List of Contributors

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 776

List of contributors

D.G. Amaral, Department of Psychiatry and Behavioral Sciences, The M.I.N.D. Institute and the
California National Primate Research Center, UC Davis, 2825 50th Street, Sacramento, CA 95817, USA
L. Acsády, Institute of Experimental Medicine, Hungarian Academy of Sciences, PO Box 67, 1450
Budapest, Hungary
C.A. Barnes, Arizona Research Laboratories Division of Neural Systems, Memory & Aging, University of
Arizona, Tucson, AZ and Evelyn F. McKnight Brain Institute, University of Arizona, Tucson, AZ and
Departments of Psychology and Neurology, University of Arizona, Tucson, AZ, USA
I. Bechmann, Institute of Clinical Neuroanatomy, J.W. Goethe-University, Theodor-Stern-Kai 7, D-60590
Frankfurt/Main, Germany
D.K. Binder, Department of Neurological Surgery, University of California, Irvine, CA 92868-3298, USA
M. Blaabjerg, Anatomy and Neurobiology, Institute of Medical Biology, University of Southern Denmark,
Winslowparken 21, DK-5000 Odense C, Denmark
C.R. Bramham, Department of Biomedicine and Mental Health Research Center, University of Bergen,
Jonas Lies vei 91, N-5009 Bergen, Norway
G. Buzsáki, Center for Molecular and Behavioral Neuroscience, Rutgers, The State University of
New Jersey, 197 University Avenue, University Heights, Newark, NJ 07102, USA
G.C. Carlson, The Children’s Hospital of Philadelphia, Abramson Pediatric Research Center, Room 410D,
3516 Civic Center Boulevard, Philadelphia, PA 19104-4318, USA
C. Chavkin, Department of Pharmacology, University of Washington, Seattle, WA 98195, USA
M.K. Chawla, Arizona Research Laboratories Division of Neural Systems, Memory & Aging, University
of Arizona, Tucson, AZ and Evelyn F. McKnight Brain Institute, University of Arizona, Tucson, AZ,
USA
B.J. Claiborne, Department of Biology, University of Texas at San Antonio, One UTSA Circle,
San Antonio, TX 78249, USA
W.F. Colmers, Department of Pharmacology, University of Alberta, 9-36 Medical Sciences Building,
Edmonton, AL T6G 2H7, Canada
D.A. Coulter, The Children’s Hospital of Philadelphia, Abramson Pediatric Research Center, Room 410D,
3516 Civic Center Boulevard, Philadelphia, PA 19104-4318, USA
D. Del Turco, Institute of Clinical Neuroanatomy, J.W. Goethe-University, Theodor-Stern-Kai 7, D-60590
Frankfurt/Main, Germany
T. Deller, Institute of Clinical Neuroanatomy, J.W. Goethe-University, Theodor-Stern-Kai 7, D-60590
Frankfurt/Main, Germany
B.E. Derrick, Department of Biology, The Cajal Neuroscience Research Institute, The University of Texas
at San Antonio, 6900 N. Loop 1604 West, San Antonio, TX 78249-0662, USA
L. DeToledo-Morrell, Department of Neurological Sciences, Rush University Medical Center, Chicago,
IL 60612, USA
C.T. Drake, Division of Neurobiology, Department of Neurology and Neuroscience, Weill-Cornell
Medical College, 411 East 69th Street, New York, NY 10021, USA

v
vi

M.R. Drew, Department of Neuroscience and Psychiatry, Division of Integrative Neuroscience, Columbia
University, New York, NY 10032, USA
F.E. Dudek, Department of Physiology, University of Utah, Salt Lake City, UT, USA
E. Förster, Institute of Anatomy and Cell Biology, University of Freiburg, Albertstr. 17, D-79104 Freiburg,
Germany
C.J. Frazier, Department of Pharmacodynamics and Department of Neuroscience, University of Florida,
College of Medicine, JHMHC 100487, 1600 S.W. Arder Road, Gainesville, FL 32610, USA
M. Frotscher, Institute of Anatomy and Cell Biology, University of Freiburg, Albertstr. 17, D-79104
Freiburg, Germany
R. Gutiérrez, Department of Physiology, Biophysics and Neurosciences, Center for Research and
Advanced Studies, Mexico City Apartado Postal 14-740, Mexico D.F. 07000, Mexico
T. Hajszan, Department of Obstetrics, Gynecology, and Reproductive Sciences, Yale University School
of Medicine, 333 Cedar Street, FMB 313, New Haven, CT 06520, USA
T. Hamilton, Department of Pharmacology, University of Alberta, 9-36 Medical Sciences Building,
Edmonton, AL T6G 2H7, Canada
C.W. Harley, Department of Psychology, Memorial University of Newfoundland, St. John’s, NL
A1B 3X9, Canada
R. Hen, Departments of Neuroscience, Psychiatry, and Pharmocology, Division of Integrative
Neuroscience, Columbia University, New York, NY 10032, USA
D.A. Henze, Neuroscience Drug Discovery Merck Research Laboratories, 770 Sunney town Pike, P.O.B. 4
WP 26A-2000, West Point, PA 19486, USA
C.R. Houser, Department of Neurobiology, David Geffen School of Medicine at UCLA, 73-235 CHS,
10833 Le-Conte Avenue, Los Angeles, CA 90095, USA
D. Hsu, Department of Neurology, University of Wisconsin, 600 Highland Avenue, H6/526, Madison,
WI 53792, USA
D.B. Jaffe, Department of Biology, University of Texas at San Antonio, One UTSA Circle, San Antonio,
TX 78249, USA
M. Joëls, Swammerdam Institute of Life Sciences, Center for NeuroScience (SILS-CNS), University of
Amsterdam, Kruislaan 320, 1098 SM Amsterdam, The Netherlands
S. Káli, Institute of Experimental Medicine, Hungarian Academy of Sciences, PO Box 67, 1450 Budapest,
Hungary
R.P. Kesner, Department of Psychology, University of Utah, 380 S. 1530 E., Room 502, Salt Lake City,
UT 84121, USA
P. Lavenex, Department of Medicine, Unit of Physiology, University of Fribourg, 1700 Fribourg,
Switzerland
C. Leranth, Department of Obstetrics, Gynecology, and Reproductive Sciences, Yale University, School
of Medicine, 333 Cedar Street, FMB 312, New Haven, CT 06520, USA
G. Li, Department of Neurology, Programs in Neuroscience, Developmental Biology, University of
California, San Francisco, 1550 4th Street, Rock Hall Room 448C, San Francisco, CA 94158, USA
J.E. Lisman, Department of Biology and Volen Center for Complex Systems, Brandeis University,
Waltham, MA 02454, USA
T.A. Milner, Department of Neurology and Neuroscience, Division of Neurobiology, Weill-Cornell
Medical College, 411 East 69th Street, New York, NY 10021, USA
R.J. Morgan, Department of Anatomy and Neurobiology, 193 Irvine Hall, University of California, Irvine,
CA 92697, USA
J.J. O’Connor, UCD School of Biomolecular and Biomedical Science, Conway Institute of Biomolecular
and Biomedical Research, University College Dublin, Belfield, Dublin 4, Ireland
vii

