression in individuals at high disorder. Am J Med Genet B Neuropsychiatr
familial risk of mood disorders using func- Genet 153B(1):120–131. doi:10.1002/
tional magnetic resonance imaging. PLoS ajmg.b.30964
One 8(3):e57357. doi:10.1371/journal. 92. Brotman MA, Deveney CM, Thomas LA,
pone.0057357 Hinton KE, Yi JY, Pine DS, Leibenluft E
83. Mannie ZN, Filippini N, Williams C, Near (2014) Parametric modulation of neural
J, Mackay CE, Cowen PJ (2014) Structural activity during face emotion processing in
and functional imaging of the hippocampus unaffected youth at familial risk for bipolar
in young people at familial risk of depres- disorder. Bipolar Disord 16(7):756–763.
sion. Psychol Med 44(14):2939–2948. doi:10.1111/bdi.12193
doi:10.1017/S0033291714000580 93. Roberts G, Green MJ, Breakspear M,
84. Miskowiak KW, Glerup L, Vestbo C, Harmer McCormack C, Frankland A, Wright A, Levy
CJ, Reinecke A, Macoveanu J, Siebner HR, F, Lenroot R, Chan HN, Mitchell PB (2013)
Kessing LV, Vinberg M (2015) Different Reduced inferior frontal gyrus activation dur-
neural and cognitive response to emotional ing response inhibition to emotional stimuli
faces in healthy monozygotic twins at risk of in youth at high risk of bipolar disorder. Biol
depression. Psychol Med 45(7):1447–1458. Psychiatry 74(1):55–61. doi:10.1016/j.
doi:10.1017/S0033291714002542 biopsych.2012.11.004
85. Yurgelun-Todd DA, Ross AJ (2006) 94. Whiteside SP, Port JD, Abramowitz JS (2004)
Functional magnetic resonance imaging A meta-analysis of functional neuroimaging in
studies in bipolar disorder. CNS Spectr obsessive-compulsive disorder. Psychiatry Res
11(4):287–297 132(1):69–79
86. Blumberg HP, Leung HC, Skudlarski P, 95. Mataix-Cols D, Rosario-Campos MC,
Lacadie CM, Fredericks CA, Harris BC, Leckman JF (2005) A multidimensional
Charney DS, Gore JC, Krystal JH, Peterson model of obsessive-compulsive disorder. Am
BS (2003) A functional magnetic resonance J Psychiatry 162(2):228–238. doi:10.1176/
imaging study of bipolar disorder: state- and appi.ajp.162.2.228
trait-related dysfunction in ventral prefrontal 96. Saxena S, Brody AL, Ho ML, Alborzian S, Ho
cortices. Arch Gen Psychiatry 60(6):601– MK, Maidment KM, Huang SC, Wu HM, Au
609. doi:10.1001/archpsyc.60.6.601 SC, Baxter LR Jr (2001) Cerebral metabolism
87. Strakowski SM, Adler CM, Holland SK, in major depression and obsessive-compulsive
Mills NP, DelBello MP, Eliassen JC (2005) disorder occurring separately and concur-
Abnormal FMRI brain activation in euthy- rently. Biol Psychiatry 50(3):159–170
mic bipolar disorder patients during a count- 97. van den Heuvel OA, Veltman DJ,
ing Stroop interference task. Am J Psychiatry Groenewegen HJ, Cath DC, van Balkom
162(9):1697–1705 AJ, van Hartskamp J, Barkhof F, van Dyck
88. Goodwin FK, Jamison KR (1990) Manic- R (2005) Frontal-striatal dysfunction dur-
depressive illness. Oxford University Press, ing planning in obsessive-compulsive disor-
New York der. Arch Gen Psychiatry 62(3):301–309.
89. Deveney CM, Connolly ME, Jenkins SE, Kim P, doi:10.1001/archpsyc.62.3.301
Fromm SJ, Brotman MA, Pine DS, Leibenluft 98. Remijnse PL, Nielen MMA, van Balkom
E (2012) Striatal dysfunction during failed AJLM, Cath DC, van Oppen P, Uylings HBM,
fMRI in Psychiatric Disorders 695
118. de Wit SJ, de Vries FE, van der Werf YD, 129. Vermetten E, Vythilingam M, Southwick SM,
Cath DC, Heslenfeld DJ, Veltman EM, van Charney DS, Bremner JD (2003) Long-term
Balkom AJ, Veltman DJ, van den Heuvel OA treatment with paroxetine increases verbal
(2012) Presupplementary motor area hyper- declarative memory and hippocampal vol-
activity during response inhibition: a candi- ume in posttraumatic stress disorder. Biol
date endophenotype of obsessive-compulsive Psychiatry 54(7):693–702
disorder. Am J Psychiatry 169(10):1100– 130. Pitman RK, Shin LM, Rauch SL (2001)
1108. doi:10.1176/appi.ajp.2012.12010073 Investigating the pathogenesis of posttrau-
119. Charney DS, Deutch AY, Krystal JH, Southwick matic stress disorder with neuroimaging.
SM, Davis M (1993) Psychobiologic mecha- J Clin Psychiatry 62(Suppl 17):47–54
nisms of posttraumatic stress disorder. Arch 131. Villarreal G, King CY (2001) Brain imaging
Gen Psychiatry 50(4):295–305 in posttraumatic stress disorder. Semin Clin
120. Charney DS (2004) Psychobiological mecha- Neuropsychiatry 6(2):131–145
nisms of resilience and vulnerability: impli- 132. Morgan MA, LeDoux JE (1995) Differential
cations for successful adaptation to extreme contribution of dorsal and ventral medial
stress. Am J Psychiatry 161(2):195–216 prefrontal cortex to the acquisition and
121. Quirk GJ, Gehlert DR (2003) Inhibition of extinction of conditioned fear in rats. Behav
the amygdala: key to pathological states? Ann Neurosci 109(4):681–688
N Y Acad Sci 985:263–272 133. Quirk GJ, Russo GK, Barron JL, Lebron K
122. Stein MB, Simmons AN, Feinstein JS, (2000) The role of ventromedial prefrontal
Paulus MP (2007) Increased amygdala and cortex in the recovery of extinguished fear.
insula activation during emotion processing J Neurosci 20(16):6225–6231
in anxiety-prone subjects. Am J Psychiatry 134. Santini E, Ge H, Ren K, Pena de Ortiz S, Quirk
164(2):318–327 GJ (2004) Consolidation of fear extinction
123. Geuze E, Vermetten E, Bremner JD (2005) requires protein synthesis in the medial pre-
MR-based in vivo hippocampal volumetrics: frontal cortex. J Neurosci 24(25):5704–5710
2. Findings in neuropsychiatric disorders. 135. Williams LM, Kemp AH, Felmingham K,
Mol Psychiatry 10(2):160–184 Barton M, Olivieri G, Peduto A, Gordon
124. Bremner JD (2002) Neuroimaging of child- E, Bryant RA (2006) Trauma modulates
hood trauma. Semin Clin Neuropsychiatry amygdala and medial prefrontal responses
7(2):104–112 to consciously attended fear. Neuroimage
125. Bremner JD, Vythilingam M, Vermetten E, 29(2):347–357
Southwick SM, McGlashan T, Nazeer A, 136. Golier JA, Yehuda R, Lupien SJ, Harvey PD,
Khan S, Vaccarino LV, Soufer R, Garg PK, Grossman R, Elkin A (2002) Memory per-
Ng CK, Staib LH, Duncan JS, Charney formance in Holocaust survivors with post-
DS (2003) MRI and PET study of defi- traumatic stress disorder. Am J Psychiatry
cits in hippocampal structure and function 159(10):1682–1688
in women with childhood sexual abuse and 137. Rauch SL, Shin LM, Phelps EA (2006)
posttraumatic stress disorder. Am J Psychiatry Neurocircuitry models of posttraumatic stress
160(5):924–932 disorder and extinction: human neuroimag-
126. Gilbertson MW, Shenton ME, Ciszewski ing research—past, present, and future. Biol
A, Kasai K, Lasko NB, Orr SP, Pitman RK Psychiatry 60(4):376–382
(2002) Smaller hippocampal volume pre- 138. Shin LM, Shin PS, Heckers S, Krangel TS,
dicts pathologic vulnerability to psychological Macklin ML, Orr SP, Lasko N, Segal E,
trauma. Nat Neurosci 5(11):1242–1247 Makris N, Richert K, Levering J, Schacter
127. Whalen PJ, Rauch SL, Etcoff NL, McInerney DL, Alpert NM, Fischman AJ, Pitman RK,
SC, Lee MB, Jenike MA (1998) Masked Rauch SL (2004) Hippocampal function in
presentations of emotional facial expressions posttraumatic stress disorder. Hippocampus
modulate amygdala activity without explicit 14(3):292–300
knowledge. J Neurosci 18(1):411–418 139. Vermetten E, Bremner JD (2002) Circuits
128. Lanius RA, Williamson PC, Bluhm RL, and systems in stress. II. Applications to neu-
Densmore M, Boksman K, Neufeld RWJ, Gati robiology and treatment in posttraumatic
JS, Menon RS (2005) Functional connectiv- stress disorder. Depress Anxiety 16(1):14–38
ity of dissociative responses in posttraumatic 140. Bremner JD, Narayan M, Staib LH,
stress disorder: a functional magnetic reso- Southwick SM, McGlashan T, Charney DS
nance imaging investigation. Biol Psychiatry (1999) Neural correlates of memories of
57(8):873–884 childhood sexual abuse in women with and
fMRI in Psychiatric Disorders 697
without posttraumatic stress disorder. Am Higher brain blood flow at amygdala and lower
J Psychiatry 156(11):1787–1795 frontal cortex blood flow in PTSD patients
141. Shin LM, McNally RJ, Kosslyn SM, with comorbid cocaine and alcohol abuse com-
Thompson WL, Rauch SL, Alpert NM, pared with normals. Psychiatry 63(1):65–74
Metzger LJ, Lasko NB, Orr SP, Pitman RK 143. Shin LM, Bush G, Milad MR, Lasko NB,
(1999) Regional cerebral blood flow during Brohawn KH, Hughes KC, Macklin ML,
script-driven imagery in childhood sexual Gold AL, Karpf RD, Orr SP, Rauch SL,
abuse-related PTSD: a PET investigation. Am Pitman RK (2011) Exaggerated activation
J Psychiatry 156(4):575–584 of dorsal anterior cingulate cortex during
142. Semple WE, Goyer PF, McCormick R, cognitive interference: a monozygotic twin
Donovan B, Muzic RF, Rugle L, McCutcheon study of posttraumatic stress disorder. Am
K, Lewis C, Liebling D, Kowaliw S, Vapenik J Psychiatry 168(9):979–985. doi:10.1176/
K, Semple MA, Flener CR, Schulz SC (2000) appi.ajp.2011.09121812
Chapter 23
Abstract
fMRI is a technology with great promise as a tool to probe abnormalities of brain activity in neurodegenerative
diseases. The detection of functional brain abnormalities may be useful, in the appropriate clinical context, for
early diagnosis, differential diagnosis, or prognostication. Prediction of response to treatment or therapeutic
monitoring may also be possible with fMRI. In addition, fMRI has the potential to provide a variety of scien-
tific insights that may have clinical relevance, including compensatory hyperactivation of brain circuits or
genetic modulation of functional brain activity.
Key words Alzheimer’s disease, Amyotrophic lateral sclerosis, Functional MRI, Huntington disease,
Magnetic resonance imaging, Neurodegenerative diseases, Parkinson’s disease
1 Introduction
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_23, © Springer Science+Business Media New York 2016
699
700 Bradford C. Dickerson et al.
Very little work has been done with task-related fMRI in patients
with bvFTD, probably in large part due to the difficulties these
patients often have in cooperating with task instructions and in the
setting of MRI. During the viewing of faces conveying emotional
expressions, bvFTD patients exhibit abnormally reduced activation in
ventrolateral prefrontal cortex, insula, and a variety of other brain
regions, but elevated activation in posterior parietal cortex [87].
fMRI has also been used to investigate the neural basis for changes in
understanding of emotion conveyed through music [88] and in rea-
soning for personal vs. impersonal moral dilemmas [89] in bvFTD.
In PPA, an early fMRI study using simple phonological and
semantic tasks demonstrated relatively normal activation within
canonical regions of the language network but with increased
recruitment in areas not typically recruited by controls; these
increases correlated with greater language impairment [90]. Sonty
et al. subsequently used dynamic causal modeling to show that
effective connectivity between canonical language regions was
impaired in PPA [91].
In a study of syntactic comprehension in nonfluent variant
PPA patients, the caudal inferior frontal cortex did not show rela-
tively greater activation for complex sentences than for simple sen-
tences, as it did in controls [92].
Vandenbulcke et al. [93] used an associative-semantic fMRI par-
adigm to demonstrate that patients with the semantic variant of PPA
show an abnormal rightward lateralization of anterior temporal acti-
vation, which is reminiscent of a similar language laterality shift that
has been reported in patients with stroke aphasia. This paradigm was
also used in nonfluent variant PPA patients to identify a similar effect
[94]. The phenomenon of surface dyslexia was investigated in seman-
tic variant PPA; in patients but not controls the inferior parietal cor-
tex was recruited for irregular words, but mid-fusiform and superior
temporal cortex was under-recruited in PPA patients [95].
Patients with the semantic variant of PPA have also been stud-
ied using a paradigm comparing meaningful to meaningless sounds.
Compared with controls, patients showed abnormal activation of
dorsolateral temporal cortical areas for both meaningless sounds as
well as for meaningful sounds (animal sounds versus tool sounds),
suggesting that aberrant processing of sounds in semantic PPA
extends to pre-semantic perceptual processing [96]; prior work by
the same authors demonstrated abnormal auditory processing in
patients with the nonfluent variant of PPA as well [97].
fMRI has been used to study abnormal patterns of brain activa-
tion in a variety of tasks in patients with Parkinson’s disease (PD),
most commonly in tasks engaging the motor system [98, 99]. fMRI
studies investigating brain activations in PD patients during self-initi-
ated movements demonstrated evidence of reduced neural responses
in the pre-supplementary motor areas (SMA), along with hyperactiva-
tion in both the lateral premotor cortex and parietal cortex
fMRI in Neurodegenerative Diseases: From Scientific Insights to Clinical Applications 707
Fig. 2 Pattern of fMRI signal during a verbal two-back working memory task in PD patients and cognitive
impairment. Significant under-recruitment occurred in (a) the bilateral anterior cingulate cortex and (b) the
right caudate in PD patients with MCI (n = 30) compared with those without MCI (n = 26; red). Mean beta
values in (c) the anterior cingulate cortex cluster and (d) right caudate, contrasting two-back with baseline
conditions for PD patients with (blue) and without (red) MCI and control individuals (green). (e) Mean beta
values in the right caudate, contrasting two-back with baseline for motor-matched (Unified Parkinson’s
disease Rating Scale III) groups for PD patients with MCI and without MCI. The group sizes were held con-
stant (n = 18 in each group) and matched by scanner model. Error bars are 1 standard error. Figure reprinted
from [107] with permission
fMRI in Neurodegenerative Diseases: From Scientific Insights to Clinical Applications 709
4.3 Compensatory Aside from neocortical hyperactivation in AD, consistent data sug-
Hyperactivation: gest that there is a phase of increased MTL activation in MCI
A Universal Adaptation (Fig. 3) [188]. This increase, which also may be present in cogni-
Response to Brain tively intact carriers of the APOE-ε4 allele (for review, see [189]),
Injury? may represent, at least in part, an attempted compensatory response
to AD neuropathology [190–192], given that some MCI individu-
als with smaller hippocampal volume perform similarly on memory
tasks to MCI individuals with larger hippocampal volume but have
relatively greater MTL activation [26, 74]. Additional studies
employing event-related fMRI paradigms [18, 24, 75] will be very
helpful in determining whether increased MTL activation in MCI
patients is specifically associated with successful memory, as
opposed to a general effect that is present regardless of success
(possibly indicating increased effort). Whether or not hyperactiva-
tion is associated with better memory performance and thus could
be viewed as behaviorally compensatory, it appears to be associated
with more prominent neurodegeneration [193] and poorer prog-
nosis [194]. It is possible that MTL hyperactivation reflects cholin-
ergic or other neurotransmitter upregulation in MCI patients
[195]. Alternatively, increased regional brain activation may be a
marker of the pathophysiologic process of AD itself, such as aber-
rant sprouting of cholinergic fibers [196] or inefficiency in synaptic
Fig. 3 A phase of compensatory hyperactivation appears to occur in the medial temporal lobe (MTL) in mild
cognitive impairment (MCI), prior to the clinical onset of Alzheimer’s disease (AD) dementia. Representative
single subjects from each group, showing normal memory-related MTL activation measured with fMRI in
normal older controls, hyperactivation, and very mild atrophy in MCI, and hypoactivation and more prominent
atrophy in mild AD (reproduced with permission from [76])
fMRI in Neurodegenerative Diseases: From Scientific Insights to Clinical Applications 715
Behavioral Abilities
Regional brain hypometabolism
4.4 The Modulatory In the last decade, there has been an explosion in literature on
Effects of Genetic Risk imaging and genetics, primarily in psychiatric disorders [228] and
Factors for Neurologic the basic science of genetic modulators of brain function [229,
Disease on Brain 230]. This is an area that is ripe for study in neurologic disease,
Activation with a number of studies having been done in populations at ele-
vated genetic risk for AD, FTD, ALS, and HD.
The APOE ε4 allele is a major genetic susceptibility factor
associated with increased risk for AD. Several fMRI studies have
718 Bradford C. Dickerson et al.
Fig. 5 APOE ε4 status reduces DMN connectivity in healthy older adults. Functional connectivity is decreased
in ε4 carriers compared to ε3 homozygotes in both the (a) anterior, and (b) posterior DMN. The bar graphs
depict the distribution of the effects across subgroups. They show the mean parameter estimates of a selected
region within the anterior DMN (the right anterior cingulate gyrus) and the left posterior cingulate cluster within
the posterior DMN, for male and female ε3 homozygotes and male and female ε4 carriers. The difference
across both genotype and gender is significant for the anterior cingulate gyrus. For the posterior cingulate
gyrus only the difference across genotype is significant. The statistical maps are overlaid on the Montreal
Neurological Institute (MNI) 152 brain; MNI coordinates (in mm) of the slices are displayed. Figure modified
from [239] with permission
4.5 fMRI Very little work has been done on the use of fMRI as a predictive
as a Predictive biomarker for prognosis. Miller et al. pursued such a study of a
Biomarker group of 25 senior citizens spanning the spectrum of MCI, none
of whom were demented at the time of baseline assessment, but
who exhibited varying degrees of mild symptoms of cognitive
impairment clinically [as measured using the CDR sum-of-boxes
(CDR-SB)] [194]. At baseline, subjects performed a visual scene-
encoding task during fMRI scanning and were clinically followed
longitudinally after scanning. Over about 6 years of follow-up after
scanning, subjects demonstrated a wide range of cognitive decline,
with some showing no change and others progressing to dementia
(change in CDR-SB ranged from 0 to 6). The degree of cognitive
decline was predicted by hippocampal activation at the time of
baseline scanning, with greater hippocampal activation predicting
greater decline (Fig. 6). This finding was present even after con-
trolling for baseline degree of impairment (CDR-SB), age, educa-
tion, and hippocampal volume, as well as gender and APOE status.
Similarly, a longitudinal fMRI study showed that healthy subjects
with more rapid cognitive decline over a 2-year period had both
the highest hippocampal activation at baseline and the greatest loss
of hippocampal activation over follow up, where the rate of activa-
tion loss correlated with the rate of cognitive decline [52]. These
data suggest that fMRI may provide a physiologic imaging bio-
marker useful for identifying the subgroup of MCI individuals at
highest risk of cognitive decline for potential inclusion in disease-
modifying clinical trials. Thus, hyperactivation might represent an
early response to AD pathology, which may predict forthcoming
hippocampal failure and memory decline.
4.6 Uses of fMRI FMRI may be particularly valuable in evaluating acute and subacute
in Understanding effects of therapeutic interventions—whether pharmacologic, non-
and Monitoring pharmacologic (e.g., transcranial magnetic stimulation), or behav-
Neurotherapeutics ioral/rehabilitative—on neural activity [261]. This may be useful
for showing that the intervention modulates targeted circuits and
may help elucidate mechanisms of action. Additionally, it may help
identify or even predict treatment responders or nonresponders.
Alterations in memory-related activation related to the admin-
istration of pharmacologic agents known to impair memory can be
detected with pharmacologic fMRI [262–264]. The effects of cog-
nitive enhancing drugs on brain activation during cognitive task
performance have shown that fMRI can detect changes after
administration of cholinesterase inhibitors in patients with AD and
MCI [265, 266]. In one of the earliest studies, after receiving a
single dose of galanthamine, AD patients demonstrated increased
fusiform activity during a face encoding task and increased prefron-
tal activity during a working memory task [265]. Another study
showed that acute dosing increased hippocampal activation during
a memory task while chronic dosing was associated with decreased
activation [267]. Although these pilot studies did not include pla-
cebo-control groups to reduce potential confounding factors, such
as learning effects, they indicate that fMRI is sensitive to both
acute and subacute medication effects, some of which relate to
behavioral change. Similar findings have been observed in subse-
quent placebo-controlled acute dosing studies [268]. Another
placebo-controlled acute dose physostigmine study demonstrated
normalization of both visual stimulus-specific activation in occipi-
tal and attention-dependent activation in frontoparietal regions in
patients with AD [269].
Pharmacological fMRI has also been used to evaluate the effects
of chronic treatment over 2–6 months in patients with AD demen-
tia. For the most part, these studies have demonstrated treatment-
related increased activation in brain regions engaged by the variety
of tasks used [270–272]. Whereas some studies have found
fMRI in Neurodegenerative Diseases: From Scientific Insights to Clinical Applications 723
Fig. 7 Behavioral significance of functional connectivity changes in the middle and posterior cingulate net-
works in AD patients after 12-week donepezil treatment. In the middle congulate connectivity network (a),
functional connectivity changes in the ventral anterior cingulate cortex and the ventral prefrontal cortex were
significantly correlated with the ADAS-cog score change. Functional connectivity change in the ventral anterior
cingulate cortex was correlated with the ADAS-cog score change also in the posterior cingulate connectivity
network (b). The blue solid circles and dotted lines represent the seed regions and the functional connections,
respectively. Figure modified from [285] with permission
6 Conclusions
References
1. Perry D et al (2015) Building a roadmap for 8. Sperling RA, Jack CR Jr, Aisen PS (2011)
developing combination therapies for Testing the right target and right drug at the
Alzheimer’s disease. Expert Rev Neurother right stage. Sci Transl Med 3(111):111cm33
15(3):327–333 9. Sperling RA et al (2011) Toward defining the
2. Selkoe DJ (2013) The therapeutics of preclinical stages of Alzheimer’s disease: rec-
Alzheimer’s disease: where we stand and ommendations from the National Institute on
where we are heading. Ann Neurol Aging and the Alzheimer’s Association work-
74(3):328–336 group. Alzheimers Dement 7(3):280–292
3. DeKosky ST, Marek K (2003) Looking back- 10. Albert MS et al (2011) The diagnosis of mild
ward to move forward: early detection of neu- cognitive impairment due to Alzheimer’s dis-
rodegenerative disorders. Science ease: recommendations from the National
302(5646):830–834 Institute on Aging-Alzheimer’s Association
4. Selkoe DJ (2002) Alzheimer’s disease is a syn- workgroups on diagnostic guidelines for
aptic failure. Science 298(5594):789–791 Alzheimer’s disease. Alzheimers Dement
5. Coleman P, Federoff H, Kurlan R (2004) A 7(3):270–279
focus on the synapse for neuroprotection in 11. McKhann GM et al (2011) The diagnosis of
Alzheimer disease and other dementias. dementia due to Alzheimer’s disease: recom-
Neurology 63(7):1155–1162 mendations from the National Institute on
6. Dickerson BC, Atri A (2014) Dementia: Aging and the Alzheimer’s Association work-
comprehensive principles and practice. group. Alzheimers Dement 7(3):263–269
Oxford University Press, New York 12. Boxer AL et al (2013) The advantages of
7. Dickerson BC (2015) Hodges’ frontotempo- frontotemporal degeneration drug develop-
ral dementia, 2nd edn. Cambridge University ment (part 2 of frontotemporal degeneration:
Press, Cambridge, UK the next therapeutic frontier). Alzheimers
Dement 9(2):189–198
fMRI in Neurodegenerative Diseases: From Scientific Insights to Clinical Applications 729
13. Paulsen JS et al (2014) Clinical and biomarker 28. Bennett CM, Miller MB (2013) fMRI reli-
changes in premanifest Huntington disease show ability: influences of task and experimen-
trial feasibility: a decade of the PREDICT-HD tal design. Cogn Affect Behav Neurosci
study. Front Aging Neurosci 6:78 13(4):690–702
14. Mills SM et al (2013) Preclinical trials in 29. Brandt DJ et al (2013) Test-retest reliability
autosomal dominant AD: implementation of fMRI brain activity during memory encod-
of the DIAN-TU trial. Rev Neurol (Paris) ing. Front Psychiatry 4:163
169(10):737–743 30. Brown GG et al (2011) Multisite reliabil-
15. Petersen RC et al (1999) Mild cognitive ity of cognitive BOLD data. Neuroimage
impairment: clinical characterization and out- 54(3):2163–2175
come. Arch Neurol 56(3):303–308 31. Cao H et al (2014) Test-retest reliability of
16. Grundman M et al (2004) Mild cogni- fMRI-based graph theoretical properties dur-
tive impairment can be distinguished from ing working memory, emotion processing,
Alzheimer disease and normal aging for clini- and resting state. Neuroimage 84:888–900
cal trials. Arch Neurol 61(1):59–66 32. Chase HW et al (2015) Accounting for
17. Sperling R, Mormino E, Johnson K (2014) dynamic fluctuations across time when exam-
The evolution of preclinical Alzheimer’s ining fMRI test-retest reliability: analysis of
disease: implications for prevention trials. a reward paradigm in the EMBARC study.
Neuron 84(3):608–622 PLoS One 10(5):e0126326
18. Dickerson BC et al (2007) Prefrontal- 33. de Bertoldi F et al (2015) Improving the
hippocampal-fusiform activity during encod- reliability of single-subject fMRI by weight-
ing predicts intraindividual differences in ing intra-run variability. Neuroimage
free recall ability: an event-related functional- 114:287–293
anatomic MRI study. Hippocampus 17(11): 34. Frassle S et al (2015) Test-retest reli-
1060–1070 ability of dynamic causal modeling for
19. Touroutoglou A et al (2014) Amygdala fMRI. Neuroimage 117:56–66
task-evoked activity and task-free connectiv- 35. Gee DG et al (2015) Reliability of an fMRI
ity independently contribute to feelings of paradigm for emotional processing in a mul-
arousal. Hum Brain Mapp 35(10):5316–5327 tisite longitudinal study. Hum Brain Mapp
20. Brewer JB et al (1998) Making memories: 36(7):2558–2579
brain activity that predicts how well visual 36. Golestani AM et al (2015) Mapping the end-
experience will be remembered. Science tidal CO2 response function in the resting-
281(5380):1185–1187 state BOLD fMRI signal: spatial specificity,
21. Kirchhoff BA et al (2000) Prefrontal-temporal test-retest reliability and effect of fMRI sam-
circuitry for episodic encoding and subsequent pling rate. Neuroimage 104:266–277
memory. J Neurosci 20(16):6173–6180 37. Lukasova K et al (2014) Test-retest reliabil-
22. Wagner AD et al (1998) Building memories: ity of fMRI activation generated by differ-
remembering and forgetting of verbal expe- ent saccade tasks. J Magn Reson Imaging
riences as predicted by brain activity. Science 40(1):37–46
281(5380):1188–1191 38. McGonigle DJ (2012) Test-retest reliabil-
23. Daselaar SM et al (2003) Neuroanatomical ity in fMRI: or how I learned to stop wor-
correlates of episodic encoding and retrieval rying and love the variability. Neuroimage
in young and elderly subjects. Brain 126(Pt 62(2):1116–1120
1):43–56 39. Plichta MM et al (2012) Test-retest reliability
24. Sperling R et al (2003) Putting names to of evoked BOLD signals from a cognitive-
faces: successful encoding of associative mem- emotive fMRI test battery. Neuroimage
ories activates the anterior hippocampal for- 60(3):1746–1758
mation. Neuroimage 20(2):1400–1410 40. Aurich NK et al (2015) Evaluating the reli-
25. Price CJ, Friston KJ (1999) Scanning patients ability of different preprocessing steps to esti-
with tasks they can perform. Hum Brain mate graph theoretical measures in resting
Mapp 8(2–3):102–108 state fMRI data. Front Neurosci 9:48
26. Dickerson BC et al (2004) Medial temporal 41. Birn RM et al (2013) The effect of scan length
lobe function and structure in mild cognitive on the reliability of resting-state fMRI con-
impairment. Ann Neurol 56(1):27–35 nectivity estimates. Neuroimage 83:550–558
27. Grady CL et al (2003) Evidence from func- 42. Braun U et al (2012) Test-retest reliability of
tional neuroimaging of a compensatory resting-state connectivity network character-
prefrontal network in Alzheimer’s disease. istics using fMRI and graph theoretical mea-
J Neurosci 23(3):986–993 sures. Neuroimage 59(2):1404–1412
730 Bradford C. Dickerson et al.
43. Patriat R et al (2013) The effect of resting 57. Johnson SC et al (2000) The relationship
condition on resting-state fMRI reliabil- between fMRI activation and cerebral atrophy:
ity and consistency: a comparison between comparison of normal aging and Alzheimer
resting with eyes open, closed, and fixated. disease. Neuroimage 11(3):179–187
Neuroimage 78:463–473 58. Saykin AJ et al (1999) Neuroanatomic substrates
44. Wisner KM et al (2013) Neurometrics of intrin- of semantic memory impairment in Alzheimer’s
sic connectivity networks at rest using fMRI: disease: patterns of functional MRI activation.
retest reliability and cross-validation using a J Int Neuropsychol Soc 5(5):377–392
meta-level method. Neuroimage 76:236–251 59. Grossman M et al (2003) Neural basis for
45. Atri A et al (2011) Test-retest reliability of semantic memory difficulty in Alzheimer’s dis-
memory task functional magnetic resonance ease: an fMRI study. Brain 126(Pt 2):292–311
imaging in Alzheimer disease clinical trials. 60. Woodard JL et al (2009) Semantic memory
Arch Neurol 68(5):599–606 activation in amnestic mild cognitive impair-
46. Turner JA et al (2012) Reliability of the ment. Brain 132(Pt 8):2068–2078
amplitude of low-frequency fluctuations in 61. Thulborn KR, Martin C, Voyvodic JT
resting state fMRI in chronic schizophrenia. (2000) Functional MR imaging using a visu-
Psychiatry Res 201(3):253–255 ally guided saccade paradigm for comparing
47. Eaton KP et al (2008) Reliability of fMRI activation patterns in patients with probable
for studies of language in post-stroke aphasia Alzheimer’s disease and in cognitively able
subjects. Neuroimage 41(2):311–322 elderly volunteers. AJNR Am J Neuroradiol
48. Kurland J et al (2004) Test-retest reliability of 21(3):524–531
fMRI during nonverbal semantic decisions in 62. Golden HL et al (2015) Functional neu-
moderate-severe nonfluent aphasia patients. roanatomy of auditory scene analysis in
Behav Neurol 15(3–4):87–97 Alzheimer’s disease. Neuroimage Clin
49. Clement F, Belleville S (2009) Test-retest 7:699–708
reliability of fMRI verbal episodic memory 63. Kato T, Knopman D, Liu H (2001)
paradigms in healthy older adults and in per- Dissociation of regional activation in mild
sons with mild cognitive impairment. Hum AD during visual encoding: a functional MRI
Brain Mapp 30(12):4033–4047 study. Neurology 57(5):812–816
50. Zanto TP, Pa J, Gazzaley A (2014) Reliability 64. Machulda MM et al (2003) Comparison
measures of functional magnetic resonance of memory fMRI response among normal,
imaging in a longitudinal evaluation of mild cog- MCI, and Alzheimer’s patients. Neurology
nitive impairment. Neuroimage 84:443–452 61(4):500–506
51. Poudel GR et al (2015) Functional changes 65. Rombouts SA et al (2000) Functional MR
during working memory in Huntington’s imaging in Alzheimer’s disease during mem-
disease: 30-month longitudinal data from ory encoding. AJNR Am J Neuroradiol
the IMAGE-HD study. Brain Struct Funct 21(10):1869–1875
220(1):501–512 66. Small SA et al (1999) Differential regional dys-
52. O’Brien JL et al (2010) Longitudinal fMRI function of the hippocampal formation among
in elderly reveals loss of hippocampal acti- elderly with memory decline and Alzheimer’s
vation with clinical decline. Neurology disease. Ann Neurol 45(4):466–472
74(24):1969–1976 67. Sperling RA et al (2003) fMRI studies of asso-
53. Scheller E et al (2014) Attempted and suc- ciative encoding in young and elderly con-
cessful compensation in preclinical and early trols and mild Alzheimer’s disease. J Neurol
manifest neurodegeneration—a review of task Neurosurg Psychiatry 74(1):44–50
FMRI studies. Front Psychiatry 5:132 68. Schwindt GC, Black SE (2009) Functional
54. Buckner RL et al (2000) Functional brain imaging studies of episodic memory in
imaging of young, nondemented, and Alzheimer’s disease: a quantitative meta-
demented older adults. J Cogn Neurosci analysis. Neuroimage 45(1):181–190
12(Suppl 2):24–34 69. Rombouts SARB et al (2005) Delayed rather
55. D’Esposito M, Deouell LY, Gazzaley A than decreased BOLD response as a marker
(2003) Alterations in the BOLD fMRI signal for early Alzheimer’s disease. Neuroimage
with ageing and disease: a challenge for neu- 26(4):1078–1085
roimaging. Nat Rev Neurosci 4(11):863–872 70. Dickerson BC et al (2005) Increased hip-
56. Grossman M et al (2003) Neural basis for verb pocampal activation in mild cognitive
processing in Alzheimer’s disease: an fMRI impairment compared to normal aging and
study. Neuropsychology 17(4):658–674 AD. Neurology 65(3):404–411
fMRI in Neurodegenerative Diseases: From Scientific Insights to Clinical Applications 731
71. Johnson SC et al (2006) Activation of brain to white matter pathology. Cereb Cortex
regions vulnerable to Alzheimer’s disease: 22(5):1038–1051
the effect of mild cognitive impairment. 85. Mormino EC et al (2012) Abeta Deposition
Neurobiol Aging 27(11):1604–1612 in aging is associated with increases in brain
72. Johnson SC et al (2004) Hippocampal adap- activation during successful memory encod-
tation to face repetition in healthy elderly and ing. Cereb Cortex 22(8):1813–1823
mild cognitive impairment. Neuropsychologia 86. Villemagne VL et al (2015) Tau imaging:
42(7):980–989 early progress and future directions. Lancet
73. Petrella JR et al (2006) Mild cognitive Neurol 14(1):114–124
impairment: evaluation with 4-T functional 87. Virani K et al (2013) Functional neural corre-
MR imaging. Radiology 240(1):177–186 lates of emotional expression processing deficits
74. Hamalainen A et al (2007) Increased in behavioural variant frontotemporal demen-
fMRI responses during encoding in mild tia. J Psychiatry Neurosci 38(3):174–182
cognitive impairment. Neurobiol Aging 88. Agustus JL et al (2015) Functional MRI of
28(12):1889–1903 music emotion processing in frontotemporal
75. Kircher T et al (2007) Hippocampal activa- dementia. Ann N Y Acad Sci 1337:232–240
tion in MCI patients is necessary for success- 89. Chiong W et al (2013) The salience network
ful memory encoding. J Neurol Neurosurg causally influences default mode network
Psychiatry 78(8):812–818 activity during moral reasoning. Brain 136(Pt
76. Dickerson BC, Sperling RA (2008) Functional 6):1929–1941
abnormalities of the medial temporal lobe 90. Sonty SP et al (2003) Primary progressive
memory system in mild cognitive impair- aphasia: PPA and the language network. Ann
ment and Alzheimer’s disease: insights from Neurol 53(1):35–49
functional MRI studies. Neuropsychologia 91. Sonty SP et al (2007) Altered effective con-
46(6):1624–1635 nectivity within the language network in
77. Dickerson BC, Sperling RA (2009) Large- primary progressive aphasia. J Neurosci
scale functional brain network abnormali- 27(6):1334–1345
ties in Alzheimer’s disease: insights from 92. Wilson SM et al (2010) Neural correlates of
functional neuroimaging. Behav Neurol syntactic processing in the nonfluent variant
21(1):63–75 of primary progressive aphasia. J Neurosci
78. Petrella JR et al (2007) Cortical deactivation 30(50):16845–16854
in mild cognitive impairment: high-field- 93. Vandenbulcke M et al (2005) Anterior tem-
strength functional MR imaging. Radiology poral laterality in primary progressive aphasia
245(1):224–235 shifts to the right. Ann Neurol 58(3):362–370
79. Vandenbulcke M et al (2007) Word reading 94. Nelissen N et al (2011) Right hemisphere
and posterior temporal dysfunction in amnes- recruitment during language processing in fron-
tic mild cognitive impairment. Cereb Cortex totemporal lobar degeneration and Alzheimer’s
17(3):542–551 disease. J Mol Neurosci 45(3):637–647
80. Vannini P et al (2012) Age and amyloid- 95. Wilson SM et al (2009) The neural basis of
related alterations in default network habitu- surface dyslexia in semantic dementia. Brain
ation to stimulus repetition. Neurobiol Aging 132(Pt 1):71–86
33(7):1237–1252
96. Goll JC et al (2012) Nonverbal sound pro-
81. Vannini P et al (2013) The ups and downs of cessing in semantic dementia: a functional
the posteromedial cortex: age- and amyloid- MRI study. Neuroimage 61(1):170–180
related functional alterations of the encod-
ing/retrieval flip in cognitively normal older 97. Goll JC et al (2010) Non-verbal sound pro-
adults. Cereb Cortex 23(6):1317–1328 cessing in the primary progressive aphasias.
Brain 133(Pt 1):272–285
82. Sperling RA et al (2009) Amyloid deposition
is associated with impaired default network 98. Elsinger CL et al (2003) Neural basis for
function in older persons without dementia. impaired time reproduction in Parkinson’s
Neuron 63(2):178–188 disease: an fMRI study. J Int Neuropsychol
Soc 9(7):1088–1098
83. Kennedy KM et al (2012) Effects of beta-
amyloid accumulation on neural function 99. Rowe JB, Siebner HR (2012) The motor system
during encoding across the adult lifespan. and its disorders. Neuroimage 61(2):464–477
Neuroimage 62(1):1–8 100. Sabatini U et al (2000) Cortical motor reorgani-
84. Hedden T et al (2012) Failure to modulate zation in akinetic patients with Parkinson’s dis-
attentional control in advanced aging linked ease: a functional MRI study. Brain 123(Pt 2):
394–403
732 Bradford C. Dickerson et al.
using echo-planar MRI. Magn Reson Med 145. Filippi M et al (2013) Functional net-
34(4):537–541 work connectivity in the behavioral vari-
130. Vincent JL et al (2007) Intrinsic functional ant of frontotemporal dementia. Cortex
architecture in the anaesthetized monkey 49(9):2389–2401
brain. Nature 447(7140):83–86 146. Agosta F et al (2012) Resting state fMRI in
131. Smith SM et al (2009) Correspondence of Alzheimer’s disease: beyond the default mode
the brain’s functional architecture during network. Neurobiol Aging 33(8):1564–1578
activation and rest. Proc Natl Acad Sci U S A 147. Lehmann M et al (2013) Intrinsic connec-
106(31):13040–13045 tivity networks in healthy subjects explain
132. Yeo BT et al (2011) The organization of the clinical variability in Alzheimer’s disease. Proc
human cerebral cortex estimated by intrin- Natl Acad Sci U S A 110(28):11606–11611
sic functional connectivity. J Neurophysiol 148. Gour N et al (2014) Functional connec-
106(3):1125–1165 tivity changes differ in early and late-onset
133. Zhou J et al (2010) Divergent network con- Alzheimer’s disease. Hum Brain Mapp
nectivity changes in behavioural variant fron- 35(7):2978–2994
totemporal dementia and Alzheimer’s disease. 149. Lehmann M et al (2015) Loss of functional
Brain 133(Pt 5):1352–1367 connectivity is greater outside the default
134. Thomas JB et al (2014) Functional connectivity mode network in nonfamilial early-onset
in autosomal dominant and late-onset Alzheimer Alzheimer’s disease variants. Neurobiol Aging
disease. JAMA Neurol 71(9):1111–1122 36(10):2678–2686
135. Binnewijzend MA et al (2012) Resting-state 150. Barnes J et al (2015) Alzheimer’s disease
fMRI changes in Alzheimer’s disease and first symptoms are age dependent: evidence
mild cognitive impairment. Neurobiol Aging from the NACC dataset. Alzheimers Dement
33(9):2018–2028 11(11):1349–1357
136. Brier MR et al (2012) Loss of intranetwork 151. Petrella JR et al (2007) Prognostic value of
and internetwork resting state functional con- posteromedial cortex deactivation in mild cog-
nections with Alzheimer’s disease progres- nitive impairment. PLoS One 2(10):e1104
sion. J Neurosci 32(26):8890–8899 152. Hedden T et al (2009) Disruption of func-
137. Zhang HY et al (2010) Resting brain con- tional connectivity in clinically normal older
nectivity: changes during the progress of adults harboring amyloid burden. J Neurosci
Alzheimer disease. Radiology 256(2):598–606 29(40):12686–12694
138. Sheline YI et al (2009) The default mode 153. Drzezga A et al (2011) Neuronal dysfunc-
network and self-referential processes in tion and disconnection of cortical hubs in
depression. Proc Natl Acad Sci U S A non-demented subjects with elevated amyloid
106(6):1942–1947 burden. Brain 134(Pt 6):1635–1646
139. Petrella JR et al (2011) Default mode net- 154. Brier MR et al (2014) Functional connectivity
work connectivity in stable vs progres- and graph theory in preclinical Alzheimer’s
sive mild cognitive impairment. Neurology disease. Neurobiol Aging 35(4):757–768
76(6):511–517 155. Wang L et al (2013) Cerebrospinal fluid
140. Li SJ et al (2002) Alzheimer disease: evalu- Abeta42, phosphorylated Tau181, and
ation of a functional MR imaging index as a resting-state functional connectivity. JAMA
marker. Radiology 225(1):253–259 Neurol 70(10):1242–1248
141. Supekar K et al (2010) Development of func- 156. Sheline YI et al (2010) Amyloid plaques dis-
tional and structural connectivity within the rupt resting state default mode network con-
default mode network in young children. nectivity in cognitively normal elderly. Biol
Neuroimage 52(1):290–301 Psychiatry 67(6):584–587
142. Damoiseaux JS et al (2012) Functional 157. Farb NA et al (2013) Abnormal network
connectivity tracks clinical deterioration connectivity in frontotemporal dementia:
in Alzheimer’s disease. Neurobiol Aging evidence for prefrontal isolation. Cortex
33(4):828.e19–30 49(7):1856–1873
143. Schwindt GC et al (2013) Modulation of 158. Whitwell JL et al (2011) Altered functional
the default-mode network between rest and connectivity in asymptomatic MAPT sub-
task in Alzheimer’s Disease. Cereb Cortex jects: a comparison to bvFTD. Neurology
23(7):1685–1694 77(9):866–874
144. Zamboni G et al (2013) Resting functional 159. Agosta F et al (2014) Disrupted brain con-
connectivity reveals residual functional activ- nectome in semantic variant of primary
ity in Alzheimer’s disease. Biol Psychiatry progressive aphasia. Neurobiol Aging
74(5):375–383 35(11):2646–2655
734 Bradford C. Dickerson et al.
160. Gardner RC et al (2013) Intrinsic connectiv- 175. Borroni B et al (2015) Structural and func-
ity network disruption in progressive supra- tional imaging study in dementia with Lewy
nuclear palsy. Ann Neurol 73(5):603–616 bodies and Parkinson’s disease dementia.