T.G. Ohm, Institute of Integrative Neuroanatomy, Department of Clinical Cell and Neurobiology, Charité
CCM, 10098 Berlin, Germany
J.M. Parent, Department of Neurology, University of Michigan Medical Center, 109 Zina Pitcher Place,
5021 BSRB, Ann Arbor, MI 48109-2200, USA
P.R. Patrylo, Department of Physiology, Southern Illinois University School of Medicine, Carbondale,
IL 62901, USA
M. Pickering, UCD School of Biomolecular and Biomedical Science, Conway Institute of Biomolecular
and Biomedical Research, University College Dublin, Belfield, Dublin 4, Ireland
S.J. Pleasure, Department of Neurology, Programs in Neuroscience, Developmental Biology, University of
California, San Francisco, 1550 4th Street, Rock Hall Room 448C, San Francisco, CA 94158, USA
B. Pöschel, Department of Cell Biology and Anatomy, New York Medical College, Valhalla, NY 10595,
USA
O. Rahimi, Department of Cellular and Structural Biology, University of Texas Health Science Center at
San Antonio, San Antonio, TX 78229, USA
A. Rappert, Institute of Clinical Neuroanatomy, J.W. Goethe-University, Theodor-Stern-Kai 7, D-60590
Frankfurt/Main, Germany
C.E. Ribak, Department of Anatomy and Neurobiology, University of California at Irvine, Irvine,
CA 92697-1275, USA
A Sahay, Departments of Neuroscience and Psychiatry, Division of Integrative Neuroscience, Columbia
University, New York, NY 10032, USA
V. Santhakumar, Department of Neurology, David Geffen School of Medicine, University of California,
Los Angeles, CA, USA
H.E. Scharfman, Departments of Pharmacology and Neurology, Columbia University, College of
Physicians and Surgeons, New York, NY, USA and Center for Neural Recovery and Rehabilitation
Research, Helen Hayes Hospital, New York State Department of Health, Rte 9W, West Haverstraw,
NY 10993-1195, USA Present address: Nathan Kline Institute for Psychiatric Research, 140 Old
Orangeburg Rd., Bldg. 35, Orangeburg, NY 10962, USA
L. Seress, Central Electron Microscopic Laboratory, Faculty of Medicine, University of Pécs, Szigeti str.
12, 7624 Pécs, Hungary
L.A. Shapiro, Department of Anatomy and Neurobiology, University of California at Irvine, Irvine,
CA 92697-1275, USA
I. Soltesz, Department of Anatomy and Neurobiology, 193 Irvine Hall, University of California, Irvine,
CA 92697, USA
G. Sperk, Department of Pharmacology, Medical University Innsbruck, Peter-Mayr-Str. 1a, 6020
Innsbruck, Austria
P.K. Stanton, Department of Cell Biology and Anatomy, New York Medical College, Valhalla, NY 10595,
USA
T.R. Stoub, Department of Neurological Sciences, Rush University Medical Center, Chicago, IL 60612,
USA
T.P. Sutula, Department of Neurology H6/570 CSC, University of Wisconsin, 600 Highland Avenue,
Madison, WI 53792, USA
M.K. Tallent, Department of Pharmacology and Physiology, Drexel University College of Medicine,
245 N. 15 Street, Philadelphia, PA 19102, USA
C. Wang, Department of Neurological Sciences, Rush University Medical Center, Chicago, IL 60612,
USA
A. Williamson, Department of Neurosurgery, Yale University School of Medicine, New Haven, CT 06518,
USA
viii

M.P. Witter, Department of Anatomy and Neurosciences, VUMC, MF-G102C, P.O.Box 7057, 1007 MB
Amsterdam, The Netherlands
S. Zhao, Institute of Anatomy and Cell Biology, University of Freiburg, Albertstr. 17, D-79104 Freiburg,
Germany
J. Zimmer, Anatomy and Neurobiology, Institute of Medical Biology, University of Southern Denmark,
Winslowparken 21, DK-5000 Odense C, Denmark
Preface

Part I. Why is a guide to the dentate gyrus necessary?


As more knowledge is acquired about the nervous system, one sometimes is lost in a sea of literature. The
current generation surveys the literature primarily from PubMed, which provide a vast resource for data,
but often there is little context for it, and even less of a guide to seminal but pre-1990 papers. Presumably
this context, and a historical overview, can be gained by courses, lectures, and from reviews, but this is often
difficult. One example of this problem is our understanding of the dentate gyrus. This part of the brain has
always been a topic of intense interest, and pioneering studies from the distant and recent past attest to it.
But where in a textbook or review can one find a comprehensive overview some basic information about the
dentate gyrus? The primary goal of the book is to address this problem by bringing together in an organized
manner a series of chapters that address, in an approachable format, the fundamental concepts about
dentate gyrus structure, function, and their implications. Some have more detail than others, some provide
distinct perspectives on a similar topic, but to some extent this is welcome because each reader may desire
different degrees of detail, and some may want to compare viewpoints of different experts in the field.
Often a region of the brain is best understood by first outlining its components, so the first section of the
book deals with the fundamental structure of the dentate gyrus. In the first chapter, David Amaral, Helen
Scharfman, and Pierre Lavenex describe the neuronal organization and intrinsic circuitry. László Seress
discusses this further by comparing the dentate gyrus of different species, emphasizing those that are
primarily used in research (mouse, rat, monkey). Menno Witter outlines the topography of arguably the
most important afferent system, the input from entorhinal cortex (the perforant path). Csaba Leranth and
Tibor Hajszan discuss other extrinsic afferents, such as those from septum, mammillary bodies, and other
areas that are poorly understood compared to the perforant path, particularly from the physiological
perspective. Morten Blaabjerg and Jens Zimmer discuss in detail one of the most impressive and complex
aspects of the dentate gyrus, the mossy fibers axons of the dentate gyrus granule cells. The remarkable
structural complexity of the mossy fiber system, including the unique specializations that characterize their
terminal arbors, is like no other in the hippocampus. Following this overview is a complementary
discussion of the expression, plasticity, and physiology of the mossy fiber system, by David Jaffe and Rafael
Gutierrez. This section then ends with two overviews about the development of the dentate gyrus, the first
by Michael Frotscher, Shanting Zhao, and Eckart Förster, and the second by Guangnan Li and Sam
Pleasure. The emphasis of both is on the unique set of signals that orchestrate the development of the
normal laminar organization and cytoarchitecture of the adult dentate gyrus.
The second section of the book addresses the major neuronal cell types in the dentate gyrus, mostly in the
rodent. Charles Ribak and Lee Shapiro begin this section with an overview of the structural characteristics
of the neuronal cell types, emphasizing their unique ultrastructural ‘‘signatures.’’ Omid Rahimi and Brenda
Claiborne discuss the granule cell from a developmental and morphological perspective. Anne Williamson
and Peter Patrylo review characteristics of granule cells from an electrophysiological viewpoint, both in
rodent and in human tissue studies. The non-granule cells are then addressed in two chapters that cover the
two primary subtypes of non-granule cells: the hilar mossy cells and the GABAergic interneurons. Darrell

ix
x

Henze and Gyuri Buzsáki provide an overview about the mossy cells. Carolyn Houser discusses the
GABAergic neurons, which are heterogeneous in anatomy, expression, and functional organization.
Section three attempts to cover the numerous transmitters and neuromodulators that influence the
dentate gyrus. Glutamatergic inputs are discussed in other parts of the book (the perforant path in Section
1, extrinsic inputs in Section 1, the CA3 input in Section 5), so this leaves the first chapter to consider
GABA (Doug Coulter and Greg Carlson), followed by a series of reviews about non-classical
neurotransmitters. Carrie Drake, Charlie Chavkin, and Teri Milner review the expression and organization
of the opioid system, a complex topic because of the numerous types of opioid peptides, receptors, and
actions. Melanie Tallent addresses the neuropeptide somatostatin, which is normally co-localized with
GABA in a subset of interneurons. Gunther + Sperk, Trevor Hamilton, and William Colmers review
neuropeptide Y and its receptors, functions, and plasticity. Carolyn Harley provides an overview of
norepinephrine, a neuromodulator that has been known to have a robust influence in the dentate gyrus for
decades. Jason Frazier addresses a topic that has been appreciated relatively recently, endocannabinoids in
the dentate gyrus. Mark Pickering and John O’Connor discuss the pro-inflammatory cytokines,
neuromodulators that have not been widely acknowledged until recent years. Marian Joëls reviews the
glucocorticoids, and Devin Binder discusses the neurotrophins, a family of ‘‘growth factors’’ that are
expressed in high concentration in the dentate gyrus, and modulate function in a profound manner. The
reproductive steroids, estrogen, progesterone, and androgen, are addressed by Tibor Hajszan, Teri Milner,
and Csaba Leranth.
One of the more remarkable characteristics of the dentate gyrus is its plasticity, and this is the focus of
the 4th part of the volume. Brian Derrick begins with a perspective that ties together studies of in vitro and
in vivo synaptic plasticity, circuitry, and the behaving animal. Clive Bramham reviews long-term
potentiation (LTP) and its candidate mechanisms. Beatrice Pöschel and Patric Stanton address long-term
depression (LTD). Other examples of plasticity follow, such as lesion-induced plasticity. One of the most
widely studied examples involves the response to a lesion of the entorhinal cortex, and this large literature is
reviewed by Thomas Deller, Domenico Del Turco, Angelika Rappert, and Ingo Bechmann. They also
provide a modern context by updating the field with what is currently understood from studies of mice. On
a separate subject, Jack Parent provides a review of one of the most exciting developments in the history of
dentate gyrus research, and could not illustrate plasticity any better: that neurogenesis occurs in the dentate
gyrus throughout the lifespan of mammals. Finally, another remarkable type of plasticity, is discussed by
Tom Sutula and Ed Dudek: mossy fiber sprouting. This reorganization of the mossy fiber pathway has
captured the attention of many, because it involves dramatic anatomical and physiological changes, and
can occur in response to many types of insults or injury.
In the 5th part of the volume, network considerations are addressed. This is a rich and diverse area of
research that is reflected by many perspectives, each suggesting distinct roles of the dentate gyrus as a
structure. Ray Kesner begins, followed by László Acsády and Szabolcs Káli. David Hsu discusses the
concept that the dentate gyrus is a ‘‘gate’’ or ‘‘filter.’’ John Lisman turns to the question of the perforant
path input, and specifically how the medial vs. lateral perforant path provide distinct roles. Helen
Scharfman discusses the evidence that CA3 is a major input to the dentate gyrus, by virtue of axon
collaterals that innervate diverse dentate neurons. This section ends with a comprehensive model of the
dentate gyrus network, presented by Robert Morgan, Vijy Santhakumar, and Ivan Soltesz.
In the last section of the book, an effort is made to review some of the literature that addresses how
abnormalities within the dentate gyrus contribute to disease, or aging. Monica Chawla and Carol Barnes
discuss how granule cell function changes with age, drawing upon the wealth of information from
electrophysiology, imaging, and other approaches. Peter Patrylo and Anne Williamson also provide a
review of this large field, but with a different perspective. The potential role of dentate gyrus neurogenesis in
depression is reviewed by Amar Sahay, Michael Drew, and Rene Hen, whose lab was one of the first to
provide evidence that deficits in neurogenesis in the dentate gyrus might be a reason for depression. Thomas
xi

Ohm discusses the diverse anatomical changes in the dentate gyrus that occur in Alzheimer’s disease. Leyla
DeToledo-Morrell, Travis Stoub, and Changsheng Wang review their studies of Alzheimer’s disease,
providing a persuasive argument that a major contributing factor is pathology that develops in the
entorhinal area. Finally, Ed Dudek and Tom Sutula review a large and active area of research:
epileptogenesis in the dentate gyrus.
Although a book such as this can never cover all the topics and all information, and certainly cannot do
justice to the multitude of contributors both from the past and present, its purpose is primarily as a guide,
and readers will be led to references that can address topics that are not covered. Particularly for those who
are not well versed in ‘‘hippocampology,’’ it is hoped that this volume can help – and in doing so, provide
new impetus to resolve the many outstanding questions that remain about the dentate gyrus.
Helen E. Scharfman
Columbia University and Helen Hayes Hospital
Part II. A guide to terminology and orientation
The dentate gyrus is a complex, three-dimensionally curved structure that is coupled to another similarly
curved structure, the hippocampus. As such, it is often difficult to understand its orientation in sections
taken from any one plane (e.g., coronal, horizontal, etc.). Moreover, some of the terms used to describe the
components of the dentate gyrus are confusing, and not completely defined in any textbook, to our
knowledge. Therefore, here we provide a general guide to the three-dimensional organization of the dentate
gyrus, and to terminology.
The dentate gyrus in rodents and in primates is a composed of a compact layer containing densely packed
granule cells. Their dendrites mainly lie on one side of the cell layer, and their axons on the other side. The
dendritic layer, called the molecular layer, is bounded by the hippocampal fissure and the granule cell layer.
The layer containing granule cell axons is a polymorphic zone, called the hilus. It includes two main classes
of non-granule cells, GABAergic interneurons, and glutamatergic mossy cells (Fig. 1A). The borders of the
hilus in the rat are shown in Fig. 1. In primates, the border between the hilus and CA3c is difficult to define,
because the hilus is relatively small, and the large mossy cells of the hilus are close to the modified
pyramidal cells of CA3.
The tri-laminar dentate gyrus is folded such that in a cross-section it looks like a V- or C-shaped
structure, having two ‘‘blades’’ and an area where these two blades meet (Fig. 1B). Moreover, it also curves
as a whole, along a dorsal-ventral and medio-lateral axis (Figs. 1C, D). In the rat, the dorsal dentate gyrus
is medial and anterior, and as it extends posteriorly, it moves to a more ventral and lateral position,
eventually curving anteriorly again in its most ventrolateral position (Fig. 1D). Because the result is a
structure that is curved in three-dimensions, it is suggested here that the best way to discuss the ‘‘blades’’ of
the dentate gyrus cell layer is not the way that is most common. As shown in Table 1, the portion of the
dentate gyrus that is closest to CA1 is often called the ‘‘upper’’ or ‘‘dorsal’’ blade. In actuality, this same
part of the granule cell layer becomes ventral and inferior as the dentate gyrus curves from dorsal to ventral
along the longitudinal axis. Therefore, it is more consistent to use the term ‘‘enclosed’’ or ‘‘hidden’’ for this
portion of the cell layer. In other words, it is always enclosed in the CA1–CA3 pyramidal cell regions of the
hippocampal formation. Similarly, the opposite part of the granule cell layer, typically referred to as a
‘‘lower’’ or ‘‘ventral’’ blade, always lies ‘‘exposed’’ or ‘‘open’’ because it is not embedded in the pyramidal
cell layers of the hippocampal formation. Therefore, the terminology that applies to all parts of the dentate
gyrus is ‘‘exposed’’ vs. ‘‘enclosed.’’
In the rodent, the curve of the tri-laminar dentate gyrus blades is slightly different in dorsal and ventral
levels. This is reflected in the fact that sections cut in the horizontal plane through the middle-third of the
dentate gyrus, show a cell layer that is folded into a ‘‘C’’ shape, yet coronal sections in the dorsal area show
a more ‘‘V’’ like structure (Fig. 1B). At the point that is most ventral and anterior (also referred to as
xii