161. Williams DR et al (2007) Pathological tau Parkinsonism Relat Disord 21(9):1049–1055
burden and distribution distinguishes pro- 176. Peraza LR et al (2015) Resting state in
gressive supranuclear palsy-parkinsonism Parkinson’s disease dementia and demen-
from Richardson’s syndrome. Brain 130(Pt tia with Lewy bodies: commonalities
6):1566–1576 and differences. Int J Geriatr Psychiatry
162. Whitwell JL et al (2011) Disrupted thalamo- 30(11):1135–1146
cortical connectivity in PSP: a resting-state 177. Peraza LR et al (2014) fMRI resting state
fMRI, DTI, and VBM study. Parkinsonism networks and their association with cognitive
Relat Disord 17(8):599–605 fluctuations in dementia with Lewy bodies.
163. Filippi M et al (2015) Progress towards a neu- Neuroimage Clin 4:558–565
roimaging biomarker for amyotrophic lateral 178. Baggio HC et al (2015) Resting-state fronto-
sclerosis. Lancet Neurol 14(8):786–788 striatal functional connectivity in Parkinson’s
164. Agosta F et al (2013) Divergent brain net- disease-related apathy. Mov Disord
work connectivity in amyotrophic lateral scle- 30(5):671–679
rosis. Neurobiol Aging 34(2):419–427 179. Yao N et al (2015) Resting activity in visual
165. Galvin JE et al (2011) Resting bold fMRI and corticostriatal pathways in Parkinson’s
differentiates dementia with Lewy bod- disease with hallucinations. Parkinsonism
ies vs Alzheimer disease. Neurology Relat Disord 21(2):131–137
76(21):1797–1803 180. Yao N et al (2014) The default mode net-
166. Kenny ER et al (2012) Functional connectiv- work is disrupted in Parkinson’s disease
ity in cortical regions in dementia with Lewy with visual hallucinations. Hum Brain Mapp
bodies and Alzheimer’s disease. Brain 135(Pt 35(11):5658–5666
2):569–581 181. Dumas EM et al (2013) Reduced functional
167. Kenny ER et al (2013) Subcortical connectivity brain connectivity prior to and after disease
in dementia with Lewy bodies and Alzheimer’s onset in Huntington’s disease. Neuroimage
disease. Br J Psychiatry 203(3):209–214 Clin 2:377–384
168. Lowther ER et al (2014) Lewy body 182. Poudel GR et al (2014) Abnormal synchrony
compared with Alzheimer dementia is of resting state networks in premanifest
associated with decreased functional connec- and symptomatic Huntington disease: the
tivity in resting state networks. Psychiatry Res IMAGE-HD study. J Psychiatry Neurosci
223(3):192–201 39(2):87–96
169. Seibert TM et al (2012) Interregional cor- 183. Werner CJ et al (2014) Altered resting-state
relations in Parkinson disease and Parkinson- connectivity in Huntington’s disease. Hum
related dementia with resting functional MR Brain Mapp 35(6):2582–2593
imaging. Radiology 263(1):226–234 184. Quarantelli M et al (2013) Default-mode
170. Rektorova I et al (2012) Default mode network network changes in Huntington’s disease:
and extrastriate visual resting state network in an integrated MRI study of functional con-
patients with Parkinson’s disease dementia. nectivity and morphometry. PLoS One
Neurodegener Dis 10(1–4):232–237 8(8):e72159
171. Chen B et al (2015) Changes in anatomical 185. Seeley WW et al (2009) Neurodegenerative
and functional connectivity of Parkinson’s diseases target large-scale human brain net-
disease patients according to cognitive status. works. Neuron 62(1):42–52
Eur J Radiol 84(7):1318–1324 186. Sanders DW et al (2014) Distinct tau
172. Amboni M et al (2015) Resting-state func- prion strains propagate in cells and mice
tional connectivity associated with mild cog- and define different tauopathies. Neuron
nitive impairment in Parkinson’s disease. 82(6):1271–1288
J Neurol 262(2):425–434 187. Filippi M et al (2013) Assessment of sys-
173. Possin KL et al (2013) Rivastigmine is asso- tem dysfunction in the brain through
ciated with restoration of left frontal brain MRI-based connectomics. Lancet Neurol
activity in Parkinson’s disease. Mov Disord 12(12):1189–1199
28(10):1384–1390 188. Sperling R (2011) Potential of functional
174. Baggio HC et al (2015) Cognitive impair- MRI as a biomarker in early Alzheimer’s dis-
ment and resting-state network connectiv- ease. Neurobiol Aging 32(Suppl 1):S37–S43
ity in Parkinson’s disease. Hum Brain Mapp 189. Wierenga CE, Bondi MW (2007) Use of
36(1):199–212 functional magnetic resonance imaging in
fMRI in Neurodegenerative Diseases: From Scientific Insights to Clinical Applications 735
the early identification of Alzheimer’s disease. 203. Davis TL et al (1998) Calibrated functional
Neuropsychol Rev 17(2):127–143 MRI: mapping the dynamics of oxida-
190. Backman L et al (1999) Brain regions asso- tive metabolism. Proc Natl Acad Sci U S A
ciated with episodic retrieval in normal 95(4):1834–1839
aging and Alzheimer’s disease. Neurology 204. Gur RC et al (1988) Effects of task diffi-
52(9):1861–1870 culty on regional cerebral blood flow: rela-
191. Becker JT et al (1996) Compensatory real- tionships with anxiety and performance.
location of brain resources supporting ver- Psychophysiology 25(4):392–399
bal episodic memory in Alzheimer’s disease. 205. Grasby PM et al (1994) A graded task
Neurology 46(3):692–700 approach to the functional mapping of brain
192. Stern Y et al (2000) Different brain networks areas implicated in auditory-verbal memory.
mediate task performance in normal aging Brain 117(Pt 6):1271–1282
and AD: defining compensation. Neurology 206. Grady CL (1996) Age-related changes in cor-
55(9):1291–1297 tical blood flow activation during perception
193. Putcha D et al (2011) Hippocampal hyper- and memory. Ann N Y Acad Sci 777:14–21
activation associated with cortical thinning 207. Rypma B, D’Esposito M (1999) The roles
in Alzheimer’s disease signature regions in of prefrontal brain regions in components of
non-demented elderly adults. J Neurosci working memory: effects of memory load and
31(48):17680–17688 individual differences. Proc Natl Acad Sci U S
194. Miller SL et al (2008) Hippocampal activation A 96(11):6558–6563
in adults with mild cognitive impairment pre- 208. Kirchhoff BA, Buckner RL (2006) Functional-
dicts subsequent cognitive decline. J Neurol anatomic correlates of individual differences
Neurosurg Psychiatry 79(6):630–635 in memory. Neuron 51(2):263–274
195. DeKosky ST et al (2002) Upregulation of 209. Haslinger B et al (2001) Event-related
choline acetyltransferase activity in hippo- functional magnetic resonance imaging in
campus and frontal cortex of elderly subjects Parkinson’s disease before and after levodopa.
with mild cognitive impairment. Ann Neurol Brain 124(Pt 3):558–570
51(2):145–155 210. Monchi O et al (2004) Neural bases of
196. Hashimoto M, Masliah E (2003) Cycles set-shifting deficits in Parkinson’s disease.
of aberrant synaptic sprouting and neuro- J Neurosci 24(3):702–710
degeneration in Alzheimer’s and demen- 211. Helmich RC et al (2007) Cerebral compensa-
tia with Lewy bodies. Neurochem Res tion during motor imagery in Parkinson’s dis-
28(11):1743–1756 ease. Neuropsychologia 45(10):2201–2215
197. Stern EA et al (2004) Cortical synaptic inte- 212. Georgiou-Karistianis N et al (2007) Increased
gration in vivo is disrupted by amyloid-beta cortical recruitment in Huntington’s dis-
plaques. J Neurosci 24(19):4535–4540 ease using a Simon task. Neuropsychologia
198. Palop JJ, Chin J, Mucke L (2006) A network 45(8):1791–1800
dysfunction perspective on neurodegenerative 213. Stanton BR et al (2007) Altered cortical acti-
diseases. Nature 443(7113):768–773 vation during a motor task in ALS: evidence
199. Bakker A et al (2012) Reduction of hippo- for involvement of central pathways. J Neurol
campal hyperactivity improves cognition in 254(9):1260–1267
amnestic mild cognitive impairment. Neuron 214. Schoenfeld MA et al (2005) Functional motor
74(3):467–474 compensation in amyotrophic lateral sclerosis.
200. Mueggler T et al (2002) Compromised J Neurol 252(8):944–952
hemodynamic response in amyloid precur- 215. Konrad C et al (2006) Subcortical reorgani-
sor protein transgenic mice. J Neurosci zation in amyotrophic lateral sclerosis. Exp
22(16):7218–7224 Brain Res 172(3):361–369
201. El Fakhri G et al (2003) MRI-guided SPECT 216. Drummond SP et al (2000) Altered brain
perfusion measures and volumetric MRI in response to verbal learning following sleep
prodromal Alzheimer disease. Arch Neurol deprivation. Nature 403(6770):655–657
60(8):1066–1072 217. Cabeza R et al (2002) Aging gracefully: com-
202. Cohen ER, Ugurbil K, Kim SG (2002) Effect pensatory brain activity in high-performing
of basal conditions on the magnitude and older adults. Neuroimage 17(3):1394–1402
dynamics of the blood oxygenation level- 218. Carey JR et al (2002) Analysis of fMRI and
dependent fMRI response. J Cereb Blood finger tracking training in subjects with
Flow Metab 22(9):1042–1053 chronic stroke. Brain 125(Pt 4):773–788
736 Bradford C. Dickerson et al.
familial frontotemporal dementia. Neurology 263. Thiel CM, Henson RN, Dolan RJ (2002)
80(9):814–823 Scopolamine but not lorazepam modulates
249. Borroni B et al (2012) Granulin mutation face repetition priming: a psychopharmacolog-
drives brain damage and reorganization from ical fMRI study. Neuropsychopharmacology
preclinical to symptomatic FTLD. Neurobiol 27(2):282–292
Aging 33(10):2506–2520 264. Leslie RA, James MF (2000) Pharmacological
250. Premi E et al (2014) Effect of TMEM106B magnetic resonance imaging: a new applica-
polymorphism on functional network con- tion for functional MRI. Trends Pharmacol
nectivity in asymptomatic GRN mutation car- Sci 21(8):314–318
riers. JAMA Neurol 71(2):216–221 265. Rombouts SA et al (2002) Alterations in
251. Lee SE et al (2014) Altered network con- brain activation during cholinergic enhance-
nectivity in frontotemporal dementia with ment with rivastigmine in Alzheimer’s
C9orf72 hexanucleotide repeat expansion. disease. J Neurol Neurosurg Psychiatry
Brain 137(Pt 11):3047–3060 73(6):665–671
252. Reading SA et al (2004) Functional brain 266. Saykin AJ et al (2004) Cholinergic enhance-
changes in presymptomatic Huntington’s dis- ment of frontal lobe activity in mild cognitive
ease. Ann Neurol 55(6):879–883 impairment. Brain 127(Pt 7):1574–1583
253. Kloppel S et al (2010) Irritability in pre-clinical 267. Goekoop R et al (2006) Cholinergic chal-
Huntington’s disease. Neuropsychologia lenge in Alzheimer patients and mild cognitive
48(2):549–557 impairment differentially affects hippocampal
254. Van den Stock J et al (2015) Functional brain activation—a pharmacological fMRI study.
changes underlying irritability in premani- Brain 129(Pt 1):141–157
fest Huntington’s disease. Hum Brain Mapp 268. Miettinen PS et al (2011) Effect of cho-
36(7):2681–2690 linergic stimulation in early Alzheimer’s
255. Wolf RC et al (2007) Dorsolateral prefron- disease—functional imaging during a rec-
tal cortex dysfunction in presymptomatic ognition memory task. Curr Alzheimer Res
Huntington’s disease: evidence from event- 8(7):753–764
related fMRI. Brain 130(Pt 11):2845–2857 269. Bentley P, Driver J, Dolan RJ (2008)
256. Wolf RC et al (2011) Longitudinal functional Cholinesterase inhibition modulates visual
magnetic resonance imaging of cognition in and attentional brain responses in Alzheimer’s
preclinical Huntington’s disease. Exp Neurol disease and health. Brain 131(Pt 2):409–424
231(2):214–222 270. Shanks MF et al (2007) Regional brain activ-
257. Wolf RC et al (2008) Altered frontostriatal ity after prolonged cholinergic enhancement
coupling in pre-manifest Huntington’s dis- in early Alzheimer’s disease. Magn Reson
ease: effects of increasing cognitive load. Eur Imaging 25(6):848–859
J Neurol 15(11):1180–1190 271. Kircher TT et al (2005) Cortical activation
258. Georgiou-Karistianis N et al (2013) during cholinesterase-inhibitor treatment
Functional and connectivity changes dur- in Alzheimer disease: preliminary findings
ing working memory in Huntington’s dis- from a pharmaco-fMRI study. Am J Geriatr
ease: 18 month longitudinal data from the Psychiatry 13(11):1006–1013
IMAGE-HD study. Brain Cogn 83(1):80–91 272. Thiyagesh SN et al (2010) Treatment effects
259. Wolf RC et al (2012) Default-mode network of therapeutic cholinesterase inhibitors on
changes in preclinical Huntington’s disease. visuospatial processing in Alzheimer’s disease:
Exp Neurol 237(1):191–198 a longitudinal functional MRI study. Dement
Geriatr Cogn Disord 29(2):176–188
260. Odish OF et al (2015) Longitudinal resting
state fMRI analysis in healthy controls and 273. McGeown WJ, Shanks MF, Venneri A (2008)
premanifest Huntington’s disease gene carri- Prolonged cholinergic enrichment influ-
ers: a three-year follow-up study. Hum Brain ences regional cortical activation in early
Mapp 36(1):110–119 Alzheimer’s disease. Neuropsychiatr Dis
Treat 4(2):465–476
261. Hampel H et al (2014) Perspective on
future role of biological markers in clinical 274. Bokde AL et al (2009) Decreased activa-
therapy trials of Alzheimer’s disease: a long- tion along the dorsal visual pathway after a
range point of view beyond 2020. Biochem 3-month treatment with galantamine in mild
Pharmacol 88(4):426–449 Alzheimer disease: a functional magnetic reso-
nance imaging study. J Clin Psychopharmacol
262. Sperling R et al (2002) Functional MRI 29(2):147–156
detection of pharmacologically induced
memory impairment. Proc Natl Acad Sci U S 275. McLaren DG et al (2012) Tracking cogni-
A 99(1):455–460 tive change over 24 weeks with longitudi-
738 Bradford C. Dickerson et al.
nal functional magnetic resonance imaging 288. Sole-Padulles C et al (2013) Donepezil treat-
in Alzheimer’s disease. Neurodegener Dis ment stabilizes functional connectivity during
9(4):176–186 resting state and brain activity during mem-
276. Venneri A, McGeown WJ, Shanks MF (2009) ory encoding in Alzheimer’s disease. J Clin
Responders to ChEI treatment of Alzheimer’s Psychopharmacol 33(2):199–205
disease show restitution of normal regional 289. Wang L et al (2014) The effect of APOE
cortical activation. Curr Alzheimer Res epsilon4 allele on cholinesterase inhibitors
6(2):97–111 in patients with Alzheimer disease: evalua-
277. Dhanjal NS et al (2013) Auditory corti- tion of the feasibility of resting state func-
cal function during verbal episodic memory tional connectivity magnetic resonance
encoding in Alzheimer’s disease. Ann Neurol imaging. Alzheimer Dis Assoc Disord
73(2):294–302 28(2):122–127
278. Dhanjal NS, Wise RJ (2014) Frontoparietal 290. Wang Z et al (2014) Acupuncture modulates
cognitive control of verbal memory resting state hippocampal functional con-
recall in Alzheimer’s disease. Ann Neurol nectivity in Alzheimer disease. PLoS One
76(2):241–251 9(3):e91160
279. Goekoop R et al (2004) Challenging the cho- 291. Jia B et al (2015) The effects of acupuncture
linergic system in mild cognitive impairment: at real or sham acupoints on the intrinsic
a pharmacological fMRI study. Neuroimage brain activity in mild cognitive impairment
23(4):1450–1459 patients. Evid Based Complement Alternat
280. Gron G et al (2006) Inhibition of hippo- Med 2015:529675
campal function in mild cognitive impair- 292. Wells RE et al (2013) Meditation’s impact
ment: targeting the cholinergic hypothesis. on default mode network and hippocampus
Neurobiol Aging 27(1):78–87 in mild cognitive impairment: a pilot study.
281. Risacher SL et al (2013) Cholinergic enhance- Neurosci Lett 556:15–19
ment of brain activation in mild cognitive 293. van Paasschen J et al (2013) Cognitive reha-
impairment during episodic memory encod- bilitation changes memory-related brain
ing. Front Psychiatry 4:105 activity in people with Alzheimer disease.
282. Petrella JR et al (2009) Effects of donepezil Neurorehabil Neural Repair 27(5):448–459
on cortical activation in mild cognitive impair- 294. Hampstead BM et al (2012) Mnemonic strat-
ment: a pilot double-blind placebo-controlled egy training partially restores hippocampal
trial using functional MR imaging. AJNR Am activity in patients with mild cognitive impair-
J Neuroradiol 30(2):411–416 ment. Hippocampus 22(8):1652–1658
283. Pa J et al (2013) Cholinergic enhancement 295. Buhmann C et al (2003) Pharmacologically
of functional networks in older adults with modulated fMRI—cortical responsiveness
mild cognitive impairment. Ann Neurol to levodopa in drug-naive hemiparkinsonian
73(6):762–773 patients. Brain 126(Pt 2):451–461
284. Goveas JS et al (2011) Recovery of hippo- 296. Kwak Y et al (2010) Altered resting state
campal network connectivity correlates with cortico-striatal connectivity in mild to mod-
cognitive improvement in mild Alzheimer’s erate stage Parkinson’s disease. Front Syst
disease patients treated with donepezil Neurosci 4:143
assessed by resting-state fMRI. J Magn Reson 297. Wu T et al (2009) Regional homogeneity
Imaging 34(4):764–773 changes in patients with Parkinson’s disease.
285. Li W et al (2012) Changes in regional cere- Hum Brain Mapp 30(5):1502–1510
bral blood flow and functional connectivity 298. Agosta F et al (2014) Cortico-striatal-
in the cholinergic pathway associated with thalamic network functional connectivity
cognitive performance in subjects with mild in hemiparkinsonism. Neurobiol Aging
Alzheimer’s disease after 12-week donepezil 35(11):2592–2602
treatment. Neuroimage 60(2):1083–1091 299. Choe IH et al (2013) Decreased and increased
286. Zaidel L et al (2012) Donepezil effects cerebral regional homogeneity in early
on hippocampal and prefrontal functional Parkinson’s disease. Brain Res 1527:230–237
connectivity in Alzheimer’s disease: pre- 300. Esposito F et al (2013) Rhythm-specific
liminary report. J Alzheimers Dis 31(Suppl modulation of the sensorimotor network in
3):S221–S226 drug-naive patients with Parkinson’s disease
287. Lorenzi M et al (2012) Effect of meman- by levodopa. Brain 136(Pt 3):710–725
tine on resting state default mode network 301. Mattay VS et al (2002) Dopaminergic modu-
activity in Alzheimer’s disease. Drugs Aging lation of cortical function in patients with
28(3):205–217
fMRI in Neurodegenerative Diseases: From Scientific Insights to Clinical Applications 739
Parkinson’s disease. Ann Neurol 51(2): movements in Parkinson’s disease. PLoS One
156–164 7(12):e50270
302. Tessitore A et al (2002) Dopamine modulates 306. Kahan J et al (2014) Resting state functional
the response of the human amygdala: a study MRI in Parkinson’s disease: the impact of
in Parkinson’s disease. J Neurosci deep brain stimulation on ‘effective’ connec-
22(20):9099–9103 tivity. Brain 137(Pt 4):1130–1144
303. Arantes PR et al (2006) Performing functional 307. Scheff SW et al (2006) Hippocampal synaptic
magnetic resonance imaging in patients with loss in early Alzheimer’s disease and mild cog-
Parkinson’s disease treated with deep brain nitive impairment. Neurobiol Aging
stimulation. Mov Disord 21(8):1154–1162 27(10):1372–1384
304. Phillips MD et al (2006) Parkinson disease: 308. Jack CR Jr et al (2013) Tracking pathophysi-
pattern of functional MR imaging activation ological processes in Alzheimer’s disease: an
during deep brain stimulation of subthalamic updated hypothetical model of dynamic bio-
nucleus—initial experience. Radiology 239(1): markers. Lancet Neurol 12(2):207–216
209–216 309. Bobholz JA et al (2007) Clinical use of func-
305. Kahan J et al (2012) Therapeutic subthalamic tional magnetic resonance imaging: reflec-
nucleus deep brain stimulation reverses tions on the new CPT codes. Neuropsychol
cortico-thalamic coupling during voluntary Rev 17(2):189–191
Chapter 24
fMRI in Epilepsy
Rachel C. Thornton, Louis André van Graan, Robert H. Powell,
and Louis Lemieux
Abstract
This chapter provides an overview of the application of functional MRI applied to the field of Epilepsy and
is divided into two sections, covering cognitive mapping and imaging of paroxysmal activity, respectively.
In addition to a review of the most scientifically and clinically relevant findings, technical and methodologi-
cal background information is provided to help the reader better understand the data acquisition process.
We show how both approaches may play a role in the presurgical evaluation of patients with drug-resistant
focal epilepsy and provide opportunities for new insights into the neuropathological processes that under-
lie both focal and generalized epilepsy.
Key words Epilepsy, Focal epilepsy, Generalized epilepsy, Interictal, Ictal, Imaging, Functional mag-
netic resonance imaging, fMRI, Electroencephalography, EEG, Multi-modal imaging, EEG-
correlated fMRI, Cognitive mapping, Functional mapping, Brain activity mapping, Language
lateralization, Memory mapping, Presurgical evaluation
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_24, © Springer Science+Business Media New York 2016
741
742 Rachel C. Thornton et al.
1.1 Language fMRI The aims of preoperative language fMRI are primarily to lateralize
and localize language functions and to use this information to pre-
1.1.1 Paradigm Design
dict and avert postoperative complications. A number of task para-
and Analysis
digms to engage anterior/expressive as well as posterior/
comprehension language areas [9–14] have been used to identify
language representation. Complementary indices provided by ver-
bal fluency, verb generation and semantic decision tasks, are vari-
ously and commonly employed in the clinical context [15]. The
most widely used tasks in language fMRI experiments are verbal
fluency tasks. These are generally strongly lateralizing and reliably
identify “expressive language functions” in the dominant inferior
frontal gyrus (IFG) (Brodmann Areas [BA], 44, 45). Specifically,
verbal fluency tasks show more prominent activity in left frontal
regions, corresponding to Broca’s area, than in the medial tempo-
ral lobe in healthy controls and TLE patients [16, 17]. Although
these tasks are usually covert (i.e. performed silently without per-
formance monitoring), they have been reliably replicated in numer-
ous studies in both normal and patient populations. Their
within-subject reproducibility has been demonstrated, with frontal
activations shown to be more reliable than temporoparietal ones
[18]. In addition they can be applied to patients with a wide range
of cognitive abilities, with language lateralization results appearing
to be relatively unaffected by patients’ performance levels [19].
744 Rachel C. Thornton et al.
Fig. 1 Verbal fluency and reading comprehension: Typical fMRI findings in a patient performing tasks of verbal
fluency (left) and reading comprehension (right) showing activation in the dominant frontal lobe and bilateral
superior temporal lobes, respectively. Areas of activation are overlaid on a distortion matched high resolution
echo-planar image
fMRI in Epilepsy 745
1.1.2 Language Focal epilepsy may be associated with disrupted lateralization and
Lateralization in Epilepsy localization of language regions; therefore, one would expect a
higher probability of abnormal language lateralization. Nevertheless,
significant differences have been reported between centers in the
relative proportions of right and left hemisphere dominant patients
using the IAT, some of which may be due to the different criteria
used for assessing dominance. The percentage of left hemisphere
dominant right-handed patients has ranged from 63 to 96 % [41]
while for left-handers a similar variation has been reported between
38 and 70 % [42, 43]. Results of fMRI studies have also shown
greater atypical language dominance in patients. In a comparison
between 100 right-handed healthy subjects and 50 right-handed
epilepsy patients, 94 % of the normal subjects were considered as left
hemisphere dominant and 6 % had bilateral representation. The epi-
lepsy group showed greater variability of language dominance, with
78 % showing left hemisphere dominance, 16 % symmetric activa-
tion, and 6 % showing right hemisphere dominance. Atypical lan-
guage dominance was associated with an earlier age of brain injury
and with weaker right hand dominance [44].
The localization of the epileptogenic lesion and epileptic
activity [45] has also been shown to influence language organiza-
tion. In a retrospective study of patients with hippocampal sclero-
sis (HS) who had undergone presurgical evaluation, atypical
speech dominance occurred in 24 % of those with left-sided HS,
whereas all those with right-sided HS had left-sided speech domi-
nance. In addition, atypical speech representation was associated
with higher spiking frequency and in those with sensory auras
suggesting ictal involvement of the lateral temporal structures.
No association was demonstrated between either age at epilepsy
onset or age at initial precipitating injury and atypical speech rep-
resentation [46].
Comparing the degree of reorganization of frontal and tempo-
ral lobe language functions has shown a significantly more left
lateralized pattern of language activation in controls and right TLE
fMRI in Epilepsy 747
1.1.3 Comparison Studies comparing fMRI and the IAT are summarized in Table 1.
of fMRI, IAT, Just as IAT protocols differ between centers, a number of fMRI
and Electrocortical paradigms to determine language dominance have been employed
Stimulation Findings but agreement of approximately 80–90 % is seen between the two
techniques [11, 58]. The remaining cases generally exhibit partial
disparity where one method shows bilateral language representa-
tion and the other lateralized language dominance and outright
disagreement between fMRI and IAT is rare. In an interesting
study that assessed the relative accuracy of Wada and fMRI in dis-
cordant cases fMRI provided a more accurate prediction of naming
at postsurgical outcome in seven patients, Wada was more accurate
in two patients. The two methods provided comparable accuracy
in one patient [58].
One study suggested that fMRI may be less reliable in left-sided
neocortical epilepsy (25 % disparity) in comparison with left-sided
medial TLE (3 % disparity) [36]. Another showed that concordance
between fMRI-based laterality and IAT was much lower in left TLE
patients than in patients with right TLE [59]. One interesting case
of false lateralization of language function in a post-ictal patient
with left HS also illustrates the need for caution in the interpreta-
tion of results in individual patients. No activation was seen in the
left temporal lobe during multiple language tasks after a cluster of
left temporal lobe seizures but in a repeat fMRI experiment 2 weeks
later, activation was seen predominantly over the left temporal
region [60]. However, bearing in mind the previously mentioned
limitations of the IAT, it is even debatable whether fMRI and IAT
are directly comparable as they probe different aspects of language.
fMRI language localization can replace Wada test in the majority of
patients. However, the Wada test is still a valuable adjunct and can
be employed when a patient cannot undergo fMRI. It can also be
used for validation of fMRI results or for the assessment of selective
language areas near structural abnormalities [61].
Comparisons have also been performed between fMRI activa-
tion maps and regions showing disruption of function during
intraoperative electrocortical stimulation (ECS). In order for fMRI
to be used instead of ECS, it must demonstrate a high predictive
power for the presence as well as the absence of critical language
function in regions of the brain. As with IAT, these studies show
strong, but incomplete agreement with fMRI, with high sensitivity
but lower specificity [69–71]. Although false-positive activation
(fMRI activation but no ECS disruption) is relatively common,
this is not surprising given that fMRI activates whole networks of
regions, not all of which are essential for the task in question. False-
negative findings (regions showing disruption by ECS but no
fMRI activation) are more critical when planning a surgical resec-
tion, and these were identified in 2 patients out of 21 reported in
two series. Activation and disruption was typically within 5 mm in
frontal regions and 10 mm in temporal areas.
fMRI in Epilepsy 749
Table 1
Concordance between fMRI language lateralization and the IAT
Sample
Authors sizea fMRI language lateralization tasks Concordance
Desmond et al. 7 Semantic decision task 100 %
(1995) [62]
Binder et al. 22 Semantic decision task r = 0.96
(1996) [63]
Hertz-Pannier 6b Verbal fluency paradigm 100 %
et al. (1997)
[64]
Yetkin et al. 13 Word generation task r = 0.93
(1998) [65]
Benson et al. 12 Verb generation task 100 %
(1999) [66]
Lehericy et al. 10 Semantic fluency Semantic fluency > story
(2000) [27] listening > sentence
Sentence repetition Story listening
repetition
Greater concordance between
IAT results and activation
asymmetry in frontal than
temporal lobes
Carpentier et al. 10 Identification of syntactic/semantic errors 80 %
(2001) [67] in target sentences
Gaillard et al. 21 Reading paradigm 85 %
(2002) [29]
Woermann 100 Word generation 91 %
et al. (2003)
[36]
Sabbah et al. 20c Word generation 95 %
(2003) [68]
Semantic decision
Benke et al. 68 Semantic decision 89 %—right TLE
(2006) [59] 40 Reading sentence comprehension/
72 %—left TLE
Arora et al. 229 auditory sentence comprehension and a
91.3 %
(2009) [11] verbal fluency task.
86 %
Janecek et al. Semantic decision/tone decision
(2013) [58]
IAT intracarotid amytal test, TLE temporal lobe epilepsy
a
Some of these studies report fMRI data on larger samples. However, only the patients with fMRI and IAT data are
included here
b
Age range 8–18
c
Patients with suspected atypical language lateralization were selected
750 Rachel C. Thornton et al.
Fig. 2 Combined MR tractography and functional mapping. Frontal lobe connections overlaid on a structural
template along with group fMRI effects for word generation (solid arrows) and reading comprehension (dashed
arrows), showing how the tracts connect together the frontal and temporal lobe functionally active regions
1.2 Memory fMRI A range of memory functions are commonly affected in epilepsy
including modality specific processes in working and long term
memory. fMRI can reliably localize and assess the impact of surgery
on memory networks [83, 84]. In addition LIs have been estab-
lished to evaluate the effects of epilepsy and surgical intervention on
memory [85]. MTL structures are associated with memory func-
tioning, and surgical resection is known to cause reduced memory
function in some cases. The study of patients following temporal
lobe surgery has provided considerable evidence supporting the
fMRI in Epilepsy 753
patients with TLE. Conversely this study showed that longer dura-
tion and higher seizure frequency were associated with greater
inefficient, extra-temporal reorganization. The severity of HS on
MRI is an important determinant, being inversely correlated with
a decline in verbal memory following left ATLR, with less severe
HS increasing the risk of memory decline [98, 103, 104].
Specifically, the extent of verbal memory decline after left ATLR is
correlated with greater BOLD activation of the diseased left hip-
pocampus and its connectivity to ipsilateral posterior cingulate
[105]. Preoperative memory performance has been related to
degree of postoperative memory impairment, with better perfor-
mance increasing the risk of memory decline [95, 106, 107]. These
risk factors reflect the functional integrity of the resected temporal
lobe and suggest that patients with residual memory function in
the pathological hippocampus are at greater risk of memory impair-
ment postoperatively. Recently, fMRI has also been shown to be a
potential predictor of postoperative material-specific memory
decline following ATLR. Comparison of pre- and postoperative
fMRI activation and correlation with better verbal memory out-
come after left ATLR indicate preoperative reorganization of ver-
bal memory function to the ipsilateral posterior medial temporal
lobe [85]. Other results indicate that visual and verbal memory
function following ATLR is correlated to activity of the contralat-
eral medial temporal lobe and its connectivity to the posterior cin-
gulate cortex ipsilateral to the damaged hippocampus [105, 108].
1.2.2 The Effect of TLE Deficits in verbal memory following left ATLR and topographical
on Memory Processes memory following right ATLR suggest a material-specific lateral-
ization of function in MTL structures. Functional imaging studies
have been used to look for lateralization of cerebral activation pat-
terns during episodic memory processes. Many have shown
material-specific lateralization in prefrontal regions but this has
been more difficult to demonstrate in the MTL [115, 119, 121,
122]. Working memory can be affected in TLE patients with HS,
with indication of altered connectivity between regions [123].
Specifically, a disruption of the regional balance between task-pos-
itive and task-negative functional networks is associated with work-
ing memory dysfunction in TLE [124]. Reduced right superior
parietal lobe activity is associated with suppression of activity in the
healthy hippocampus in the context of an increasing WM load with
fMRI in Epilepsy 757
Table 2
fMRI memory studies in TLE
0.4
0.3
fMRI activation
0.2
Left
0.1
Right
0
−0.1
−0.2
words pictures faces
Material type
C P
0.4
0.3
0.2
0.1
−0.1
−0.2
−0.3
−0.4
0.5
0.4
0.3
0.2
0.1
−0.1
−0.2
−0.3
−0.4
−0.5
L R
Fig. 5 fMRI memory encoding experiment: Left TLE patients vs. healthy controls. Regions showing significant
differences in activation between left temporal lobe epilepsy (TLE) patients and controls are highlighted.
Contrast estimates are shown on the right of the images. Controls (C) are on the left and patients (P) on the
right. A reorganization of function is seen in the left TLE patients with reduced activation in the left hippocam-
pus, and greater activation in the right hippocampus, compared with healthy controls
760 Rachel C. Thornton et al.
1.3 Challenges When designing paradigms for patients with neurological deficits, it
of Clinical Cognitive is important to use tasks that they are able to perform. A differential
fMRI pattern of activation between patients and normal subjects is only
interpretable if patients are performing the task adequately [135].
In addition, one must be aware of differences in the questions
being asked by cognitive neuroscientists and clinicians, which can
lead to different approaches to data analysis. Generally, neuroscien-
tists look at groups of matched controls performing the same task
and determine which brain regions are commonly activated across
the group. The emphasis is on avoiding false-positive results (Type
I errors) and conservative statistical thresholds need to be used,
which may lead to an under representation of brain areas truly
involved. Conversely, clinicians are considering individual patients
where the priority is to identify all brain regions involved in a task,
i.e., avoiding false negatives (Type II errors). As a result, less strin-
gent statistical thresholds are required and indeed thresholds used
may need to vary on an individual basis.
fMRI in Epilepsy 761
1.4 Cognitive fMRI fMRI is a noninvasive and widely available tool, which has had a
in Epilepsy: Summary dramatic impact on cognitive neuroscience. Much of the progress
made will benefit clinical neuroimaging, although some problems
exist in the application of fMRI to patients with neurological defi-
cits. fMRI allows the noninvasive assessment of language function
to be performed and offers a valid alternative to the IAT for estab-
lishing language dominance. By tailoring paradigms towards the
localization of the specific language skills most at risk following
temporal and frontal resections, it will be possible to map relevant
language functions in the epilepsy surgery population. This in turn
will allow better assessment of the risks posed by surgery in each
individual patient.
Considerable effort is also being made in the development of
memory paradigms that can lateralize MTL functions and provide
meaningful data at the single subject level. This information, in
combination with structural MRI to evaluate hippocampal pathol-
ogy and baseline neuropsychology, will enable preoperative predic-
tion of likely material specific memory impairments seen following
unilateral ATLR to be made with greater accuracy. In consequence,
it will be possible to modify surgical approaches in those patients
most at risk and to improve preoperative patient counseling.
Clinically it is what happens to individual patients that is
important and the next step in the validation of these techniques
will involve similar studies with larger numbers of patients. These
should include more heterogeneous samples, including both left
and right TLE undergoing ATLR. As well as showing group level
correlations either at the voxel-level or within a predefined ROI, it
will be important to establish methods for using this data to pre-
dict language and memory changes in individual cases. Investigating
how the brain sustains memory postoperatively also requires fur-
ther investigation. Longitudinal fMRI studies with pre- and post-
operative imaging, including correlations with neuropsychological
measures of language and memory, will be required to look at
functional reorganization following surgery, and it is anticipated
that these will offer valuable insights into brain plasticity.
2.1 EEG-Correlated The recording of EEG inside the MR scanner still presents safety,
fMRI in Epilepsy: image data quality and EEG data quality challenges. Historically, the
Technical Issues issue of EEG data quality has been the determining factor in the
technique’s evolution, from interleaved to simultaneous EEG-
fMRI. This reflects in part the fact that a gradual degradation in
EEG quality mainly linked to cardiac activity can be readily observed
in most subjects as they are moved inside the MR scanner (without
scanning), posing an immediate challenge ahead of any other con-
siderations such as safety (albeit this should also be at the forefront
of the considerations of investigators introducing any new equip-
ment in the scanner room) or the effect of scanning on EEG quality
and the possible impact of the EEG recording equipment on image
quality. In the following, we provide an overview of the state of
EEG-fMRI technology, which remains an active area of research in
particular in the area of EEG quality, although mostly for the pur-
pose of evoked response recordings. The focus will be on the impli-
cations for studies in epilepsy and in particular at field strengths
commonly used in neurological studies (≤3.0 T); the reader inter-
ested in the implementation of combined EEG and fMRI record-
ings at higher field strengths (e.g. 7 T) is directed towards two recent
specialized reports that address data quality and safety [143, 144].
The electro-magnetic processes that take place during MR image
2.1.1 Physical Principles acquisition and that are susceptible to interactions with the EEG
of EEG-MR System system are: strong static magnetic field (∼1.0–3.0 T), switching
Interactions magnetic gradient fields (∼100 T/m/s), and radio frequency
(RF) pulses (∼10 μT and 100 MHz). In addition, although MR
scanners are designed to optimize the magnetic component of the
RF pulses, an electrical component is unavoidable. This may lead
to linear antenna effects with possible safety implications [145].
EEG recording, on the other hand, requires electrodes and leads
to be placed within the imaging field of view and electronic
components, depending on the exact equipment and setup, in
proximity to the scanner coils and antenna(s).
Four main mechanisms are at the origin of EEG-MR instru-
mentation interactions:
1. Magnetic induction: any change in magnetic flux (essentially
the component of the magnetic field that is perpendicular to a
surface) over time through a conducting medium (loop, sur-
face, volume) gives rise to an electromotive force in the mate-
764 Rachel C. Thornton et al.
2.1.2 Safety Health hazards not normally encountered when MR or EEG are
performed separately can arise due to induced currents flowing
through loops or the heating of EEG components in proximity or
contact with the subject. For a specific 1.5 T scanner, and based on
a worst case scenario, this study recommended that one 10 kΩ
current-limiting resistor be inserted serially at each electrode lead
and the possibility of large (EEG lead-electrode-head-electrode-
lead-amplifier circuit) loops being formed reduced to a minimum
by lead twisting. In experiments using a different custom-made
EEG system, no significant heating was observed [146]. An impor-
tant general consideration when placing wires in contact with the
body is the type of RF transmit coil used and length of wire exposed
to the electrical component of the RF field [147]. A number of
MR-compatible EEG system or electrode cap vendors have incor-
porated current-limiting resistors in their product design. To the
authors’ knowledge, no adverse incident linked specifically to
EEG-fMRI data acquisition has been formally reported to date.