Fig. 1. The three-dimensional organization of the dentate gyrus. (A) The fundamental organization of the neurons in the dentate gyrus
is shown, with the main categories of neurons (granule cells and non-granule cells, the latter including GABAergic interneurons and
glutamatergic mossy cells). (B) The two-dimensional perspective of the dentate gyrus changes with septotemporal position and plane of
section. In the dorsal, coronal plane, the layers are folded similar to a ‘‘V’’, with the apex or crest as the location of the point of the V.
In the middle of the hippocampus cut in horizontal section, the granule cell layer is curved. At extreme septal locations, dorsal sections
demonstrate an arc to the ‘‘V’’; at extreme temporal sites, the exposed blade is relatively small. The hilus of the rodent is illustrated
schematically for two orientations. (C) Coronal sections from rostral to caudal levels illustrate the changes in the blades and illustrate
the lack of consistency of terms such as ‘‘dorsal blade’’ across the entire length of the hippocampus. (D) Comparative views of the
rodent and primate hippocampus illustrate differences in relative locations schematically.

temporal pole), the dentate appears as if almost no exposed blade is present. The most dorsal and anterior
region (also referred to as septal pole) is distinct again, because the curve is flattened and the crest twisted
upward, toward the dorsal aspect of the brain. Some of these changes impact the way the layers are
discriminated, particularly the separation between the hilus and CA3. In the dorsal dentate gyrus, the CA3c
xiii

Table 1

Formal term Synonyms

A. Common nomenclature
Stratum moleculare Molecular layer
Stratum granulosum Granule cell layer
Polymorphic layer (zone) Hilus, hilar region, CA4, zone 4 of Amaral (1978)
Subgranular layer (zone) 50–100 mm of the hilus, next to the granule cell layer
Granule cell Granular cell
GABAergic neuron Interneuron
Glutamatergic hilar neuron Mossy cell
Granule cell axons Mossy fibers
Enclosed blade Upper, dorsal, inner, hidden, superior blade
Exposed blade Lower, ventral, outer, open, inferior blade

Rodents Primates

B. Species-specific nomenclature
Dorsal, septal Dorsal, posterior
Ventral, temporal Ventral, temporal
Rostral, anterior, dorsal pole Rostral, anterior, temporal pole

pyramidal cell layer extends far into the dentate gyrus, almost reaching the point where the blades meet (the
apex or crest). At temporal levels, the CA3c pyramidal cell layer curves toward the enclosed blade. These
issues stress the importance of clear descriptive criteria to differentiate between subfields and layers based
on Nissl-stained sections, in conjunction with other staining methods, as needed.
Another confusing set of terms describes the parts of the longitudinal axis of the dentate gyrus (and the
entire hippocampal formation, for that matter). In rodents, this axis is most commonly referred to as the
septal-to-temporal axis or dorsal-to-ventral axis (Fig. 1). In primates, the longitudinal axis is generally
referred to as the posterior-to-anterior axis, where posterior is comparable to dorsal in rodents. The
confusion lies in the fact that the dentate is positioned differently in the primate, with the more anterior tip
located in the ventral part of the brain (Fig. 1).
Helen Scharfman
Columbia University and Helen Hayes Hospital
Menno Witter
Vrije Universiteit, Amsterdam and Norwegian University of Science and Technology, Trondheim
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 1

The dentate gyrus: fundamental neuroanatomical


organization (dentate gyrus for dummies)

David G. Amaral1,, Helen E. Scharfman2 and Pierre Lavenex3

1
Department of Psychiatry and Behavioral Sciences, The M.I.N.D. Institute and the California National Primate Research
Center, UC Davis, Davis, CA, USA
2
Departments of Pharmacology and Neurology, Columbia University, New York, NY 10032 and the Center for Neural
Recovery and Rehabilitation Research, Helen Hayes Hospital, New York State Department of Health, Rte 9W, West
Haverstraw, NY 10993-1195, USA
3
Department of Medicine, Unit of Physiology, University of Fribourg, 1700 Fribourg, Switzerland

Abstract: The dentate gyrus is a simple cortical region that is an integral portion of the larger functional
brain system called the hippocampal formation. In this review, the fundamental neuroanatomical organ-
ization of the dentate gyrus is described, including principal cell types and their connectivity, and a
summary of the major extrinsic inputs of the dentate gyrus is provided. Together, this information provides
essential information that can serve as an introduction to the dentate gyrus — a ‘‘dentate gyrus for
dummies.’’

Keywords: neuroanatomy; circuits; connections; cell types

Introduction One of the unique features of the hippocampal


formation is that many of its connections are uni-
The dentate gyrus is a simple cortical region that is directional (Fig. 1). Thus, the entorhinal cortex
an integral portion of the larger functional brain provides the major input to the dentate gyrus via
system called the hippocampal formation (Fig. 1) fibers called the perforant path (see Chapter 3).
(Amaral and Lavenex, 2007). The regular organ- However, the dentate gyrus does not return a pro-
ization of its principal cell layers coupled with the jection to the entorhinal cortex. Since the ent-
highly ordered laminar distribution of many of its orhinal cortex is the source of much of the cortical
inputs has encouraged its use as a model system sensory information that the hippocampal forma-
for many facets of modern neurobiology. In this tion uses to carry out its functions, and since the
chapter, we present the fundamental principles of dentate gyrus is the major termination of projec-
neuroanatomical organization of the dentate tions from the entorhinal cortex, it is reasonable to
gyrus, its principal cell types and their connectiv- consider the dentate gyrus as the first step in the
ity, and a summary of the major extrinsic inputs of processing of information processing that ulti-
the dentate gyrus. mately leads to the production of episodic mem-
ories. Moreover, the unique neuroanatomy of the
Corresponding author. Tel.: +1 916 703 0225; dentate gyrus predicts that it carries out a specific
Fax: +1 916 703 0287; E-mail: dgamaral@ucdavis.edu information-processing task with the information