2.1.3 Image Quality Image quality remains an important issue throughout the field of
MRI and the subject of investigation, particularly for echo-planar
imaging (EPI), which is particularly prone to distortion and local
signal dropout [148]. Artifacts caused by electrodes and leads were
observed in early EEG-fMRI experiments [138]. Therefore, one
must consider carefully the choice of materials and components
placed within the field-of-view [146, 149–152]. It has been shown
that the presence of high-density (256 channels) EEG caps can
significantly impair structural MR imaging [153].
fMRI in Epilepsy 765
2.1.4 EEG Quality In the literature it is common to categorize the artifacts observed
on EEG recorded inside the MR scanner into two types: heart
beat-related (whether scanning is taking place or not) and MR
image acquisition-related. These can be considered distinct in
terms of their generating mechanism, and deserve to be addressed
separately in terms of remedies to minimize them, as reflected in
the structure of this section. However, in practice they are linked
by a third phenomenon, widely recognized as a nuisance in fMRI,
namely subject motion. This is in part because the heart beat-related
artifact is thought to mostly originate from the body motion linked
to the heart beat, but also because subject (and EEG electrode and
lead) motion can have an important impact on the ability to cor-
rect both types of artifact, depending on the approach taken.
Therefore body motion is nefarious for EEG recording quality
(and almost without saying, fMR image quality) and should be
minimized. This will be the subject of Sect. 2.1.4.3.
EEG Quality: Pulse The first attempts at recording EEG inside MR scanners revealed
Artifact, Reduction, and the presence of pulse-related artifacts delayed in relation to the
Correction Methods QRS complexes on ECG [138]. This effect has been shown to be
common across subjects and has a slight frontal emphasis [154].
The pulse artifacts can have amplitude of the order 50 μV (at 1.5 T)
and resemble epileptic spikes. Because of natural heart beat variabil-
ity, it is considered a more challenging problem than that of image
acquisition artifacts. EEG artifacts linked to subject movement are
also amplified in the scanner’s strong static magnetic field.
The precise mechanism through which the circulatory system
exposed to a strong magnetic field gives rise to these artifacts
remains uncertain, but it is thought to represent a combination of
the motion of the electrodes and leads (induction) and the Hall
effect (voltage induced by flow of conducting blood in proximity
of electrodes) [155]. Electrode motion can result from local arte-
rial pulsation, brain and head motion or whole-body motion (bal-
listocardiogram, or BCG, in the latter case) [156, 157].
Methods to reduce artifacts at the source include: careful lay-
ing out and immobilization of the leads, twisting of the leads,
bipolar electrode chain arrangement [158], head vacuum cushion
[159] and the introduction of a reference electrode layer insulated
from the EEG-measuring electrodes to capture and subtract the
artifact from the EEG prior to amplification [160]. Such measures
do not eliminate the problem completely resulting in degraded
EEG quality, impeding the identification of epileptiform dis-
charges. The first pulse artifact reduction algorithm published, and
to this day still the gold standard against which most methods are
compared, is based on subtraction of a running average estimate of
the artifact based on automatic QRS detection, and is commonly
referred to as the average artifact subtraction (AAS) method [154].
Using this method, the residual artifact is of the order of a few
microvolts. The reliance of the algorithm on ECG is a common,
though not universal, feature among subsequently developed
766 Rachel C. Thornton et al.
techniques (some of which use the signal from the standard scan-
ner pulse oxymeter). The method has been and continues to be
used successfully in our lab allowing the satisfactory identification
of ictal and interictal epileptiform discharges (IED) in real time (at
1.5 T) [161] and for the purpose of source analysis [162], and has
been implemented in widely used commercial MR-compatible
EEG recording systems (see Fig. 6).
The artifact amplitude is theoretically directly proportional to
the scanner static field strength (B0). This phenomenon, and an
increasing interest in recording evoked potentials in the MR scan-
ner, has motivated an important research effort towards improving
existing pulse-related artifact reduction methods and the develop-
ment of new ones. Variants of the AAS method have been pro-
posed, ranging from different ways of estimating the artifact
waverform, for example to account for a greater degree of inter-
beat variability [158, 166–168], more general motion effects [151,
169, 170], to improving QRS detection [171] and removing the
need for ECG recording [172].
Fig. 6 IED-related BOLD pattern in patient with drug-resistant focal epilepsy. The patient had refractory focal epilepsy,
lateralized to the right with a normal structural MRI. Frequent mid and posterior temporal sharp waves were recorded
on EEG.(a) Representative segment of 32-channel EEG showing a sharp wave, maximum at the right mid-posterior
temporal region, recorded during two 20-min fMRI sessions. Top left: EEG prior to artifact correction [154, 163].
Fig. 6 (continued) (b) Design matrix: BOLD signal changes related to 40 sharp waves were modeled by convolution
of the EEG event onsets with a canonical HRF and its time-derivative. Signal changes linked to head motion and
heartbeat were modeled as nuisance effects [164, 165]. (c) Top: SPM showing significant sharp wave-related BOLD
response in glass brain display (p < 0.05 corrected for multiple comparisons). The red arrow marks the global maxi-
mum, located in the BA 28 (superior temporal gyrus). Bottom: BOLD response overlaid onto the patient’s normalized
T1-weighted volumetric scan. Intracranial recording confirmed a right posterior temporal lobe onset. No significant
sharp wave-related deactivation was revealed. The activation clusters were labeled using the Talairach Daemon,
http://ric.uthscsa.edu/project/talairachdaemon.html
768 Rachel C. Thornton et al.
EEG Quality: Image In the absence of any special measures, the EEG recorded inside
Acquisition Artifact, the MR scanner becomes un-interpretable during image acquisi-
Special EEG Recording tion because of the presence of repetitive artifact waveforms caused
Equipment, and by the time-varying fields employed in the scanning process super-
Correction Methods imposed on the physiological signal [163, 183]. In addition, on
some MR instruments, the helium cooling pump can introduce
significant amount of noise in the signals measured using the EEG
equipment, and a method to remedy this problem has been pro-
posed in cases where the pump cannot be switched off for the
duration of the scan [184].
One way of circumventing this problem is to leave time gaps in
the fMRI acquisition (e.g. between EPI volumes) of sufficient
duration to capture the EEG features of interest (assuming suffi-
cient data quality, e.g., following pulse artifact removal); this is
interleaved EEG-fMRI [158, 185, 186]. This approach relies on
artifact not persisting following each acquisition (e.g. due to ampli-
fier saturation). Interleaved EEG-fMRI can be most useful to study
predictable events (evoked responses) or slowly varying phenom-
ena, such as brain rhythms. EEG-triggered fMRI and in particular
spike-triggered fMRI, which involves limiting fMRI acquisition to
single or multi-volume blocks, each triggered following the
identification of an EEG event of interest is a form of interleaved
EEG-fMRI with obvious relevance to epilepsy [183, 187–190].
Although interleaved EEG-fMRI is capable of providing useful
data in many circumstances, it imposes a limit on experimental effi-
ciency due to EEG quality degradation during scanning.
We now review the technical developments that have made it
possible to record EEG of sufficient quality throughout fMRI
acquisition (so-called continuous EEG-fMRI), by the image acquisi-
tion artifact to be corrected. For all practical purposes, and assum-
ing that the time gap between volumes is the same as between slices,
the artifact’s spectral signature ranges from 1/TR (TR: slice acqui-
sition repetition time) to around 1 kHz (corresponding to the read-
out gradient). In fact it extends into the mega-hertz (RF) range,
well beyond the recording capability of any EEG equipment. It can
appear artificially benign when captured using standard EEG
fMRI in Epilepsy 769
referred to as AAS) method [163] (see Fig. 6). It relies on the lack of
correlation between physiological signals and the artifacts, enabling
the latter to be estimated by averaging the EEG over a number of
epochs, corresponding to individual scan repetitions, for example.
The success of the artifact (template) estimation and its subsequent
subtraction from the ongoing EEG depend critically on the sam-
pling rate, the number of averaging epochs, and the precision of
their timing. In Allen’s original implementation, this is addressed by
the use of the scanner’s scan trigger pulse to mark each scan acquisi-
tion and interpolation. Following subtraction, residual artifacts are
reduced using Adaptive Noise Cancellation. The method can be
used in real time, allowing continuous EEG-fMRI studies in patients
with epilepsy [196–199]. Possibly the most important practical
development has been the demonstration that synchronized MR
acquisition and EEG digitization lead to significantly improved EEG
quality, and in particular over a wider frequency range, when com-
bined with an AAS-like method [225]. The method has been found
to perform well for spiral EPI [226]. As noted previously, changes in
the artifact waveform due to subject motion will lead to suboptimal
template estimation. To address this, refinements of Allen’s method
which incorporate PCA of the residual artifact have been proposed
[171, 227]. The shape of the image acquisition artifact may be cap-
tured in a separate experiment for subsequent subtraction [192]. In
the study by Wan et al. [228], a method designed to bypass the
requirement for a slice acquisition signal from the scanner is pro-
posed. As is commonly the case for artifact reduction methods based
on ICA, identification of the components containing artifact is
mainly done visually [179]. A difficulty encountered when compar-
ing the various methods available for artifact reduction is the range
of methodologies used. This has been addressed to some degree in
a rigorous comparative study [229].
EEG Quality: Subject (head) motion, and almost inevitably motion of the EEG
The Impact of Subject recording circuit (formed by the head, electrodes and leads) which
Motion results from it, will cause fluctuations in the recorded signal addi-
tive to the effects discussed in the previous sections. In most situa-
tions, this motion is not synchronized with either the scanning
process or heartbeat; due to the associated change in geometry of
the EEG circuit in relation to the scanner, it therefore can intro-
duce random fluctuations in the magnitude of both types of arti-
fact. In terms of the AAS algorithm, this phenomenon imposes a
limit on the quality of the artifact correction: the greater and more
random the motion, the less efficient AAS will be. The previously
mentioned measures to reduce the pulse artifacts based on subject
immobilization and minimization of EEG loops are obviously
worth reiterating at this point [158, 159]. More recently, a method
to measure head motion and use the information to reduce pulse
and scanning-related artifacts has been proposed [230, 231].
fMRI in Epilepsy 771
2.2.1 Ictal fMRI in Focal As mentioned previously, a number of case studies of ictal events
Epilepsy captured using fMRI alone were published prior to the advent of
EEG-fMRI [139–142]. Despite the fact that these contain inter-
esting observations, particularly with regard to the signal change
around the time of seizure onset, the availability of simultaneously
recorded EEG would have contributed important information.
The development of the ability to record good quality EEG inside
the MR scanner has offered the possibility of improved models of
ictal fMRI signal changes by the inclusion of precisely timed
Table 3 EEG-fMRI studies of paroxysmal activity: early milestones and important series
Number of subjects/no.
Study in which IED recorded Results Conclusion Comment
Focal epilepsy 1/1 frequent IED Bilateral activation where EEG suggested left “We cannot make conclusions about the source of the
Warach et al. temporal localization and anterior cingulate discharge from the present data”
1996 [183] activation in relation to generalized
epileptiform activity
Seeck et al. 1998 1/1 frequent IED Multiple areas of signal enhancement on The combination of EEG-triggered fMRI and 3D EEG
[200] fMRI. Confirmed on 3D-EEG source source analysis, represents a promising additional tool
localization with evidence of a focal onset. for presurgical epilepsy evaluation allowing precise
Focus later confirmed on subdural recordings noninvasive identification of the epileptic foci
Symms et al. 1999 1/1 frequent IED Reproducible and concordant activation across four sessions
[201]
Patel et al. 1999 20/10 frequent IED 9/10 overall reported as showing “activation corresponding to the EEG focus”
[202]
Krakow et al. 10 frequent IED Reproducible activations (same lobe and overlapping) obtained in 6/10 patients in close spatial relation
1999 [187] to EEG focus
Krakow et al. 1 frequent IED Focal activation within a large malformation of cortical development in response to focal epileptiform
1999 [203] discharges
Lazeyras et al. 11 frequent IED Activation confirmed clinical diagnosis in 7/11. In 5/6 intracranial EEG confirmed result
2000 [204]
Lazeyras et al. 1/1 frequent IED Area of signal enhancement concordant with hyperintensity seen on ictal FLAIR images in a patient with
2000 [205] nonlesional partial epilepsy
Krakow et al. 24/14 frequent IED 12/24 patients showed activations concordant with EEG focus, 7/12 of which also had concordant
2001 [189] structural lesions. 2/24 were discordant and 10/24 showed no significant activation
Lemieux et al. 1/1 frequent, In a case with stereotyped frequent IED, BOLD Localization of BOLD activation was consistent with First description of
2001 [197] stereotyped IED signal change concordant with the seizure previous findings and EEG source modeling application of
onset zone was recorded continuous
EEG-fMRI
Jager et al. 2002 10/5 frequent IED, Focal activation in 5/5 patients, concordant with EEG amplitude mapping. Mean signal increase was
[206] focal epilepsy 15 ± 9 %. Spike amplitude correlated with volume of activation
Benar et al. 2002 4/4 frequent IED The average HRF presented a wider positive There was no clear correlation between the amplitudes
[207] lobe in three patients and a longer undershoot of individual BOLD responses and EEG spikes
in two.
Al-Asmi et al. 48/31 frequent IED BOLD activation in 39 % of studies. Concordant Combining EEG and fMRI in focal epilepsy yields
2003 [190] with seizure focus in almost all. four patients regions of activation that are presumably the source
had concordant intracranial recording (by of spiking activity and these are high
lobe)
Benar et al. 2006 5/5 presurgical When an intracranial electrode is in the vicinity of an EEG or fMRI peak, it usually includes one active Largest series of
[208] candidates having contact intracranial
sEEG EEG correlated
fMRI activation
Aghakhani et al. 64/40 focal epilepsy A positive thalamic response was seen in 12.5 % The thalamus is involved in partial epilepsy during
2006 [209] with frequent of studies with unilateral and 55 % with interictal discharges. This involvement and also
unilateral or bilateral spikes. Cortical acitvation was more cortical deactivation are more commonly seen with
bilateral IED concordant with focus than deactivation bilateral spikes than focal discharges
Salek-Hadaddi 63/34 focal epilepsy Significant hemodynamic correlates were These findings provide important new information on
et al. 2006 frequent IED detectable in over 68 % of patients and were the optimal use and interpretation of EEG-fMRI in
[199] highly, but not entirely, concordant with site focal epilepsy
of presumed seizure onset
Ziljmans et al. 29/15 focal epilepsy, 8/15 subjects: IED correlated BOLD response EEG-fMRI provides additional information about the First evaluation of
2007 [210] declined for surgery at site of focus. Multifocal in 4, unifocal in 4. epileptic source in the presurgical work-up of impact on
Thornton et al. 23/12 focal epilepsy, Concordant with IC data in 2 complex cases presurgical
2011 [211] focal cortical 11/12 showed significant IED-correlated Widely distributed discordant regions of IED-related evaluation
Pittau et al. 2012 dysplasia, BOLD. BOLD matched icEEG SOZ and hemodynamic change appear to be associated with a EEG-fMRI (with
[212] intracranial EEG outcome was > 50 % seizure reduction in 5; widespread SOZ and poor postsurgical outcome video) of
Chaudhary et al. and surgery BOLD was widespread in 6 and outcome EEG-fMRI may contribute to the localization of the seizures can be
2012 [213] 43/33 focal epilepsy poor in 5/6 interictal epileptic generator in patients with focal done with
Coan et al. 2015 20/15 seizures in 21/33 IED-related BOLD contributed to the epilepsy acceptable risk,
[214] focal epilepsy delineation of the focus compared to scalp Preictal and ictal haemodynamic changes in refractory and provides
30 patients who EEG; icEEG validation was positive in 12/14 focal seizures can noninvasively localize seizure important new
underwent surgery patients onset at sublobar/gyral level when ictal scalp localizing
for TLE; in 14 Widespread preictal BOLD changes followed by electroencephalography is not helpful information
cases with no IED more focused early ictal and spread Interictal EEG-fMRI retrospectively confirmed the
during fMRI, a Good surgical outcome in 13/16 patients with epileptogenic zone of TLE patients
topographic concordant BOLD changes and in 3/14
method was used patients with discordant BOLD
[215]
(continued)
Table 3 (continued)
Number of subjects/no.
Study in which IED recorded Results Conclusion Comment
Tousseyn et al. 28/27 refractory High congruence between maps derived from Hemodynamic changes related to seizures and spikes Largest
2015 [216] focal epilepsy ictal the two techniques; some discrepancies varied spatially within a common network. Overlap comparison of
SPECT and observed nearby and distant from discharge origin interictal
interictal EEG-fMRI and
EEG-fMRI ictal SPECT
Generalized 1/1 prolonged GSW Thalamic activation and widespread cortical Supports thalamo-cortical model of GSW
epilepsy epochs (ictal) deactivation
Salek-Haddadi
et al. 2003
[217]
Baudewig et al. 1/1 frequent GSW Unilateral insular activation shown in relation to Strategy resulted in robust BOLD MRI responses to
2001 [218] generalized epileptiform discharges epileptic activity that resemble those commonly
observed for functional challenges
Hamandi et al. 46/30 interictal GSW Thalamic activation and cortical deactivation Observed cortical deactivation may represent correlate
2006 [196] in IGE and SGE observed at group level. Deactivation in the of clinical absence seizure.
default brain areas. Cortical pattern mixed at
individual level
Aghakhani et al. 15/14 interictal GSW Bilateral thalamic activation in 80 % of BOLD Cortical deactivation mediated by hyperpolarization of
2004 [219] response. Cortical deactivation in 93 % the thalamus
Hamandi et al. 4/4 interictal GSW Qualitatively reproducible BOLD and blood Consistent with preserved neurovascular coupling in
2007 perfusion patterns; Cortical deactivation GSW and decreased cortical activity
corresponds to decrease in blood flow
Yang et al. 2013 10/10 drug-naïve GSW discharge-related alterations in the default Interictal GSWDs can cause dysfunction in specific First study
[220] CAE patients; mode network (DMN), cognitive control networks important for psychosocial function specifically
interictal GSW network (CCN), and affective networks focused on
functional
connectivity
alterations
during GSWD
in drug-naive
patients
Vaudano et al., 15 with eyelid Elevated eye closure-related BOLD signal in the Supports concept of EMA as a distinct epileptic
2014 [221] myoclonus with visual cortex, posterior thalamus, and eye condition
absences motor control areas in EMA group compared
(EMA) + 14 with to IGE
IGE
Children 6/6 focal epilepsy Concordant activation with presumed focus in EEG-fMRI is a promising tool to noninvasively First series
De Tiege et al. (lesional and four cases. IC recording corroborative in 1 localize epileptogenic regions in children with specifically
2007 [222] nonlesional) pharmacoresistant focal epilepsy addressing use
of EEG-fMRI in
children
Jacobs et al. 2007 9/9, mixed focal All had BOLD activation or deactivation EEG-fMRI at 3 T could be useful to localize focus in
[223] epilepsy, sedated concordant with seizure focus. Deactivation children with focal epilepsies. Nature and origin of
appears more common in adults the negative BOLD requires investigation
Jacobs et al. 2007 13/13 symptomatic Activation corresponding with the lesion was “Good results could be obtained from the EEG-fMRI
[224] (lesional) epilepsy), seen in 20 % and deactivation in 52 % of the recordings, performed in sedated children”
sedated studies
776 Rachel C. Thornton et al.
2.2.2 Interictal fMRI: Although the study of ictal events may be useful in pursuing the
EEG-fMRI aim of noninvasively identifying the seizure onset zone and thereby
contributing most to presurgical evaluation, its acquisition remains
challenging. Attention has, therefore, focused on interictal activity
and the insight gained from it into the function of epileptic net-
works. Spike-triggered fMRI acquisition was used in the first appli-
cations of EEG-fMRI in epilepsy, which tended to focus on the
study of subjects pre-selected for the large number of IED observed
in prior routine surface EEG or as part of presurgical assessment.
The acquisition and analysis of spike-triggered fMRI data generally
rests on the assumption that interictal spikes are associated with a
BOLD signal change pattern similar to the so-called canonical
HRF. In these series, regions of positive BOLD signal change asso-
ciated with IED were observed in approximately 50 % of the cases
overall, and occasional negative changes [183, 187, 189, 190,
200–202, 204, 239, 240]. Using bursts of BOLD EPI scans,
Krakow et al. made an initial attempt at estimating the shape of the
IED-related HRF [189]. We note that in the work by Seeck and
colleagues, Clonazepam was used to suppress interictal discharges,
thereby creating a control scan state [200, 204]. Interleaved EEG-
fMRI, whereby fMRI data are acquired in blocks with inter-block
gaps of a sufficient duration to allow interpretation of part of the
EEG was employed in a patient with IGE for the mapping of
BOLD changes related to spike-wave complexes [218].
Following the implementation of image acquisition artifact
removal [163], the technique of continuous, simultaneous EEG-
fMRI acquisition was demonstrated along with estimation of the
shape of the IED-related HRF in a subject with Raussmussen’s
encephalitis and stereotyped high amplitude sharp waves on the
fMRI in Epilepsy 777
2.2.3 The HRF in Focal The normal hemodynamic response associated with neuronal activ-
Epilepsy ity arising from brief external stimuli in humans has a characteristic
shape with a peak at around 5–6 s following the event, the so-
called canonical HRF, with a significant degree of inter-subject
variability [249]. The shape of the HRF is a key element of fMRI
signal modeling with an important impact on sensitivity, and it is
therefore important to attempt to characterize it. In epilepsy, par-
ticularly in relation to deactivations it has been proposed that neu-
rovascular coupling is abnormal with possible consequences on the
shape of the HRF, and this was suggested to be one possible expla-
nation for the significant widespread BOLD deactivations observed
in both focal and generalized epilepsy [250, 251].
This has lead to increased interest in estimating the shape of
the HRF in epilepsy. An efficient way of estimating the shape of the
hemodynamic changes linked with epileptiform discharges is by
using a set of functions (sometimes referred to as “basis set”) that
can, by linear superposition, fit signal changes with almost any time
course. Various such schemes have been used in studies of epilepsy
including sets of gamma response functions, Fourier basis sets and
a linear combination of canonical HRF, its time derivative and a
dispersion derivative [197, 199, 207, 252, 253]. Although a
degree of variability has been observed, the HRF linked to IED
was found to be principally canonical in shape [254]. In some cases
deviation from the norm at locations distant to the IED may reflect
artifacts, while in others, deviant activation in proximity to the pre-
sumed focus, early responses may reflect brain activity that system-
atically precedes the event captured on the scalp or propagation
[252, 255]. In a recent study, the development of penicillin-
induced epileptic activity was correlated with increase in BOLD
signal prior to the onset of IED [256]. These studies coupled with
those combining EEG-fMRI with multiple source analysis [257]
suggest that in some cases the maximum BOLD response may be
detected just prior to seizure or even IED onset. It may be, how-
ever, that this reflects BOLD changes linked to interictal abnor-
malities, which are not included in the model by virtue of the fact
they are not seen on the scalp EEG.
2.2.4 Clinical Relevance Having shown that EEG-fMRI is capable of providing a unique
of fMRI of Paroxysmal form of localizing information, the issue of its clinical value arises.
Activity in Focal Epilepsy The evaluation of new noninvasive imaging modalities in the pre-
surgical assessment of patients with drug-resistant epilepsy is a
complex issue in part due to the lack of an established method-
ological consensus (baseline) across centers. Added value and clini-
cal relevance are a function of the new test’s sensitivity and
specificity. In the case of EEG-fMRI, neither has been properly
assessed to date. Thus far the field has focused on proof of principle
demonstrations, usually in patients selected based on high rates of
EEG abnormalities, with a success rate of roughly 50–60 % [183,
fMRI in Epilepsy 779
187, 189, 199, 202, 204, 240]. Therefore, one may anticipate a
lower success rate in the most clinically challenging cases for which
the need for noninvasive assessment is most pressing.
When available, the new localizing information must be evalu-
ated relative to a surgically confirmed irritative and epileptogenic
zone [258] using the current gold standard of invasive recording
and outcome data. In practice, a gradual approach to validation is
often taken, whereby the face validity of new localizing informa-
tion is tested against other existing techniques in cases with well
characterized syndromes, such as mesial TLE.
In our laboratory, we have assessed the value of the fMRI find-
ings by comparing the localization of the BOLD cluster containing
the most significant activation to the seizure onset zone defined
electro-clinically, when possible, and found a very good degree of
concordance at the lobar level [187, 189, 199]. Additional regions
of activation were observed in roughly 50 % of the cases with
significant activation. The finding of localized BOLD activation in
cases with poor electroclinical localization suggests a means of
obtaining target areas for intracranial EEG [199]. In TLE, the
yield has been characterized as relatively high [259] and the degree
of concordance of BOLD activations with the presumed focus gen-
erally good in one study [199], but more varied in another [259].
Comparison of EEG-fMRI with source localization suggests
that IED-related BOLD activation are often in proximity to those
detected by conventional EEG source localization, but some stud-
ies have suggested a distance of up to 50 mm. The possible sources
of discrepancy between BOLD and electrical (or magnetic) source
localization include: differences in the nature of the observed phe-
nomena and neurovascular coupling, vascular architecture and
scanner field-related effects on sensitivity, instrumental and physi-
ological noise, source reconstruction limitations, fMRI sensitivity
limitations [208, 260–262].
Presurgically, intracerebral EEG data are widely recognized as
the localization gold standard. Comparison of scalp EEG-fMRI
against invasive EEG has been performed in small groups and using
widely varying fMRI analysis techniques and comparison criteria
[190, 200, 250], noting a degree of concordance between the epi-
leptogenic zone and the area of maximal BOLD activation in some,
but not all subjects [204, 208]. The potential role of EEG-fMRI in
presurgical evaluation was assessed in a series of patients with focal
epilepsy in whom surgery was not offered following conventional
electroclinical evaluation [210]. The impact of the EEG-fMRI find-
ings was assessed in eight cases with unclear foci or suspected muti-
focality (based on the center’s usual battery of tests) in whom
significant IED-related BOLD activation were observed: IED-
related changes suggested a more restricted seizure onset zone in
four, and multifocality in a further patient. The EEG-fMRI finding
was concordant with intracranial recordings in two. The authors
780 Rachel C. Thornton et al.
2.2.5 Relationship Focal epilepsy can be divided by pathological subtype. Given the
with Pathology in Focal knowledge that the irritative and epileptogenic zones may extend
Epilepsy beyond the area of pathological abnormality and that animal models
suggest abnormal subpopulations of neurons within dysplastic areas,
there have been studies of EEG-fMRI aimed at evaluating the hemo-
dynamic response in abnormal tissue revealed on MRI [203, 211,
244, 263–266]. In a series of 14 cases with heterotopia notable vari-
ability of the BOLD response across abnormal tissue was observed,
but the area of BOLD signal increase was often concordant with the
area of pathology, with a more mixed pattern of deactivations [266].
In MCDs and particularly Taylor type focal cortical dysplasia, BOLD
signal increase was observed within the lesion while peri-close and
distant from the lesional activity displayed a negative BOLD response
in four out of six cases [237]. Other studies in cases with MCD have
supported these findings. IED-related BOLD signal changes have
been observed in patients with cavernomas [267] close and distant
from the lesion. The frequently observed negative BOLD responses,
particularly in MCD have been attributed to loss of neuronal inhibi-
tion (in the presence of normal neurovascular coupling) in the
regions surrounding the abnormality or abnormalities in neurovas-
cular coupling itself. The significance of these deactivations will be
discussed in more detail later.
2.2.6 fMRI EEG-fMRI has been used to attempt to localize sources in specific
in the Investigation epilepsy syndromes, in addition to adding evidence to the under-
of Epilepsy Syndromes standing of the differences in subtypes of focal epilepsy. TLE is of
particular interest in this context as surface EEG may not detect
the deep sources involved in TLE and surgery for TLE, where it is
correctly localized, is associated with excellent outcome [268].
fMRI in Epilepsy 781
2.2.8 The Significance BOLD signal deactivation has been observed in the monkey visual
of BOLD Deactivation system and found to have a linear relationship with CBF in the
in Epilepsy same way as positive BOLD signal change in normal physiological
conditions [283]. The pattern of cortical deactivation commonly
observed in relation to GSW using EEG-fMRI in humans has been
discussed earlier. Gotman et al. proposed that the reason for
observing the widespread cortical deactivation may be hypersyn-
chronization of the thalamus, supporting Avoli’s proposed model
of the thalamus driving the cortex during GSW [273]. The results
of Hamandi et al. using EEG-ASL are consistent with GSW-related
deactivation reflecting decreased cortical activity [277].
Focal IED-related deactivations are less common than activa-
tions and seem particularly linked to the presence of activation,
possibly reflecting a smaller hemodynamic effect of individual
events [199, 251]. In cases with MCD IED-related negative
BOLD signal change adjacent to the seizure onset zone or area of
dysplastic cortex, has led to several possible explanations including
“vascular steal” from more metabolically active regions, abnormal
neuronal coupling, or perhaps more likely, loss of inhibitory neu-
ronal activity in these regions supporting the link between BOLD
deactivation and underlying decrease in neuronal activity [251,
264].
Deactivation of the default mode areas is a typical feature of
IED-related BOLD changes in TLE in contrast to cases with extra-
TLE [243]. The observed pattern may reflect IED-related effects
in areas involved in cognition, with a possible link to transient cog-
nitive impairment [284].
2.2.9 fMRI Although limited by the sluggish BOLD response, we have seen
and the Neurobiology that EEG-fMRI is able to reveal multiple regions more or less
of Epileptic Networks simultaneously activated or deactivated in relation to EEG events
fMRI in Epilepsy 783
2.2.10 Limitations, The study of spontaneous, pathological brain activity using fMRI
Challenges, and Future presents the experimentalist numerous challenges. The advent of
Work simultaneously recorded EEG greatly facilitates this endeavor. EEG-
fMRI was originally held to be an excellent tool incorporating the
spatial coverage and resolution of imaging techniques and the tem-
poral resolution of EEG. However, a number of challenges remain.
Technically, postprocessing methods are now available to offer
EEG quality sufficient to allow a degree of abnormality identifica-
tion reliability comparable to that of routine clinical EEG. Although
EEG quality is sufficient to detect most IED with a good degree of
certainty (at 1.5 T) [161], pulse-related artifacts can sometimes
interfere with EEG interpretation particularly at 3 T. Improved
artifact correction and the use of techniques based on EEG source
reconstruction may yield more localized and reproducible results
[162, 257]. In addition, very little work has been done on this
aspect of EEG interpretation and in particular the differences
between the clinical and experimental approaches, with the former
using a summary of the abnormalities observed over the whole
EEG rather than categorization and quantification of all IED
required for fMRI. Concerning the effects of motion, which is par-
ticularly problematic in patients with epilepsy (and other neuro-
logical conditions), on MR image and EEG quality, the advent of
prospective motion correction offers a possible way forward [285].
Reliance on scalp EEG for fMRI modeling is both a strength,
as it allows to answer the question “what, if any, are the BOLD
correlates of EEG pattern X?”, and weakness because it is bur-
dened with certain limiting aspects of scalp EEG, such as sensitivity
bias subjectivity in interpretation, and the unpredictable nature of
the phenomena of interest. This presents the investigator with sig-
nificant challenges in terms of unpredictable experimental effi-
ciency (and consequently yield) and EEG interpretation for the
purpose of GLM building. The former may require a more aggres-
sive approach possibly using drug management as a tool for modu-
lating EEG activity [200]. Although EEG event classification
remains problematic, possible solutions may give the opportunity
of using fMRI to inform EEG interpretation [162].
784 Rachel C. Thornton et al.
2.2.11 fMRI of The possibility of recording EEG of good quality during fMRI has
Paroxysmal Activity: created a new instrument, which to date has been mainly used in
Conclusion an exploratory fashion.
EEG-fMRI of ictal and interictal activity in focal epilepsy has
demonstrated the capability to provide new localization information
in a large proportion of cases. In focal epilepsy, varied patterns have
been identified often suggestive of “functional lesions” homologous
to abnormalities seen on structural imaging, but additionally the
involvement of possibly less disease-specific regions. The clinical
value of this information remains uncertain and is the subject of
ongoing investigations. At the very least, EEG-fMRI may provide
complementary data useful for planning of invasive recordings where
conventional localization of the seizure focus is unsuccessful.
However, there are signs that it may be able to offer more.
In generalized epilepsy, EEG-fMRI has revealed activation and
deactivation patterns that are mainly suggestive of less specific
effects linked to generalized spike-wave, rather than reflecting syn-
drome, the spatial distribution of the EEG generators or more spe-
cifically a putative focus responsible for initiating GSW due in part
to the averaging effect of BOLD.
2.2.12 Data Acquisition Some of the brain regions of most interest in epilepsy are subject
and Modeling Challenges to susceptibility artifact, leading to geometric distortions and sig-
in Clinical fMRI nal loss during fMRI acquisition. Ideally in the absence of an
applied gradient, the magnetic field would be homogenous
throughout the bore of an MRI scanner. Unfortunately, the differ-
ent magnetic properties of bone, tissue, and air introduce inhomo-
geneities in the field when a head is introduced into the bore. Brain
regions closest to borders between sinuses and brain or bone and
brain, for example the inferior frontal and MTLs, are most affected,
and therefore especially likely to suffer geometric distortions or
signal loss [290]. This can result in reduced sensitivity and ana-
tomical uncertainties when interpreting the images. Most epilepsy
studies have been performed on 1.5 T clinical MRI scanners.
Scanning at higher field strength improves signal-to-noise ratio but
increases distortions and dropout [291].
786 Rachel C. Thornton et al.
Acknowledgments
Thanks to Suejen Perani for help with the second edition and to
Philip Allen for his comments on parts of the manuscript and to
Dr. Anna Vaudano and Dr. Serge Vulliemoz for supplying some of
the illustrations. Some of the work reported in this chapter was
funded through a grant from the Medical Research Council (MRC
grant number G0301067) and by the Wellcome Trust. We are
grateful to the Big Lottery Fund, Wolfson Trust, and National
Society for Epilepsy for supporting the NSE MRI scanner. This
work was carried out under the auspices of the UCL/UCLH
Biomedical Comprehensive Research Centre.
fMRI in Epilepsy 787
References
1. Wagner DD, Sziklas V, Garver KE, Jones- 13. Bonelli SB, Powell R, Thompson PJ,
Gotman M (2009) Material-specific later- Yogarajah M, Focke NK, Stretton J, Vollmar
alization of working memory in the medial C et al (2011) Hippocampal activation corre-
temporal lobe. Neuropsychologia 47:112–122 lates with visual confrontation naming: fMRI
2. Campo P, Garrido MI, Moran RJ, Garcia- findings in controls and patients with tempo-
Morales I, Poch C, Toledano R, Gil-Nagel ral lobe epilepsy. Epilepsy Res 95:246–254
A et al (2013) Network reconfiguration and 14. Gartus A, Foki T, Geissler A, Beisteiner R
working memory impairment in mesial tem- (2009) Improvement of clinical language
poral lobe epilepsy. NeuroImage 72:48–54 localization with an overt semantic and syntac-
3. Stretton J, Winston G, Sidhu M, Centeno tic language functional MR imaging paradigm.
M, Vollmar C, Bonelli S, Symms M et al AJNR Am J Neuroradiol 30:1977–1985
(2012) Neural correlates of working memory 15. Sanjuan A, Bustamante JC, Forn C, Ventura-
in Temporal Lobe Epilepsy--an fMRI study. Campos N, Barros-Loscertales A, Martinez JC,
NeuroImage 60:1696–1703 Villanueva V et al (2010) Comparison of two
4. Hoppe C, Elger CE, Helmstaedter C (2007) fMRI tasks for the evaluation of the expressive
Long-term memory impairment in patients language function. Neuroradiology 52:407–415
with focal epilepsy. Epilepsia 48(Suppl 16. Friedman L, Kenny JT, Wise AL, Wu D, Stuve
9):26–29 TA, Miller DA, Jesberger JA et al (1998)
5. Waites AB, Briellmann RS, Saling MM, Abbott Brain activation during silent word genera-
DF, Jackson GD (2006) Functional connec- tion evaluated with functional MRI. Brain
tivity networks are disrupted in left temporal Lang 64:231–256
lobe epilepsy. Ann Neurol 59:335–343 17. Bonelli SB, Thompson PJ, Yogarajah M,
6. Vlooswijk MC, Jansen JF, Majoie HJ, Vollmar C, Powell RH, Symms MR, McEvoy
Hofman PA, de Krom MC, Aldenkamp AP, AW et al (2012) Imaging language networks
Backes WH (2010) Functional connectiv- before and after anterior temporal lobe resec-
ity and language impairment in cryptogenic tion: results of a longitudinal fMRI study.
localization-related epilepsy. Neurology Epilepsia 53:639–650
75:395–402 18. Fernandez G, Specht K, Weis S, Tendolkar
7. Baxendale S (2002) The role of functional MRI I, Reuber M, Fell J, Klaver P et al (2003)
in the presurgical investigation of temporal Intrasubject reproducibility of presurgical
lobe epilepsy patients: a clinical perspective and language lateralization and mapping using
review. J Clin Exp Neuropsychol 24:664–676 fMRI. Neurology 60:969–975
8. Kirsch HE, Walker JA, Winstanley FS, 19. Weber B, Wellmer J, Schur S, Dinkelacker V,
Hendrickson R, Wong ST, Barbaro NM, Laxer Ruhlmann J, Mormann F, Axmacher N et al
KD et al (2005) Limitations of Wada memory (2006) Presurgical language fMRI in patients
asymmetry as a predictor of outcomes after with drug-resistant epilepsy: effects of task
temporal lobectomy. Neurology 65:676–680 performance. Epilepsia 47:880–886
9. Abbott DF, Waites AB, Lillywhite LM, 20. Duncan J (2009) The current status of neu-
Jackson GD (2010) fMRI assessment of lan- roimaging for epilepsy. Curr Opin Neurol
guage lateralization: an objective approach. 22:179–184
NeuroImage 50:1446–1455 21. Rosazza C, Ghielmetti F, Minati L, Vitali P,
10. Appel S, Duke ES, Martinez AR, Khan OI, Giovagnoli AR, Deleo F, Didato G et al (2013)
Dustin IM, Reeves-Tyer P, Berl MB et al Preoperative language lateralization in tempo-
(2012) Cerebral blood flow and fMRI BOLD ral lobe epilepsy (TLE) predicts peri-ictal, pre-
auditory language activation in temporal lobe and post-operative language performance: an
epilepsy. Epilepsia 53:631–638 fMRI study. NeuroImage Clin 3:73–83
11. Arora J, Pugh K, Westerveld M, Spencer 22. Hamberger MJ, McClelland S 3rd, McKhann
S, Spencer DD, Todd Constable R (2009) GM 2nd, Williams AC, Goodman RR (2007)
Language lateralization in epilepsy patients: Distribution of auditory and visual naming
fMRI validated with the Wada procedure. sites in nonlesional temporal lobe epilepsy
Epilepsia 50:2225–2241 patients and patients with space-occupying
12. Binder JR, Gross WL, Allendorfer JB, Bonilha temporal lobe lesions. Epilepsia 48:531–538
L, Chapin J, Edwards JC, Grabowski TJ et al 23. Hermann BP, Wyler AR (1988) Effects of
(2011) Mapping anterior temporal lobe lan- anterior temporal lobectomy on language
guage areas with fMRI: a multicenter norma- function: a controlled study. Ann Neurol
tive study. NeuroImage 54:1465–1475 23:585–588
788 Rachel C. Thornton et al.
24. Hamberger MJ, Seidel WT (2009) 35. You X, Adjouadi M, Wang J, Guillen MR,
Localization of cortical dysfunction based Bernal B, Sullivan J, Donner E et al (2013) A
on auditory and visual naming performance. decisional space for fMRI pattern separation
J Int Neuropsychol Soc 15:529–535 using the principal component analysis--a com-
25. Specht K, Osnes B, Hugdahl K (2009) parative study of language networks in pediatric
Detection of differential speech-specific pro- epilepsy. Hum Brain Mapp 34:2330–2342
cesses in the temporal lobe using fMRI and a 36. Woermann FG, Jokeit H, Luerding R, Freitag
dynamic “sound morphing” technique. Hum H, Schulz R, Guertler S, Okujava M et al
Brain Mapp 30:3436–3444 (2003) Language lateralization by Wada
26. Schlosser MJ, Aoyagi N, Fulbright RK, Gore test and fMRI in 100 patients with epilepsy.
JC, McCarthy G (1998) Functional MRI Neurology 61:699–701
studies of auditory comprehension. Hum 37. Wang J, You X, Wu W, Guillen MR, Cabrerizo
Brain Mapp 6:1–13 M, Sullivan J, Donner E et al (2014)
27. Lehericy S, Cohen L, Bazin B, Samson S, Classification of fMRI patterns--a study of
Giacomini E, Rougetet R, Hertz-Pannier L the language network segregation in pediat-
et al (2000) Functional MR evaluation of ric localization related epilepsy. Hum Brain
temporal and frontal language dominance Mapp 35:1446–1460
compared with the Wada test. Neurology 38. You X, Adjouadi M, Guillen MR, Ayala M,
54:1625–1633 Barreto A, Rishe N, Sullivan J et al (2011)
28. Gaillard WD, Balsamo L, Xu B, McKinney C, Sub-patterns of language network reorganiza-
Papero PH, Weinstein S, Conry J et al (2004) tion in pediatric localization related epilepsy: a
fMRI language task panel improves determi- multisite study. Hum Brain Mapp 32:784–799
nation of language dominance. Neurology 39. Friston KJ, Price CJ, Fletcher P, Moore C,
63:1403–1408 Frackowiak RS, Dolan RJ (1996) The trou-
29. Gaillard WD, Balsamo L, Xu B, Grandin CB, ble with cognitive subtraction. NeuroImage
Braniecki SH, Papero PH, Weinstein S et al 4:97–104
(2002) Language dominance in partial epi- 40. Price CJ, Friston KJ (1997) Cognitive con-
lepsy patients identified with an fMRI reading junction: a new approach to brain activation
task. Neurology 59:256–265 experiments. NeuroImage 5:261–270
30. Adcock JE, Wise RG, Oxbury JM, Oxbury 41. Risse GL, Gates JR, Fangman MC (1997) A
SM, Matthews PM (2003) Quantitative reconsideration of bilateral language repre-
fMRI assessment of the differences in later- sentation based on the intracarotid amobar-
alization of language-related brain activa- bital procedure. Brain Cogn 33:118–132
tion in patients with temporal lobe epilepsy. 42. Serafetinides EA, Hoare RD, Driver M
NeuroImage 18:423–438 (1965) Intracarotid sodium amylobarbitone
31. Branco DM, Suarez RO, Whalen S, O’Shea and cerebral dominance for speech and con-
JP, Nelson AP, da Costa JC, Golby AJ (2006) sciousness. Brain 88:107–130
Functional MRI of memory in the hip- 43. Rasmussen T, Milner B (1977) The role of
pocampus: Laterality indices may be more early left-brain injury in determining lateral-
meaningful if calculated from whole voxel ization of cerebral speech functions. Ann N Y
distributions. NeuroImage 32:592–602 Acad Sci 299:355–369
32. Liegeois F, Connelly A, Cross JH, Boyd SG, 44. Springer JA, Binder JR, Hammeke TA,
Gadian DG, Vargha-Khadem F, Baldeweg T Swanson SJ, Frost JA, Bellgowan PS, Brewer
(2004) Language reorganization in children CC et al (1999) Language dominance in
with early-onset lesions of the left hemi- neurologically normal and epilepsy sub-
sphere: an fMRI study. Brain 127:1229–1236 jects: a functional MRI study. Brain 122(Pt
33. Karunanayaka P, Kim KK, Holland SK, 11):2033–2046
Szaflarski JP (2011) The effects of left or right 45. Janszky J, Mertens M, Janszky I, Ebner A,
hemispheric epilepsy on language networks Woermann FG (2006) Left-sided interictal
investigated with semantic decision fMRI epileptic activity induces shift of language lat-
task and independent component analysis. eralization in temporal lobe epilepsy: an fMRI
Epilepsy BehavB 20:623–632 study. Epilepsia 47:921–927
34. Mbwana J, Berl MM, Ritzl EK, Rosenberger 46. Janszky J, Jokeit H, Heinemann D, Schulz
L, Mayo J, Weinstein S, Conry JA et al (2009) R, Woermann FG, Ebner A (2003) Epileptic
Limitations to plasticity of language network activity influences the speech organiza-
reorganization in localization related epilepsy. tion in medial temporal lobe epilepsy. Brain
Brain 132:347–356 126:2043–2051
fMRI in Epilepsy 789
47. Thivard L, Hombrouck J, du Montcel ST, 58. Janecek JK, Swanson SJ, Sabsevitz DS,
Delmaire C, Cohen L, Samson S, Dupont S Hammeke TA, Raghavan M, E Rozman M
et al (2005) Productive and perceptive lan- M, Binder JR (2013) Language lateralization
guage reorganization in temporal lobe epi- by fMRI and Wada testing in 229 patients
lepsy. NeuroImage 24:841–851 with epilepsy: rates and predictors of discor-
48. Berl MM, Balsamo LM, Xu B, Moore EN, dance. Epilepsia 54:314–322
Weinstein SL, Conry JA, Pearl PL et al 59. Benke T, Koylu B, Visani P, Karner E,
(2005) Seizure focus affects regional lan- Brenneis C, Bartha L, Trinka E et al (2006)
guage networks assessed by fMRI. Neurology Language lateralization in temporal lobe epi-
65:1604–1611 lepsy: a comparison between fMRI and the
49. Duke ES, Tesfaye M, Berl MM, Walker Wada Test. Epilepsia 47:1308–1319
JE, Ritzl EK, Fasano RE, Conry JA et al 60. Jayakar P, Bernal B, Santiago Medina L,
(2012) The effect of seizure focus on Altman N (2002) False lateralization of lan-
regional language processing areas. Epilepsia guage cortex on functional MRI after a clus-
53:1044–1050 ter of focal seizures. Neurology 58:490–492
50. Wilke M, Pieper T, Lindner K, Dushe T, 61. Wagner K, Hader C, Metternich B, Buschmann
Staudt M, Grodd W, Holthausen H et al F, Schwarzwald R, Schulze-Bonhage A (2012)
(2011) Clinical functional MRI of the lan- Who needs a Wada test? Present clinical indi-
guage domain in children with epilepsy. Hum cations for amobarbital procedures. J Neurol
Brain Mapp 32:1882–1893 Neurosurg Psychiatry 83:503–509
51. Weber B, Wellmer J, Reuber M, Mormann F, 62. Desmond JE, Sum JM, Wagner AD, Demb
Weis S, Urbach H, Ruhlmann J et al (2006) JB, Shear PK, Glover GH, Gabrieli JD et al
Left hippocampal pathology is associated with (1995) Functional MRI measurement of lan-
atypical language lateralization in patients guage lateralization in Wada-tested patients.