DOI: 10.1016/S0079-6123(07)63001-5 3
4
5

that it receives from the entorhinal cortex and number of interneurons that reside in the molec-
ultimately conveys to the CA3 field of the hippo- ular layer and fibers from a variety of extrinsic
campus. inputs terminate there. The principal cell layer, the
In this chapter, we will first give a general over- granule cell layer, is made up largely of densely
view of the neuroanatomical organization of the packed granule cells. The thickness of the granule
dentate gyrus. We will then put the spotlight on cell layer ranges from 4 to 8 neurons or 60 mm.
three of its most important neurons: the dentate While the granule cell layer is mainly made up of
granule cell, the dentate pyramidal basket cell and granule cells, there are some other neurons that are
the mossy cell. We will summarize the major located at the boundary of the granule and pol-
structural features of these neurons as well as the ymorphic layers. The cell body of the dentate py-
connections they make within and beyond the ramidal basket cell, for example, is often located
dentate gyrus. We will then very briefly discuss just within the granule cell layer at its border with
some of the other neurons located in the dentate the polymorphic layer. The granule cell layer en-
gyrus. Finally, we briefly summarize the origin and closes a cellular region, the polymorphic cell layer,
termination of the major extrinsic inputs to the which constitutes the third layer of the dentate
dentate gyrus. Many of these topics will be dis- gyrus (Fig. 1). A number of cell types are located
cussed in greater detail in other chapters of this in the polymorphic layer but the most prominent is
book. But, this chapter provides a synoptic view in the mossy cell that we will describe below.
order to serve as an introduction to the dentate The dentate gyrus along with many other fields
gyrus — this is dentate gyrus for dummies. of the hippocampal formation create a banana-
shaped structure that, in the rat, runs from the
septal nuclei rostrally to the temporal cortex cau-
Overall organization dally. Thus, the long axis is typically called the
septotemporal axis. The axis at right angles to the
The dentate gyrus has three layers (Figs. 1–2). septotemporal axis is typically called the transverse
There is a relatively cell-free layer called the mo- axis. The dentate gyrus has a relatively similar
lecular layer which, in the rat, is approximately structure at all septotemporal levels of the hippo-
250 mm thick and is occupied by, among other campal formation and is not generally divided into
things, the dendrites of the dentate granule cells. subregions. However, it does tend to have more of
The other major occupants of the molecular layer a ‘‘V’’ shape septally and more of a ‘‘U’’ shape
are the fibers of the perforant path that originate temporally (Fig. 3). In discussing features of the
in the entorhinal cortex. There are also a small dentate gyrus, it is often useful to refer to a

Fig. 1. The rat hippocampal formation. (A) Nissl-stained horizontal section through the hippocampal formation of the rat. The major
fields are indicated. Projections (1) originate from layer II of the entorhinal cortex (EC) and terminate in the molecular layer of the
dentate gyrus (DG) and in the stratum lacunosum-moleculare of the CA3 field of the hippocampus. An additional component of the
perforant path originates in layer III and terminates in the CA1 field of the hippocampus and the subiculum. Granule cells of the DG
give rise to the mossy fibers (2) that terminate both within the polymorphic layer of the DG and within stratum lucidum of the CA3
field of the hippocampus. The CA3 field, in turn, gives rise to the Schaffer collaterals (3) that innervate the CA1 field of the
hippocampus. Pyramidal cells in CA1 project to the subiculum and to the deep layers of the EC. The subiculum also gives rise to
projections to the deep layers of the EC. (B) Line drawing to illustrate the major regional and laminar organization of the DG. The DG
is divided into a molecular layer (ml) a granule cell layer (gcl) and a polymorphic layer (pl). The molecular layer is divided into three
sublayers based on the laminar organization of inputs. The hippocampus is divided into CA3, CA2 and CA1 subfields. Within CA3, a
number of layers are defined. The main cell layer is the pyramidal cell layer (pcl). Deep to the pyramidal cell layer is the stratum oriens
(so); deep to this is the white matter of the alveus (al). Superficial to the pyramidal cell layer is stratum lucidum (sl), stratum radiatum
(sr) and stratum lacunosum-moleculare (sl-m). Fields CA2 and CA1 have the same layers as CA3 except for stratum lucidum. The
remainder of the hippocampal formation is made up of the subiculum (Sub), presubiculum (Pre), parasubiculum (Para) and EC. Layers
of these latter structures are indicated with roman numerals. Additional abbreviations: ab, angular bundle; fi, fimbria; hf, hippocampal
fissure. (C) Schematic illustration of DG and hippocampus to illustrate position of suprapyramidal blade, infrapyramidal blade and
crest of the DG. This model is used in subsequent illustrations to demonstrate the major cell types and connections of the DG. (See
Color Plate 1.1 in color plate section.)
6

Fig. 2. Comparative Nissl and Timm’s staining of the rat and monkey dentate gyrus. Approximately the same portions of the rat (A &
B) and monkey (C & D) hippocampal formation are illustrated. The magnification of the paired images are different. The increased
complexity of the hilar region in the monkey compared to the rat is obvious in these photomicrographs. In the monkey, the CA3 field
inserts more deeply into the dentate gyrus and accounts for a much greater area of the hilar region. While the darkly stained
polymorphic layer is pronounced in the rat, it is much thinner in the monkey brain. The condensed layer of mossy fibers is also much
more pronounced in the rat compared to the monkey. Also note that the strict lamination of the molecular layer into thirds in the rat is
not as sharp as in the monkey. This corresponds to a more gradient-like termination of the medial and lateral perforant paths in the
monkey compared to the rat. Calibration, 250 mm (A, B); 350 mm (C, D).

particular transverse portion of the ‘‘V’’- or ‘‘U’’- relative to other hippocampal structures such as
shaped structure. The portion of the granule cell the CA1 field of the hippocampus or the ent-
layer that is located between the CA3 field and the orhinal cortex, there has not been substantial
CA1 field (separated by the hippocampal fissure) is phylogenetic modification. For example, in the rat,
called the suprapyramidal (above CA3) blade and there are approximately one million granule cells
the portion opposite to this, the infrapyramidal in the dentate gyrus. There are approximately ten
(below CA3) blade. The region bridging the two times more granule cells in the monkey than in the
blades (at the apex of the ‘‘V’’ or ‘‘U’’) is called the rat, a ratio that parallels the overall volume differ-
crest (Fig. 3). ences. But there are only 15 times more dentate
granule cells in humans when compared to rats,
while the volume of the dentate gyrus plus the
Comparative neuroanatomy of the dentate gyrus hippocampus is 100 times larger in humans than
in rats (reviewed in more detail in Amaral and
The comparative neuroanatomy of the dentate Lavenex, 2007).
gyrus is covered in detail in Chapter 2. However, Despite the similarities, there are also some
we would like to raise a few issues for which there obvious differences. In the rat, the CA3 field of the
remains substantial confusion in the literature. hippocampus inserts into the dentate gyrus and
First, the basic trilaminar structure of the dentate abuts the cells of the polymorphic layer. By using
gyrus is common across all species studied. And, stains such as the Timm’s stain that identifies
7

Fig. 3. Horizontal sections through the rat hippocampal formation. This figure illustrates a more dorsally situated (A) and a more
ventrally situated (B) horizontal section through the rat hippocampal formation. The approximate level of the section is illustrated on a
3D reconstruction of a magnetic resonance image series of the rat brain. Subtle differences in the cytoarchitectonic organization are
seen throughout the hippocampal formation. The dentate gyrus, takes on a more ‘‘V’’ shape dorsally and a more ‘‘U’’ shape ventrally.
Calibration bar ¼ 250 mm. (See Color Plate 1.3 in color plate section.)

depositions of metals such as zinc (Fig. 2), the 1987). Similarly, the mossy cells of the rat have
heavily labeled polymorphic layer can be easily very rare dendrites that enter the molecular layer.
differentiated from the CA3 field. This differenti- Yet, in the macaque monkey, many of the mossy
ation is not so clear in the nonhuman primate or in cells give rise to numerous dendrites that enter the
the human dentate gyrus. This is because the CA3 molecular layer (Buckmaster and Amaral, 2001).
field is much more prominent within the confines The implication of this is that rat mossy cells get
of the dentate gyrus and the polymorphic layer has little if any perforant path input whereas monkey
thinned to a very narrow subgranular region mossy cells could get a quite substantial perforant
(Fig. 2). In the early nomenclature espoused by path input.
Lorente de Nó, he used a term ‘‘CA4’’ that was Unfortunately, many of the connections of the
vaguely defined but appears to have been intended monkey and human dentate gyrus have yet to be
for the part of CA3 that inserts within the dent- investigated. One example of a projection that has
ate gyrus. However, sometimes Lorente de Nó ap- been studied in both species is the commissural
pears to have used the CA4 term for the connection. While cells of the polymorphic layer
polymorphic layer. We have recommended that (mainly the mossy cells) give rise to a very robust
this term be abandoned and that even the part of commissural projection to the molecular layer of
CA3 that enters the limbs of the dentate gyrus the contralateral dentate gyrus in the rat, this pro-
be called CA3. jection does not exist in the nonhuman primate
There are a variety of other differences in the brain or in the human brain. As connections of the
rat, monkey and human dentate gyrus. The gran- primate and human dentate gyrus are better stud-
ule cells, which we will describe in much more de- ied, it is undoubtedly the case that there will be
tail shortly, only have apical dendrites in the rat. other fundamental differences that will affect not
But in the monkey and human, many granule cells only normal function but also pathological proc-
also have basal dendrites (Seress and Mrzljak, esses. This raises the caveat that while the rodent
8

dentate gyrus is an extremely valuable tool for spiny apical dendrites. The branches extend
evaluating function and potential mechanisms of throughout the molecular layer and the distal tips
pathology, caution must be exercised when these of the dendritic tree end just at the hippocampal
results are extrapolated to the human dentate fissure or at the ventricular surface. The total
gyrus. For the remainder of this chapter, we will length of the dendritic trees of granule cells lo-
focus on neuroanatomical findings from the rat cated in the suprapyramidal blade are, on average,
brain. larger than those of cells in the infrapyramidal
blade (3500 mm vs. 2800 mm, respectively). Dend-
rites of cells in the suprapyramidal blade have
Major cell types of the dentate gyrus — and their 1.6 spines/mm, whereas dendrites in the infrapy-
connections ramidal blade have 1.3 spines/mm (Desmond and
Levy, 1985). Thus, an estimate for the number of
The dentate granule cell spines on the average suprapyramidal granule cell
would be around 5600 and for an infrapyramidal
The principal cell type of the dentate gyrus is the cell 3640. Since virtually all of the excitatory inputs
granule cell (Figs. 4 and 5). The dentate granule to granule cells are on these dendritic spines, these
cell has an elliptical cell body with a width of ap- numbers indicate the approximate number of ex-
proximately 10 mm and a height of 18 mm (Clai- citatory synapses that dentate granule cells receive
borne et al., 1990). The granule cell bodies are from all sources. As noted earlier, the total
tightly packed together and, in most cases, there is number of granule cells in one dentate gyrus of
no glial sheath interposed between cells. The gran- the rat is 1.2  106 (West et al., 1991; Rapp
ule cell has a characteristic cone-shaped tree of and Gallagher, 1996). Although neurogenesis in

Fig. 4. The dentate granule cell. The characteristic features of the dentate granule cell are illustrated, including its axonal arbor. A
collateral plexus gives rise to numerous (200) typical synapses on cells located within the polymorphic layer. Most of these synapses
are onto the dendrites of inhibitory interneurons. Some of the large mossy fiber expansions are also distributed in the polymorphic
layer. Many of these terminate on the proximal dendrites of mossy cells. The mossy fiber axons ultimately enter the CA3 field where
they travel through the full transverse extent of the field. On their course, they terminate with mossy fiber expansions on a small
number (15–20) of CA3 pyramidal cells. Additional abbreviations: gc, granule cell; pc, pyramidal cell. (See Color Plate 1.4 in color
plate section.)
9

Fig. 5. The dentate granule cell. A photomicrograph (A) and line drawing (B) of a prototypical granule cell that was filled with Lucifer
yellow in a hippocampal slice. The dendrites arise primarily from the apical surface of the cell body, and the axon emerges from the
basal surface. The spiny dendrites extend into the molecular layer until the hippocampal fissure, and the axon collateralizes profusely
within the polymorphic layer. Calibration bar ¼ 25 mm (A), 20 mm (B).

the dentate gyrus persists into adulthood, and fibers. The mossy fibers have unusually large
appears to be under environmental control, mod- boutons that form en passant synapses with the
ern stereological studies have shown that the total mossy cells of the polymorphic layer and with the
number of granule cells does not vary in adult CA3 pyramidal cells of the hippocampus. Less
animals (Rapp and Gallagher, 1996). This implies appreciated is the fact that the mossy fiber axons
that there is a steady state turnover of granule cells give rise to a distinctive set of collaterals that
rather than a continuous accretion. heavily innervate cells within the polymorphic
The packing number of granule cells and their layer of the dentate gyrus. Each principal mossy
ratio to CA3 pyramidal cells varies along the fiber (which is on the order of 0.2–0.5 mm in di-
septotemporal axis (Gaarskjaer, 1978b); the pack- ameter) gives rise to about seven thinner collater-
ing density is higher septally than temporally. als within the polymorphic layer before entering
Since the packing density of CA3 pyramidal cells the CA3 field of the hippocampus. As much as
follows an inverse gradient, the net result is that at 2300 mm of collateral axonal plexus is generated by
septal levels of the hippocampal formation, the a single mossy fiber in the polymorphic layer
ratio of granule cells to CA3 pyramidal cells is (Claiborne et al., 1986). Within the polymorphic
something on the order of 12:1, whereas at the layer, the mossy fiber collaterals branch exten-
temporal pole the ratio drops to 2:3. Since the CA3 sively and the daughter branches bear two types of
pyramidal cells are the major recipients of granule synaptic varicosities. Numerous small (approxi-
cell innervation, and the number of mossy fiber mately 2 mm) spherical synaptic varicosities are
synapses is roughly the same along the septotem- distributed unevenly along these collaterals. There
poral axis, contact probability is much lower are 160–200 of these varicosities distributed
septally than temporally. throughout the axonal collateral plexus of a
single granule cell, and these form contacts on
dendrites located in the polymorphic layer. At the
The mossy fibers — projections to the polymorphic end of many of the collateral branches there are
layer also larger (3–5 mm diameter), irregularly shaped
varicosities that resemble, although are smaller
The granule cells give rise to distinctive unmyelin- than, the mossy fiber boutons found in CA3. The
ated axons which Ramón y Cajal called mossy mossy fiber terminals in the polymorphic layer
10