with focal epilepsy. Brain 129:346–351 Brain 118(Pt 6):1411–1419
52. Briellmann RS, Labate A, Harvey AS, Saling 63. Binder JR, Swanson SJ, Hammeke TA, Morris
MM, Sveller C, Lillywhite L, Abbott DF et al GL, Mueller WM, Fischer M, Benbadis S et al
(2006) Is language lateralization in tem- (1996) Determination of language domi-
poral lobe epilepsy patients related to the nance using functional MRI: a comparison
nature of the epileptogenic lesion? Epilepsia with the Wada test. Neurology 46:978–984
47:916–920 64. Hertz-Pannier L, Gaillard WD, Mott SH,
53. Jensen EJ, Hargreaves IS, Pexman PM, Cuenod CA, Bookheimer SY, Weinstein S,
Bass A, Goodyear BG, Federico P (2011) Conry J et al (1997) Noninvasive assessment
Abnormalities of lexical and semantic process- of language dominance in children and ado-
ing in left temporal lobe epilepsy: an fMRI lescents with functional MRI: a preliminary
study. Epilepsia 52:2013–2021 study. Neurology 48:1003–1012
54. Fakhri M, Oghabian MA, Vedaei F, Zandieh 65. Yetkin FZ, Swanson S, Fischer M, Akansel G,
A, Masoom N, Sharifi G, Ghodsi M et al Morris G, Mueller W, Haughton V (1998)
(2013) Atypical language lateralization: an Functional MR of frontal lobe activation:
fMRI study in patients with cerebral lesions. comparison with Wada language results.
Funct Neurol 28:55–61 AJNR Am J Neuroradiol 19:1095–1098
55. Wellmer J, Weber B, Urbach H, Reul J, 66. Benson RR, FitzGerald DB, LeSueur LL,
Fernandez G, Elger CE (2009) Cerebral Kennedy DN, Kwong KK, Buchbinder BR, Davis
lesions can impair fMRI-based language lat- TL et al (1999) Language dominance deter-
eralization. Epilepsia 50:2213–2224 mined by whole brain functional MRI in patients
56. Sanjuan A, Bustamante JC, Garcia-Porcar with brain lesions. Neurology 52:798–809
M, Rodriguez-Pujadas A, Forn C, Martinez 67. Carpentier A, Pugh KR, Westerveld M,
JC, Campos A et al (2013) Bilateral inferior Studholme C, Skrinjar O, Thompson JL,
frontal language-related activation correlates Spencer DD et al (2001) Functional MRI of
with verbal recall in patients with left tempo- language processing: dependence on input
ral lobe epilepsy and typical language distribu- modality and temporal lobe epilepsy. Epilepsia
tion. Epilepsy Res 104:118–124 42:1241–1254
57. Everts R, Harvey AS, Lillywhite L, Wrennall J, 68. Sabbah P, Chassoux F, Leveque C, Landre E,
Abbott DF, Gonzalez L, Kean M et al (2010) Baudoin-Chial S, Devaux B, Mann M et al
Language lateralization correlates with verbal (2003) Functional MR imaging in assessment
memory performance in children with focal of language dominance in epileptic patients.
epilepsy. Epilepsia 51:627–638 NeuroImage 18:460–467
790 Rachel C. Thornton et al.
69. FitzGerald DB, Cosgrove GR, Ronner S, in language processing following left hemi-
Jiang H, Buchbinder BR, Belliveau JW, Rosen sphere injury. Brain 129:754–766
BR et al (1997) Location of language in the 81. Powell HW, Parker GJ, Alexander DC,
cortex: a comparison between functional Symms MR, Boulby PA, Wheeler-Kingshott
MR imaging and electrocortical stimulation. CA, Barker GJ et al (2006) Hemispheric
AJNR Am J Neuroradiol 18:1529–1539 asymmetries in language-related pathways: a
70. Schlosser MJ, Luby M, Spencer DD, Awad combined functional MRI and tractography
IA, McCarthy G (1999) Comparative local- study. NeuroImage 32:388–399
ization of auditory comprehension by using 82. Powell HW, Parker GJ, Alexander DC,
functional magnetic resonance imaging and Symms MR, Boulby PA, Wheeler-Kingshott
cortical stimulation. J Neurosurg 91:626–635 CA, Barker GJ et al (2007) Abnormalities of
71. Pouratian N, Bookheimer SY, Rex DE, Martin language networks in temporal lobe epilepsy.
NA, Toga AW (2002) Utility of preoperative NeuroImage 36:209–221
functional magnetic resonance imaging for 83. Stretton J, Thompson PJ (2012) Frontal lobe
identifying language cortices in patients with function in temporal lobe epilepsy. Epilepsy
vascular malformations. J Neurosurg 97:21–32 Res 98:1–13
72. Rutten GJ, Ramsey NF, van Rijen PC, 84. Centeno M, Thompson PJ, Koepp MJ,
Noordmans HJ, van Veelen CW (2002) Helmstaedter C, Duncan JS (2010) Memory in
Development of a functional magnetic reso- frontal lobe epilepsy. Epilepsy Res 91:123–132
nance imaging protocol for intraoperative 85. Bonelli SB, Thompson PJ, Yogarajah M,
localization of critical temporoparietal lan- Powell RH, Samson RS, McEvoy AW, Symms
guage areas. Ann Neurol 51:350–360 MR et al (2013) Memory reorganization
73. Davies KG, Bell BD, Bush AJ, Hermann following anterior temporal lobe resection:
BP, Dohan FC Jr, Jaap AS (1998) Naming a longitudinal functional MRI study. Brain
decline after left anterior temporal lobectomy 136:1889–1900
correlates with pathological status of resected 86. Scoville WB, Milner B (2000) Loss of recent
hippocampus. Epilepsia 39:407–419 memory after bilateral hippocampal lesions.
74. Saykin AJ, Stafiniak P, Robinson LJ, Flannery 1957. J Neuropsychiatry Clin Neurosci
KA, Gur RC, O’Connor MJ, Sperling MR 12:103–113
(1995) Language before and after temporal 87. Ivnik RJ, Sharbrough FW, Laws ER Jr (1987)
lobectomy: specificity of acute changes and rela- Effects of anterior temporal lobectomy on
tion to early risk factors. Epilepsia 36:1071–1077 cognitive function. J Clin Psychol 43:128–137
75. Hermann BP, Perrine K, Chelune GJ, Barr 88. Spiers HJ, Burgess N, Maguire EA, Baxendale
W, Loring DW, Strauss E, Trenerry MR SA, Hartley T, Thompson PJ, O’Keefe
et al (1999) Visual confrontation nam- J (2001) Unilateral temporal lobectomy
ing following left anterior temporal lobec- patients show lateralized topographical and
tomy: a comparison of surgical approaches. episodic memory deficits in a virtual town.
Neuropsychology 13:3–9 Brain 124:2476–2489
76. Devinsky O, Perrine K, Llinas R, Luciano DJ, 89. Penfield W, Milner B (1958) Memory defi-
Dogali M (1993) Anterior temporal language cit produced by bilateral lesions in the hip-
areas in patients with early onset of temporal pocampal zone. AMA Arch Neurol Psychiatry
lobe epilepsy. Ann Neurol 34:727–732 79:475–497
77. Schwartz TH, Devinsky O, Doyle W, Perrine 90. Warrington EK, Duchen LW (1992) A re-
K (1998) Preoperative predictors of ante- appraisal of a case of persistent global amne-
rior temporal language areas. J Neurosurg sia following right temporal lobectomy: a
89:962–970 clinico-pathological study. Neuropsychologia
78. Sabsevitz DS, Swanson SJ, Hammeke TA, 30:437–450
Spanaki MV, Possing ET, Morris GL 3rd, 91. Loring DW, Hermann BP, Meador KJ, Lee
Mueller WM et al (2003) Use of preoperative GP, Gallagher BB, King DW, Murro AM et al
functional neuroimaging to predict language (1994) Amnesia after unilateral temporal lobec-
deficits from epilepsy surgery. Neurology tomy: a case report. Epilepsia 35:757–763
60:1788–1792
92. Alessio A, Pereira FR, Sercheli MS, Rondina
79. Noppeney U, Price CJ, Duncan JS, Koepp JM, Ozelo HB, Bilevicius E, Pedro T et al
MJ (2005) Reading skills after left anterior (2013) Brain plasticity for verbal and visual
temporal lobe resection: an fMRI study. Brain memories in patients with mesial temporal
128:1377–1385 lobe epilepsy and hippocampal sclerosis: an
80. Voets NL, Adcock JE, Flitney DE, Behrens fMRI study. Hum Brain Mapp 34:186–199
TE, Hart Y, Stacey R, Carpenter K et al 93. Banks SJ, Sziklas V, Sodums DJ, Jones-Gotman
(2006) Distinct right frontal lobe activation M (2012) fMRI of verbal and nonverbal mem-
fMRI in Epilepsy 791
ory processes in healthy and epileptogenic medial 105. McCormick C, Quraan M, Cohn M, Valiante
temporal lobes. Epilepsy Behav 25:42–49 TA, McAndrews MP (2013) Default mode
94. Chelune GJ (1995) Hippocampal adequacy network connectivity indicates episodic mem-
versus functional reserve: predicting memory ory capacity in mesial temporal lobe epilepsy.
functions following temporal lobectomy. Epilepsia 54:809–818
Arch Clin Neuropsychol 10:413–432 106. Jokeit H, Ebner A, Holthausen H,
95. Chelune GJ, Naugle RI, Luders H, Awad IA Markowitsch HJ, Moch A, Pannek H,
(1991) Prediction of cognitive change as a Schulz R et al (1997) Individual prediction
function of preoperative ability status among of change in delayed recall of prose passages
temporal lobectomy patients seen at 6-month after left-sided anterior temporal lobectomy.
follow-up. Neurology 41:399–404 Neurology 49:481–487
96. Kneebone AC, Chelune GJ, Dinner DS, 107. Helmstaedter C, Elger CE (1996) Cognitive conse-
Naugle RI, Awad IA (1995) Intracarotid quences of two-thirds anterior temporal lobectomy
amobarbital procedure as a predictor of on verbal memory in 144 patients: a three-month
material-specific memory change after anterior follow-up study. Epilepsia 37:171–180
temporal lobectomy. Epilepsia 36:857–865 108. Cheung MC, Chan AS, Lam JM, Chan YL
97. Sass KJ, Spencer DD, Kim JH, Westerveld M, (2009) Pre- and postoperative fMRI and
Novelly RA, Lencz T (1990) Verbal memory clinical memory performance in temporal
impairment correlates with hippocampal pyra- lobe epilepsy. J Neurol Neurosurg Psychiatry
midal cell density. Neurology 40:1694–1697 80:1099–1106
98. Trenerry MR, Jack CR Jr, Ivnik RJ, 109. Corkin S, Amaral DG, Gonzalez RG, Johnson
Sharbrough FW, Cascino GD, Hirschorn KA, Hyman BT (1997) H. M’.s medial tem-
KA, Marsh WR et al (1993) MRI hippocam- poral lobe lesion: findings from magnetic res-
pal volumes and memory function before onance imaging. J Neurosci 17:3964–3979
and after temporal lobectomy. Neurology 110. Fernandez G, Effern A, Grunwald T, Pezer
43:1800–1805 N, Lehnertz K, Dumpelmann M, Van Roost
99. Guedj E, Bettus G, Barbeau EJ, Liegeois- D et al (1999) Real-time tracking of memory
Chauvel C, Confort-Gouny S, Bartolomei formation in the human rhinal cortex and
F, Chauvel P et al (2011) Hyperactivation of hippocampus. Science 285:1582–1585
parahippocampal region and fusiform gyrus 111. Ojemann JG, Akbudak E, Snyder AZ,
associated with successful encoding in medial McKinstry RC, Raichle ME, Conturo TE
temporal lobe epilepsy. Epilepsia 52:1100–1109 (1997) Anatomic localization and quantitative
100. Sidhu MK, Stretton J, Winston GP, Bonelli analysis of gradient refocused echo-planar fMRI
S, Centeno M, Vollmar C, Symms M et al susceptibility artifacts. NeuroImage 6:156–167
(2013) A functional magnetic resonance 112. Greicius MD, Krasnow B, Boyett-Anderson
imaging study mapping the episodic memory JM, Eliez S, Schatzberg AF, Reiss AL,
encoding network in temporal lobe epilepsy. Menon V (2003) Regional analysis of hip-
Brain 136:1868–1888 pocampal activation during memory encod-
101. Bonelli SB, Powell RH, Yogarajah M, Samson ing and retrieval: fMRI study. Hippocampus
RS, Symms MR, Thompson PJ, Koepp MJ 13:164–174
et al (2010) Imaging memory in temporal 113. Lipschutz B, Friston KJ, Ashburner J,
lobe epilepsy: predicting the effects of tempo- Turner R, Price CJ (2001) Assessing study-
ral lobe resection. Brain 133:1186–1199 specific regional variations in fMRI signal.
102. Sidhu MK, Stretton J, Winston GP, Symms NeuroImage 13:392–398
M, Thompson PJ, Koepp MJ, Duncan JS 114. Craik FIM, Lockhart RS (1972) Levels of pro-
(2015) Memory fMRI predicts verbal mem- cessing: A framework for memory research.
ory decline after anterior temporal lobe resec- J Verbal Learn Verbal Behav 11:671–684
tion. Neurology 84:1512–1519 115. Kelley WM, Miezin FM, McDermott KB,
103. Hermann BP, Wyler AR, Somes G, Berry AD Buckner RL, Raichle ME, Cohen NJ, Ollinger
3rd, Dohan FC Jr (1992) Pathological status JM et al (1998) Hemispheric specialization in
of the mesial temporal lobe predicts memory human dorsal frontal cortex and medial tem-
outcome from left anterior temporal lobectomy. poral lobe for verbal and nonverbal memory
Neurosurgery 31:652–656, discussion 656-657 encoding. Neuron 20:927–936
104. Sass KJ, Westerveld M, Buchanan CP, Spencer 116. Demb JB, Desmond JE, Wagner AD, Vaidya
SS, Kim JH, Spencer DD (1994) Degree of CJ, Glover GH, Gabrieli JD (1995) Semantic
hippocampal neuron loss determines sever- encoding and retrieval in the left inferior pre-
ity of verbal memory decrease after left frontal cortex: a functional MRI study of task
anteromesiotemporal lobectomy. Epilepsia difficulty and process specificity. J Neurosci
35:1179–1186 15:5870–5878
792 Rachel C. Thornton et al.
117. Wagner AD, Schacter DL, Rotte M, Koutstaal to unilateral hippocampal sclerosis. Epilepsia
W, Maril A, Dale AM, Rosen BR et al (1998) 48:1512–1525
Building memories: remembering and for- 130. Richardson MP, Strange BA, Thompson PJ,
getting of verbal experiences as predicted by Baxendale SA, Duncan JS, Dolan RJ (2004)
brain activity. Science 281:1188–1191 Pre-operative verbal memory fMRI predicts
118. Buckner RL, Kelley WM, Petersen SE (1999) post-operative memory decline after left tem-
Frontal cortex contributes to human memory poral lobe resection. Brain 127:2419–2426
formation. Nat Neurosci 2:311–314 131. Richardson MP, Strange BA, Duncan JS,
119. Golby AJ, Poldrack RA, Brewer JB, Spencer Dolan RJ (2006) Memory fMRI in left hippo-
D, Desmond JE, Aron AP, Gabrieli JD (2001) campal sclerosis: optimizing the approach to
Material-specific lateralization in the medial predicting postsurgical memory. Neurology
temporal lobe and prefrontal cortex during 66:699–705
memory encoding. Brain 124:1841–1854 132. Powell HW, Richardson MP, Symms MR,
120. Wagner AD, Koutstaal W, Schacter DL (1999) Boulby PA, Thompson PJ, Duncan JS, Koepp
When encoding yields remembering: insights MJ (2008) Preoperative fMRI predicts memory
from event-related neuroimaging. Phil Trans decline following anterior temporal lobe resec-
Roy Soc Lond B Biol Sci 354:1307–1324 tion. J Neurol Neurosurg Psychiatry 79:686–693
121. Powell HW, Koepp MJ, Symms MR, Boulby 133. Rabin ML, Narayan VM, Kimberg DY,
PA, Salek-Haddadi A, Thompson PJ, Duncan Casasanto DJ, Glosser G, Tracy JI, French JA
JS et al (2005) Material-specific lateralization et al (2004) Functional MRI predicts post-
of memory encoding in the medial temporal surgical memory following temporal lobec-
lobe: blocked versus event-related design. tomy. Brain 127:2286–2298
NeuroImage 27:231–239 134. Janszky J, Jokeit H, Kontopoulou K, Mertens
122. Detre JA, Maccotta L, King D, Alsop DC, M, Ebner A, Pohlmann-Eden B, Woermann
Glosser G, D’Esposito M, Zarahn E et al (1998) FG (2005) Functional MRI predicts memory
Functional MRI lateralization of memory in performance after right mesiotemporal epi-
temporal lobe epilepsy. Neurology 50:926–932 lepsy surgery. Epilepsia 46:244–250
123. Doucet G, Osipowicz K, Sharan A, Sperling 135. Price CJ, Friston KJ (1999) Scanning patients
MR, Tracy JI (2013) Hippocampal functional with tasks they can perform. Hum Brain
connectivity patterns during spatial working Mapp 8:102–108
memory differ in right versus left temporal 136. von Helmholtz HLF (2004) Some laws con-
lobe epilepsy. Brain Connect 3:398–406 cerning the distribution of electric currents
124. Stretton J, Winston GP, Sidhu M, Bonelli in volume conductors with applications to
S, Centeno M, Vollmar C, Cleary RA et al experiments on animal electricity. Proc IEEE
(2013) Disrupted segregation of working 92:868–870
memory networks in temporal lobe epilepsy. 137. Geselowitz DB (2004) Introduction to
NeuroImage Clin 2:273–281 “some laws concerning the distribution of
125. Bellgowan PS, Binder JR, Swanson SJ, electric currents in volume conductors with
Hammeke TA, Springer JA, Frost JA, Mueller applications to experiments on animal elec-
WM et al (1998) Side of seizure focus pre- tricity”. Proc IEEE 92:864–867
dicts left medial temporal lobe activation dur- 138. Ives JR, Warach S, Schmitt F, Edelman
ing verbal encoding. Neurology 51:479–484 RR, Schomer DL (1993) Monitoring
126. Jokeit H, Okujava M, Woermann FG (2001) the patient’s EEG during echo planar
Memory fMRI lateralizes temporal lobe epi- MRI. Electroencephalogr Clin Neurophysiol
lepsy. Neurology 57:1786–1793 87:417–420
127. Golby AJ, Poldrack RA, Illes J, Chen D, 139. Detre JA, Alsop DC, Aguirre GK, Sperling
Desmond JE, Gabrieli JD (2002) Memory later- MR (1996) Coupling of cortical and thalamic
alization in medial temporal lobe epilepsy assessed ictal activity in human partial epilepsy: dem-
by functional MRI. Epilepsia 43:855–863 onstration by functional magnetic resonance
128. Richardson MP, Strange BA, Duncan JS, Dolan imaging. Epilepsia 37:657–661
RJ (2003) Preserved verbal memory function in 140. Krings T, Topper R, Reinges MH, Foltys
left medial temporal pathology involves reorgan- H, Spetzger U, Chiappa KH, Gilsbach JM
isation of function to right medial temporal lobe. et al (2000) Hemodynamic changes in sim-
NeuroImage 20(Suppl 1):S112–S119 ple partial epilepsy: a functional MRI study.
129. Powell HW, Richardson MP, Symms MR, Neurology 54:524–527
Boulby PA, Thompson PJ, Duncan JS, Koepp 141. Connelly A (1995) Ictal imaging using func-
MJ (2007) Reorganization of verbal and non- tional magnetic resonance. Magn Reson
verbal memory in temporal lobe epilepsy due Imaging 13:1233–1237
fMRI in Epilepsy 793
142. Jackson GD, Connelly A, Cross JH, Gordon pulse artifact and a method for its subtraction.
I, Gadian DG (1994) Functional mag- NeuroImage 8:229–239
netic resonance imaging of focal seizures. 155. Wendt RE 3rd, Rokey R, Vick GW 3rd,
Neurology 44:850–856 Johnston DL (1988) Electrocardiographic
143. Jorge J, Grouiller F, Ipek O, Stoermer R, gating and monitoring in NMR imaging.
Michel CM, Figueiredo P, van der Zwaag Magn Reson Imaging 6:89–95
W et al (2015) Simultaneous EEG-fMRI at 156. Poncelet BP, Wedeen VJ, Weisskoff RM,
ultra-high field: artifact prevention and safety Cohen MS (1992) Brain parenchyma motion:
assessment. NeuroImage 105:132–144 measurement with cine echo-planar MR
144. Arrubla J, Neuner I, Dammers J, Breuer L, imaging. Radiology 185:645–651
Warbrick T, Hahn D, Poole MS et al (2014) 157. Tenforde TS, Gaffey CT, Moyer BR, Budinger
Methods for pulse artefact reduction: expe- TF (1983) Cardiovascular alterations in
riences with EEG data recorded at 9.4 T Macaca monkeys exposed to stationary mag-
static magnetic field. J Neurosci Methods netic fields: experimental observations and
232:110–117 theoretical analysis. Bioelectromagnetics 4:1–9
145. Lemieux L, Allen PJ, Franconi F, Symms 158. Goldman RI, Stern JM, Engel J Jr, Cohen
MR, Fish DR (1997) Recording of EEG dur- MS (2000) Acquiring simultaneous EEG
ing fMRI experiments: patient safety. Magn and functional MRI. Clin Neurophysiol
Reson Med 38:943–952 111:1974–1980
146. Mirsattari SM, Lee DH, Jones D, Bihari F, 159. Benar C, Aghakhani Y, Wang Y, Izenberg
Ives JR (2004) MRI compatible EEG elec- A, Al-Asmi A, Dubeau F, Gotman J (2003)
trode system for routine use in the epilepsy Quality of EEG in simultaneous EEG-fMRI
monitoring unit and intensive care unit. Clin for epilepsy. Clin Neurophysiol 114:569–580
Neurophysiol 115:2175–2180 160. Chowdhury ME, Mullinger KJ, Glover
147. Konings MK, Bartels LW, Smits HF, Bakker P, Bowtell R (2014) Reference layer arte-
CJ (2000) Heating around intravascular fact subtraction (RLAS): a novel method of
guidewires by resonating RF waves. J Magn minimizing EEG artefacts during simultane-
Reson Imaging 12:79–85 ous fMRI. NeuroImage 84:307–319
148. Fischer H, Ladebeck R (1998) Echo-planar 161. Salek-Haddadi A, Lemieux L, Merschhemke
imaging image artifacts. In: Schmitt F, M, Diehl B, Allen PJ, Fish DR (2003) EEG
Stehling MK, Turner R (eds) Echo-planar quality during simultaneous functional MRI
imaging: theory, technique, and application. of interictal epileptiform discharges. Magn
Springer, Berlin, pp 179–200 Reson Imaging 21:1159–1166
149. Krakow K, Allen PJ, Symms MR, Lemieux 162. Liston AD, De Munck JC, Hamandi K, Laufs
L, Josephs O, Fish DR (2000) EEG record- H, Ossenblok P, Duncan JS, Lemieux L (2006)
ing during fMRI experiments: image quality. Analysis of EEG-fMRI data in focal epilepsy
Hum Brain Mapp 10:10–15 based on automated spike classification and Signal
150. Bonmassar G, Anami K, Ives J, Belliveau JW Space Projection. NeuroImage 31:1015–1024
(1999) Visual evoked potential (VEP) mea- 163. Allen PJ, Josephs O, Turner R (2000) A
sured by simultaneous 64-channel EEG and method for removing imaging artifact from
3T fMRI. Neuroreport 10:1893–1897 continuous EEG recorded during functional
151. Bonmassar G, Purdon PL, Jaaskelainen IP, MRI. NeuroImage 12:230–239
Chiappa K, Solo V, Brown EN, Belliveau JW 164. Lemieux L, Salek-Haddadi A, Lund TE, Laufs
(2002) Motion and ballistocardiogram artifact H, Carmichael D (2007) Modelling large
removal for interleaved recording of EEG and motion events in fMRI studies of patients with
EPs during MRI. NeuroImage 16:1127–1141 epilepsy. Magn Reson Imaging 25:894–901
152. Scarff CJ, Reynolds A, Goodyear BG, 165. Liston AD, Lund TE, Salek-Haddadi A,
Ponton CW, Dort JC, Eggermont JJ (2004) Hamandi K, Friston KJ, Lemieux L (2006)
Simultaneous 3-T fMRI and high-density Modelling cardiac signal as a confound in
recording of human auditory evoked poten- EEG-fMRI and its application in focal epi-
tials. NeuroImage 23:1129–1142 lepsy studies. NeuroImage 30:827–834
153. Klein C, Hanggi J, Luechinger R, Jancke L 166. Ellingson ML, Liebenthal E, Spanaki MV,
(2015) MRI with and without a high-den- Prieto TE, Binder JR, Ropella KM (2004)
sity EEG cap--what makes the difference? Ballistocardiogram artifact reduction in the
NeuroImage 106:189–197 simultaneous acquisition of auditory ERPS
154. Allen PJ, Polizzi G, Krakow K, Fish DR, and fMRI. NeuroImage 22:1534–1542
Lemieux L (1998) Identification of EEG 167. Kruggel F, Wiggins CJ, Herrmann CS, von
events in the MR scanner: the problem of Cramon DY (2000) Recording of the event-
794 Rachel C. Thornton et al.
related potentials during functional MRI at gram artifact removing. Magn Reson Imaging
3.0 Tesla field strength. Magn Reson Med 24:393–400
44:277–282 179. Mantini D, Perrucci MG, Cugini S, Ferretti A,
168. Sijbersa J, Van Audekerke J, Verhoye M, Van Romani GL, Del Gratta C (2007) Complete
der Linden A, Van Dyck D (2000) Reduction artifact removal for EEG recorded during
of ECG and gradient related artifacts in continuous fMRI using independent compo-
simultaneously recorded human EEG/MRI nent analysis. NeuroImage 34:598–607
data. Magn Reson Imaging 18:881–886 180. Maggioni E, Arrubla J, Warbrick T, Dammers
169. Kim KH, Yoon HW, Park HW (2004) J, Bianchi AM, Reni G, Tosetti M et al (2014)
Improved ballistocardiac artifact removal Removal of pulse artefact from EEG data
from the electroencephalogram recorded in recorded in MR environment at 3T. Setting
fMRI. J Neurosci Methods 135:193–203 of ICA parameters for marking artefactual
170. Wan X, Iwata K, Riera J, Ozaki T, Kitamura components: application to resting-state data.
M, Kawashima R (2006) Artifact reduction PLoS One 9, e112147
for EEG/fMRI recording: nonlinear reduc- 181. Debener S, Strobel A, Sorger B, Peters J,
tion of ballistocardiogram artifacts. Clin Kranczioch C, Engel AK, Goebel R (2007)
Neurophysiol 117:668–680 Improved quality of auditory event-related
171. Niazy RK, Beckmann CF, Iannetti GD, Brady potentials recorded simultaneously with 3-T
JM, Smith SM (2005) Removal of FMRI fMRI: removal of the ballistocardiogram arte-
environment artifacts from EEG data using fact. NeuroImage 34:587–597
optimal basis sets. NeuroImage 28:720–737 182. Harrison AH, Noseworthy MD, Reilly JP,
172. In MH, Lee SY, Park TS, Kim TS, Cho MH, Connolly JF (2014) Ballistocardiogram
Ahn YB (2006) Ballistocardiogram artifact correction in simultaneous EEG/ fMRI
removal from EEG signals using adaptive filter- recordings: a comparison of average artifact
ing of EOG signals. Physiol Meas 27:1227–1240 subtraction and optimal basis set methods
173. Eichele T, Specht K, Moosmann M, Jongsma using two popular software tools. Crit Rev
ML, Quiroga RQ, Nordby H, Hugdahl K Biomed Eng 42:95–107
(2005) Assessing the spatiotemporal evolution 183. Warach S, Ives JR, Schlaug G, Patel MR,
of neuronal activation with single-trial event- Darby DG, Thangaraj V, Edelman RR et al
related potentials and functional MRI. Proc (1996) EEG-triggered echo-planar functional
Natl Acad Sci U S A 102:17798–17803 MRI in epilepsy. Neurology 47:89–93
174. Otzenberger H, Gounot D, Foucher JR 184. Rothlubbers S, Relvas V, Leal A, Murta
(2005) P300 recordings during event-related T, Lemieux L, Figueiredo P (2015)
fMRI: a feasibility study. Brain Res Cogn Characterisation and reduction of the EEG
Brain Res 23:306–315 artefact caused by the helium cooling pump
175. Otzenberger H, Gounot D, Foucher JR in the MR environment: validation in epilepsy
(2007) Optimisation of a post-processing patient data. Brain Topogr 28:208–220
method to remove the pulse artifact from EEG 185. Bonmassar G, Schwartz DP, Liu AK,
data recorded during fMRI: an application to Kwong KK, Dale AM, Belliveau JW (2001)
P300 recordings during e-fMRI. Neurosci Spatiotemporal brain imaging of visual-evoked
Res 57:230–239 activity using interleaved EEG and fMRI
176. Srivastava G, Crottaz-Herbette S, Lau KM, recordings. NeuroImage 13:1035–1043
Glover GH, Menon V (2005) ICA-based 186. Huang-Hellinger FR, Breiter HC,
procedures for removing ballistocardiogram McCormack G, Cohen MS, Kwong KK,
artifacts from EEG data acquired in the MRI Sutton JP, Savoy RL et al (1995) Simultaneous
scanner. NeuroImage 24:50–60 functional magnetic resonance imaging and
177. Nakamura W, Anami K, Mori T, Saitoh O, electrophysiological recording. Hum Brain
Cichocki A, Amari S (2006) Removal of bal- Mapp 3:13–23
listocardiogram artifacts from simultaneously 187. Krakow K, Woermann FG, Symms MR, Allen
recorded EEG and fMRI data using inde- PJ, Lemieux L, Barker GJ, Duncan JS et al
pendent component analysis. IEEE Trans (1999) EEG-triggered functional MRI of
Biomed Eng 53:1294–1308 interictal epileptiform activity in patients with
178. Briselli E, Garreffa G, Bianchi L, Bianciardi partial seizures. Brain 122(Pt 9):1679–1688
M, Macaluso E, Abbafati M, Grazia Marciani 188. Krakow K, Allen PJ, Lemieux L, Symms
M et al (2006) An independent component MR, Fish DR (2000) Methodology: EEG-
analysis-based approach on ballistocardio- correlated fMRI. Adv Neurol 83:187–201
fMRI in Epilepsy 795
189. Krakow K, Lemieux L, Messina D, Scott CA, Non-invasive epileptic focus localization using
Symms MR, Duncan JS, Fish DR (2001) EEG-triggered functional MRI and electro-
Spatio-temporal imaging of focal interictal magnetic tomography. Electroencephalogr
epileptiform activity using EEG-triggered Clin Neurophysiol 106:508–512
functional MRI. Epileptic Disord 3:67–74 201. Symms MR, Allen PJ, Woermann FG, Polizzi
190. Al-Asmi A, Benar CG, Gross DW, Khani YA, G, Krakow K, Barker GJ, Fish DR et al
Andermann F, Pike B, Dubeau F et al (2003) (1999) Reproducible localization of interictal
fMRI activation in continuous and spike- epileptiform discharges using EEG-triggered
triggered EEG-fMRI studies of epileptic fMRI. Phys Med Biol 44:N161–N168
spikes. Epilepsia 44:1328–1339 202. Patel MR, Blum A, Pearlman JD, Yousuf
191. Anami K, Mori T, Tanaka F, Kawagoe N, Ives JR, Saeteng S, Schomer DL et al
Y, Okamoto J, Yarita M, Ohnishi T et al (1999) Echo-planar functional MR imaging
(2003) Stepping stone sampling for retriev- of epilepsy with concurrent EEG monitoring.
ing artifact-free electroencephalogram dur- AJNR Am J Neuroradiol 20:1916–1919
ing functional magnetic resonance imaging. 203. Krakow K, Wieshmann UC, Woermann FG,
NeuroImage 19:281–295 Symms MR, McLean MA, Lemieux L, Allen
192. Garreffa G, Carni M, Gualniera G, Ricci GB, PJ et al (1999) Multimodal MR imaging:
Bozzao L, De Carli D, Morasso P et al (2003) functional, diffusion tensor, and chemical
Real-time MR artifacts filtering during con- shift imaging in a patient with localization-
tinuous EEG/fMRI acquisition. Magn Reson related epilepsy. Epilepsia 40:1459–1462
Imaging 21:1175–1189 204. Lazeyras F, Blanke O, Perrig S, Zimine I,
193. Mullinger KJ, Yan WX, Bowtell R (2011) Golay X, Delavelle J, Michel CM et al (2000)
Reducing the gradient artefact in simultane- EEG-triggered functional MRI in patients
ous EEG-fMRI by adjusting the subject’s with pharmacoresistant epilepsy. J Magn
axial position. NeuroImage 54:1942–1950 Reson Imaging 12:177–185
194. Hoffmann A, Jager L, Werhahn KJ, 205. Lazeyras F, Blanke O, Zimine I, Delavelle
Jaschke M, Noachtar S, Reiser M (2000) J, Perrig SH, Seeck M (2000) MRI, (1)
Electroencephalography during functional H-MRS, and functional MRI during and
echo-planar imaging: detection of epileptic after prolonged nonconvulsive seizure activ-
spikes using post-processing methods. Magn ity. Neurology 55:1677–1682
Reson Med 44:791–798 206. Jager L, Werhahn KJ, Hoffmann A, Berthold
195. Sijbers J, Michiels I, Verhoye M, Van S, Scholz V, Weber J, Noachtar S et al (2002)
Audekerke J, Van der Linden A, Van Dyck D Focal epileptiform activity in the brain: detec-
(1999) Restoration of MR-induced artifacts tion with spike-related functional MR imaging-
in simultaneously recorded MR/EEG data. -preliminary results. Radiology 223:860–869
Magn Reson Imaging 17:1383–1391 207. Benar CG, Gross DW, Wang Y, Petre V, Pike
196. Hamandi K, Salek-Haddadi A, Laufs H, B, Dubeau F, Gotman J (2002) The BOLD
Liston A, Friston K, Fish DR, Duncan JS response to interictal epileptiform discharges.
et al (2006) EEG-fMRI of idiopathic and sec- NeuroImage 17:1182–1192
ondarily generalized epilepsies. NeuroImage 208. Benar CG, Grova C, Kobayashi E, Bagshaw
31:1700–1710 AP, Aghakhani Y, Dubeau F, Gotman J (2006)
197. Lemieux L, Salek-Haddadi A, Josephs O, Allen EEG-fMRI of epileptic spikes: concordance
P, Toms N, Scott C, Krakow K et al (2001) with EEG source localization and intracranial
Event-related fMRI with simultaneous and con- EEG. NeuroImage 30:1161–1170
tinuous EEG: description of the method and 209. Aghakhani Y, Kobayashi E, Bagshaw AP, Hawco
initial case report. NeuroImage 14:780–787 C, Benar CG, Dubeau F, Gotman J (2006)
198. Salek-Haddadi A, Merschhemke M, Lemieux Cortical and thalamic fMRI responses in partial
L, Fish DR (2002) Simultaneous EEG- epilepsy with focal and bilateral synchronous
correlated Ictal fMRI. NeuroImage 16:32–40 spikes. Clin Neurophysiol 117:177–191
199. Salek-Haddadi A, Diehl B, Hamandi K, 210. Zijlmans M, Huiskamp G, Hersevoort
Merschhemke M, Liston A, Friston K, M, Seppenwoolde JH, van Huffelen AC,
Duncan JS et al (2006) Hemodynamic cor- Leijten FS (2007) EEG-fMRI in the preop-
relates of epileptiform discharges: an EEG- erative work-up for epilepsy surgery. Brain
fMRI study of 63 patients with focal epilepsy. 130:2343–2353
Brain Res 1088:148–166 211. Thornton R, Vulliemoz S, Rodionov R,
200. Seeck M, Lazeyras F, Michel CM, Blanke O, Carmichael DW, Chaudhary UJ, Diehl B,
Gericke CA, Ives J, Delavelle J et al (1998) Laufs H et al (2011) Epileptic networks
796 Rachel C. Thornton et al.
their role in intra- and inter-subject variation dant with the epileptogenic region deter-
in fMRI. NeuroImage 26:960–964 mined by intracranial EEG. Magn Reson
235. Kobayashi E, Hawco CS, Grova C, Dubeau F, Imaging 24:367–371
Gotman J (2006) Widespread and intense 246. Grouiller F, Thornton R, Groening K, Spinelli
BOLD changes during brief focal electro- L, Duncan JS, Schaller K, Siniatchkin M et al
graphic seizures. Neurology 66:1049–1055 (2011) Localization of focal epileptic activity
236. Baumgartner C, Serles W, Leutmezer F, with EEG-FMRI informed by EEG voltage
Pataraia E, Aull S, Czech T, Pietrzyk U et al maps. Epilepsia 52:169–169
(1998) Preictal SPECT in temporal lobe epi- 247. Morgan VL, Price RR, Arain A, Modur P,
lepsy: regional cerebral blood flow is increased Abou-Khalil B (2004) Resting functional
prior to electroencephalography-seizure MRI with temporal clustering analysis for
onset. J Nucl Med 39:978–982 localization of epileptic activity without
237. Federico P, Abbott DF, Briellmann RS, EEG. NeuroImage 21:473–481
Harvey AS, Jackson GD (2005) Functional 248. Hamandi K, Salek Haddadi A, Liston A,
MRI of the pre-ictal state. Brain Laufs H, Fish DR, Lemieux L (2005) fMRI
128:1811–1817 temporal clustering analysis in patients with
238. Meletti S, Vaudano AE, Tassi L, Caruana F, frequent interictal epileptiform discharges:
Avanzini P (2015) Intracranial time-frequency comparison with EEG-driven analysis.
correlates of seizure-related negative BOLD NeuroImage 26:309–316
response in the sensory-motor network. Clin 249. Aguirre GK, Zarahn E, D’Esposito M (1998)
Neurophysiol 126:847–849 The variability of human, BOLD hemody-
239. Archer JS, Briellman RS, Abbott DF, namic responses. NeuroImage 8:360–369
Syngeniotis A, Wellard RM, Jackson GD 250. Salek-Haddadi A, Friston KJ, Lemieux L,
(2003) Benign epilepsy with centro-temporal Fish DR (2003) Studying spontaneous EEG
spikes: spike triggered fMRI shows somato- activity with fMRI. Brain Res Brain Res Rev
sensory cortex activity. Epilepsia 44:200–204 43:110–133
240. Archer JS, Briellmann RS, Syngeniotis A, 251. Kobayashi E, Bagshaw AP, Grova C, Dubeau
Abbott DF, Jackson GD (2003) Spike- F, Gotman J (2006) Negative BOLD
triggered fMRI in reading epilepsy: involve- responses to epileptic spikes. Hum Brain
ment of left frontal cortex working memory Mapp 27:488–497
area. Neurology 60:415–421 252. Lemieux L, Laufs H, Carmichael D, Paul JS,
241. Tousseyn S, Dupont P, Robben D, Goffin K, Walker MC, Duncan JS (2008) Noncanonical
Sunaert S, Van Paesschen W (2014) A reliable spike-related BOLD responses in focal epi-
and time-saving semiautomatic spike- lepsy. Hum Brain Mapp 29:329–345
template-based analysis of interictal EEG- 253. Lu Y, Bagshaw AP, Grova C, Kobayashi E,
fMRI. Epilepsia 55:2048–2058 Dubeau F, Gotman J (2006) Using voxel-
242. Bagshaw AP, Hawco C, Benar CG, Kobayashi specific hemodynamic response function in
E, Aghakhani Y, Dubeau F, Pike GB et al EEG-fMRI data analysis. NeuroImage
(2005) Analysis of the EEG-fMRI response 32:238–247
to prolonged bursts of interictal epileptiform 254. Watanabe S, An D, Safi-Harb M, Dubeau F,
activity. NeuroImage 24:1099–1112 Gotman J (2014) Hemodynamic response
243. Laufs H, Hamandi K, Salek-Haddadi A, function (HRF) in epilepsy patients with hip-
Kleinschmidt AK, Duncan JS, Lemieux L pocampal sclerosis and focal cortical dysplasia.
(2007) Temporal lobe interictal epileptic dis- Brain Topogr 27:613–619
charges affect cerebral activity in “default 255. Hawco CS, Bagshaw AP, Lu Y, Dubeau F,
mode” brain regions. Hum Brain Mapp Gotman J (2007) BOLD changes occur prior
28:1023–1032 to epileptic spikes seen on scalp
244. Diehl B, Salek-haddadi A, Fish DR, Lemieux EEG. NeuroImage 35:1450–1458
L (2003) Mapping of spikes, slow waves, and 256. Makiranta M, Ruohonen J, Suominen K,
motor tasks in a patient with malformation of Niinimaki J, Sonkajarvi E, Kiviniemi V,
cortical development using simultaneous Seppanen T et al (2005) BOLD signal
EEG and fMRI. Magn Reson Imaging increase preceeds EEG spike activity--a
21:1167–1173 dynamic penicillin induced focal epilepsy in
245. Laufs H, Hamandi K, Walker MC, Scott C, deep anesthesia. NeuroImage 27:715–724
Smith S, Duncan JS, Lemieux L (2006) EEG- 257. Siniatchkin M, Moeller F, Jacobs J, Stephani
fMRI mapping of asymmetrical delta activity U, Boor R, Wolff S, Jansen O et al (2007)
in a patient with refractory epilepsy is concor- Spatial filters and automated spike detection
798 Rachel C. Thornton et al.
based on brain topographies improve sensitiv- 269. Raichle ME, MacLeod AM, Snyder AZ,
ity of EEG-fMRI studies in focal epilepsy. Powers WJ, Gusnard DA, Shulman GL
NeuroImage 37:834–843 (2001) A default mode of brain function.
258. Rosenow F, Luders H (2001) Presurgical Proc Natl Acad Sci U S A 98:676–682
evaluation of epilepsy. Brain 124:1683–1700 270. Lengler U, Kafadar I, Neubauer BA, Krakow
259. Kobayashi E, Bagshaw AP, Benar CG, K (2007) fMRI correlates of interictal epilep-
Aghakhani Y, Andermann F, Dubeau F, tic activity in patients with idiopathic benign
Gotman J (2006) Temporal and extratempo- focal epilepsy of childhood. A simultaneous
ral BOLD responses to temporal lobe interic- EEG-functional MRI study. Epilepsy Res
tal spikes. Epilepsia 47:343–354 75:29–38
260. Lemieux L, Krakow K, Fish DR (2001) 271. Engel J Jr, International League Against E
Comparison of spike-triggered functional MRI (2001) A proposed diagnostic scheme for
BOLD activation and EEG dipole model local- people with epileptic seizures and with epi-
ization. NeuroImage 14:1097–1104 lepsy: report of the ILAE Task Force on
261. Bagshaw AP, Kobayashi E, Dubeau F, Pike Classification and Terminology. Epilepsia
GB, Gotman J (2006) Correspondence 42:796–803
between EEG-fMRI and EEG dipole localisa- 272. Archer JS, Abbott DF, Waites AB, Jackson
tion of interictal discharges in focal epilepsy. GD (2003) fMRI “deactivation” of the poste-
NeuroImage 30:417–425 rior cingulate during generalized spike and
262. Boor R, Jacobs J, Hinzmann A, Bauermann wave. NeuroImage 20:1915–1922
T, Scherg M, Boor S, Vucurevic G et al 273. Gotman J, Grova C, Bagshaw A, Kobayashi
(2007) Combined spike-related functional E, Aghakhani Y, Dubeau F (2005) Generalized
MRI and multiple source analysis in the non- epileptic discharges show thalamocortical
invasive spike localization of benign rolandic activation and suspension of the default state
epilepsy. Clin Neurophysiol 118:901–909 of the brain. Proc Natl Acad Sci U S A
263. Salek-Haddadi A, Lemieux L, Fish DR (2002) 102:15236–15240
Role of functional magnetic resonance imag- 274. Laufs H, Lengler U, Hamandi K, Kleinschmidt
ing in the evaluation of patients with malfor- A, Krakow K (2006) Linking generalized
mations caused by cortical development. spike-and-wave discharges and resting state
Neurosurg Clin N Am 13:63–69, viii brain activity by using EEG/fMRI in a patient
264. Federico P, Archer JS, Abbott DF, Jackson with absence seizures. Epilepsia 47:444–448
GD (2005) Cortical/subcortical BOLD 275. Meeren HK, Pijn JP, Van Luijtelaar EL,
changes associated with epileptic discharges: Coenen AM, Lopes da Silva FH (2002)
an EEG-fMRI study at 3 T. Neurology Cortical focus drives widespread corticotha-
64:1125–1130 lamic networks during spontaneous absence
265. Kobayashi E, Bagshaw AP, Jansen A, seizures in rats. J Neurosci 22:1480–1495
Andermann F, Andermann E, Gotman J, 276. Steriade M, Dossi RC, Nunez A (1991)
Dubeau F (2005) Intrinsic epileptogenicity in Network modulation of a slow intrinsic oscil-
polymicrogyric cortex suggested by EEG- lation of cat thalamocortical neurons impli-
fMRI BOLD responses. Neurology cated in sleep delta waves: cortically induced
64:1263–1266 synchronization and brainstem cholinergic
266. Kobayashi E, Bagshaw AP, Grova C, Gotman suppression. J Neurosci 11:3200–3217
J, Dubeau F (2006) Grey matter heterotopia: 277. Hamandi K, Laufs H, Noth U, Carmichael
what EEG-fMRI can tell us about epileptoge- DW, Duncan JS, Lemieux L (2008) BOLD
nicity of neuronal migration disorders. Brain and perfusion changes during epileptic gener-
129:366–374 alised spike wave activity. NeuroImage
267. Kobayashi E, Bagshaw AP, Gotman J, Dubeau 39:608–618
F (2007) Metabolic correlates of epileptic 278. Moeller F, Siebner HR, Wolff S, Muhle H,
spikes in cerebral cavernous angiomas. Granert O, Jansen O, Stephani U et al (2008)
Epilepsy Res 73:98–103 Simultaneous EEG-fMRI in drug-naive chil-
268. Wiebe S, Blume WT, Girvin JP, Eliasziw M dren with newly diagnosed absence epilepsy.
(2001) Effectiveness and G efficiency of sur- Epilepsia 49:1510–1519
gery for temporal lobe epilepsy study. A ran- 279. Moeller F, Maneshi M, Pittau F, Gholipour T,
domized, controlled trial of surgery for Bellec P, Dubeau F, Grova C et al (2011)
temporal-lobe epilepsy. N Engl J Med Functional connectivity in patients with idiopathic
345:311–318 generalized epilepsy. Epilepsia 52:515–522
fMRI in Epilepsy 799
280. Vaudano AE, Laufs H, Kiebel SJ, Carmichael during 3-Hz spike-and-wave complexes in
DW, Hamandi K, Guye M, Thornton R et al generalized epilepsy. Magn Reson Imaging
(2009) Causal hierarchy within the thalamo- 22:1441–1444
cortical network in spike and wave discharges. 289. Daunizeau J, Grova C, Marrelec G, Mattout
PLoS One 4, e6475 J, Jbabdi S, Pelegrini-Issac M, Lina JM et al
281. Moeller F, LeVan P, Muhle H, Stephani U, (2007) Symmetrical event-related EEG/
Dubeau F, Siniatchkin M, Gotman J (2010) fMRI information fusion in a variational
Absence seizures: individual patterns revealed Bayesian framework. NeuroImage
by EEG-fMRI. Epilepsia 51:2000–2010 36:69–87
282. Moeller F, Muthuraman M, Stephani U, 290. Jezzard P, Clare S (1999) Sources of distor-
Deuschl G, Raethjen J, Siniatchkin M (2013) tion in functional MRI data. Hum Brain
Representation and propagation of epileptic Mapp 8:80–85
activity in absences and generalized photopar- 291. Bagshaw AP, Torab L, Kobayashi E, Hawco
oxysmal responses. Hum Brain Mapp 34(8): C, Dubeau F, Pike GB, Gotman J (2006)
1896–1909 EEG-fMRI using z-shimming in patients with
283. Shmuel A, Augath M, Oeltermann A, temporal lobe epilepsy. J Magn Reson
Logothetis NK (2006) Negative functional Imaging 24:1025–1032
MRI response correlates with decreases in 292. Jezzard P, Balaban RS (1995) Correction for
neuronal activity in monkey visual area V1. geometric distortion in echo planar images
Nat Neurosci 9:569–577 from B0 field variations. Magn Reson Med
284. Binnie CD (2003) Cognitive impairment 34:65–73
during epileptiform discharges: is it ever justi- 293. Hutton C, Bork A, Josephs O, Deichmann R,
fiable to treat the EEG? Lancet Neurology Ashburner J, Turner R (2002) Image distor-
2:725–730 tion correction in fMRI: A quantitative evalu-
285. Todd N, Josephs O, Callaghan MF, Lutti A, ation. NeuroImage 16:217–240
Weiskopf N (2015) Prospective motion cor- 294. Niendorf T (1999) On the application of
rection of 3D echo-planar imaging data for susceptibility-weighted ultra-fast low-angle
functional MRI using optical tracking. RARE experiments in functional MR imag-
NeuroImage 113:1–12 ing. Magn Reson Med 41:1189–1198
286. Rodionov R, De Martino F, Laufs H, 295. Deichmann R, Josephs O, Hutton C, Corfield
Carmichael DW, Formisano E, Walker M, DR, Turner R (2002) Compensation of
Duncan JS et al (2007) Independent compo- susceptibility-induced BOLD sensitivity losses
nent analysis of interictal fMRI in focal epi- in echo-planar fMRI imaging. NeuroImage
lepsy: comparison with general linear 15:120–135
model-based EEG-correlated 296. Deichmann R, Gottfried JA, Hutton C,
fMRI. NeuroImage 38:488–500 Turner R (2003) Optimized EPI for fMRI
287. Merlet I, Gotman J (2001) Dipole modeling studies of the orbitofrontal cortex.
of scalp electroencephalogram epileptic dis- NeuroImage 19:430–441
charges: correlation with intracerebral fields. 297. Glover GH, Li TQ, Ress D (2000) Image-
Clin Neurophysiol 112:414–430 based method for retrospective correction of
288. Liston AD, Salek-Haddadi A, Kiebel SJ, physiological motion effects in fMRI:
Hamandi K, Turner R, Lemieux L (2004) RETROICOR. Magn Reson Med
The MR detection of neuronal depolarization 44:162–167
Chapter 25
fMRI in Neurosurgery
Oliver Ganslandt, Christopher Nimsky, Michael Buchfelder,
and Peter Grummich
Abstract
Functional magnetic resonance imaging has evolved from a basic research application to a useful clinical
tool that also has found its place in modern neurosurgery. The localization of functional important brain
areas as language and sensorimotor cortex has been the focus of numerous investigations and can now be
implemented in neurosurgical planning. Since the neurosurgeon must have detailed knowledge about the
individual anatomy and related neurological function to resect a brain tumor with the highest safety, the
need for individualized maps of brain function is essential. Advanced fMRI techniques and modern imag-
ing methods contribute significantly to brain mapping as do already established concepts of electrophysi-
ological monitoring and the Wada test. The implementation of functional maps into neuronavigation
systems enables the surgeon to superimpose anatomy and function to the surgical site. This chapter
describes our experience with the use of fMRI in neurosurgery.
1 Introduction
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_25, © Springer Science+Business Media New York 2016
801
802 Oliver Ganslandt et al.
[14, 16, 17], which also performed comparisons with direct motor
stimulation. Language fMRI has been found to be an alternative to
the invasive Wada test [18–20] for language lateralization.
Furthermore, fMRI has been used to predict memory localization
[21]. Concerning the reliability of fMRI-localization of speech
areas in the frontotemporal cortex, as compared with direct electri-
cal stimulation, the neurosurgical community is still reluctant to
rely on fMRI language alone, since inconsistent agreement has
been found between activation sites by fMRI naming and verb-
generation tasks and cortical stimulation [5]. As language fMRI
and intraoperative electrophysiological monitoring use different
physiological mechanisms the results of language fMRI are not per
se to be considered wrong. The ongoing clinical use of language
fMRI in our department has shown that this modality can be used
with the same results as awake craniotomy.
2 Methods
Fig. 1 fMRI-guided epilepsy surgery for ganglioglioma in the left hippocampus. We localized activity for factual
knowledge in the parahippocampal gyrus (crosshair). The segmented area is depicted in green in the right
image showing the view through the navigation microscope. Also shown in purple are the fiber tracts of the
visual pathway and occipitofrontal fasciculus (middle)
fMRI in Neurosurgery 805
(a) In case of tumor location in the inferior parietal area close to the
intraparietal sulcus, we use an arithmetic task so that besides
Wernicke’s area (activated by reading, adding numbers, and for-
mulating the result) the cortex for calculation in the intraparietal
sulcus is also activated and so can be spared during surgery.
In case of tumor location close to Broca’s area, we select
language tasks that are also expressive and demand grammati-
cal abilities, because these may increase activity in Broca’s area.
This happens during the verb-generation task, but also the
verb conjugation task gives suitable Broca’s area activations.
(b) For patients with reduced abilities, simple tasks are selected to
obtain reliable results. Especially in patients who suffer from
word finding disorders, we avoid the picture-naming task. For
patients with better cognitive performance, we select one or
more complex tasks such as verb-generation task, because these
are reported to show a more clear lateralization, whereas in
patients who have difficulties in this complex task the activa-
tion is usually worse than with a simple paradigm.
For motion correction, we apply an image-based prospective
acquisition correction by applying interpolation in the k-space
[22]. We produce activation maps by analyzing the correlation
between signal intensity and a square wave reference function for
each pixel according to the paradigm. Pixels exceeding a signifi-
cance threshold (typical correlations above a threshold of 0.3 with
p < 0.000045) are displayed, if at least six contiguous voxels con-
stitute a cluster, to eliminate isolated voxels. We align the func-
tional slices to magnetization prepared rapid acquisition gradient
echo (MPRAGE) images (160 slices of 1 mm slice thickness).
3 Results
3.1 Localizations Since 2002, we have investigated preoperative fMRI with motor or
sensory stimulation in 515 cases. Of these patients, 205 underwent
tumor resection and 75 had stereotactic brain biopsy. In five addi-
tional patients, invasive electrodes were implanted by fMRI guid-
ance for chronic recording of epileptic discharges.
For language and memory testing, we examined 623 cases and
used additional information from MEG studies. Of them, 465
underwent tumor resection and 53 had stereotactic brain biopsy.
The remaining patients either obtained radiation therapy, endovas-
cular treatment, or were just enrolled in a “wait-and-see” protocol.
It was possible to localize the primary motor and sensory cor-
tex as well as the supplementary motor area (SMA) in all examined
cases (Figs. 2 and 3). Only in one case, the motor activity of the toe
was not detectable by fMRI because of tumor infiltration. However,
in this patient it was possible to obtain motor activation from
nearby muscle representations of the motor homunculus.
806 Oliver Ganslandt et al.
Fig. 2 fMRI activations during movement of left foot in a patient with an oligoastrocytoma (WHO III) in the right
parietal lobe. In front of the activation of the motor cortex (posterior wall of precentral gyrus), activation of the
supplementary motor area (SMA) is also evident
Fig. 3 Comparison of fMRI motor activations during arm movement and sensory
stimulation of the arm (oligoastrocytoma WHO III, same patient as Fig. 2)
Fig. 4 fMRI activations during movement of toe, leg, arm, and fingers (note that the lesion, a cavernoma in the
left motor cortex, is located between the cortical representation of the arm and that of the leg in the precentral
gyrus)
3.2 Laterality Although it is generally accepted that the majority of people has left
hemispheric language dominance, the true number of atypical (right)
dominance is unknown. Studies using the Wada test showed an inci-
dence of left hemispheric dominance in right handers in a range of
63–96 % and a right hemispheric dominance for left handers and ambi-
dextrous patients in 48–75 % [23]. Furthermore, it is thought that
there are varying degrees of language dominance in the population.
In certain circumstances, activity can be located in both hemi-
spheres or reorganization to the other hemisphere could have been
occurred. FMRI is a useful method to clarify this. If activity is only
found in one hemisphere or the activity on one side is much stron-
ger than the activity on the other side, then it is clear that the active
area has to be spared during surgery.
It is important to know that certain stimulation tasks and modali-
ties show more lateralized activations than others. In case of complex
motor tasks, the ipsilateral hemisphere may also show activation.
For the localization of language activity, we found that visual
stimulation shows a more accentuated lateralization than acoustic
stimulation [24]. Stimulation with words, especially in a complex task,
shows a stronger lateralization than a picture-naming task, a finding
that was also described by Herholz et al. [25, 26]. In rare cases, it can
occur that not all language areas are located on the same side.
Fig. 5 Microscopic view with neuronavigation markers showing the sensory activation of the arm area in light
blue and the pyramidal tract in purple (oligoastrocytoma WHO III, same patient as Fig. 2)
fMRI in Neurosurgery 809
Fig. 6 fMRI activation of Broca’s and Wernicke’s areas and primary motor cortex
after a reading paradigm. The functional mapping was requested to plan surgery
of a cavernoma, which was located between Broca’s area and the motor cortex
Fig. 7 Same patient as Fig. 5. Segmentation lines indicating Broca’s area (left)
and motor cortex of tongue (right). The figure shows the beginning of the corti-
cotomy on a trajectory that spared the eloquent cortices (cross). Postoperatively
the patient was neurologically intact
Fig. 8 Intraoperative MRI showing the outcome of the fMRI-guided tumor resection. Note that the tumor was
removed sparing the sensory cortex (oligoastrocytoma WHO III right, same patient as Fig. 2)
fMRI in Neurosurgery 811
3.4 Problems Sometimes the BOLD activations are not clearly visible in spite of
with the BOLD Effect the fact that the function is there, as confirmed by MEG
812 Oliver Ganslandt et al.
Fig. 9 Wernicke and Broca activity during reading of fragmentary sentences with mistakes. Comparison
between fMRI (orange) and MEG beamformer localizations at 500 ms (light blue). With fMRI a bilateral activa-
tion in the operculum frontale and in the superior temporal sulcus can be seen. With MEG, activity is only seen
in the right hemisphere. In the first and second image in the lower row, activity of the insula can be seen with
MEG only. Left-handed patient with astrocytoma WHO II
fMRI in Neurosurgery 813
they were only found in the right hemisphere. The activity detected
by MEG in the right insula was not found by fMRI.
Other reasons that might lead to suboptimal fMRI results are
continuous brain activation during rest or a very short activation of
brain areas. This may be the reason why memory activity in the
hippocampus is difficult to find by fMRI.
4 Conclusions
References
1. Penfield W, Rasmussen T (1950) The cerebral 5. Roux FE, Boulanouar K, Lotterie JA, Mejdoubi
cortex of man. A clinical study of localization M, LeSage JP et al (2003) Language functional
of function. Macmillan, New York magnetic resonance imaging in preoperative
2. Berger MS, Rostomily RC (1997) Low grade assessment of language areas: correlation with
gliomas: functional mapping resection strate- direct cortical stimulation. Neurosurgery
gies, extent of resection, and outcome. 52:1335–1345, discussion 1345–1337
J Neurooncol 34:85–101 6. Nimsky C, Ganslandt O, Kober H, Moller M,
3. Duffau H, Capelle L, Denvil D, Sichez N, Ulmer S et al (1999) Integration of functional
Gatignol P et al (2003) Usefulness of intraop- magnetic resonance imaging supported by
erative electrical subcortical mapping during magnetoencephalography in functional neuro-
surgery for low-grade gliomas located within navigation. Neurosurgery 44:1249–1255, dis-
eloquent brain regions: functional results in a cussion 1255–1246
consecutive series of 103 patients. J Neurosurg 7. Rutten GJ, Ramsey N, Noordmans HJ,
98:764–778 Willems P, van Rijen P et al (2003) Toward
4. Kober H, Moller M, Nimsky C, Vieth J, functional neuronavigation: implementation of
Fahlbusch R et al (2001) New approach to functional magnetic resonance imaging data in
localize speech relevant brain areas and hemi- a surgical guidance system for intraoperative
spheric dominance using spatially filtered mag- identification of motor and language cortices.
netoencephalography. Hum Brain Mapp Technical note and illustrative case. Neurosurg
14:236–250 Focus 15:E6
814 Oliver Ganslandt et al.
8. Rossler K, Sommer B, Grummich P, Hamer uation of temporal and frontal language domi-
HM, Pauli E et al (2015) Risk reduction in nance compared with the Wada test. Neurology
dominant temporal lobe epilepsy surgery com- 54:1625–1633
bining fMRI/DTI maps, neuronavigation and 20. Stippich C, Rapps N, Dreyhaupt J, Durst A, Kress
intraoperative 1.5-Tesla MRI. Stereotact Funct B et al (2007) Localizing and lateralizing lan-
Neurosurg 93:168–177 guage in patients with brain tumors: feasibility of
9. Zhang J, Chen X, Zhao Y, Wang F, Li F et al routine preoperative functional MR imaging in 81
(2015) Impact of intraoperative magnetic reso- consecutive patients. Radiology 243:828–836
nance imaging and functional neuronavigation 21. Branco DM, Suarez RO, Whalen S, O’Shea JP,
on surgical outcome in patients with gliomas Nelson AP et al (2006) Functional MRI of mem-
involving language areas. Neurosurg Rev ory in the hippocampus: laterality indices may be
38:319–330, discussion 330 more meaningful if calculated from whole voxel
10. Sun GC, Chen XL, Yu XG, Zhang M, Liu G distributions. Neuroimage 32:592–602
et al (2015) Functional neuronavigation- 22. Thesen S, Heid O, Mueller E, Schad R (2000)
guided transparieto-occipital cortical resection Prospective acquisition correction for head
of meningiomas in trigone of lateral ventricle. motion with image-base tracking for real-time
World Neurosurg 84(3):756–765 fMRI. Magn Reson Med 44:457–465
11. Duffau H, Denvil D, Capelle L (2002) Long 23. Springer JA, Binder JR, Hammeke TA,
term reshaping of language, sensory, and Swanson SJ, Frost JA et al (1999) Language
motor maps after glioma resection: a new dominance in neurologically normal and epi-
parameter to integrate in the surgical strategy. lepsy subjects: a functional MRI study. Brain
J Neurol Neurosurg Psychiatry 72:511–516 122(Pt 11):2033–2046
12. Grummich P, Nimsky C, Fahlbusch R, 24. Grummich P, Nimsky C, Pauli E, Buchfelder
Ganslandt O (2005) Observation of unaver- M, Ganslandt O (2006) Combining fMRI and
aged giant MEG activity from language areas MEG increases the reliability of presurgical lan-
during speech tasks in patients harboring brain guage localization: a clinical study on the dif-
lesions very close to essential language areas: ference between and congruence of both
expression of brain plasticity in language pro- modalities. Neuroimage 32:1793–1803
cessing networks? Neurosci Lett 380:143–148 25. Herholz K, Reulen HJ, von Stockhausen HM,
13. Tieleman A, Vandemaele P, Seurinck R, Thiel A, Ilmberger J et al (1997) Preoperative acti-
Deblaere K, Achten E (2007) Comparison vation and intraoperative stimulation of language-
between functional magnetic resonance imag- related areas in patients with glioma. Neurosurgery
ing at 1.5 and 3 Tesla: effect of increased field 41:1253–1260, discussion 1260–1262
strength on 4 paradigms used during presurgi- 26. Lazar RM, Marshall RS, Pile-Spellman J, Duong
cal work-up. Invest Radiol 42:130–138 HC, Mohr JP et al (2000) Interhemispheric
14. Matthews PM, Jezzard P (2004) Functional transfer of language in patients with left frontal
magnetic resonance imaging. J Neurol cerebral arteriovenous malformation.
Neurosurg Psychiatry 75:6–12 Neuropsychologia 38:1325–1332
15. Tharin S, Golby A (2007) Functional brain map- 27. Duffau H, Lopes M, Arthuis F, Bitar A, Sichez
ping and its applications to neurosurgery. JP et al (2005) Contribution of intraoperative
Neurosurgery 60:185–201, discussion 201–202 electrical stimulations in surgery of low grade
16. Majos A, Tybor K, Stefanczyk L, Goraj B gliomas: a comparative study between two series
(2005) Cortical mapping by functional mag- without (1985–96) and with (1996–2003) func-
netic resonance imaging in patients with brain tional mapping in the same institution. J Neurol
tumors. Eur Radiol 15:1148–1158 Neurosurg Psychiatry 76:845–851
17. Roux FE, Boulanouar K, Ibarrola D, Tremoulet 28. Berman JI, Berger MS, Chung SW, Nagarajan
M, Chollet F et al (2000) Functional MRI and SS, Henry RG (2007) Accuracy of diffusion
intraoperative brain mapping to evaluate brain plas- tensor magnetic resonance imaging tractogra-
ticity in patients with brain tumours and hemipare- phy assessed using intraoperative subcortical
sis. J Neurol Neurosurg Psychiatry 69:453–463 stimulation mapping and magnetic source
18. Desmond JE, Sum JM, Wagner AD, Demb JB, imaging. J Neurosurg 107:488–494
Shear PK et al (1995) Functional MRI mea- 29. Haglund MM, Berger MS, Shamseldin M,
surement of language lateralization in Wada- Lettich E, Ojemann GA (1994) Cortical local-
tested patients. Brain 118(Pt 6):1411–1419 ization of temporal lobe language sites in
19. Lehericy S, Cohen L, Bazin B, Samson S, patients with gliomas. Neurosurgery 34:567–
Giacomini E et al (2000) Functional MR eval- 576, discussion 576
fMRI in Neurosurgery 815
30. Hamzei F, Knab R, Weiller C, Roether 32. Holodny AI, Schulder M, Liu WC, Wolko J,
J (2002) Intra- und extrakranielle Maldjian JA et al (2000) The effect of brain
Gefäßstenosen beeinflussen BOLD Antwort. tumors on BOLD functional MR imaging acti-
Aktuelle Neurologie 29:231 vation in the adjacent motor cortex: implica-
31. Holodny AI, Schulder M, Liu WC, Maldjian tions for image-guided neurosurgery. AJNR
JA, Kalnin AJ (1999) Decreased BOLD func- Am J Neuroradiol 21:1415–1422
tional MR activation of the motor and sen- 33. Schreiber A, Hubbe U, Ziyeh S, Hennig
sory cortices adjacent to a glioblastoma J (2000) The influence of gliomas and nonglial
multiforme: implications for image-guided space-occupying lesions on blood-oxygen-
neurosurgery. AJNR Am J Neuroradiol level-dependent contrast enhancement. AJNR
20:609–612 Am J Neuroradiol 21:1055–1063
Chapter 26
Abstract
Increasing societal expectations for new drugs, lack of confidence in short-term endpoints related to long-term
outcomes for chronic neurological and psychiatric diseases and rising costs of development in an increasing
cost-constrained market all have created a sense of crisis in CNS drug development. New approaches are
needed. For some time, the potential of clinical functional imaging for more confident progression from pre-
clinical to clinical development stages has been recognized. Pharmacological functional MRI (fMRI), which
refers specifically to applications of fMRI to questions in drug development, provides one set of these tools.
With related structural MRI measures, relatively high resolution data concerning target, disease-relevant
pathophysiology and effects of therapeutic interventions can be related to brain functional anatomy. In this
chapter, current and potential applications of pharmacological fMRI for target validation, patient stratification
and characterization of therapeutic molecule pharmacokinetics and pharmacodynamics are reviewed.
Challenges to better realizing the promise of pharmacological fMRI will be discussed. The review concludes
that there is a strong rationale for greater use of pharmacological fMRI particularly for early phase studies, but
also outlines the need for preclinical and early clinical development to be more seamlessly integrated, for
greater harmonization of clinical imaging methodologies and for sharing of data to facilitate these goals.
1 Introduction
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_26, © Springer Science+Business Media New York 2016
817
818 Paul M. Matthews
Disease
selection
Gene Taget to PoC to
Lead to FTIH to Phase File and Life cycle
function lead Preclinical commit to
candidate PoC III launch management
Target to target compound Phase III
family
selection
Fig. 1 The “critical path” for drug development. Pharmacological MRI has the potential to enhance the effi-
ciency of early clinical development with better translation of biological concepts from preclinical to clinical
studies, providing a new pharmacodynamic measure and enhancing potential in proof-of-mechanism studies
(FTIH first time in human study, PoC proof of concept study)
3 Target Validation
4 Patient Stratification
5 Pharmacodynamics
Fig. 2 Pharmacological fMRI can be performed in both animals and humans to assess correspondences in
tests of mechanisms. (a) Pharmacological fMRI results with metamphetamine challenge of a rodent, identify-
ing major regions in the monamine network as sites of direct or indirect action (Mctx motor cortex, PrL pre-
limbic medial prefrontal cortex, thal thalamus, SSctx somatosensory cortex, AcbSh shell of the nucleus
accumbens, VTA ventral tegmental area) (Images courtesy of Dr. A. Bifone, GSK, Verona). (b) A similar pharma-
cological fMRI experiment with acute amphetamine infusion in human subjects performed using “mind racing”
as a behavioral index of drug effects identified comparable elements of the core response network (OFC
orbitofrontal cortex, ACC anterior cingulate cortex, NAC nucleus accumbens)
Acknowledgements
References
1. Trusheim MR, Berndt ER, Douglas FL (2007) Drug Discov Today Technol
Stratified medicine: strategic and economic 10(3):e343–e350
implications of combining drugs and clinical bio- 3. Uppoor RS et al (2008) The use of imaging in
markers. Nat Rev Drug Discov 6(4):287–293 the early development of neuropharmacologi-
2. Matthews PM et al (2013) Technologies: cal drugs: a survey of approved NDAs. Clin
preclinical imaging for drug development. Pharmacol Ther 84(1):69–74
Pharmacological Applications of fMRI 829
4. Borsook D, Becerra L, Fava M (2013) Use of 18. Goense JB, Logothetis NK (2008)
functional imaging across clinical phases in Neurophysiology of the BOLD fMRI signal in
CNS drug development. Transl Psychiatry awake monkeys. Curr Biol 18(9):631–640
3:e282 19. Smith SM (2012) The future of FMRI connec-
5. Mathiesen C et al (1998) Modification of tivity. Neuroimage 62(2):1257–1266
activity-dependent increases of cerebral blood 20. FSL—FslWiki (2015) http://fsl.fmrib.ox.ac.
flow by excitatory synaptic activity and spikes in uk/fsl/fslwiki/%5D
rat cerebellar cortex. J Physiol 512(Pt 21. Iannetti GD, Wise RG (2007) BOLD func-
2):555–566 tional MRI in disease and pharmacological
6. Logothetis NK (2003) The underpinnings of studies: room for improvement? Magn Reson
the BOLD functional magnetic resonance Imaging 25(6):978–988
imaging signal. J Neurosci 23(10):3963–3971 22. Beckmann CF, Smith SM (2005) Tensorial
7. Mukamel R et al (2005) Coupling between extensions of independent component analysis
neuronal firing, field potentials, and FMRI in for multisubject FMRI analysis. Neuroimage
human auditory cortex. Science 25(1):294–311
309(5736):951–954 23. Hodkinson DJ et al (2012) Differential effects
8. Caesar K, Thomsen K, Lauritzen M (2003) of anaesthesia on the phMRI response to acute
Dissociation of spikes, synaptic activity, and ketamine challenge. Br J Med Med Res
activity-dependent increments in rat cerebellar 2(3):373–385
blood flow by tonic synaptic inhibition. Proc 24. Littlewood CL et al (2006) Using the BOLD
Natl Acad Sci U S A 100(26):16000–16005 MR signal to differentiate the stereoisomers of
9. Minzenberg MJ (2012) Pharmacological MRI ketamine in the rat. Neuroimage
approaches to understanding mechanisms of 32(4):1733–1746
drug action. Curr Top Behav Neurosci 25. Roberts TJ, Williams SC, Modo M (2008) A
11:365–388 pharmacological MRI assessment of dizocil-
10. Girouard H, Iadecola C (2006) Neurovascular pine (MK-801) in the 3-nitroproprionic acid-
coupling in the normal brain and in hyperten- lesioned rat. Neurosci Lett 444(1):42–47
sion, stroke, and Alzheimer disease. J Appl 26. Miyamoto S et al (2000) Effects of ketamine,
Physiol (1985) 100(1):328–335 MK-801, and amphetamine on regional brain
11. Suri S et al (2015) Reduced cerebrovascular 2-deoxyglucose uptake in freely moving mice.
reactivity in young adults carrying the APOE Neuropsychopharmacology 22(4):400–412
epsilon4 allele. Alzheimers Dement 11(6):648– 27. Homayoun H, Jackson ME, Moghaddam B
657.e1 (2005) Activation of metabotropic glutamate
12. Glodzik L et al (2013) Cerebrovascular reactiv- 2/3 receptors reverses the effects of NMDA
ity to carbon dioxide in Alzheimer’s disease. receptor hypofunction on prefrontal cortex
J Alzheimers Dis 35(3):427–440 unit activity in awake rats. J Neurophysiol
13. Kwong KK et al (1992) Dynamic magnetic 93(4):1989–2001
resonance imaging of human brain activity dur- 28. Deakin JF et al (2008) Glutamate and the neu-
ing primary sensory stimulation. Proc Natl ral basis of the subjective effects of ketamine: a
Acad Sci U S A 89(12):5675–5679 pharmaco-magnetic resonance imaging study.
14. Ogawa S et al (1990) Oxygenation-sensitive Arch Gen Psychiatry 65(2):154–164
contrast in magnetic resonance image of rodent 29. Gottesman II, Gould TD (2003) The endo-
brain at high magnetic fields. Magn Reson phenotype concept in psychiatry: etymology
Med 14(1):68–78 and strategic intentions. Am J Psychiatry
15. Belliveau JW et al (1991) Functional mapping 160(4):636–645
of the human visual cortex by magnetic reso- 30. Callicott JH et al (2005) Variation in DISC1
nance imaging. Science 254(5032):716–719 affects hippocampal structure and function and
16. Mezue M et al (2014) Optimization and reli- increases risk for schizophrenia. Proc Natl Acad
ability of multiple postlabeling delay pseudo- Sci U S A 102(24):8627–8632
continuous arterial spin labeling during rest 31. Egan MF et al (2004) Variation in GRM3
and stimulus-induced functional task activa- affects cognition, prefrontal glutamate, and
tion. J Cereb Blood Flow Metab risk for schizophrenia. Proc Natl Acad Sci U S
34(12):1919–1927 A 101(34):12604–12609
17. Brookes MJ et al (2011) Investigating the elec- 32. Egan MF et al (2001) Effect of COMT
trophysiological basis of resting state networks Val108/158 Met genotype on frontal lobe
using magnetoencephalography. Proc Natl function and risk for schizophrenia. Proc Natl
Acad Sci U S A 108(40):16783–16788 Acad Sci U S A 98(12):6917–6922
830 Paul M. Matthews
33. David SP et al (2005) Ventral striatum/nucleus functional MRI. Proc Natl Acad Sci U S A
accumbens activation to smoking-related pic- 101(13):4637–4642
torial cues in smokers and nonsmokers: a func- 48. Raichle ME, Snyder AZ (2007) A default
tional magnetic resonance imaging study. Biol mode of brain function: a brief history of an
Psychiatry 58(6):488–494 evolving idea. Neuroimage 37(4):1083–1090,
34. Myrick H et al (2004) Differential brain discussion 1097–1099
activity in alcoholics and social drink- 49. Petrella JR et al (2011) Default mode network
ers to alcohol cues: relationship to craving. connectivity in stable vs progressive mild cog-
Neuropsychopharmacology 29(2):393–402 nitive impairment. Neurology 76(6):511–517
35. Reuter J et al (2005) Pathological gambling is 50. Sheline YI, Raichle ME (2013) Resting
linked to reduced activation of the mesolimbic state functional connectivity in preclini-
reward system. Nat Neurosci 8(2):147–148 cal Alzheimer’s disease. Biol Psychiatry
36. Paulus MP, Tapert SF, Schuckit MA 74(5):340–347
(2005) Neural activation patterns of 51. Szewczyk-Krolikowski K et al (2014)
methamphetamine-dependent subjects dur- Functional connectivity in the basal ganglia
ing decision making predict relapse. Arch Gen network differentiates PD patients from con-
Psychiatry 62(7):761–768 trols. Neurology 83(3):208–214
37. Kaufman JN et al (2003) Cingulate hypoactiv- 52. Honey GD et al (2003) The functional neu-
ity in cocaine users during a GO-NOGO task as roanatomy of schizophrenic subsyndromes.
revealed by event-related functional magnetic res- Psychol Med 33(6):1007–1018
onance imaging. J Neurosci 23(21):7839–7843 53. Baghai TC, Moller HJ, Rupprecht R (2006)
38. Forman SD et al (2004) Opiate addicts lack Recent progress in pharmacological and non-
error-dependent activation of rostral anterior pharmacological treatment options of major
cingulate. Biol Psychiatry 55(5):531–537 depression. Curr Pharm Des 12(4):503–515
39. Heinz A et al (2004) Correlation between 54. Canli T et al (2005) Amygdala reactivity to
dopamine D(2) receptors in the ventral stria- emotional faces predicts improvement in major
tum and central processing of alcohol cues and depression. Neuroreport 16(12):1267–1270
craving. Am J Psychiatry 161(10):1783–1789 55. Killgore WD, Yurgelun-Todd DA (2006)
40. Cole DM et al (2012) Orbitofrontal connec- Ventromedial prefrontal activity correlates
tivity with resting-state networks is associated with depressed mood in adolescent children.
with midbrain dopamine D3 receptor availabil- Neuroreport 17(2):167–171
ity. Cereb Cortex 22(12):2784–2793 56. Anand A et al (2005) Antidepressant effect on
41. Egan MF et al (2003) The BDNF val66met connectivity of the mood-regulating circuit:
polymorphism affects activity-dependent secre- an FMRI study. Neuropsychopharmacology
tion of BDNF and human memory and hip- 30(7):1334–1344
pocampal function. Cell 112(2):257–269 57. Allen P et al (2012) Transition to psycho-
42. Lu B, Nagappan G, Lu Y (2014) BDNF and sis associated with prefrontal and subcorti-
synaptic plasticity, cognitive function, and dys- cal dysfunction in ultra high-risk individuals.
function. Handb Exp Pharmacol 220:223–250 Schizophr Bull 38(6):1268–1276
43. Rabiner EA et al (2011) Pharmacological dif- 58. Schmidt A et al (2014) Approaching a network
ferentiation of opioid receptor antagonists by connectivity-driven classification of the psychosis
molecular and functional imaging of target occu- continuum: a selective review and suggestions
pancy and food reward-related brain activation for future research. Front Hum Neurosci 8:1047
in humans. Mol Psychiatry 16(8):826–835, 785 59. del Campo N, Muller U, Sahakian BJ (2012)
44. Ziauddeen H et al (2013) Effects of the mu- Neural and behavioral endophenotypes in
opioid receptor antagonist GSK1521498 on ADHD. Curr Top Behav Neurosci 11:65–91
hedonic and consummatory eating behaviour: a 60. Hasler G, Northoff G (2011) Discovering
proof of mechanism study in binge-eating obese imaging endophenotypes for major depression.
subjects. Mol Psychiatry 18(12):1287–1293 Mol Psychiatry 16(6):604–619
45. Engel RH, Kaklamani VG (2007) HER2- 61. Savitz JB, Drevets WC (2009) Imaging phe-
positive breast cancer: current and future treat- notypes of major depressive disorder: genetic
ment strategies. Drugs 67(9):1329–1341 correlates. Neuroscience 164(1):300–330
46. Matthews PM et al (2014) The emerging 62. Keener MT, Phillips ML (2007) Neuroimaging
agenda of stratified medicine in neurology. Nat in bipolar disorder: a critical review of current
Rev Neurol 10(1):15–26 findings. Curr Psychiatry Rep 9(6):512–520
47. Greicius MD et al (2004) Default-mode 63. Lee MC et al (2013) Amygdala activity con-
network activity distinguishes Alzheimer’s tributes to the dissociative effect of cannabis on
disease from healthy aging: evidence from pain perception. Pain 154(1):124–134
Pharmacological Applications of fMRI 831
64. Lee MC, Tracey I (2013) Imaging pain: a modulation by rivastigmine. Brain 126(Pt
potent means for investigating pain mecha- 12):2750–2760
nisms in patients. Br J Anaesth 111(1):64–72 79. Matthews PM, Johansen-Berg H, Reddy H
65. Mazzola V et al (2010) Affective response to (2004) Non-invasive mapping of brain func-
a loved one’s pain: insula activity as a func- tions and brain recovery: applying lessons from
tion of individual differences. PLoS One cognitive neuroscience to neurorehabilitation.