establish contacts with the proximal dendrites of the parent cell body. In the proximal portion of
the mossy cells (Ribak et al., 1985), the basal CA3 (closer to the dentate gyrus), mossy fibers
dendrites of the pyramidal basket cells as well as are distributed below, within and above the py-
with other, unidentified cells. Acsady et al. (1998) ramidal cell layer. The fibers located below the
reported that the vast majority of mossy fiber col- layer, i.e., those that are in the area occupied pri-
laterals in the polymorphic cell layer terminate on marily by basal dendrites, are generally called the
GABAergic interneurons. Since there are 160–200 infrapyramidal bundle (Fig. 2). The fibers located
such varicosities (compared with 20 of the larger within the pyramidal cell layer are called the in-
thorny excrescences), mossy fiber axons synapse trapyramidal bundle and those located above the
on a larger number of interneurons than mossy pyramidal cell layer (in the area occupied mainly
cells or CA3 pyramidal cells. Mossy fiber collat- by proximal apical dendrites) the suprapyramidal
erals occasionally enter the granule cell layer, but bundle. The suprapyramidal bundle occupies the
they rarely enter the molecular layer under normal stratum lucidum. At mid and distal portions of
conditions. The collaterals that enter the granule CA3, the intra- and infrapyramidal bundles are
cell layer appear to terminate preferentially on the largely eliminated and those fibers that were in
apical dendritic shafts of pyramidal basket cells. these regions cross the pyramidal cell layer and
The lack of mossy fiber innervation of the molec- join the other mossy fibers within stratum lucid-
ular layer changes dramatically in pathological um.
conditions and sprouting of mossy fibers is one of Granule cells at all transverse positions within
the major hallmarks of temporal lobe epilepsy (see the granule cell layer generate mossy fibers that
Chapter 29). extend for the full transverse extent of CA3
(Gaarskjaer, 1981). Cells located in the infrapy-
ramidal blade of the granule cell layer have axons
The mossy fibers — projections to CA3 that tend to enter CA3 in the infrapyramidal
bundle, but ultimately cross the pyramidal cell
The rat dentate gyrus does not project to any brain layer to enter the deep portion of stratum lucidum.
region other than the CA3 field of the hippocam- The axons of granule cells located in the crest of
pus (but see Chapter by Zimmer for exceptions the dentate gyrus tend to enter CA3 in the intra-
in other species). The mossy fibers terminate in a pyramidal bundle and also ultimately ascend into
relatively narrow zone mainly located just above stratum lucidum. Cells located in the suprapyram-
the CA3 pyramidal cell layer (Blackstad et al., idal blade of the dentate gyrus give rise to axons
1970; Swanson et al., 1978; Gaarskjaer, 1978a; that enter CA3 in the stratum lucidum and
Claiborne et al., 1986). In the proximal portion of continue within the most superficial portion of
CA3, mossy fibers are also located below and stratum lucidum (Claiborne et al., 1986).
within the pyramidal cell layer. The layer of mossy Early Golgi anatomists indicated that the
fiber termination located just above the pyramidal mossy fiber axons were mainly oriented perpen-
cell layer is called stratum lucidum. There is no dicular to the long axis of the hippocampus.
indication that dentate neurons other than the Blackstad and colleagues confirmed the largely
granule cells project to CA3. In particular, cells transverse trajectory of the mossy fibers using de-
in the polymorphic layer do not project to the generation track tracing methods. Mossy fibers
hippocampus, at least in the rodent. The dentate originating from any particular septotemporal
projection to CA3 stops near the border of CA3 level travel through the full transverse extent of
with CA2, and the lack of granule cell input is one CA3 in a very narrow septotemporal zone of
of the main features that distinguishes CA3 from 1 mm, which has been described as a lamella.
CA2 pyramidal cells. There is one peculiarity of the mossy fiber pro-
As far as can be determined, all dentate granule jection, however, that has eluded explanation.
cells appear to project to CA3 and the axon tra- Lorente de Nó (1934) and McLardy (McLardy
jectory is partially correlated with the position of and Kilmer, 1970) indicated that the mossy fibers
11

actually do change their course and take a lon- The dentate pyramidal basket cell
gitudinal direction, but not until they reach the
distal portion of CA3. The extent of this distal As with most other brain regions, some neurons of
longitudinal projection was further clarified by the dentate gyrus, such as the granule cells, are
Swanson and colleagues, using the autoradio- excitatory whereas others are inhibitory. The most
graphic method of neural tract tracing (Swanson intensively studied type of inhibitory interneuron
et al., 1978). They showed that granule cells in the dentate gyrus is the pyramidal basket cell
located at septal levels give rise to mossy fibers (Figs. 6 and 7; Ribak et al., 1978; Ribak and
that travel throughout most of the transverse Seress, 1983). These cells are generally located
extent of CA3 at the same septotemporal level. along the interface between the granule cell layer
But just at the CA3/CA2 border, they abruptly and the polymorphic layer. They have pyramidal-
change course and travel toward the temporal shaped cell bodies that are substantially larger
pole for nearly 2 mm. The extent of this longitu- (25–35 mm in diameter) than the granule cells
dinal component, however, appears to depend on (10–18 mm). Ramón y Cajal first described the
the septotemporal location of the cells of origin. pyramidal basket cells as having a single, principal
Granule cells in the mid to temporal portions of aspiny apical dendrite directed into the molecular
the dentate gyrus have mossy fibers that exhibit layer that divides into several aspiny branches,
only a slight temporal inclination at their distal and several basal dendrites that divide and extend
extremity. And mossy fibers that originate at the into the polymorphic cell layer. The number of
extreme temporal pole of the dentate gyrus barely basket cells is not constant throughout either the
extend to the CA3/CA2 border and have little or transverse or septotemporal extents of the dentate
no longitudinal component. On the face of it, this gyrus (Seress and Pokorny, 1981). At septal levels,
indicates that some CA3 pyramidal cells that are the ratio of basket cells to granule cells is 1:100
located very close to the border with CA2 may be in the suprapyramidal blade and 1:180 in the
contacted by mossy fiber axons from granule cells infrapyramidal blade. At temporal levels, the
spread out over a much broader septotemporal number is 1:150 for the suprapyramidal blade
extent of the dentate gyrus. They might, therefore, and 1:300 for the infrapyramidal blade.
form a special class of integrator CA3 cells. An The basket portion of the name refers to the
elegant neuroanatomical verification of this has fact that the axon of these cells forms pericellular
come from the work of Acsady et al. (1998) plexuses, like the covering of a chianti bottle, that
who stained dentate granule cells by using in vivo form encompassing synapses with the cell bodies
intracellular techniques. Not only do mossy of granule cells. The dentate pyramidal basket
fibers travel temporally for 2–3 mm, but they also cell along with other basket cells located just be-
demonstrate typical mossy fiber expansions along low the granule cell layer contribute to a very
this portion of their trajectory. dense terminal plexus that is confined to the
There is substantial evidence indicating that the granule cell layer (Struble et al., 1978; Sik et al.,
granule cells use glutamate as their primary trans- 1997). The terminals in this basket plexus are
mitter, and the asymmetric contacts made between GABAergic and form symmetric, inhibitory con-
the mossy fiber expansions and the thorny excres- tacts, located primarily on the cell bodies and
cences would tend to confirm this notion. The proximal dendritic shafts of the apical dendrites
mossy fibers are, nonetheless, also immunoreactive of the granule cells. Analysis of Golgi-stained
for several other neuroactive substances. At least axonal plexuses from single basket cells in the rat
some of the mossy fibers demonstrate immunore- indicates that they extend for distances greater
activity for the opioid peptide dynorphin and they than 900 mm in the transverse axis and 1.5 mm
are also immunoreactive for GABA (Walker et al., in the septotemporal axis. This widely distributed
2002). The chemical neuroanatomy of the dentate axonal plexus would allow a single basket cell to
granule cell will be dealt with in more detail in influence as many as 10,000 (1%) of the granule
Chapter 6. cells.
12

Fig. 6. The pyramidal basket cell. The cell body of the pyramidal basket cell is located at the interface between the granule cell layer
and the polymorphic cell layer. The axon (arrow) emerges from the apical dendrite. Collaterals of this axon form a curtain of terminals
that synapse with the granule cell bodies. Additional abbreviations: pbc, pyramidal basket cell. (See Color Plate 1.6 in color plate
section.)