5(12):e15268 Restor Neurol Neurosci 22(3–5):245–260
66. Derbyshire SW, Whalley MG, Oakley DA 80. Borsook D et al (2012) Decision-making using
(2009) Fibromyalgia pain and its modulation fMRI in clinical drug development: revisiting
by hypnotic and non-hypnotic suggestion: an NK-1 receptor antagonists for pain. Drug
fMRI analysis. Eur J Pain 13(5):542–550 Discov Today 17(17–18):964–973
67. Fairhurst M et al (2012) An fMRI study 81. Leknes S et al (2013) The importance of con-
exploring the overlap and differences between text: when relative relief renders pain pleasant.
neural representations of physical and recalled Pain 154(3):402–410
pain. PLoS One 7(10):e48711 82. Schwarz AJ et al (2007) In vivo mapping of
68. Murray D, Stoessl AJ (2013) Mechanisms and functional connectivity in neurotransmitter sys-
therapeutic implications of the placebo effect tems using pharmacological MRI. Neuroimage
in neurological and psychiatric conditions. 34(4):1627–1636
Pharmacol Ther 140(3):306–318 83. Batterham RL et al (2007) PYY modulation of
69. Denk F, McMahon SB, Tracey I (2014) Pain cortical and hypothalamic brain areas predicts
vulnerability: a neurobiological perspective. feeding behaviour in humans. Nature
Nat Neurosci 17(2):192–200 450(7166):106–109
70. Baliki MN et al (2012) Corticostriatal func- 84. Borsook D, Becerra L, Hargreaves R (2006) A
tional connectivity predicts transition to chronic role for fMRI in optimizing CNS drug devel-
back pain. Nat Neurosci 15(8):1117–1119 opment. Nat Rev Drug Discov 5(5):411–424
71. Smucny J, Wylie KP, Tregellas JR (2014) 85. Cummings JL (2010) Integrating ADNI
Functional magnetic resonance imag- results into Alzheimer’s disease drug develop-
ing of intrinsic brain networks for transla- ment programs. Neurobiol Aging
tional drug discovery. Trends Pharmacol Sci 31(8):1481–1492
35(8):397–403 86. Prentice RL (1989) Surrogate endpoints in
72. Vollm BA et al (2004) Methamphetamine clinical trials: definition and operational crite-
activates reward circuitry in drug naive ria. Stat Med 8(4):431–440
human subjects. Neuropsychopharmacology 87. Hoge RD et al (1999) Linear coupling between
29(9):1715–1722 cerebral blood flow and oxygen consumption
73. Gerdelat-Mas A et al (2005) Chronic admin- in activated human cortex. Proc Natl Acad Sci
istration of selective serotonin reuptake inhibi- U S A 96(16):9403–9408
tor (SSRI) paroxetine modulates human 88. Rombouts SA et al (2003) Loss of frontal
motor cortex excitability in healthy subjects. fMRI activation in early frontotemporal
Neuroimage 27(2):314–322 dementia compared to early AD. Neurology
74. Pariente J et al (2001) Fluoxetine modulates 60(12):1904–1908
motor performance and cerebral activation of 89. Floyer-Lea A, Matthews PM (2004) Changing
patients recovering from stroke. Ann Neurol brain networks for visuomotor control with
50(6):718–729 increased movement automaticity.
75. Goekoop R et al (2004) Challenging the cho- J Neurophysiol 92(4):2405–2412
linergic system in mild cognitive impairment: 90. Cader S et al (2006) Reduced brain functional
a pharmacological fMRI study. Neuroimage reserve and altered functional connectivity in
23(4):1450–1459 patients with multiple sclerosis. Brain 129(Pt
76. Farahani K et al (1999) Contemporaneous pos- 2):527–537
itron emission tomography and MR imaging at 91. Lachaux JP et al (2007) Relationship between
1.5 T. J Magn Reson Imaging 9(3):497–500 task-related gamma oscillations and BOLD sig-
77. Wilkinson D, Halligan P (2004) The relevance nal: new insights from combined fMRI and
of behavioural measures for functional-imaging intracranial EEG. Hum Brain Mapp
studies of cognition. Nat Rev Neurosci 28(12):1368–1375
5(1):67–73 92. Merlo Pich E et al (2014) Imaging as a bio-
78. Parry AM et al (2003) Potentially adaptive marker in drug discovery for Alzheimer’s dis-
functional changes in cognitive processing for ease: is MRI a suitable technology? Alzheimers
patients with multiple sclerosis and their acute Res Ther 6(4):51
Chapter 27
Abstract
Motor deficits contribute to disability in a number of neurological conditions. A wide range of emerging
restorative therapies have the potential to reduce this by favorably modifying function. In many medical
contexts, a study of target organ function improves efficacy of a therapeutic intervention. However, the
optimal methods to prescribe a restorative therapy in the setting of central nervous system (CNS) disease
are not clear. Brain mapping studies have the potential to provide useful insights in this regard. Examples of
restorative therapies are provided, and human trials are summarized whereby brain mapping data have
proven useful in promoting motor improvements in subjects with a neurological condition. A number of
forms of brain mapping metrics are under study, including those emphasizing network connectivity obtained
using resting-state fMRI. In some cases, brain mapping findings that correlate with better outcome with
spontaneous behavioral recovery correspond to findings that predict better treatment response in the con-
text of a clinical trial. Similarities across CNS conditions, such as stroke and multiple sclerosis, are discussed.
Further studies are needed to understand which methods have the greatest value to monitor, predict, triage,
and dose restorative therapies in trials that aim to reduce motor, and other neurological, deficits.
Key words Functional neuroimaging, Brain mapping, Stroke, Motor system, Recovery, Repair,
Plasticity, Treatment
833
834 Steven C. Cramer and Jessica M. Cassidy
2 Stroke
2.1 Use of Functional One study used functional neuroimaging in a clinical trial of a
Neuroimaging to Guide restorative intervention to extract data from an fMRI scan in order
a Restorative to guide details of decision-making during therapy [24, 52]. An
Intervention in fMRI scan was used to identify the centroid of ipsilesional primary
Patients with Stroke motor cortex activation when patients with stroke moved the
affected hand. This information then guided neurosurgical place-
ment of an investigational epidural cortical stimulation device over
ipsilesional motor cortex. Using this approach, patients receiving
stimulation plus rehabilitation therapy showed significantly greater
arm motor gains than patients receiving rehabilitation therapy
alone. A similar approach was used in studies based on transcranial
magnetic stimulation (TMS) to identify the optimal physiological
representation site for hand motor function. These studies found
repetitive TMS to be useful for improving motor function after
stroke [20, 21].
pair of studies also hints at the potential use of human brain map-
ping measures to identify the dose of a restorative therapy in indi-
vidual patients. For example, could a TMS or fMRI measure of
brain function inform a clinician of the likelihood that the brain is
receptive to further change that supports behavioral gains? In this
regard, note that a probe of brain plasticity, such as might be used
to predict treatment response to a restorative intervention, can be
developed even in the setting of severe deficits, such as complete
plegia [61].
Another study found that fMRI had independent value for pre-
dicting treatment response in a restorative therapy trial in patients
with chronic stroke [57]. This study used a multivariate model to
examine the specific ability of a baseline fMRI to predict trial-related
behavioral gains, and compared this fMRI predictive ability directly
to a number of other baseline measures. Patients in this study each
underwent baseline clinical and functional MRI assessments,
received 6 weeks of rehabilitation therapy with or without investi-
gational motor cortex stimulation, then had repeat assessments.
Across all patients, univariate analyses found that several baseline
measures had predictive value for trial-related gains. However, mul-
tiple linear regression modeling found that only two variables
remained significant predictors: degree of motor cortex activation
on fMRI (lower motor cortex activation predicted larger gains) and
arm motor function (greater arm function predicted larger gains).
This study emphasized that an assessment of brain function can be
a unique source of information for clinical decision-making in the
setting of restorative therapy after stroke. Interestingly, clinical
gains during study participation were paralleled by boosts in motor
cortex activity, the latter detected via serial fMRI scanning, suggest-
ing that lower baseline cortical activity in some patients likely repre-
sents under-use of an available cortical resource.
Burke Quinlan and colleagues [60] also utilized a multivariate
model that encompassed various demographic, behavioral, and
neuroimaging measures to determine which metrics best predicted
behavioral gains following a three-week upper-extremity robotic
therapy program in individuals with chronic stroke. Bivariate
screening revealed significant correlations between improvement
in upper-extremity status with baseline MRI and diffusion tensor
imaging measures of brain injury (i.e. infarct volume, cortical
injury, percentage injury to corticospinal tract), task-evoked fMRI
measures of cortical function (i.e. ipsilesional primary motor cor-
tex area (M1) activation), and resting fMRI measures of ipsile-
sional/contralesional M1 connectivity. Subsequent multivariate
linear regression modeling revealed that the percentage injury to
corticospinal tract and ipsilesional/contralesional M1 connectiv-
ity accounted for 44 % of the variance in treatment gains. Brain-
based measures, therefore, depicted better predictive quality than
the more conventional behavioral and demographic measures.
Application of fMRI to Monitor Motor Rehabilitation 837
2.3 Use of Functional Carey et al. [62] found that a population of subjects with chronic
Neuroimaging to stroke, when performing a finger tracking task with the stroke-
Investigate the affected hand, had activation within contralesional brain regions,
Biological i.e., regions that were primarily ipsilateral to movement. After
Mechanisms of training at this task, the normal pattern of laterality of brain activa-
Restorative tion was restored, with activation shifting to ipsilesional brain
Intervention Effects regions, i.e., contralateral to movement, and thereby more closely
in Patients with Stroke resembling findings in healthy control subjects. In this landmark
study, functional neuroimaging provided insights into the mecha-
nistic effects of treatment. Since then, other studies have shown
varying modulatory effects of cortical activation following Botox
[63], constraint-induced movement therapy (CIMT) [64], visuo-
motor tracking task practice [65], implicit motor learning [66],
real-time fMRI feedback training [67], and noninvasive brain stim-
ulation in individuals with chronic stroke [68].
Two meta-analyses [69, 70] extend these results by examining
studies that have employed functional neuroimaging as a biological
marker of treatment effects targeting the motor system after stroke.
A review of 24 studies utilizing sensorimotor tasks in 255 patients
found higher activation in the contralesional M1 (relative to
healthy controls) that decreased over time but was unrelated to
motor outcome. Reorganization consistent with increased
ipsilesional M1 and medial premotor cortex activation was associ-
ated with positive recovery; whereas, increased activation of the
cerebellar vermis was associated with negative recovery. These con-
clusions highlight both beneficial and detrimental examples of
neuroplastic reorganization following stroke as demonstrated by
shifts in premotor and cerebellar vermis activation, respectively. An
earlier meta-analysis that reviewed 13 studies of 121 patients per-
mitted drawing a number of conclusions [70]. Motor deficits have
been most often studied, in part because of their substantial contri-
bution to overall disability after stroke, and in part because of their
relatively high prevalence. Most published studies have focused on
patients with good to excellent outcome at baseline since they were
more able to perform the motor tasks required to probe brain
function. Consequently, less is known about the functional
838 Steven C. Cramer and Jessica M. Cassidy
3 Multiple Sclerosis
3.1 Use of Functional The extent to which these spontaneous changes in brain function
Neuroimaging to after MS can be used to monitor therapeutic interventions has
Investigate the been assessed in several small studies. One study tested the effects
Biological Mechanism of increased cholinergic tone on the pattern of fMRI activation
of Restorative during performance of a cognitive task, the Stroop test. At base-
Intervention Effects line, patients with MS and moderate disability had similar behav-
in Patients with MS ioral performance as compared to controls, but on fMRI showed
840 Steven C. Cramer and Jessica M. Cassidy
4.1 Use of Functional There has been limited study of the CNS mechanisms underlying
Neuroimaging to spontaneous motor improvement during the months following
Investigate the SCI. Jurkiewicz et al. [103] examined the acute-to-chronic time-
Biological Mechanism course of post-SCI sensorimotor reorganization in four individuals
of Restorative with tetraplegia over a 12-month period. Shortly after injury, sub-
Intervention Effects jects with SCI demonstrated a similar volume of contralateral M1
in Patients with SCI activation as healthy controls when attempting ankle dorsiflexion
movements. However, with increasing time post injury and persist-
ing paralysis, contralateral M1 activation decreased along with pre-
frontal, premotor, supplementary, primary somatosensory, and
posterior parietal cortices and cingulate motor area activation.
These results depict a progressive shift in cortical reorganization
further influenced by lower-extremity disuse. A related study in
individuals with chronic SCI found cortical thinning in the leg area
of the M1 and primary sensory cortex compared to healthy con-
trols [104]. Moreover, subjects with SCI demonstrated increased
activation of the left M1 leg area during right handgrip task com-
pared to controls that was associated with smaller cervical cord area
and impaired upper-extremity function. Lundell et al. [105] also
found associations between spinal cord atrophy, motor function,
and ipsilateral M1, somatosensory, and premotor cortical activa-
tion during ankle dorsiflexion movements in individuals with
chronic SCI. Additional investigation is needed to further substan-
tiate the relationship between neuroplastic reorganization, severity
of SCI, and ensuing motor function.
Studies to date have more been focused on the nature of brain
motor systems function in the chronic state, with some divergence
of results to date. Some studies have found a broad decrease in acti-
vation [106–108], particularly in primary sensorimotor cortex,
whereas others have found supranormal activation [109]. The basis
for these discrepancies remains unclear but could be due to differ-
ences in age or injury pattern of the population studied, years post-
SCI at the time of study, amount of motor function at the time of
study, or the nature of the task used to probe motor system function,
some uncovering deficient processing and others emphasizing supra-
normal efforts to compensate [107, 110]. A commonly described
feature is a change in somatotopic organization within primary sen-
sorimotor cortex contralateral to sensory or motor events, with rep-
resentation of supralesional body regions expanding at the expense
of infralesional body regions [111–115]. Spontaneous changes in
laterality, so prominent in studies of stroke or MS, as above, are
842 Steven C. Cramer and Jessica M. Cassidy
generally not prominent after SCI [116], perhaps due in part to the
fact that injury typically affects the CNS bilaterally or perhaps due in
part to the fact that SCI spares brain commissural fibers whose integ-
rity helps maintain normal hemispheric balance. As such, laterality is
unlikely to be a useful variable to examine in brain mapping studies
of treatment effects in the setting of SCI.
At least two studies have evaluated changes in brain function in
relation to therapy after SCI. Winchester et al. [117] studied body
weight supported treadmill training in four patients with motor
incomplete SCI. These authors compared fMRI during attempted
unilateral foot and toe movement before vs. after training. This
therapy was associated with increased activation within several
bilateral areas, including primary sensorimotor cortex and cerebel-
lum, though to a variable extent. The authors observed that,
although all participants demonstrated a change in the BOLD
signal following training, only those patients who demonstrated a
substantial increase in activation of the cerebellum demonstrated
an improvement in their ability to walk over ground, suggesting
that this measure in this brain region, at least when examined using
this task during fMRI, might be useful as a biological marker of
successful treatment effect.
Another form of intervention that has been evaluated after SCI is
motor imagery. Motor imagery normally activates many of the same
brain regions as motor execution, and has been associated with
improvements in motor performance [118, 119]. The effects of
1-week of motor imagery training to tongue and to foot were evalu-
ated in ten subjects with chronic, complete tetra-/paraplegia plus ten
healthy controls [61]. The behavioral outcome measure was speed of
performance of a complex sequence. Motor imagery training was
associated with a significant improvement in this behavior in non-
paralyzed muscles (tongue for both groups, right foot for healthy
subjects). In both the healthy controls and the subjects with SCI,
serial fMRI scanning (before vs. after training) during attempted
right foot movement was associated with increased fMRI activation
in left putamen, an area associated with motor learning, despite foot
movements being present in controls and absent in subjects with
SCI. Behavioral training can thus result in measurable brain plasticity
that is not accompanied by outward behavioral gains, a finding that
might be important for designing biological markers in trials target-
ing severely disabled patient populations. Note that this fMRI change
was absent in a second healthy control group serially imaged without
training. The main conclusion from this study is that motor imagery
training improves brain function whether or not sensorimotor func-
tion is present in the trained limb. An additional conclusion is that
motor imagery, by virtue of its favorable effects on brain motor sys-
tem organization, might have value as an adjunct motor restorative
therapy. Another key point from this study is that brain plasticity
related to plegic limbs can be studied in subjects with chronic SCI.
Application of fMRI to Monitor Motor Rehabilitation 843
5 Conclusions
References
1. Dobkin B (2003) The clinical science of neu- lowing focal stroke. Neuroreport
rologic rehabilitation. Oxford University 9(7):1441–1445
Press, New York 9. Schabitz WR, Berger C, Kollmar R, Seitz M,
2. Chen J, Cui X, Zacharek A, Jiang H, Roberts Tanay E, Kiessling M et al (2004) Effect of
C, Zhang C et al (2007) Niaspan increases brain-derived neurotrophic factor treatment
angiogenesis and improves functional recov- and forced arm use on functional motor
ery after stroke. Ann Neurol 62(1):49–58 recovery after small cortical ischemia. Stroke
3. Li L, Jiang Q, Zhang L, Ding G, Gang Zhang 35(4):992–997
Z, Li Q et al (2007) Angiogenesis and 10. Wang L, Zhang Z, Wang Y, Zhang R, Chopp
improved cerebral blood flow in the ischemic M (2004) Treatment of stroke with erythro-
boundary area detected by MRI after admin- poietin enhances neurogenesis and angiogen-
istration of sildenafil to rats with embolic esis and improves neurological function in
stroke. Brain Res 1132(1):185–192 rats. Stroke 35(7):1732–1737
4. Chen P, Goldberg D, Kolb B, Lanser M, 11. Tsai PT, Ohab JJ, Kertesz N, Groszer M,
Benowitz L (2002) Inosine induces axonal Matter C, Gao J et al (2006) A critical role of
rewiring and improves behavioral outcome erythropoietin receptor in neurogenesis and
after stroke. Proc Natl Acad Sci U S A post-stroke recovery. J Neurosci
99(13):9031–9036 26(4):1269–1274
5. Freret T, Valable S, Chazalviel L, Saulnier R, 12. Schneider UC, Schilling L, Schroeck H, Nebe
Mackenzie ET, Petit E et al (2006) Delayed CT, Vajkoczy P, Woitzik J (2007) Granulocyte-
administration of deferoxamine reduces brain macrophage colony-stimulating factor-
damage and promotes functional recovery induced vessel growth restores cerebral blood
after transient focal cerebral ischemia in the supply after bilateral carotid artery occlusion.
rat. Eur J Neurosci 23(7):1757–1765 Stroke 38(4):1320–1328
6. Papadopoulos CM, Tsai SY, Cheatwood JL, 13. Kolb B, Morshead C, Gonzalez C, Kim M,
Bollnow MR, Kolb BE, Schwab ME et al (2006) Gregg C, Shingo T et al (2007) Growth
Dendritic plasticity in the adult rat following factor-stimulated generation of new cortical
middle cerebral artery occlusion and Nogo-a tissue and functional recovery after stroke
neutralization. Cereb Cortex 16(4):529–536 damage to the motor cortex of rats. J Cereb
7. Kawamata T, Dietrich W, Schallert T, Gotts J, Blood Flow Metab 27(5):983–997
Cocke R, Benowitz L et al (1997) 14. Zhao LR, Berra HH, Duan WM, Singhal S,
Intracisternal basic fibroblast growth factor Mehta J, Apkarian AV et al (2007) Beneficial
(bFGF) enhances functional recovery and effects of hematopoietic growth factor ther-
upregulates the expression of a molecular apy in chronic ischemic stroke in rats. Stroke
marker of neuronal sprouting following focal 38(10):2804–2811
cerebral infarction. Proc Natl Acad Sci U S A 15. Ehrenreich H, Hasselblatt M, Dembowski C,
94:8179–8184 Cepek L, Lewczuk P, Stiefel M et al (2002)
8. Kawamata T, Ren J, Chan T, Charette M, Erythropoietin therapy for acute stroke is
Finklestein S (1998) Intracisternal osteogenic both safe and beneficial. Mol Med
protein-1 enhances functional recovery fol- 8(8):495–505
Application of fMRI to Monitor Motor Rehabilitation 845
16. Savitz SI, Dinsmore JH, Wechsler LR, 28. Volpe BT, Ferraro M, Lynch D, Christos P,
Rosenbaum DM, Caplan LR (2004) Cell Krol J, Trudell C et al (2005) Robotics and
therapy for stroke. NeuroRx 1(4):406–414 other devices in the treatment of patients
17. Keirstead HS, Nistor G, Bernal G, Totoiu M, recovering from stroke. Curr Neurol Neurosci
Cloutier F, Sharp K et al (2005) Human Rep 5(6):465–470
embryonic stem cell-derived oligodendrocyte 29. Reinkensmeyer D, Emken J, Cramer S (2004)
progenitor cell transplants remyelinate and Robotics, motor learning, and neurologic
restore locomotion after spinal cord injury. recovery. Annu Rev Biomed Eng 6:497–525
J Neurosci 25(19):4694–4705 30. Deutsch JE, Lewis JA, Burdea G (2007)
18. Cummings BJ, Uchida N, Tamaki SJ, Salazar Technical and patient performance using a
DL, Hooshmand M, Summers R et al (2005) virtual reality-integrated telerehabilitation
Human neural stem cells differentiate and system: preliminary finding. IEEE Trans
promote locomotor recovery in spinal cord- Neural Syst Rehabil Eng 15(1):30–35
injured mice. Proc Natl Acad Sci U S A 31. Duncan P, Studenski S, Richards L, Gollub S,
102(39):14069–14074 Lai S, Reker D et al (2003) Randomized clini-
19. Shen LH, Li Y, Chen J, Zacharek A, Gao Q, cal trial of therapeutic exercise in subacute
Kapke A et al (2006) Therapeutic benefit of stroke. Stroke 34(9):2173–2180
bone marrow stromal cells administered 1 32. Woldag H, Hummelsheim H (2002)
month after stroke. J Cereb Blood Flow Evidence-based physiotherapeutic concepts
Metab 27:6–13 for improving arm and hand function in
20. Khedr EM, Ahmed MA, Fathy N, Rothwell stroke patients: a review. J Neurol
JC (2005) Therapeutic trial of repetitive tran- 249(5):518–528
scranial magnetic stimulation after acute isch- 33. French B, Thomas L, Leathley M, Sutton C,
emic stroke. Neurology 65(3):466–468 McAdam J, Forster A et al (2007) Repetitive
21. Kim YH, You SH, Ko MH, Park JW, Lee KH, task training for improving functional ability
Jang SH et al (2006) Repetitive transcranial after stroke. Cochrane Database Syst Rev
magnetic stimulation-induced corticomotor (4):CD006073
excitability and associated motor skill acquisi- 34. Kwakkel G, Wagenaar R, Twisk J, Lankhorst
tion in chronic stroke. Stroke G, Koetsier J (1999) Intensity of leg and arm
37(6):1471–1476 training after primary middle-cerebral-artery
22. Malcolm MP, Triggs WJ, Light KE, Gonzalez stroke: a randomised trial. Lancet
Rothi LJ, Wu S, Reid K et al (2007) Repetitive 354(9174):191–196
transcranial magnetic stimulation as an 35. Van Peppen RP, Kwakkel G, Wood-Dauphinee
adjunct to constraint-induced therapy: an S, Hendriks HJ, Van der Wees PJ, Dekker
exploratory randomized controlled trial. Am J (2004) The impact of physical therapy on
J Phys Med Rehabil 86(9):707–715 functional outcomes after stroke: what’s the
23. Hummel F, Celnik P, Giraux P, Floel A, Wu evidence? Clin Rehabil 18(8):833–862
WH, Gerloff C et al (2005) Effects of non- 36. Luft A, McCombe-Waller S, Whitall J,
invasive cortical stimulation on skilled motor Forrester L, Macko R, Sorkin J et al (2004)
function in chronic stroke. Brain 128(Pt Repetitive bilateral arm training and motor
3):490–499 cortex activation in chronic stroke: a random-
24. Brown JA, Lutsep HL, Weinand M, Cramer ized controlled trial. JAMA
SC (2006) Motor cortex stimulation for the 292(15):1853–1861
enhancement of recovery from stroke: a pro- 37. Wolf SL, Winstein CJ, Miller JP, Taub E,
spective, multicenter safety study. Uswatte G, Morris D et al (2006) Effect of
Neurosurgery 58(3):464–473 constraint-induced movement therapy on
25. Ring H, Rosenthal N (2005) Controlled upper extremity function 3 to 9 months after
study of neuroprosthetic functional electrical stroke: the EXCITE randomized clinical trial.
stimulation in sub-acute post-stroke rehabili- JAMA 296(17):2095–2104
tation. J Rehabil Med 37(1):32–36 38. Rathore S, Hinn A, Cooper L, Tyroler H,
26. Sheffler LR, Chae J (2007) Neuromuscular Rosamond W (2002) Characterization of
electrical stimulation in neurorehabilitation. incident stroke signs and symptoms: findings
Muscle Nerve 35(5):562–590 from the atherosclerosis risk in communities
27. Kwakkel G, Kollen BJ, Krebs HI (2007) study. Stroke 33(11):2718–2721
Effects of robot-assisted therapy on upper 39. Gresham G, Duncan P, Stason W, Adams H,
limb recovery after stroke: a systematic review. Adelman A, Alexander D et al (1995) Post-
Neurorehabil Neural Repair 22:111–121 stroke rehabilitation. U.S. Department of
846 Steven C. Cramer and Jessica M. Cassidy
Health and Human Services. Public Health motor cortex after stroke. Brain 127(Pt
Service, Agency for Health Care Policy and 4):747–758
Research, Rockville, MD 52. Cramer S, Benson R, Himes D, Burra V,
40. Duncan P, Goldstein L, Horner R, Landsman Janowsky J, Weinand M et al (2005) Use of
P, Samsa G, Matchar D (1994) Similar motor functional MRI to guide decisions in a clinical
recovery of upper and lower extremities after stroke trial. Stroke 36(5):e50–e52
stroke. Stroke 25(6):1181–1188 53. Platz T, Kim I, Engel U, Kieselbach A,
41. Duncan P, Goldstein L, Matchar D, Divine Mauritz K (2002) Brain activation pattern as
G, Feussner J (1992) Measurement of motor assessed with multi-modal EEG analysis pre-
recovery after stroke. Stroke 23:1084–1089 dict motor recovery among stroke patients
42. Nakayama H, Jorgensen H, Raaschou H, with mild arm paresis who receive the Arm
Olsen T (1994) Recovery of upper extremity Ability Training. Restor Neurol Neurosci
function in stroke patients: the Copenhagen 20(1–2):21–35
Stroke Study. Arch Phys Med Rehabil 54. Dong Y, Dobkin BH, Cen SY, Wu AD,
75(4):394–398 Winstein CJ (2006) Motor cortex activation
43. Wade D, Langton-Hewer R, Wood V, during treatment may predict therapeutic
Skilbeck C, Ismail H (1983) The hemiplegic gains in paretic hand function after stroke.
arm after stroke: measurement and recovery. Stroke 37(6):1552–1555
J Neurol Neurosurg Psychiatry 55. Fritz SL, Light KE, Patterson TS, Behrman
46(6):521–524 AL, Davis SB (2005) Active finger extension
44. Yozbatiran N, Cramer SC (2006) Imaging predicts outcomes after constraint-induced
motor recovery after stroke. NeuroRx movement therapy for individuals with hemi-
3(4):482–488 paresis after stroke. Stroke 36(6):1172–1177
45. Ward NS, Cohen LG (2004) Mechanisms 56. Koski L, Mernar T, Dobkin B (2004)
underlying recovery of motor function after Immediate and long-term changes in cortico-
stroke. Arch Neurol 61(12):1844–1848 motor output in response to rehabilitation:
46. Baron J, Cohen L, Cramer S, Dobkin B, correlation with functional improvements in
Johansen-Berg H, Loubinoux I et al (2004) chronic stroke. Neurorehabil Neural Repair
Neuroimaging in stroke recovery: a position 18(4):230–249
paper from the First International workshop 57. Cramer S, Parrish T, Levy R, Stebbins G,
on neuroimaging and stroke recovery. Ruland S, Lowry D et al (2007) An assess-
Cerebrovasc Dis (Basel, Switzerland) ment of brain function predicts functional
18(3):260–267 gains in a clinical stroke trial. Stroke 38:520
47. Lotze M, Markert J, Sauseng P, Hoppe J, (abstract)
Plewnia C, Gerloff C (2006) The role of mul- 58. Stinear CM, Barber PA, Smale PR, Coxon JP,
tiple contralesional motor areas for complex Fleming MK, Byblow WD (2007) Functional
hand movements after internal capsular potential in chronic stroke patients depends
lesion. J Neurosci 26(22):6096–6102 on corticospinal tract integrity. Brain 130(Pt
48. Winhuisen L, Thiel A, Schumacher B, Kessler 1):170–180
J, Rudolf J, Haupt WF et al (2005) Role of 59. Burke E, Dobkin BH, Noser EA, Enney LA,
the contralateral inferior frontal gyrus in Cramer SC (2014) Predictors and biomarkers
recovery of language function in poststroke of treatment gains in a clinical stroke trial tar-
aphasia: a combined repetitive transcranial geting the lower extremity. Stroke
magnetic stimulation and positron emission 45(8):2379–2384
tomography study. Stroke 36(8):1759–1763 60. Burke Quinlan E, Dodakian L, See J,
49. Johansen-Berg H, Rushworth M, Bogdanovic McKenzie A, Le V, Wojnowicz M et al (2015)
M, Kischka U, Wimalaratna S, Matthews P Neural function, injury, and stroke subtype
(2002) The role of ipsilateral premotor cortex predict treatment gains after stroke. Ann
in hand movement after stroke. Proc Natl Neurol 77(1):132–145
Acad Sci U S A 99(22):14518–14523 61. Cramer SC, Orr EL, Cohen MJ, Lacourse
50. Werhahn K, Conforto A, Kadom N, Hallett MG (2007) Effects of motor imagery training
M, Cohen L (2003) Contribution of the ipsi- after chronic, complete spinal cord injury.
lateral motor cortex to recovery after chronic Exp Brain Res 177(2):233–242
stroke. Ann Neurol 54(4):464–472 62. Carey J, Kimberley T, Lewis S, Auerbach E,
51. Fridman E, Hanakawa T, Chung M, Hummel Dorsey L, Rundquist P et al (2002) Analysis of
F, Leiguarda R, Cohen L (2004) fMRI and finger tracking training in subjects
Reorganization of the human ipsilesional pre- with chronic stroke. Brain 125(Pt 4):773–788
Application of fMRI to Monitor Motor Rehabilitation 847
63. Veverka T, Hlustik P, Hok P, Otruba P, Tudos induced movement therapy in chronic stroke.
Z, Zapletalova J et al (2014) Cortical activity Eur J Neurol 19(4):578–586
modulation by botulinum toxin type A in 74. Vukusic S, Confavreux C (2007) Natural his-
patients with post-stroke arm spasticity: real tory of multiple sclerosis: risk factors and
and imagined hand movement. J Neurol Sci prognostic indicators. Curr Opin Neurol
346(1–2):276–283 20(3):269–274
64. Laible M, Grieshammer S, Seidel G, Rijntjes 75. Vollmer T (2007) The natural history of
M, Weiller C, Hamzei F (2012) Association relapses in multiple sclerosis. J Neurol Sci
of activity changes in the primary sensory cor- 256(Suppl 1):S5–S13
tex with successful motor rehabilitation of the 76. Rocca MA, Filippi M (2007) Functional MRI
hand following stroke. Neurorehabil Neural in multiple sclerosis. J Neuroimaging
Repair 26(7):881–888 17(Suppl 1):36S–41S
65. Bosnell RA, Kincses T, Stagg CJ, Tomassini 77. Rocca MA, Colombo B, Falini A, Ghezzi A,
V, Kischka U, Jbabdi S et al (2011) Motor Martinelli V, Scotti G et al (2005) Cortical
practice promotes increased activity in brain adaptation in patients with MS: a cross-
regions structurally disconnected after sub- sectional functional MRI study of disease
cortical stroke. Neurorehabil Neural Repair phenotypes. Lancet Neurol 4(10):618–626
25(7):607–616
78. Wang J, Hier DB (2007) Motor reorganiza-
66. Meehan SK, Randhawa B, Wessel B, Boyd LA tion in multiple sclerosis. Neurol Res
(2011) Implicit sequence-specific motor 29(1):3–8
learning after subcortical stroke is associated
with increased prefrontal brain activations: an 79. Pantano P, Mainero C, Caramia F (2006)
fMRI study. Hum Brain Mapp Functional brain reorganization in multiple
32(2):290–303 sclerosis: evidence from fMRI studies.
J Neuroimaging 16(2):104–114
67. Sitaram R, Veit R, Stevens B, Caria A, Gerloff
C, Birbaumer N et al (2012) Acquired con- 80. Ward N, Brown M, Thompson A, Frackowiak
trol of ventral premotor cortex activity by R (2003) Neural correlates of outcome after
feedback training: an exploratory real-time stroke: a cross-sectional fMRI study. Brain
FMRI and TMS study. Neurorehabil Neural 126(Pt 6):1430–1448
Repair 26(3):256–265 81. Cramer SC, Crafton KR (2006) Somatotopy
68. Stagg CJ, Bachtiar V, O’Shea J, Allman C, and movement representation sites following
Bosnell RA, Kischka U et al (2012) Cortical cortical stroke. Exp Brain Res
activation changes underlying stimulation- 168(1–2):25–32
induced behavioural gains in chronic stroke. 82. Lenzi D, Conte A, Mainero C, Frasca V,
Brain 135(Pt 1):276–284 Fubelli F, Totaro P et al (2007) Effect of cor-
69. Favre I, Zeffiro TA, Detante O, Krainik A, pus callosum damage on ipsilateral motor
Hommel M, Jaillard A (2014) Upper limb activation in patients with multiple sclerosis: a
recovery after stroke is associated with ipsile- functional and anatomical study. Hum Brain
sional primary motor cortical activity: a meta- Mapp 28(7):636–644
analysis. Stroke 45(4):1077–1083 83. Rocca MA, Gallo A, Colombo B, Falini A,
70. Hodics T, Cohen LG, Cramer SC (2006) Scotti G, Comi G et al (2004) Pyramidal tract
Functional imaging of intervention effects in lesions and movement-associated cortical
stroke motor rehabilitation. Arch Phys Med recruitment in patients with MS. Neuroimage
Rehabil 87(12 Suppl):36–42 23(1):141–147
71. Schaechter J, Kraft E, Hilliard T, Dijkhuizen 84. Reddy H, Narayanan S, Matthews P, Hoge R,
R, Benner T, Finklestein S et al (2002) Motor Pike G, Duquette P et al (2000) Relating axo-
recovery and cortical reorganization after nal injury to functional recovery in
constraint-induced movement therapy in MS. Neurology 54(1):236–239
stroke patients: a preliminary study. 85. Mezzapesa DM, Rocca MA, Rodegher M,
Neurorehabil Neural Repair 16(4):326–338 Comi G, Filippi M (2007) Functional cortical
72. Johansen-Berg H, Dawes H, Guy C, Smith S, changes of the sensorimotor network are
Wade D, Matthews P (2002) Correlation associated with clinical recovery in multiple
between motor improvements and altered sclerosis. Hum Brain Mapp 29(5):562–573
fMRI activity after rehabilitative therapy. 86. Cramer S, Nelles G, Benson R, Kaplan J,
Brain 125(Pt 12):2731–2742 Parker R, Kwong K et al (1997) A functional
73. Kononen M, Tarkka IM, Niskanen E, MRI study of subjects recovered from hemi-
Pihlajamaki M, Mervaala E, Pitkanen K et al paretic stroke. Stroke 28(12):2518–2527
(2012) Functional MRI and motor behav- 87. Filippi M, Rocca MA, Mezzapesa DM, Falini A,
ioral changes obtained with constraint- Colombo B, Scotti G et al (2004) A functional
848 Steven C. Cramer and Jessica M. Cassidy
MRI study of cortical activations associated with 100. Jayaraman A, Gregory CM, Bowden M,
object manipulation in patients with Stevens JE, Shah P, Behrman AL et al (2006)
MS. Neuroimage 21(3):1147–1154 Lower extremity skeletal muscle function in
88. Colorado RA, Shukla K, Zhou Y, Wolinsky persons with incomplete spinal cord injury.
JS, Narayana PA (2012) Multi-task functional Spinal Cord 44(11):680–687
MRI in multiple sclerosis patients without 101. DeVivo MJ, Richards JS (1992) Community
clinical disability. Neuroimage reintegration and quality of life following spi-
59(1):573–581 nal cord injury. Paraplegia 30(2):108–112
89. Giorgio A, Portaccio E, Stromillo ML, 102. Frankel HL, Coll JR, Charlifue SW, Whiteneck
Marino S, Zipoli V, Battaglini M et al (2010) GG, Gardner BP, Jamous MA et al (1998)
Cortical functional reorganization and its Long-term survival in spinal cord injury: a fifty
relationship with brain structural damage in year investigation. Spinal Cord 36(4):266–274
patients with benign multiple sclerosis. Mult 103. Jurkiewicz MT, Mikulis DJ, Fehlings MG,
Scler 16(11):1326–1334 Verrier MC (2010) Sensorimotor cortical
90. Parry AM, Scott RB, Palace J, Smith S, activation in patients with cervical spinal cord
Matthews PM (2003) Potentially adaptive injury with persisting paralysis. Neurorehabil
functional changes in cognitive processing for Neural Repair 24(2):136–140
patients with multiple sclerosis and their acute 104. Freund P, Weiskopf N, Ward NS, Hutton C,
modulation by rivastigmine. Brain 126(Pt Gall A, Ciccarelli O et al (2011) Disability, atro-
12):2750–2760 phy and cortical reorganization following spinal
91. Sadato N, Campbell G, Ibanez V, Deiber M, cord injury. Brain 134(Pt 6):1610–1622
Hallett M (1996) Complexity affects regional 105. Lundell H, Christensen MS, Barthelemy D,
cerebral blood flow change during sequential fin- Willerslev-Olsen M, Biering-Sorensen F,
ger movements. J Neurosci 16(8):2691–2700 Nielsen JB (2011) Cerebral activation is cor-
92. Verstynen T, Diedrichsen J, Albert N, related to regional atrophy of the spinal cord
Aparicio P, Ivry RB (2005) Ipsilateral motor and functional motor disability in spinal cord
cortex activity during unimanual hand move- injured individuals. Neuroimage
ments relates to task complexity. 54(2):1254–1261
J Neurophysiol 93(3):1209–1222 106. Cramer SC, Lastra L, Lacourse MG, Cohen
93. Mainero C, Inghilleri M, Pantano P, Conte A, MJ (2005) Brain motor system function after
Lenzi D, Frasca V et al (2004) Enhanced chronic, complete spinal cord injury. Brain
brain motor activity in patients with MS after 128(Pt 12):2941–2950
a single dose of 3,4-diaminopyridine. 107. Jurkiewicz MT, Mikulis DJ, McIlroy WE,
Neurology 62(11):2044–2050 Fehlings MG, Verrier MC (2007)
94. Tomassini V, Johansen-Berg H, Jbabdi S, Sensorimotor cortical plasticity during recov-
Wise RG, Pozzilli C, Palace J et al (2012) ery following spinal cord injury: a longitudi-
Relating brain damage to brain plasticity in nal fMRI study. Neurorehabil Neural Repair
patients with multiple sclerosis. Neurorehabil 21(6):527–538
Neural Repair 26(6):581–593 108. Sabbah P, de Schonen S, Leveque C, Gay S,
95. Morgen K, Kadom N, Sawaki L, Tessitore A, Pfefer F, Nioche C et al (2002) Sensorimotor
Ohayon J, McFarland H et al (2004) Training- cortical activity in patients with complete spinal
dependent plasticity in patients with multiple cord injury: a functional magnetic resonance
sclerosis. Brain 127(Pt 11):2506–2517 imaging study. J Neurotrauma 19(1):53–60
96. Facts and Figures at a Glance—June 2006 109. Alkadhi H, Brugger P, Boendermaker S,
(2007) www.spinalcord.uab.edu Crelier G, Curt A, Hepp-Reymond M et al
97. Geisler F, Dorsey F, Coleman W (1991) (2005) What disconnection tells about motor
Recovery of motor function after spinal-cord imagery: evidence from paraplegic patients.
injury—a randomized, placebo-controlled Cereb Cortex 15(2):131–140
trial with GM-1 ganglioside. N Engl J Med 110. Humphrey D, Mao H, Schaeffer E (eds)
324(26):1829–1838 (2000) Voluntary activation of ineffective cere-
98. Ditunno J, Stover S, Freed M, Ahn J (1992) bral motor areas in short- and long-term para-
Motor recovery of the upper extremities in plegics. Society for Neuroscience (abstract)
traumatic quadriplegia: a multicenter study. 111. Topka H, Cohen L, Cole R, Hallett M (1991)
Arch Phys Med Rehabil 73(5):431–436 Reorganization of corticospinal pathways
99. Kirshblum S, Millis S, McKinley W, Tulsky D following spinal cord injury. Neurology
(2004) Late neurologic recovery after trau- 41(8):1276–1283
matic spinal cord injury. Arch Phys Med 112. Bruehlmeier M, Dietz V, Leenders K, Roelcke
Rehabil 85(11):1811–1817 U, Missimer J, Curt A (1998) How does the
Application of fMRI to Monitor Motor Rehabilitation 849
human brain deal with a spinal cord injury? 124. Park CH, Chang WH, Ohn SH, Kim ST,
Eur J Neurosci 10(12):3918–3922 Bang OY, Pascual-Leone A et al (2011)
113. Mikulis D, Jurkiewicz M, McIlroy W, Staines W, Longitudinal changes of resting-state func-
Rickards L, Kalsi-Ryan S et al (2002) Adaptation tional connectivity during motor recovery
in the motor cortex following cervical spinal after stroke. Stroke 42(5):1357–1362
cord injury. Neurology 58(5):794–801 125. Carter AR, Patel KR, Astafiev SV, Snyder AZ,
114. Turner J, Lee J, Martinez O, Medlin A, Rengachary J, Strube MJ et al (2012) Upstream
Schandler S, Cohen M (2001) Somatotopy of dysfunction of somatomotor functional connec-
the motor cortex after long-term spinal cord tivity after corticospinal damage in stroke.
injury or amputation. IEEE Trans Neural Syst Neurorehabil Neural Repair 26(1):7–19
Rehabil Eng 9(2):154–160 126. Varkuti B, Guan C, Pan Y, Phua KS, Ang KK,
115. Corbetta M, Burton H, Sinclair R, Conturo Kuah CW et al (2013) Resting state changes
T, Akbudak E, McDonald J (2002) Functional in functional connectivity correlate with
reorganization and stability of somatosensory- movement recovery for BCI and robot-
motor cortical topography in a tetraplegic assisted upper-extremity training after stroke.
subject with late recovery. Proc Natl Acad Sci Neurorehabil Neural Repair 27(1):53–62
U S A 99(26):17066–17071 127. Petsas N, Tomassini V, Filippini N, Sbardella
116. Sabre L, Tomberg T, Korv J, Kepler J, Kepler E, Tona F, Piattella MC et al (2015) Impaired
K, Linnamagi U et al (2013) Brain activation functional connectivity unmasked by simple
in the acute phase of traumatic spinal cord repetitive motor task in early relapsing-
injury. Spinal Cord 51(8):623–629 remitting multiple sclerosis. Neurorehabil
117. Winchester P, McColl R, Querry R, Foreman Neural Repair 29(6):557–565
N, Mosby J, Tansey K et al (2005) Changes in 128. Basile B, Castelli M, Monteleone F, Nocentini
supraspinal activation patterns following U, Caltagirone C, Centonze D et al (2013)
robotic locomotor therapy in motor- Functional connectivity changes within specific
incomplete spinal cord injury. Neurorehabil networks parallel the clinical evolution of mul-
Neural Repair 19(4):313–324 tiple sclerosis. Mult Scler 20(8):1050–1057
118. Lacourse MG, Turner JA, Randolph-Orr E, 129. Filippi M, Agosta F, Spinelli EG, Rocca MA
Schandler SL, Cohen MJ (2004) Cerebral (2013) Imaging resting state brain function in
and cerebellar sensorimotor plasticity follow- multiple sclerosis. J Neurol 260(7):1709–1713
ing motor imagery-based mental practice of a 130. Chisholm AE, Peters S, Borich MR, Boyd
sequential movement. J Rehabil Res Dev LA, Lam T (2015) Short-term cortical plas-
41(4):505–524 ticity associated with feedback-error learning
119. Sharma N, Pomeroy VM, Baron JC (2006) after locomotor training in a patient with
Motor imagery: a backdoor to the motor incomplete spinal cord injury. Phys Ther
system after stroke? Stroke 95(2):257–266
37(7):1941–1952 131. Hou JM, Sun TS, Xiang ZM, Zhang JZ,
120. Kleim J, Kleim E, Cramer S (2007) Systematic Zhang ZC, Zhao M et al (2014) Alterations
assessment of training-induced changes in of resting-state regional and network-level
corticospinal output to hand using frameless neural function after acute spinal cord injury.
stereotaxic transcranial magnetic stimulation. Neuroscience 277:446–454
Nat Protoc 2:1675–1684 132. Newton JM, Ward NS, Parker GJ, Deichmann
121. Kleim JA, Chan S, Pringle E, Schallert K, R, Alexander DC, Friston KJ et al (2006)
Procaccio V, Jimenez R et al (2006) BDNF val- Non-invasive mapping of corticofugal fibres
66met polymorphism is associated with modi- from multiple motor areas—relevance to
fied experience-dependent plasticity in human stroke recovery. Brain 129(Pt 7):1844–1858
motor cortex. Nat Neurosci 9(6):735–737 133. Crafton K, Mark A, Cramer S (2003)
122. Wadden KP, Woodward TS, Metzak PD, Improved understanding of cortical injury by
Lavigne KM, Lakhani B, Auriat AM et al incorporating measures of functional anat-
(2015) Compensatory motor network con- omy. Brain 126(Pt 7):1650–1659
nectivity is associated with motor sequence 134. Heiss W, Emunds H, Herholz K (1993)
learning after subcortical stroke. Behav Brain Cerebral glucose metabolism as a predictor of
Res 286:136–145 rehabilitation after ischemic stroke. Stroke
123. Golestani AM, Tymchuk S, Demchuk A, 24(12):1784–1788
Goodyear BG (2013) Longitudinal evalua- 135. Cappa S, Perani D, Grassi F, Bressi S, Alberoni
tion of resting-state FMRI after acute stroke M, Franceschi M et al (1997) A PET follow-
with hemiparesis. Neurorehabil Neural Repair up study of recovery after stroke in acute
27(2):153–163 aphasics. Brain Lang 56(1):55–67
Part IV
Abstract
Recent years have witnessed a rapid growth of interest in moving functional magnetic resonance imaging
(fMRI) beyond simple scan-length averages and into approaches that can integrate structural MRI measures
and capture rich multimodal interactions. It is becoming increasingly clear that multimodal fusion is able to
provide more information for individual subjects by exploiting covariation between modalities, rather an
analysis of each modality alone. Multimodal fusion is a more complicated endeavor that must be approached
carefully and efficient methods should be developed to draw generalized and valid conclusions out of high
dimensional data with a limited number of subjects, such as patients with brain disorders. Numerous research
efforts have been reported in the field based on various statistical models, including independent component
analysis (ICA), canonical correlation analysis (CCA), and partial least squares (PLS). In this chapter, we sur-
vey a number of methods previously shown in multimodal fusion reports, performed with or without prior
information, and with their possible strengths and limitations addressed. To examine the function–structure
associations of the brain in a more comprehensive and integrated manner, we also reviewed a number of
multimodal studies that combined fMRI and structural (sMRI and/or diffusion tensor MRI) measures,
which could reveal important brain alterations that may not be fully detected by employing separate analysis
of individual modalities, and also enable us to identify potential brain illness biomarkers.
Key words Multimodal fusion methods, Data driven, Functional magnetic resonance imaging,
Structural MRI, Diffusion MRI, Independent component analysis, Canonical correlation analysis
1 Introduction
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_28, © Springer Science+Business Media New York 2016
853
854 Jing Sui and Vince D. Calhoun
Fig. 1 Summary of the current multimodal data analysis approaches related to fMRI
856 Jing Sui and Vince D. Calhoun
Fig. 2 A summary of seven blind and semi-blind data-driven methods for multimodal fusion. Figure modified
and reprinted with permission from Sui et al. [4]
858 Jing Sui and Vince D. Calhoun
3.1 Joint ICA Joint ICA (jICA) is a second-level fMRI analysis method that
assumes two or more features (modalities) share the same mixing
matrix and maximizes the independence among joint components
[36]. This is a straightforward yet effective method by performing
ICA on the horizontally concatenated features (along voxels). It is
suitable for examining the common connection among modalities
and requires acceptance of the likelihood of changes in one data
type (e.g., GM) being related to another one, such as functional
activation. Joint ICA is feasible to many paired combinations of
features, such as fMRI, sMRI, and DT MRI, or three-way data
fusion [37–41]. In order to control for intensity differences in MR
images based on scanner, template, and population variations, usu-
ally each feature matrix (dimension: number of subject by number
of voxel) is normalized to a study specific template [42, 43].
Figure 3 [36], shows analyzed data collected from groups of
schizophrenia patients and healthy controls using the jICA
approach. The main finding was that group differences in bilateral
parietal and frontal as well as posterior temporal regions in GM
matter distinguished groups. A finding of less patient GM and less
hemodynamic activity for target detection in these bilateral anterior
temporal lobe regions was consistent with previous work. An unex-
pected corollary to this finding was that, in the regions showing the
largest group differences, GM concentrations were larger in patients
vs. controls, suggesting that more GM may be related to less func-
tional connectivity during performance of an auditory oddball task.
Fig. 3 Auditory oddball/gray matter jICA analysis. Only one component demonstrated a significant difference
between patients and controls. The joint source map for the auditory oddball fMRI data (left) and gray matter
(middle) data is presented along with the loading parameters for patients and controls (far right). Reproduced
with permission from Calhoun et al. [36]
Multimodal Fusion of Structural and Functional Brain Imaging Data 859
3.2 Multimodal CCA Multimodal CCA allows a different mixing matrix for each modality
and is used to find the transformed coordinate system that maxi-
mizes these inter-subject covariations across the two data sets [44].
This method decomposes each dataset into a set of components
and their corresponding mixing profile, which is called canonical
variants (CVs). The CVs have varying levels of activations for dif-
ferent subjects and are linked if they modulate similarly across sub-
jects. After decomposition, the CVs correlate each other only on
the same indices and their corresponding correlation values are
called canonical correlation coefficients. Compared to jICA that
constrains two features to have the same mixing matrix, mCCA is
flexible in that it allows common as well as distinct level of connec-
tion between two features, as shown in Fig. 4, but the associated
source maps may not be spatially sparse, especially when the canon-
ical correlation coefficients are not sufficiently distinct [4].
Multimodal CCA is invariant to differences in the range of the
data types and can be used to jointly analyze very diverse data
Fig. 4 MCCA + jICA enables people to capture components of interest that are either common or distinct across
modalities. For example, when examining group differences across three modalities, joint ICs are significantly
group-discriminative in more than two modalities (green framed), while modality-specific discriminative ICs
(pink framed), i.e., fMRI_IC4, DTI_IC3, and DTI_IC7 only show significant group difference in a single modality.
Reproduced with permission from Sui et al. [5]
860 Jing Sui and Vince D. Calhoun
3.3 Partial Least Partial Least Squares (PLS) as part of a family of multivariate data
Squares analyses, is based on the definition of a linear relationship between
a dependent variable and predictor variables, and hence the goal is
to determine which aspect of a set of observations (e.g., imaging
data) are related directly to another set of data (e.g., experimental
design, behavior) [46]. PLS was first applied to multimodal fusion
by Martinen-Montes et al. [28], where the multiway PLS (N-PLS)
was proposed to find correlations between fMRI time courses
(dependent variables) and the spectral components of the EEG
data (independent variables) from a single subject. Chen et al. [27]
further proposed a multimodal PLS (MMPLS) to simultaneously
characterize the linkage between patterns of PET and GM. The
multimodal PLS can be performed either informed to the variable
of interest such as age or agnostic to this additional information
(with or without priori). Investigators may want to pre-specify
which of these two methods to use in the data analysis. The agnos-
tic MMPLS was used to identify the linkage between PET (depen-
dent data block) and sMRI (independent data block), which is
mainly oriented for extraction of covariance patterns and related
latent variables rather than for classifications.
Even though PLS has some similarity to CCA in that they both
maximize between-set correlations, PLS is based on the definition
of a dependency and works well especially when the dependence
among the constituents of the datasets is explicitly assessed [47];
while CCA does not assign independent/dependent labels to either
of the modalities and treats both equally [48]. Hence PLS is par-
ticularly suited to the analysis of relationships between measures of
brain activity and of behavior or experimental design [49], while
interpretational difficulties in PLS often arise when the effects iden-
tified do not correspond to the a priori expectations of the researcher.
3.4 mCCA + jICA According to many previous findings in brain connectivity studies
which combined function and structure [19, 50], it is plausible to
assume the components decomposed from each modality have some
degree of correlation between their mixing profiles among subjects.
mCCA + jICA is a blind data-driven model that is optimized for this
situation [26, 51] and also to have excellent performance for achiev-
ing both flexible modal association and source separation. It takes
advantage of two complementary approaches: mCCA and jICA,
thus allowing both strong and weak inter-modality connection as
well as the joint independent components. mCCA makes the jICA
Multimodal Fusion of Structural and Functional Brain Imaging Data 861
3.5 Linked ICA Linked ICA is a fully probabilistic approach based on a modular
Bayesian framework, which is designed for simultaneously model-
ing and discovering common characteristics across multiple modal-
ities [24]. The combined modalities can potentially have different
units, noise level, spatial smoothness, and intensity distributions.
Each modality group is modeled using a Bayesian tensor ICA
model [52]. Note that the Bayesian ICA differs from standard
methods like FastICA [53] and Infomax [54] in that it incorpo-
rates dimensionality reduction into the ICA method itself by the
use of automatic relevance determination priors on components
[55] and works on the full-dimensionality data directly. Linked
ICA is good at detecting and isolating single-modality noise; how-
ever, it is much more computationally demanding than the above
four methods, and the spatial maps of the decomposed compo-
nents tend to be scattered. In a real application, linked ICA was
applied to high functioning autism by combining DT MRI, voxel-
based morphometry (VBM), and resting-state fMRI connectivity
to assess differences in brain structure and function [56].
5 Conclusions
Table 1
Summary of studies combining measures of functional and structural MRI
References
27. Chen K, Reiman EM, Huan Z et al (2009) 39. Calhoun VD, Adali T, Liu J (2006) A feature-
Linking functional and structural brain images based approach to combine functional MRI,
with multivariate network analyses: a novel structural MRI and EEG brain imaging data.
application of the partial least square method. Conf Proc IEEE Eng Med Biol Soc 1:3672–3675
Neuroimage 47(2):602–610 40. Teipel SJ, Bokde AL, Meindl T et al (2010)
28. Martinez-Montes E, Valdes-Sosa PA, White matter microstructure underlying
Miwakeichi F et al (2004) Concurrent EEG/ default mode network connectivity in the
fMRI analysis by multiway Partial Least human brain. Neuroimage 49(3):2021–2032
Squares. Neuroimage 22(3):1023–1034 41. Eichele T, Calhoun VD, Debener S (2009) Mining
29. Correa NM, Eichele T, Adali T et al (2010) EEG-fMRI using independent component analy-
Multi-set canonical correlation analysis for the sis. Int J Psychophysiol 73(1):53–61
fusion of concurrent single trial ERP and func- 42. Xu L, Groth KM, Pearlson G et al (2009)
tional MRI. Neuroimage 50(4):1438–1445 Source-based morphometry: the use of
30. Liu J, Pearlson G, Windemuth A et al (2009) independent component analysis to iden-
Combining fMRI and SNP data to investi- tify gray matter differences with applica-
gate connections between brain function and tion to schizophrenia. Hum Brain Mapp
genetics using parallel ICA. Hum Brain Mapp 30(3):711–724
30(1):241–255 43. Shenton ME, Dickey CC, Frumin M et al
31. Sui J, Adali T, Pearlson GD et al (2009) A (2001) A review of MRI findings in schizo-
method for accurate group difference detec- phrenia. Schizophr Res 49(1–2):1–52
tion by constraining the mixing coefficients 44. Correa NM, Li YO, Adali T et al (2008)
in an ICA framework. Hum Brain Mapp Canonical correlation analysis for feature-
30(9):2953–2970 based fusion of biomedical imaging modalities
32. Sui J, Adali T, Pearlson GD et al (2009) An and its application to detection of associative
ICA-based method for the identification networks in schizophrenia. IEEE J Sel Top
of optimal FMRI features and components Signal Process 2(6):998–1007
using combined group-discriminative tech- 45. Li Y-O, Adali T, Wang W et al (2009) Joint
niques. Neuroimage 46(1):73–86 blind source separation by multiset canoni-
33. Caprihan A, Pearlson GD, Calhoun VD (2008) cal correlation analysis. IEEE Trans Signal
Application of principal component analysis to Process 57(10):3918–3929
distinguish patients with schizophrenia from 46. Lin FH, McIntosh AR, Agnew JA et al (2003)
healthy controls based on fractional anisotropy Multivariate analysis of neuronal interac-
measurements. Neuroimage 42(2):675–682 tions in the generalized partial least squares
34. Liu J, Xu L, Calhoun VD (2008). Extracting framework: simulations and empirical studies.
principle components for discriminant analy- Neuroimage 20(2):625–642
sis of FMRI images. ICASSP 449–452 47. Krishnan A, Williams LJ, McIntosh AR et al
35. Ulloa AE, Chen J, Vergara VM et al (2014) (2010) Partial Least Squares (PLS) meth-
Association between copy number varia- ods for neuroimaging: a tutorial and review.
tion losses and alcohol dependence across Neuroimage 56(2):455–475
African American and European American 48. Correa NM, Adali T, Li YO et al (2010)
ethnic groups. Alcohol Clin Exp Res Canonical correlation analysis for data fusion
38(5):1266–1274 and group inferences: examining applications
36. Calhoun VD, Adali T, Giuliani NR et al of medical imaging data. IEEE Signal Process
(2006) Method for multimodal analysis of Mag 27(4):39–50
independent source differences in schizophre- 49. McIntosh AR, Lobaugh NJ (2004) Partial
nia: combining gray matter structural and least squares analysis of neuroimaging data:
auditory oddball functional data. Hum Brain applications and advances. Neuroimage
Mapp 27(1):47–62 23(Suppl 1):S250–S263
37. Xu L, Pearlson G, Calhoun VD (2009) Joint 50. Camara E, Rodriguez-Fornells A, Munte
source based morphometry identifies linked TF (2010) Microstructural brain differences
gray and white matter group differences. predict functional hemodynamic responses
Neuroimage 44(3):777–789 in a reward processing task. J Neurosci
38. Franco AR, Ling J, Caprihan A et al (2008) 30(34):11398–11402
Multimodal and multi-tissue measures of 51. Sui J, He H, Yu Q et al (2013) Combination
connectivity revealed by joint independent of resting state fMRI, DTI, and sMRI data
component analysis. IEEE J Sel Top Signal to discriminate schizophrenia by N-way
Process 2(6):986–997 MCCA + jICA. Front Hum Neurosci 7:235
Multimodal Fusion of Structural and Functional Brain Imaging Data 867
52. Beckmann CF, Smith SM (2005) Tensorial 65. Kim J, Lee JH (2012) Integration of struc-
extensions of independent component analysis tural and functional magnetic resonance
for multisubject FMRI analysis. Neuroimage imaging improves mild cognitive impairment
25(1):294–311 detection. Magn Reson Imaging
53. Hyvarinen A, Oja E (2000) Independent 31:718–732
component analysis: algorithms and applica- 66. Fan Y, Resnick SM, Wu X et al (2008)
tions. Neural Netw 13(4–5):411–430 Structural and functional biomarkers of pro-
54. Bell AJ, Sejnowski TJ (1995) An information- dromal Alzheimer’s disease: a high-
maximization approach to blind separation dimensional pattern classification study.
and blind deconvolution. Neural Comput Neuroimage 41(2):277–285
7(6):1129–1159 67. Wee CY, Yap PT, Zhang D et al (2011)
55. Bishop CM (1999) Variational principal com- Identification of MCI individuals using
ponents. Artif Neural Netw 7:509, structural and functional connectivity net-
Conference Publication No. 470 works. Neuroimage 59(3):2045–2056
56. Liu J, Chen J, Ehrlich S et al (2014) 68. Westlye LT, Walhovd KB, Bjornerud A et al
Methylation patterns in whole blood corre- (2009) Error-related negativity is mediated
late with symptoms in schizophrenia patients. by fractional anisotropy in the posterior cin-
Schizophr Bull 40(4):769–776 gulate gyrus—a study combining diffusion
57. Segall JM, Allen EA, Jung RE et al (2012) tensor imaging and electrophysiology in
Correspondence between structure and func- healthy adults. Cereb Cortex 19(2):293–304
tion in the human brain at rest. Front 69. Sui J, Yu Q, He H et al (2012) A selective
Neuroinform 6:10 review of multimodal fusion methods in
58. Khullar S, Michael AM, Cahill ND et al schizophrenia. Front Hum Neurosci 6:27
(2011) ICA-fNORM: spatial normalization 70. Zhang H, Liu L, Wu H, Fan Y (2012) Feature
of fMRI data using intrinsic group-ICA net- selection and SVM classification of multiple
works. Front Syst Neurosci 5:93 modality images for predicting MCI, in
59. Salgado-Pineda P, Junque C, Vendrell P et al OHBM, Beijing, China.
(2004) Decreased cerebral activation during 71. Kim DI, Sui J, Rachakonda S et al (2010)
CPT performance: structural and functional Identification of imaging biomarkers in
deficits in schizophrenic patients. Neuroimage schizophrenia: a coefficient-constrained inde-
21(3):840–847 pendent component analysis of the mind
60. Salgado-Pineda P, Fakra E, Delaveau P et al multi-site schizophrenia study.
(2011) Correlated structural and functional Neuroinformatics 8(4):213–229
brain abnormalities in the default mode net- 72. Tian L, Meng C, Yan H et al (2011)
work in schizophrenia patients. Schizophr Convergent evidence from multimodal imag-
Res 125(2–3):101–109 ing reveals amygdala abnormalities in schizo-
61. Skudlarski P, Jagannathan K, Anderson K et al phrenic patients and their first-degree
(2010) Brain connectivity is not only lower relatives. PLoS One 6(12):e28794
but different in schizophrenia: a combined 73. Casey BJ, Tottenham N, Liston C et al (2005)
anatomical and functional approach. Biol Imaging the developing brain: what have we
Psychiatry 68(1):61–69 learned about cognitive development? Trends
62. Schlosser RG, Nenadic I, Wagner G et al Cogn Sci 9(3):104–110
(2007) White matter abnormalities and brain 74. Smieskova R, Allen P, Simon A et al (2012)
activation in schizophrenia: a combined DTI Different duration of at-risk mental state asso-
and fMRI study. Schizophr Res ciated with neurofunctional abnormalities. A
89(1–3):1–11 multimodal imaging study. Hum Brain Mapp
63. Staempfli P, Reischauer C, Jaermann T et al 33(10):2281–2294
(2008) Combining fMRI and DTI: a frame- 75. Rasser PE, Johnston P, Lagopoulos J et al
work for exploring the limits of fMRI-guided (2005) Functional MRI BOLD response to
DTI fiber tracking and for verifying DTI- Tower of London performance of first-
based fiber tractography results. Neuroimage episode schizophrenia patients using cortical
39(1):119–126 pattern matching. Neuroimage
64. Koch K, Wagner G, Schachtzabel C et al 26(3):941–951
(2011) Neural activation and radial diffusivity 76. Fusar-Poli P, Broome MR, Woolley JB et al
in schizophrenia: combined fMRI and diffu- (2011) Altered brain function directly related
sion tensor imaging study. Br J Psychiatry to structural abnormalities in people at ultra
198(3):223–229 high risk of psychosis: longitudinal VBM-
fMRI study. J Psychiatr Res 45(2):190–198
868 Jing Sui and Vince D. Calhoun
77. Michael AM, Baum SA, White T et al (2010) 89. Wang F, Kalmar JH, He Y et al (2009)
Does function follow form?: methods to fuse Functional and structural connectivity
structural and functional brain images show between the perigenual anterior cingulate and
decreased linkage in schizophrenia. amygdala in bipolar disorder. Biol Psychiatry
Neuroimage 49(3):2626–2637 66(5):516–521
78. Michael AM, King MD, Ehrlich S et al (2011) A 90. Schonberg T, Pianka P, Hendler T et al
data-driven investigation of gray matter-function (2006) Characterization of displaced white
correlations in schizophrenia during a working matter by brain tumors using combined DTI
memory task. Front Hum Neurosci 5:71 and fMRI. Neuroimage 30(4):1100–1111
79. Rektorova I, Mikl M, Barrett J et al (2012) 91. Voss HU, Schiff ND (2009) MRI of neuronal
Functional neuroanatomy of vocalization in network structure, function, and plasticity.
patients with Parkinson’s disease. J Neurol Sci Prog Brain Res 175:483–496
313(1–2):7–12 92. Palacios EM, Sala-Llonch R, Junque C et al
80. Harms MP, Wang L, Csernansky JG et al (2012) White matter integrity related to
(2012) Structure-function relationship of functional working memory networks in trau-
working memory activity with hippocampal matic brain injury. Neurology
and prefrontal cortex volumes. Brain Struct 78(12):852–860
Funct 218(1):173–186 93. Jacobson S, Kelleher I, Harley M et al (2009)
81. Choi K, Yang Z, Hu X et al (2008) A com- Structural and functional brain correlates of
bined functional-structural connectivity anal- subclinical psychotic symptoms in 11–13 year
ysis of major depression using joint old schoolchildren. Neuroimage
independent components analysis. Psychiatric 49(2):1875–1885
MRI/MRS 3555 94. Supekar K, Uddin LQ, Prater K et al (2010)
82. Camchong J, MacDonald AW III, Bell C et al Development of functional and structural
(2011) Altered functional and anatomical connectivity within the default mode network
connectivity in schizophrenia. Schizophr Bull in young children. Neuroimage
37(3):640–650 52(1):290–301
83. Kim DJ, Park B, Park HJ (2012) Functional 95. Pomarol-Clotet E, Canales-Rodriguez EJ,
connectivity-based identification of subdivi- Salvador R et al (2010) Medial prefrontal cor-
sions of the basal ganglia and thalamus using tex pathology in schizophrenia as revealed by
multilevel independent component analysis of convergent findings from multimodal imag-
resting state fMRI. Hum Brain Mapp ing. Mol Psychiatry 15(8):823–830
34:1371–1385 96. Sexton CE, Allan CL, Le Masurier M et al
84. Olesen PJ, Nagy Z, Westerberg H et al (2003) (2012) Magnetic resonance imaging in late-
Combined analysis of DTI and fMRI data life depression: multimodal examination of
reveals a joint maturation of white and grey network disruption. Arch Gen Psychiatry
matter in a fronto-parietal network. Brain Res 69(7):680–689
Cogn Brain Res 18(1):48–57 97. Qiu MG, Ye Z, Li QY et al (2011) Changes of
85. Matthews SC, Strigo IA, Simmons AN et al brain structure and function in ADHD chil-
(2011) A multimodal imaging study in U.S. dren. Brain Topogr 24(3–4):243–252
veterans of Operations Iraqi and Enduring 98. Sui J, He H, Liu J et al (2012) Three-way
Freedom with and without major depression FMRI-DTI-methylation data fusion based on
after blast-related concussion. Neuroimage mCCA + jICA and its application to schizo-
54(Suppl 1):S69–S75 phrenia. Conf Proc IEEE Eng Med Biol Soc
86. Zhou Y, Shu N, Liu Y et al (2008) Altered 2012:2692–2695
resting-state functional connectivity and ana- 99. Zhang D, Wang Y, Zhou L et al (2011)
tomical connectivity of hippocampus in Multimodal classification of Alzheimer’s dis-
schizophrenia. Schizophr Res ease and mild cognitive impairment.
100(1–3):120–132 Neuroimage 55(3):856–867
87. Yan H, Tian L, Yan J et al (2012) Functional 100. Groves AR, Smith SM, Fjell AM et al (2012)
and anatomical connectivity abnormalities in Benefits of multi-modal fusion analysis on a
cognitive division of anterior cingulate cortex large-scale dataset: life-span patterns of inter-
in schizophrenia. PLoS One 7(9):e45659 subject variability in cortical morphometry and
88. Soldner J, Meindl T, Koch W et al (2011) white matter microstructure. Neuroimage
Structural and functional neuronal connectiv- 63(1):365–380
ity in Alzheimer’s disease: a combined DTI 101. Sui J, Castro E, Hao H et al (2014) Combination
and fMRI study. Nervenarzt 83(7):878–887 of FMRI-SMRI-EEG data improves discrimina-
Multimodal Fusion of Structural and Functional Brain Imaging Data 869
tion of schizophrenia patients by ensemble fea- 106. Jagannathan K, Calhoun VD, Gelernter J et al
ture selection. The 36th Annual International (2010) Genetic associations of brain structural
Conference of the IEEE engineering in medi- networks in schizophrenia: a preliminary
cine and biology society (EMBC‘14) Chicago, study. Biol Psychiatry 68(7):657–666
Illinois, USA, no. August 26–30 107. Jamadar S, Powers NR, Meda SA et al (2010)
102. Eichele T, Calhoun VD, Moosmann M et al Genetic influences of cortical gray matter in
(2008) Unmixing concurrent EEG-fMRI language-related regions in healthy controls and
with parallel independent component analy- schizophrenia. Schizophr Res 129:141–148
sis. Int J Psychophysiol 67(3):222–234 108. Hao X, Xu D, Bansal R et al (2011) Multimodal
103. Haller S, Xekardaki A, Delaloye C et al (2011) magnetic resonance imaging: the coordinated
Combined analysis of grey matter voxel-based use of multiple, mutually informative probes to
morphometry and white matter tract-based understand brain structure and function. Hum
spatial statistics in late-life bipolar disorder. Brain Mapp 34(2):253–271
J Psychiatry Neurosci 36(6):391–401 109. Fusar-Poli P, McGuire P, Borgwardt S (2011)
104. Chen Z, Cui L, Li M et al (2011) Voxel based Mapping prodromal psychosis: a critical
morphometric and diffusion tensor imaging review of neuroimaging studies. Eur
analysis in male bipolar patients with first- Psychiatry 27(3):181–191
episode mania. Prog Neuropsychopharmacol 110. Meda SA, Gill A, Stevens MC et al (2012)
Biol Psychiatry 36(2):231–238 Differences in resting-state functional mag-
105. Meda SA, Jagannathan K, Gelernter J et al netic resonance imaging functional network
(2010) A pilot multivariate parallel ICA study connectivity between schizophrenia and psy-
to investigate differential linkage between chotic bipolar probands and their unaffected
neural networks and genetic profiles in schizo- first-degree relatives. Biol Psychiatry
phrenia. Neuroimage 53(3):1007–1015 71(10):881–889
Chapter 29
Abstract
Evidence to date shows that fMRI of the spinal cord (spinal fMRI) can reliably demonstrate regions
involved with sensation of tactile, thermal, and painful stimuli, and with motor tasks. Spinal fMRI acquisi-
tion methods based on BOLD contrast have been recently optimized. Results have demonstrated the
ability of spinal fMRI to provide objective assessments of sensory and motor function, and discriminate
responses when modulated by cognitive/emotional factors. Studies have been also carried out with patients
with cord trauma, and in people with multiple sclerosis (MS). The availability of essentially automated
analysis, large extent coverage of the spinal cord, and spatial normalization to permit comparisons with
reference results and labeling of active regions are being implemented with the aim to translate the method
into a practical clinical assessment tool.
The research completed so far indicates that spinal fMRI will be able to demonstrate where the neu-
ronal activity is altered at any level (cervical, thoracic, lumbar, or sacral), whether or not information is
reaching the cord from the periphery, and whether or not there is descending modulation of the response.
It may also be able to provide an objective measure of pain, and to demonstrate the extent and mechanism
of changes over time after an injury.
Key words Spinal fMRI, Blood oxygen level dependent, Multiple sclerosis, Cord trauma, Pain
1 Introduction
Evidence to date shows that fMRI of the spinal cord (spinal fMRI)
can reliably demonstrate regions involved with sensation of tactile,
thermal, and painful stimuli, and with motor tasks. There is also
reliable evidence of the descending modulation of activity in the
spinal cord. While spinal fMRI has not yet been applied or verified
in a clinical setting, its value is expected to be in its ability to pro-
vide objective assessments of sensory and motor function, and dis-
criminate responses when modulated by cognitive/emotional
factors, or even detect responses when a patient cannot feel the
stimulus or is even conscious. Studies have been carried out with
patients with cord trauma, and in people with multiple sclerosis
(MS) to investigate the clinical utility of the results. Robust meth-
ods for analysis, and for displaying the results in an effective m
anner
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1_29, © Springer Science+Business Media New York 2016
871
872 Patrick Stroman and Massimo Filippi
Fig. 1 Anatomy of the spinal cord (reproduced with permission from Blumenfeld H., Neuroanatomy Through
Clinical Cases. Sunderland, MA: Sinauer Associates; 2002, p 22)
874 Patrick Stroman and Massimo Filippi
3.1 Gradient-Echo As mentioned above, the earliest results with BOLD methods [3,
Methods 6–9, 18] showed promise that spinal fMRI is feasible. Yoshizawa
et al. demonstrated areas of activity with a hand motor task (average
signal changes 4.8 %) corresponded with consistent areas of the
spinal cord gray matter; they also showed that the rostral-caudal
distribution of activity corresponded well with the neuroanatomy
[3]. This was the first spinal fMRI study, and it set the standard for
the studies which followed by Stroman et al. [6, 9], Madi et al. [7],
and Backes et al. [8], which had a number of similarities. As in the
study by Yoshizawa et al., each of these employed gradient-echo
imaging (fast gradient-echo or echo-planar encoding), relatively
thick (5–10 mm) transverse slices, with the echo-time (TE) set for
BOLD sensitivity. All of these studies investigated activity with
motor tasks, two investigated activity with sensory stimuli as well
[6, 9], and three of them [6–8] compared results obtained with
sagittal (4–8 mm) and transverse slices.
The results of these studies showed a number of consistent fea-
tures (Fig. 2). Signal intensity changes with hand motor tasks were
consistently in the range of 4.3–4.8 % by Yoshizawa et al. [3] and
Stroman et al. [6], and 0.5–7.5 % with graded force tasks by Madi
et al. [7], and 8–12 % by Backes et al. [8], although there were
Fig. 2 Example of spinal fMRI results obtained with gradient-recalled echo EPI in transverse slices of the cervi-
cal spinal cord (reproduced from Stroman and Ryner. Magn. Reson. Imaging 19: 27–32, 2001) [9]. (a) The left
side of the body is at the top of this image and the red marks indicate the locations which underwent intensity
changes in response to sensory stimulation of the right hand. The location of the slice is indicated in relation
to the sagittal view of the cervical spine shown in the larger image. (b) Five transverse slices corresponding to
the slices indicated in the sagittal view
876 Patrick Stroman and Massimo Filippi
3.2 Spin-Echo One of the earliest spinal fMRI studies [10] was a comparison of
Methods the signal intensity time-course properties obtained with T2-
weighted and T2*-weighted acquisitions. The intent was to inves-
tigate whether the BOLD effect occurred in the spinal cord as in
the brain. The T2-weighted data had signal changes that were as
large, or larger, than the T2*-weighted data at approximately the
same echo time, and so the observations were not entirely consis-
tent with the BOLD model. More importantly, the results indi-
cated that spinal fMRI is feasible with both motor and sensory
stimulation at 1.5 T, and can be achieved with good image quality
with spin-echo imaging methods. A series of studies followed, to
investigate the biophysical nature of the underlying contrast mech-
anism, and consistently showed significant BOLD effects, as well as
a contribution from a proton-density change which was greater at
shorter echo times [12–16, 27–30]. However, a key observation
across these studies was that the spatial encoding method—EPI or
fast spin-echo—is a critical factor in the choice of methods. A
detailed analysis of the methods, including characterizations of the
noise, physiological motion, and sensitivity to neuronal activity,
demonstrated several key findings [17]. These included that it is
important to avoid EPI methods for spinal fMRI, and that optimal
sensitivity is obtained with spin-echo fMRI with an echo time (TE)
of 75 ms (at 3 T), which corresponds with the T2 of the spinal cord
tissue. This finding agrees with the established BOLD theory. This
increase in TE, compared to methods used in earlier studies, repre-
sented a small (20 %) but significant increase in sensitivity. These
findings also confirmed that the previous studies with an echo time
of 38 ms were likely dominated by BOLD contrast, with only a
small contribution from the proton-density change [31–37]. This
point is important because it means that earlier studies done with a
shorter echo time, and the more recent studies that are optimized
for BOLD contrast, are dominated by the same contrast mecha-
nism and the results are comparable.
878 Patrick Stroman and Massimo Filippi
The spinal fMRI methods that have been used in studies to date
appear to have converged on two widely-used methods, one based
on spin-echo to avoid EPI spatial encoding, and the other based on
gradient-echo EPI. In terms of the magnitude of the BOLD response
that is expected, when the two methods are both optimized, they
can be expected to be approximately equal, as detailed in Table 1.
In Table 1 the terms Δ(1/T2*) and Δ(1/T2) refer to the changes
in transverse relaxation rates between two conditions, such as a
“baseline” state and “active” state. It is well established that, under
the same experimental conditions, Δ(1/T2*) is three to four times
larger than Δ(1/T2) and hence T2*-weighted imaging is most fre-
quently used for fMRI [11, 28]. However, another important fac-
tor is that the echo time, TE, for optimal BOLD contrast is equal
to the transverse relaxation time (i.e., T2* or T2) which is roughly
three times larger for spin-echo than gradient-echo. The net effect
is that the magnitude of the BOLD response detected with spin-
echo methods is nearly equal to that with gradient-echo methods,
when both methods are set for optimal BOLD sensitivity.
The signal-to-noise ratio and image quality are also important
considerations for fMRI methods. The SNR can be compared
between imaging methods with the expression [53]:
Table 1
Comparison of the expected BOLD response magnitude detected with
gradient-echo and spin-echo methods
Gradient-echo Spin-echo
DS æ 1 ö DS æ 1 ö
@ -TE D ç * ÷ @ -TE D ç ÷
S è T2 ø S è T2 ø
DS DS
@ -0.025 s ´ -1.22s -1 @ -0.075 s ´ -0.37 s -1
S S
DS DS
@ 0.031 @ 0.028
S S
Estimates are for data acquired at 3 T, in the spinal cord, with TE values set for optimal
BOLD contrast. The values for the expected changes in relaxation rates, Δ(1/T2*) and
Δ(1/T2), are taken from [28]
erfc -1 ( p ) 2
eff =
SNR NR (1 - R )
where “eff” is the effect size, “p” is the statistical threshold used,
the function “erfc−1” is the inverse complementary error function,
and “R” is the proportion of time spent in the stimulation condi-
tion, assuming a block design with only two conditions. For the
purposes of this comparison we can set R = 0.5, and p = 10−6, which
corresponds to erfc−1(p) = 3.46. An estimate of the corresponding
t-value is given by √2 erfc−1(p), which is equal to 5.0. Using these
numbers we can compare the relative sensitivities of the methods,
in terms of the % BOLD signal change that can be detected with a
fixed acquisition duration (Table 2).
These estimates show that the faster sampling of the gradient-
echo EPI method offsets its lower SNR and improves its sensitivity,
but it does not reach the sensitivity of the spin-echo HASTE method.
With either of these methods the duration of the fMRI acquisitions
can be increased to provide greater sensitivity, within practical limits.
882 Patrick Stroman and Massimo Filippi
Table 2
Estimates of SNR for commonly used gradient-echo and spin-echo spinal fMRI acquisitions
Gradient-echo spinal
Typical brain fMRI fMRI Spin-echo spinal fMRI
Imaging 3.3 mm × 3.3 mm 1 mm × 1 mm 1.5 mm × 1.5 mm
parameters 3.3 mm thick slice 5 mm thick slice 2 mm thick slice
200 kHz bandwidth 200 kHz bandwidth 151 kHz bandwidth
64 × 64 matrix 128 × 128 matrix 192 × 144 matrix
TE = T2* TE: 43 ms (~1.7 T2*) TE: 75 ms (T2)
Acceleration factor = 1 (no acceleration factor = 2 acceleration factor = 1
parallel imaging assumed)
Estimated SNR 150 15 56
Acquisition time/ 3s 1.1 s 6.75 s
volume
Estimated effect 0.42 % 2.6 % 1.7 %
size
p = 10−6, 12 min
acquisition
SNR values are estimated compared to a typical brain fMRI method which is assumed to have an SNR of approximately
150 at 3 T
5 Recent Developments
Fig. 4 Comparison of image quality obtained with gradient-echo EPI and spin-echo HASTE sequences for
spinal fMRI. Gradient-echo EPI images were acquired in contiguous axial slices (left panel) and were reformat-
ted into sagittal views (middle frame) for comparison with spin-echo HASTE images acquired in sagittal planes
(right frame). Selected axial slices are shown for comparison, and the slice positions are indicated relative to
the caudal medulla (top slice)
Fig. 5 A 3D-normalized reference template for the cervical spinal cord, brain-
stem, and medial regions of the thalamus and corpus collosum. Image views
through the center of the volume are shown in axial (upper left), sagittal (upper
right), and coronal (lower right) slices
method has been developed for spinal cord fMRI data, written in
MatLab® (The Mathworks Inc.), based on prior descriptions [66,
67]. The method is based on identifying coordinated BOLD
responses between regions, accounting for the fact that input to one
region may arise simultaneously from multiple other regions [68,
69]. The BOLD signal time-courses in each region are expressed as
a linear combination of the BOLD signal time-courses in other
regions, and the weighting factors for the linear combination (i.e.
the “connectivity strengths”) are determined for a complete net-
work [70]. This analysis requires a predefined anatomical model of
all plausible connections between regions, which is provided by the
normalized temperature and region mask described above. Possible
connectivity relationships between the regions were identified based
on the extensive description of the regions/networks involved in
pain processing provided by Millan [71]. An example of SEM results
from Bosma and Stroman [40] is shown in Fig. 6.
Recent advances in the applications of spinal fMRI serve to fur-
ther demonstrate the reliability and sensitivity of the results, and their
potential value for future clinical applications. The first detailed rest-
ing-state study was carried out by Barry et al. [72] and used a 3D
multishot gradient-echo sequence at 7 T. They demonstrated func-
tional connectivity between right- and left-side gray matter in the spi-
nal cord, in the resting state. This is an important finding for spinal
cord fMRI in general, because of the variations in the baseline state
that may occur, even when a stimulus is not applied. Related studies
have also been carried out using spin-echo methods, by using either
thought directed at a particular region of the body [73] or images
displayed to the participant [74, 75]. In each of these studies, signal
variations were detected in the spinal cord in response to the cogni-
tive/emotional stimuli, demonstrating the influence of descending
input to the spinal cord. These findings may be related to the recent
demonstrations of effects such as placebo [23]. Detailed studies of
886 Patrick Stroman and Massimo Filippi
Fig. 6 SEM results obtained from healthy participants with a block-design stimulus at 45 °C, and a step-wise
increase in temperature to 45 °C. Arrows indicate the direction of the influence, while the line thickness indi-
cates the strength of the SEM connectivity. Solid lines represent positive path coefficients and dotted lines
represent negative coefficients. Abbreviations are as follows: Cord right dorsal region of the C8 spinal cord
segment, PBN parabrachial nucleus, LC locus coeruleus, NRM nucleus raphe magnus, NTS nucleus tractus
solitarius, NGC nucleus gigantocellularis, DRt dorsal reticular nucleus, PAG periaqueductal gray matter, HYP
hypothalamus, Thal thalamus. This figure is reproduced from Bosma and Stroman, J Magn Reson Imaging
2014 (DOI: 10.1002/jmri.24656) [41]
pain processing have also been carried out, showing effects of pain
modulation by changes in attention focus and by listening to music
[23, 24, 42]. Individual differences in pain processing have also been
investigated, and demonstrate correlations between BOLD responses
and individual pain ratings in the PAG, PBN, and spinal cord dorsal
horn [76]. This study provides evidence that spinal fMRI methods are
adequately sensitive to provide characterizations of the pain state in
individuals. Moreover, a study of the effects of spinal cord injury on
thermal sensory processing has demonstrated plasticity, with signifi-
cant differences detected in individuals compared to a group of healthy
control participants [50]. This body of recent work shows that spinal
fMRI methods and applications are rapidly developing and expand-
ing. Their future clinical potential is also further demonstrated.
The evidence to date shows that spinal fMRI has developed rap-
idly over the past several years, and highly sensitive results have
been obtained. Details of responses to sensory and motor para-
digms, pain processing, and descending modulation related to
cognitive and emotional factors have all been demonstrated.
Detailed anatomical mapping of responses has also been demon-
strated. There is no question that spinal fMRI is feasible and effec-
tive for research applications. It has also been shown to provide
valuable information about multiple-sclerosis and traumatic spinal
cord injury. However, it still has many limitations and will require
more development before clinical applications can be considered.
While the technical challenges and limitations of spatial and tem-
poral resolution have been identified, and efforts to overcome
these challenges are proceeding, one key limitation for the prog-
ress of spinal fMRI is the lack of “critical mass” of researchers
working on it. Divisions over the best methods to use have further
limited the pace of development. Consensus over the acquisition
methods would contribute to the development of common soft-
ware methods for analysis. This would reduce the burden of time
and effort for new groups to apply spinal fMRI to new research
questions, and would allow new developments to be shared more
easily between groups. Fortunately, efforts are underway to facili-
tate sharing of ideas such as “Spinal Cord Hack 2014” organized
by Dr. Paul Summers as a satellite meeting of the International
Society for Magnetic Resonance in Medicine (ISMRM) 2014
annual meeting, and the 2015 version being organized by Dr
Julien Cohen-Adad. In addition, the International Spinal Research
Trust and the Wings for Life Foundation have organized interna-
tional imaging workshops [80, 81], and are promoting a multisite
diagnostic trial using spinal cord imaging.
Functional MRI of the Spinal Cord 889
References
1. Figley CR, Stroman PW (2007) Investigation 13. Figley CR, Stroman PW (2012)
of human cervical and upper thoracic spinal Measurement and characterization of the
cord motion: implications for imaging spinal human spinal cord SEEP response using
cord structure and function. Magn Reson Med event-related spinal fMRI. Magn Reson
58(1):185–189 Imaging 30(4):471–484
2. Figley CR, Yau D, Stroman PW (2008) 14. Stroman PW, Krause V, Malisza KL,
Attenuation of lower-thoracic, lumbar, and Frankenstein UN, Tomanek B (2002)
sacral spinal cord motion: implications for imag- Extravascular proton-density changes as a
ing human spinal cord structure and function. non- BOLD component of contrast in fMRI
AJNR Am J Neuroradiol 29(8):1450–1454 of the human spinal cord. Magn Reson Med
3. Yoshizawa T, Nose T, Moore GJ, Sillerud LO 48(1):122–127
(1996) Functional magnetic resonance imag- 15. Stroman PW, Lee AS, Pitchers KK, Andrew
ing of motor activation in the human cervical RD (2008) Magnetic resonance imaging of
spinal cord. Neuroimage 4(3 Pt 1):174–182 neuronal and glial swelling as an indicator of
4. Menon RS, Ogawa S, Kim SG, Ellermann function in cerebral tissue slices. Magn Reson
JM, Merkle H, Tank DW, Ugurbil K (1992) Med 59(4):700–706
Functional brain mapping using magnetic res- 16. Stroman PW, Malisza KL, Onu M (2003)
onance imaging. Signal changes accompany- Functional magnetic resonance imaging at 0.2
ing visual stimulation. Invest Radiol 27(Suppl Tesla. Neuroimage 20(2):1210–1214
2):S47–S53 17. Bosma RL, Stroman PW (2014) Assessment of
5. Ogawa S, Tank DW, Menon R, Ellermann JM, data acquisition parameters, and analysis tech-
Kim SG, Merkle H, Ugurbil K (1992) Intrinsic niques for noise reduction in spinal cord fMRI
signal changes accompanying sensory stimula- data. Magn Reson Imaging 32(5):473–481
tion: functional brain mapping with magnetic 18. Komisaruk BR, Mosier KM, Liu WC,
resonance imaging. Proc Natl Acad Sci U S A Criminale C, Zaborszky L, Whipple B, Kalnin
89(13):5951–5955 A (2002) Functional localization of brainstem
6. Stroman PW, Nance PW, Ryner LN (1999) and cervical spinal cord nuclei in humans with
BOLD MRI of the human cervical spinal cord fMRI. AJNR Am J Neuroradiol 23(4):609–617
at 3 tesla. Magn Reson Med 42(3):571–576 19. Cohen-Adad J, Gauthier CJ, Brooks JC,
7. Madi S, Flanders AE, Vinitski S, Herbison GJ, Slessarev M, Han J, Fisher JA, Rossignol S,
Nissanov J (2001) Functional MR imaging Hoge RD (2010) BOLD signal responses to
of the human cervical spinal cord. AJNR Am controlled hypercapnia in human spinal cord.
J Neuroradiol 22(9):1768–1774 Neuroimage 50(3):1074–1084
8. Backes WH, Mess WH, Wilmink JT (2001) 20. Maieron M, Iannetti GD, Bodurka J, Tracey
Functional MR imaging of the cervical spi- I, Bandettini PA, Porro CA (2007) Functional
nal cord by use of median nerve stimulation responses in the human spinal cord dur-
and fist clenching. AJNR Am J Neuroradiol ing willed motor actions: evidence for side-
22(10):1854–1859 and rate-dependent activity. J Neurosci
9. Stroman PW, Ryner LN (2001) Functional 27(15):4182–4190
MRI of motor and sensory activation in the 21. Summers PE, Ferraro D, Duzzi D, Lui F,
human spinal cord. Magn Reson Imaging Iannetti GD, Porro CA (2010) A quantitative
19(1):27–32 comparison of BOLD fMRI responses to nox-
10. Stroman PW, Krause V, Malisza KL, ious and innocuous stimuli in the human spinal
Frankenstein UN, Tomanek B (2001) cord. Neuroimage 50(4):1408–1415
Characterization of contrast changes in func- 22. Nash P, Wiley K, Brown J, Shinaman R, Ludlow
tional MRI of the human spinal cord at 1.5 D, Sawyer AM, Glover G, Mackey S (2013)
T. Magn Reson Imaging 19(6):833–838 Functional magnetic resonance imaging identi-
11. Bandettini PA, Wong EC, Jesmanowicz A, fies somatotopic organization of nociception in
Hinks RS, Hyde JS (1994) Spin-echo and the human spinal cord. Pain 154(6):776–781
gradient-echo EPI of human brain activation 23. Eippert F, Finsterbusch J, Bingel U, Buchel
using BOLD contrast: a comparative study at C (2009) Direct evidence for spinal cord
1.5 T. NMR Biomed 7(1–2):12–20 involvement in placebo analgesia. Science
12. Figley CR, Leitch JK, Stroman PW (2010) 326(5951):404
In contrast to BOLD: signal enhancement by 24. Sprenger C, Eippert F, Finsterbusch J, Bingel
extravascular water protons as an alternative U, Rose M, Buchel C (2012) Attention modu-
mechanism of endogenous fMRI signal change. lates spinal cord responses to pain. Curr Biol
Magn Reson Imaging 28(8):1234–1243 22(11):1019–1022
890 Patrick Stroman and Massimo Filippi
25. Geuter S, Buchel C (2013) Facilitation of pain innocuous thermal sensory stimuli and study-
in the human spinal cord by nocebo treatment. related emotional influences. Magn Reson
J Neurosci 33(34):13784–13790 Imaging 27(10):1333–1346
26. van de Sand MF, Sprenger C, Buchel C (2015) 38. Lawrence JM, Stroman PW, Kollias SS (2008)
BOLD responses to itch in the human spinal Functional magnetic resonance imaging of the
cord. Neuroimage 108:138–143 human spinal cord during vibration stimulation
27. Figley CR, Stroman PW (2011) The role(s) of of different dermatomes. Neuroradiology
astrocytes and astrocyte activity in neurome- 50(3):273–280
tabolism, neurovascular coupling, and the pro- 39. Stroman PW, Bosma RL, Tsyben A (2012)
duction of functional neuroimaging signals. Somatotopic arrangement of thermal sensory
Eur J Neurosci 33(4):577–588 regions in the healthy human spinal cord deter-
28. Stroman PW, Krause V, Frankenstein UN, mined by means of spinal cord functional
Malisza KL, Tomanek B (2001) Spin-echo ver- MRI. Magn Reson Med 68(3):923–931
sus gradient-echo fMRI with short echo times. 40. Bosma RL, Stroman PW (2014) Spinal cord
Magn Reson Imaging 19(6):827–831 response to stepwise and block presentation of
29. Stroman PW, Tomanek B, Krause V, thermal stimuli: a functional MRI study.
Frankenstein UN, Malisza KL (2003) J Magn Reson Imaging 41(5):1318–1325
Functional magnetic resonance imaging of the 41. Stroman PW, Coe BC, Munoz DP (2011)
human brain based on signal enhancement by Influence of attention focus on neural activity
extravascular protons (SEEP fMRI). Magn in the human spinal cord during thermal sen-
Reson Med 49(3):433–439 sory stimulation. Magn Reson Imaging
30. Stroman PW, Kornelsen J, Lawrence J, Malisza 29(1):9–18
KL (2005) Functional magnetic resonance 42. Dobek CE, Beynon ME, Bosma RL, Stroman
imaging based on SEEP contrast: response PW (2014) Music modulation of pain percep-
function and anatomical specificity. Magn tion and pain-related activity in the brain,
Reson Imaging 23(8):843–850 brainstem, and spinal cord: an fMRI study.
31. Cahill CM, Stroman PW (2011) Mapping of J Pain 15(10):1057–1068
neural activity produced by thermal pain in the 43. Agosta F, Valsasina P, Absinta M, Sala S,
healthy human spinal cord and brain stem: a Caputo D, Filippi M (2009) Primary progres-
functional magnetic resonance imaging study. sive multiple sclerosis: tactile-associated func-
Magn Reson Imaging 29(3):342–352 tional MR activity in the cervical spinal cord.
32. Ghazni NF, Cahill CM, Stroman PW (2010) Radiology 253(1):209–215
Tactile sensory and pain networks in the human 44. Agosta F, Valsasina P, Caputo D, Rocca MA,
spinal cord and brain stem mapped by means of Filippi M (2009) Tactile-associated fMRI
functional MR imaging. AJNR Am recruitment of the cervical cord in healthy sub-
J Neuroradiol 31(4):661–667 jects. Hum Brain Mapp 30(1):340–345
33. Kozyrev N, Figley CR, Alexander MS, Richards 45. Agosta F, Valsasina P, Caputo D, Stroman PW,
JS, Bosma RL, Stroman PW (2012) Neural Filippi M (2008) Tactile-associated recruit-
correlates of sexual arousal in the spinal cords ment of the cervical cord is altered in patients
of able-bodied men: a spinal fMRI investiga- with multiple sclerosis. Neuroimage
tion. J Sex Marital Ther 38(5):418–435 39(4):1542–1548
34. Lawrence JM, Kornelsen J, Stroman PW 46. Agosta F, Valsasina P, Rocca MA, Caputo D,
(2011) Noninvasive observation of cervical spi- Sala S, Judica E, Stroman PW, Filippi M (2008)
nal cord activity in children by functional MRI Evidence for enhanced functional activity of
during cold thermal stimulation. Magn Reson cervical cord in relapsing multiple sclerosis.
Imaging 29(6):813–818 Magn Reson Med 59(5):1035–1042
35. Rempe T, Wolff S, Riedel C, Baron R, Stroman 47. Valsasina P, Agosta F, Absinta M, Sala S,
PW, Jansen O, Gierthmuhlen J (2014) Spinal Caputo D, Filippi M (2010) Cervical cord
fMRI reveals decreased descending inhibition functional MRI changes in relapse-onset MS
during secondary mechanical hyperalgesia. patients. J Neurol Neurosurg Psychiatry
PLoS One 9(11):e112325 81(4):405–408
36. Rempe T, Wolff S, Riedel C, Baron R, Stroman 48. Valsasina P, Agosta F, Caputo D, Stroman
PW, Jansen O, Gierthmuhlen J (2014) Spinal PW, Filippi M (2008) Spinal fMRI during
and supraspinal processing of thermal stimuli: proprioceptive and tactile tasks in healthy sub-
an fMRI study. J Magn Reson Imaging jects: activity detected using cross-correlation,
41(4):1046–1055 general linear model and independent compo-
37. Stroman PW (2009) Spinal fMRI investigation nent analysis. Neuroradiology
of human spinal cord function over a range of 50(10):895–902
Functional MRI of the Spinal Cord 891
49. Valsasina P, Rocca MA, Absinta M, Agosta F, 62. Lang J, Bartram CT (1982) Fila radicularia of
Caputo D, Comi G, Filippi M (2012) Cervical the ventral and dorsal radices of the human spi-
cord FMRI abnormalities differ between the nal cord. Gegenbaurs Morphol Jahrb
progressive forms of multiple sclerosis. Hum 128(4):417–462
Brain Mapp 33(9):2072–2080 63. Gray H (1995) Gray’s anatomy: the anatomical
50. Cadotte DW, Bosma R, Mikulis D, Nugaeva basis of medicine and surgery. In: Williams PL,
N, Smith K, Pokrupa R, Islam O, Stroman PW, Bannister LH, Berry MM, Collins P, Dyson M,
Fehlings MG (2012) Plasticity of the injured Dussek JE, Ferguson MWJ (eds) Gray’s anatomy:
human spinal cord: insights revealed by spinal the anatomical basis of medicine and surgery.
cord functional MRI. PLoS One 7(9):e45560 Churchill-Livingstone, London, pp 975–1011
51. Stroman PW, Kornelsen J, Bergman A, Krause 64. Talairach J, Tournoux P (1988) Co-planar ste-
V, Ethans K, Malisza KL, Tomanek B (2004) rotaxic atlas of the human brain. Thieme
Noninvasive assessment of the injured human Medical Publishers Inc, New York
spinal cord by means of functional magnetic 65. Naidich TP, Duvernoy HM, Delman BN,
resonance imaging. Spinal Cord 42(2):59–66 Sorensen AG, Kollias SS, Haacke EM (2009)
52. Kornelsen J, Stroman PW (2007) Detection of Internal architecture of the brain stem with key
the neuronal activity occurring caudal to the axial sections. Duvernoy’s atlas of the human
site of spinal cord injury that is elicited during brain stem and cerebellum. Springer,
lower limb movement tasks. Spinal Cord New York, pp 79–82
45(7):485–490 66. McArdle JJ, McDonald RP (1984) Some alge-
53. Stroman PW (2011) Essentials of functional braic properties of the Reticular Action Model
MRI. Taylor & Francis Group, LLC, Boca for moment structures. Br J Math Stat Psychol
Raton, FL 37(Pt 2):234–251
54. Murphy K, Bodurka J, Bandettini PA (2007) 67. Craggs JG, Staud R, Robinson ME, Perlstein
How long to scan? The relationship between WM, Price DD (2012) Effective connectivity
fMRI temporal signal to noise ratio and neces- among brain regions associated with slow tem-
sary scan duration. Neuroimage 34(2):565–574 poral summation of C-fiber-evoked pain in
55. Finsterbusch J, Eippert F, Buchel C (2012) fibromyalgia patients and healthy controls.
Single, slice-specific z-shim gradient pulses J Pain 13(4):390–400
improve T2*-weighted imaging of the spinal 68. Buchel C, Friston K (2001) Interactions
cord. Neuroimage 59(3):2307–2315 among neuronal systems assessed with func-
56. Figley CR, Stroman PW (2009) Development tional neuroimaging. Rev Neurol 157(8–9 Pt
and validation of retrospective spinal cord 1):807–815
motion time-course estimates (RESPITE) for 69. Buchel C, Friston KJ (1997) Modulation of
spin-echo spinal fMRI: improved sensitivity connectivity in visual pathways by attention:
and specificity by means of a motion- cortical interactions evaluated with structural
compensating general linear model analysis. equation modelling and fMRI. Cereb Cortex
Neuroimage 44(2):421–427 7(8):768–778
57. Brooks JC, Beckmann CF, Miller KL, Wise 70. Bollen KA (1989) Structural equations with
RG, Porro CA, Tracey I, Jenkinson M (2008) latent variables. Wiley, New York
Physiological noise modelling for spinal func- 71. Millan MJ (2002) Descending control of pain.
tional magnetic resonance imaging studies. Prog Neurobiol 66(6):355–474
Neuroimage 39(2):680–692 72. Barry RL, Smith SA, Dula AN, Gore JC (2014)
58. Verma T, Cohen-Adad J (2014) Effect of respi- Resting state functional connectivity in the
ration on the B0 field in the human spinal cord human spinal cord. eLife 3:e02812
at 3T. Magn Reson Med 72(6):1629–1636 73. Kashkouli Nejad K, Sugiura M, Thyreau B,
59. Myronenko A, Song XB (2009) Image regis- Nozawa T, Kotozaki Y, Furusawa Y, Nishino
tration by minimization of residual complexity. K, Nukiwa T, Kawashima R (2014) Spinal
Proc Cvpr IEEE. pp 49–56 fMRI of interoceptive attention/awareness in
60. Myronenko A, Song XB (2010) Intensity- experts and novices. Neural Plast 2014:679509
based image registration by minimizing resid- 74. Smith SD, Kornelsen J (2011) Emotion-
ual complexity. IEEE Trans Med Imag dependent responses in spinal cord neurons: a
29(11):1882–1891 spinal fMRI study. Neuroimage 58(1):269–274
61. Stroman PW, Figley CR, Cahill CM (2008) 75. Kornelsen J, Smith SD, McIver TA (2014) A
Spatial normalization, bulk motion correction neural correlate of visceral emotional responses:
and coregistration for functional magnetic res- evidence from fMRI of the thoracic spinal cord.
onance imaging of the human cervical spinal Soc Cogn Affect Neurosci 10(4):584–588
cord and brainstem. Magn Reson Imaging 76. Khan HS, Bosma RL, Beynon M, Dobek C,
26(6):809–814 McIver T, Stroman PW (2013) Pain processing
892 Patrick Stroman and Massimo Filippi
networks in the brain and spinal cord mapped 80. Stroman PW, Wheeler-Kingshott C, Bacon M,
using Functional Magnetic Resonance Schwab JM, Bosma R, Brooks J, Cadotte D,
Imaging. Program number II6 66.01. 2013 Carlstedt T, Ciccarelli O, Cohen-Adad J, Curt
Meeing Planner San Diego, CA; Society for A, Evangelou N, Fehlings MG, Filippi M, Kelley
Neuroscience BJ, Kollias S, Mackay A, Porro CA, Smith S,
77. McIver TA, Kornelsen J, Smith SD (2013) Strittmatter SM, Summers P, Tracey I (2014)
Limb-specific emotional modulation of cervical The current state-of-the-art of spinal cord imag-
spinal cord neurons. Cogn Affect Behav ing: methods. Neuroimage 84:1070–1081
Neurosci 13(3):464–472 81. Wheeler-Kingshott CA, Stroman PW, Schwab
78. Cox RW (1996) AFNI: software for analysis JM, Bacon M, Bosma R, Brooks J, Cadotte
and visualization of functional magnetic reso- DW, Carlstedt T, Ciccarelli O, Cohen-Adad J,
nance neuroimages. Comput Biomed Res Curt A, Evangelou N, Fehlings MG, Filippi M,
29(3):162–173 Kelley BJ, Kollias S, Mackay A, Porro CA,
79. Friston KJ, Josephs O, Zarahn E, Holmes AP, Smith S, Strittmatter SM, Summers P,
Rouquette S, Poline J (2000) To smooth or not Thompson AJ, Tracey I (2014) The current
to smooth? Bias and efficiency in fMRI time- state-of-the-art of spinal cord imaging: applica-
series analysis. Neuroimage 12(2):196–208 tions. Neuroimage 84:1082–1093
Chapter 30
Abstract
Network-based analysis of brain functional connections has provided a novel instrument to study the
human brain in healthy and diseased individuals. Graph theory provides a powerful tool to describe quan-
titatively the topological organization of brain connectivity. Using such a framework, the brain can be
depicted as a set of nodes connected by edges. Distinct modifications of brain network topology have been
identified during development and normal aging, whereas disrupted functional connectivity has been asso-
ciated to several neurological and psychiatric conditions, including multiple sclerosis, dementia, amyo-
trophic lateral sclerosis, and schizophrenia. Such an assessment has contributed to explain part of the
clinical manifestations usually observed in these patients, including disability and cognitive impairment.
Future network-based research might reveal different stages of the different diseases, subtypes for cogni-
tive impairments, and connectivity profiles associated with different clinical outcomes.
Key words Brain networks, Structural connectivity, Functional connectivity, Graph theory, Multiple
sclerosis, Dementias, Psychiatric conditions
1 Introduction
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods vol. 119,
DOI 10.1007/978-1-4939-5611-1_30, © Springer Science+Business Media New York 2016
893
894 Massimo Filippi and Maria A. Rocca
2 Normal Development
3 Normal Aging
Fig. 1 Developmental changes in interregional functional connectivity. (a) Children had significantly greater
subcortical-primary sensory, subcortical-association, subcortical-paralimbic, and lower paralimbic-
association, paralimbic-limbic, association-limbic connectivity than young-adults (p < 0.01, indicated by **).
(b) Graphical representation of developmental changes in functional connectivity along the posterior-anterior
and ventral-dorsal axes, highlighting higher subcortical connectivity (subcortical nodes are shown in green)
and lower paralimbic connectivity (paralimbic nodes are shown in gold) in children, compared to young-adults.
Brain regions are plotted using the y and z coordinates of their centroids (in mm) in the MNI space. 430 pairs
of anatomical regions showed significantly higher correlations in children and 321 pairs showed significantly
higher correlations in young-adults (p < 0.005, FDR corrected). For illustration purposes, the plot shows dif-
ferential connectivity that were most significant, 105 pairs higher in children (indicated in red) and 53 higher
in young-adults (indicated in blue), (p < 0.0001, FDR corrected). From [6] with permission
Clinical Applications of the Functional Connectome 897
4 Multiple Sclerosis
Fig. 3 Functional hub distribution in healthy controls and multiple sclerosis patients. Brain hubs (a left hemi-
sphere, b right hemisphere) of the functional networks of healthy controls (HCs) and patients with MS as a
whole and according to the presence/absence of cognitive impairment. Hubs were identified as brain regions
having either integrated nodal degree or betweenness centrality one standard deviation greater than the net-
work average. Hubs present in HC only are reported in red, hubs present in MS patients only are reported in
blue, and hubs present in all groups are reported in black. CP cognitively preserved, CI cognitively impaired,
ACC anterior cingulate cortex, Caud caudate nucleus, Cereb cerebellum, ITG inferior temporal gyrus, Ling lin-
gual gyrus, MCC middle cingulate cortex, MTG middle temporal gyrus, OFC orbitofrontal cortex, Pall pallidus,
Precun precuneus, Put putamen, SFG superior frontal gyrus, Sup TP superior temporal pole, Thal thalamus.
From [18] with permission
5 Neurodegnerative Diseases
5.1 Alzheimer’s Many studies used graph theoretical analysis in patients with the
Disease and Other most prevalent type of dementia, Alzheimer’s disease (AD). A cor-
Dementias relation between the site of amyloid-β deposition in AD patients
and the location of major hubs as defined by graph theoretical
analysis of RS FC in healthy adults has been demonstrated [21].
These regions include the posterior cingulate cortex/precuneus,
the inferior parietal lobule, and the medial frontal cortex, implying
that the hubs are preferentially affected in the progression of
AD. There is also convergent evidence from methodologically
disparate MRI studies that AD is associated with perturbations of
brain small-world network organization [22–24].
Although studies showed considerable variability in reported
group differences of most graph properties, the average
900 Massimo Filippi and Maria A. Rocca
5.2 Amyotrophic Consistent motor and extra-motor brain pathology supports the
Lateral Sclerosis notion of amyotrophic lateral sclerosis (ALS) as a system failure.
Thus, a simplistic (motor-based) approach to ALS is no longer ten-
able. Overall functional organization of the motor network was
unchanged in patients with ALS compared to healthy controls;
however, patients with a stronger and more interconnected motor
network had a more progressive disease course [28].
6 Psychiatric Conditions
Fig. 4 Resting-state synchronization in healthy controls and Alzheimer’s disease patients. (a) Matrix of signifi-
cant differences of synchronization between Alzheimer’s disease (AD) and controls (2-tail t-test, p < 0.05
uncorrected). The white and black dots represent brain areas pairs with increased and decreased synchroniza-
tion in AD respectively. (b-d) A subset of connectional differences corresponding to the matrix (a) are plotted
at three superior-to-inferior levels through the anatomical automatic labeled brain template. Lines depict
synchronization between pairs of regions: solid lines = enhanced synchronization; dashed lines = reduced
synchronization. Note the pattern of generalized posterior (parietal and occipital) synchronization reductions
and increased frontal synchronization. From [23] with permission
7 Conclusions
The extensive application during the past few years of graph theoreti-
cal approaches to define brain network topology in healthy and dis-
eased people has undoubtedly provided a novel instrument to
characterize functional abnormalities associated with different neuro-
logical and psychiatric conditions and to test hypothesized discon-
nectivity effects in these diseases. Disrupted functional brain
connectivity is present in the major neurological and psychiatric con-
ditions discussed in this chapter and their assessment has contributed
to explain part of the clinical manifestations of these patients.
However, there are also inconsistencies between existing studies,
which might be attributable to the clinical heterogeneity of the
patient groups as well as to differences in imaging modality and ana-
lytic methods. Future network-based research might reveal different
stages of the different diseases, subtypes for cognitive impairments,
and connectivity profiles associated with different clinical outcomes.
References
1. Bullmore E, Sporns O (2009) Complex brain “local to distributed” organization. PLoS
networks: graph theoretical analysis of struc- Comput Biol 5:e1000381
tural and functional systems. Nat Rev Neurosci 6. Supekar K, Musen M, Menon V (2009)
10:186–198 Development of large-scale functional brain
2. Filippi M, van den Heuvel MP, Fornito A et al networks in children. PLoS Biol 7:e1000157
(2013) Assessing brain system dysfunction 7. Hwang K, Hallquist MN, Luna B (2013) The
through MRI-based connectomics. Lancet development of hub architecture in the human
Neurol 12:1189–1199 functional brain network. Cereb Cortex
3. Fair DA, Dosenbach NU, Church JA et al 23:2380–2393
(2007) Development of distinct control net- 8. Dosenbach NU, Nardos B, Cohen AL et al
works through segregation and integration. (2010) Prediction of individual brain maturity
Proc Natl Acad Sci U S A 104:13507–13512 using fMRI. Science 329:1358–1361
4. Fair DA, Cohen AL, Dosenbach NU et al 9. Wu K, Taki Y, Sato K et al (2013) Topological
(2008) The maturing architecture of the organization of functional brain networks in
brain’s default network. Proc Natl Acad Sci U healthy children: differences in relation to age,
S A 105:4028–4032 sex, and intelligence. PLoS One 8:e55347
5. Fair DA, Cohen AL, Power JD et al (2009) 10. van den Heuvel MP, van Soelen IL, Stam CJ,
Functional brain networks develop from a Kahn RS, Boomsma DI, Hulshoff Pol HE
Clinical Applications of the Functional Connectome 903
(2013) Genetic control of functional brain net- analysis of FMRI resting-state functional con-
work efficiency in children. Eur nectivity. PLoS One 5:e13788
Neuropsychopharmacol 23:19–23 24. Supekar K, Menon V, Rubin D, Musen M,
11. Fransson P, Aden U, Blennow M, Lagercrantz Greicius MD (2008) Network analysis of intrin-
H (2011) The functional architecture of the sic functional brain connectivity in Alzheimer’s
infant brain as revealed by resting-state disease. PLoS Comput Biol 4:e1000100
fMRI. Cereb Cortex 21:145–154 25. Tijms BM, Wink AM, de Haan W et al (2013)
12. Gao W, Gilmore JH, Giovanello KS et al Alzheimer’s disease: connecting findings from
(2011) Temporal and spatial evolution of brain graph theoretical studies of brain networks.
network topology during the first two years of Neurobiol Aging 34:2023–2036
life. PLoS One 6:e25278 26. Zhao X, Liu Y, Wang X et al (2012) Disrupted
13. Achard S, Bullmore E (2007) Efficiency and small-world brain networks in moderate
cost of economical brain functional networks. Alzheimer’s disease: a resting-state FMRI
PLoS Comput Biol 3:e17 study. PLoS One 7:e33540
14. Tomasi D, Volkow ND (2012) Aging and 27. Agosta F, Sala S, Valsasina P et al (2013) Brain
functional brain networks. Mol Psychiatry network connectivity assessed using graph the-
17(471):549–458 ory in frontotemporal dementia. Neurology
15. Meier TB, Desphande AS, Vergun S et al (2012) 9:134–143
Support vector machine classification and charac- 28. Verstraete E, van den Heuvel MP, Veldink JH
terization of age-related reorganization of func- et al (2010) Motor network degeneration in amy-
tional brain networks. Neuroimage 60:601–613 otrophic lateral sclerosis: a structural and func-
16. Rocca M, Valsasina P, Martinelli V et al (2012) tional connectivity study. PLoS One 5:e13664
Large-scle neuronal network dysfunction in 29. Bullmore E, Sporns O (2012) The economy of
relapsing-remitting multiple sclerosis. brain network organization. Nat Rev Neurosci
Neurology 79:1449–1457 13:336–349
17. Schoonheim MM, Hulst HE, Landi D et al 30. van den Heuvel MP, Mandl RC, Stam CJ,
(2012) Gender-related differences in func- Kahn RS, Hulshoff Pol HE (2010) Aberrant
tional connectivity in multiple sclerosis. Mult frontal and temporal complex network struc-
Scler 18:164–173 ture in schizophrenia: a graph theoretical anal-
18. Rocca MA, Valsasina P, Meani A, Falini A, ysis. J Neurosci 30:15915–15926
Comi G, Filippi M (2016) Impaired functional 31. Lynall ME, Bassett DS, Kerwin R et al (2010)
integration in multiple sclerosis: a graph theory Functional connectivity and brain networks in
study. Brain Struct Funct 221:115–131 schizophrenia. J Neurosci 30:9477–9487
19. Gamboa OL, Tagliazucchi E, von Wegner F 32. Bassett DS, Nelson BG, Mueller BA,
et al (2014) Working memory performance of Camchong J, Lim KO (2012) Altered resting
early MS patients correlates inversely with mod- state complexity in schizophrenia. Neuroimage
ularity increases in resting state functional con- 59:2196–2207
nectivity networks. Neuroimage 94:385–395 33. Lo CY, Su TW, Huang CC et al (2015)
20. Richiardi J, Gschwind M, Simioni S et al Randomization and resilience of brain func-
(2012) Classifying minimally disabled multiple tional networks as systems-level endopheno-
sclerosis patients from resting state functional types of schizophrenia. Proc Natl Acad Sci U S
connectivity. Neuroimage 62:2021–2033 A 112:9123–9128
21. Buckner RL, Sepulcre J, Talukdar T et al 34. Alexander-Bloch AF, Gogtay N, Meunier D et al
(2009) Cortical hubs revealed by intrinsic (2010) Disrupted modularity and local connec-
functional connectivity: mapping, assessment tivity of brain functional networks in childhood-
of stability, and relation to Alzheimer’s disease. onset schizophrenia. Front Syst Neurosci 4:147
J Neurosci 29:1860–1873 35. Yu Q, Erhardt EB, Sui J et al (2015) Assessing
22. He Y, Chen Z, Evans A (2008) Structural dynamic brain graphs of time-varying connec-
insights into aberrant topological patterns of tivity in fMRI data: application to healthy con-
large-scale cortical networks in Alzheimer’s trols and patients with schizophrenia.
disease. J Neurosci 28:4756–4766 Neuroimage 107:345–355
23. Sanz-Arigita EJ, Schoonheim MM, 36. Rubinov M, Bullmore E (2013) Schizophrenia
Damoiseaux JS et al (2010) Loss of ‘small- and abnormal brain network hubs. Dialogues
world’ networks in Alzheimer’s disease: graph Clin Neurosci 15:339–349
INDEX
A B
Acceleration factor............................50, 51, 84–89, 465, 476, Bayesian inference ............................................ 242, 244–247
477, 479, 877, 881, 882 Behaviourally relevant coding ...................................601–602
Activation maps ................................139, 143, 326, 430, 534, Between-group comparison ...................................... 142, 144
651, 748, 805, 808, 856 Biomarkers .......................................345–348, 666, 699–702,
Active noise cancellation (ANC) ........................ 55, 575, 579 711, 717, 721–722, 726–727, 823, 827, 837
Active tasks............................................... 138, 140, 524, 645 Biophysical models ....................209, 211, 241, 242, 250–252
Adjacency matrix ..............................................................297 Bipolar disorder (BD)...............................659, 660, 671, 672,
Alzheimer’s disease (AD) .......................... 121, 266, 271, 437, 674–677, 687, 688, 862
700–705, 707, 709–711, 713–718, 721–727, 818, 819, Block design ............................................................... 70, 758
823, 824, 862, 864, 899–901 Blood oxygen level dependent (BOLD)
Amnesic syndrome ................................................... 742, 753 effect ............................................................. 22, 114, 146
Amyotrophic lateral sclerosis (ALS)........................ 700, 709, initial dip .................................25, 26, 128, 130, 326, 642
712, 716, 717, 719, 720, 900 response ....................................25, 26, 63, 130, 146–147,
ANALYZE .............................................................. 156, 157 323, 460, 469, 470, 483
Anatomical connectivity ........................... 243, 293, 545, 546 Brain activation ........................ 25, 74, 76, 98, 121, 125, 126,
Anterior temporal lobe .................................... 363, 365, 437, 129, 133, 138–140, 142, 331, 332, 342, 347, 368, 375,
687, 741, 858 394, 440, 457–459, 482, 483, 486, 501, 503, 511, 512,
resection (ATLR) ...............................741, 743, 744, 750, 524, 547, 574, 626, 647, 651, 667, 676, 703–705, 709
751, 753, 754, 756, 757, 761 Brain atlases
Anxiety disorders ...............................658, 659, 661, 677–686 deformation atlases .............................................268–270
Arterial spin labeling (ASL) .............................41, 44, 64, 88, density-based ..............................................................268
122, 334, 335, 347, 515, 657–659, 724, 782, 820, 828 disease-specific ............................266, 268, 270–271, 785
Ascending sensory pathways .................................... 389, 390 genetic atlases .............................................................271
Attention label-based ..........................................................268–269
nonselective ................................................................389 Montreal Neurological Institute (MNI) ............ 157, 273,
preparatory .........................................................400–404 289, 463, 719
reflexive.......................................................................388 Talairach & Tournoux (T&T) ...................................... 57
selective...............................................................389–397 Brain-behavior relationships............................ 317, 318, 327,
spatial..................................................343, 388, 392–400, 335, 340, 348, 728
403, 406–411, 548 Brain imaging data
spatial selective ................................... 392, 393, 400, 402 functional and structural MRI .................... 861–862, 864
visual ...................................................336, 387–413, 595, joint ICA (jICA) ................................................ 856, 858
704, 707, 710, 716 multimodal CCA ................................................859–861
voluntary ............................................. 388, 407, 411–413 multimodal MRI data.........................................854–857
voluntary visual ................................... 388, 407, 411–413 multivariate approaches ......................................856–857
Attentional control mechanisms physiological or behavioral features ....................853–854
EEG-fMRI ........................................................407–412 Partial Least Squares (PLS)........................................860
top-down control ........................................................400 Brain networks
Attention control network ........................................ 397, 403 graph projections ........................................................298
Auditory system........................................ 360, 390, 573–602 topological properties .........................................300–308
Automated anatomical labelling (AAL) ...........................290 Brain plasticity.................................... vii, 609, 612, 616, 626,
Automated normalization process ....................................884 836, 838, 839, 842, 843
Average artefact subtraction (AAS).................. 765, 768, 770 Brain systems ............................. 253, 278, 344, 355, 366, 396
Massimo Filippi (ed.), fMRI Techniques and Protocols, Neuromethods, vol. 119,
DOI 10.1007/978-1-4939-5611-1, © Springer Science+Business Media New York 2016
905
FMRI TECHNIQUES AND PROTOCOLS
906 Index
P Q
Quantitative relaxometry ..............................................70–74
Paced Auditory Serial Addition Test (PASAT) ........ 616, 617
Quenching ..........................................................................39
Paced Visual Serial Addition Test (PVSAT) ....................616
Pain
R
assessment ..........................................................498–500
chronic ..........496–498, 506, 508, 510–512, 514, 515, 826 Radiofrequency (RF)
empathy ...................................................... 512–513, 824 coils................................... 47, 48, 54, 103, 104, 115–117,
modulation .................. 507–508, 510, 511, 515, 630, 886 126, 130, 131, 133
pathways ............................................................. 510, 631 pulse...................................................................... 19, 820
sharp pricking .............................................................500 Random field theory (RFT)..........221, 224, 227–231, 244–246
slow burning ...............................................................500 Random noise................................................... 167, 212, 872
thresholds ........................................... 497, 498, 501, 509 Real time fMRI (rt-fMRI) ............................... 613, 614, 837
FMRI TECHNIQUES AND PROTOCOLS
910 Index