Fig. 7. The pyramidal basket cell. A prototypical pyramidal basket cell is shown after intracellular injection of Neurobiotin. The
montage that was created after visualization of the cell (A) and line drawing (B) illustrate the characteristics of this cell type. An arrow
points to the axon. Calibration bar ¼ 25 mm (A), 40 mm (B).
13

Fig. 8. The mossy cell. A line drawing of a mossy cell (mc) in the polymorphic layer. The axon (arrow) develops a plexus within the
polymorphic layer, and also an ipsilateral projection to the inner molecular layer, known as the associational pathway. The main axon
also projects contralaterally to the inner molecular layer, forming the commissural pathway. The ipsilateral projection increases in
density with distance from the cell body of origin. (See Color Plate 1.8 in color plate section.)

Fig. 9. The mossy cell. A montage of several focal planes (A), and line drawing (B) of a classic mossy cell, illustrating the characteristic
thorny excrescences and dendritic tree of this cell type. Thorny excrescences are present proximal to the soma. The dendrites of the cell
extend throughout nearly the entire polymorphic region, but few enter either the granule cell or molecular layers. Calibration
bar ¼ 25 mm (A), 50 mm (B).

The mossy cell located in his ‘‘subzone of fusiform cells’’, and is


undoubtedly what Lorente de Nó referred to as
The polymorphic layer harbors a variety of neu- ‘‘modified pyramids’’. The mossy cell received its
ronal cell types. The most common, and certainly current name from Amaral (1978) who studied this
the most impressive, is the mossy cell (Figs. 8 and and other neurons of the polymorphic layer using
9) This cell type is probably what Ramón y Cajal the Golgi method. The cell bodies of the mossy
referred to as the ‘‘stellate or triangular’’ cells cells are large (25–35 mm) and are often triangular
14

or multipolar in shape. Three or more thick dend- cells and GABAergic interneurons. It should also
rites originate from the cell body and extend be noted for the sake of completeness that a small
for long distances within the polymorphic layer. component of the commissural connection in the
Each principal dendrite bifurcates once or twice rat does appear to arise from GABAergic neurons
and generally gives rise to a few side branches. (Ribak et al., 1986).
While most of the dendritic branches remain There are a number of interesting features about
within the polymorphic layer, an occasional dend- the ‘‘feedback’’ projection from the mossy cells to
rite pierces the granule cell layer and enters the the granule cells. First, the projection from mossy
molecular layer. The mossy cell dendrites virtually cells located at any particular level of the dentate
never enter the adjacent CA3 field in the rat. gyrus is distributed widely along the longitudinal
The most distinctive feature of the mossy cell is axis, both septally and temporally from the point
that all of its proximal dendrites are covered by of origin. Axons from any particular septotempo-
very large and complex spines, the so-called thorny ral point in the dentate gyrus may innervate as
excrescences (Fig. 9). These are the distinctive sites much as 75% of the long axis of the dentate gyrus
of termination of the mossy fiber axons. While (Amaral and Witter, 1989). Second, the projection
thorny excrescences are also observed on the prox- to the molecular layer at the septotemporal level of
imal dendrites of CA3 pyramidal cells, they are origin is very weak, but gets increasingly stronger
never as dense or as complex as the ones on the at levels that are progressively more distant from
mossy cells. The distal dendrites of the mossy cell the cells of origin. Remembering that mossy cells
have typical pedunculate spines that appear to be are the recipients of massive innervation from the
less dense than those on the distal dendrites of the granule cells at their same level (via the mossy fiber
hippocampal pyramidal cells. collaterals into the polymorphic layer), it would
appear that the mossy cells pass on the collective
output of granule cells from one septotemporal
Mossy cell projections level to granule cells located at distant levels of the
dentate gyrus.
The inner third of the molecular layer (Fig. 2) re-
ceives a projection that originates exclusively from
neurons in the polymorphic layer (Laurberg and Other neurons of the dentate gyrus
Sorensen, 1981; Frotscher et al., 1991; Buckmaster
et al., 1992, 1996). Since, in the rat, this projection There has been an explosion in the number of in-
originates both in the ipsilateral and contralateral terneurons identified in the rat hippocampal for-
sides of the dentate gyrus, it has been called the mation. A detailed overview of the characteristics
associational/commissural projection. As noted of the various hippocampal interneurons has been
above, the commissural portion of this projection published by Freund and Buzsaki (1996) and is
does not exist in the primate brain. There is sub- also covered in Chapter 13 of this book. Many of
stantial evidence from neural track tracing studies the cell types can be distinguished on the basis of
that this projection arises almost exclusively from the distribution of their axonal plexus. Some have
cells in the polymorphic layer and mainly from the axons that terminate on cell bodies, whereas others
mossy cells. The fact that the mossy cells are have axons that terminate exclusively on the initial
immunoreactive for glutamate (Soriano and segments of other axons. Interneurons have also
Frotscher, 1994) adds credence to the notion that been distinguished on the basis of their inputs.
the associational/commissural projection is excita- Some are preferentially innervated, for example,
tory. A direct electrophysiological demonstration by the serotonergic fibers originating from the
of the excitatory nature of the mossy cell synapse raphe nuclei. Interneurons can also be differenti-
on the granule cell has been provided by Scharf- ated from principal cells on the basis of their
man (1994). Scharfman (1995) has also demon- electrophysiological characteristics. At least some
strated that mossy cells innervate both granule interneurons have high rates of spontaneous
15

activity and fire in relation to the theta rhythm. the location of the cell body and to the region
For this reason, interneurons are often called theta where the axon is distributed. Unfortunately, this
cells (Sik et al., 1997). terminology has not been embraced and more
Within the same subgranular region occupied by widely applied to neurons of the hippocampal
the cell bodies and dendrites of the pyramidal formation.
basket cells are several other cell types with dis- Frotscher and colleagues have described a
tinctly different somal shapes, as well as different second type of neuron in the molecular layer that
dendritic and axonal configurations. Some of these resembles the chandelier or axo-axonic cell origi-
cells are multipolar with several aspiny dendrites nally found in the neocortex (Soriano and Frotsc-
entering the molecular and polymorphic layers, her, 1989). These cells are generally located
while others tend to be more fusiform-shaped with immediately adjacent or even within the superfi-
a similar dendritic distribution. As Ribak and cial portion of the granule cell layer. The axo-
colleagues have pointed out, many of these cells axonic cell is named for the fact that its axon
share fine structural characteristics such as infold- descends from the molecular layer into the granule
ed nuclei, extensive perikaryal cytoplasm with cell layer, collateralizes profusely and then termi-
large Nissl bodies and intranuclear rods. More- nates, with symmetric synaptic contacts, exclu-
over, it appears that all of these cells give rise to sively on the axon initial segments of granule cells.
axons that contribute to the basket plexus in the Thus, their shape resembles that of a chandelier.
granule cell layer. Most of these neurons are Each axo-axonic cell may innervate the axon in-
immunoreactive for GABA, form symmetrical itial segments of as many as 1000 granule cells.
synaptic contacts with the cell bodies, proximal Since these cells are immunoreactive for markers
dendrites and occasionally with axon initial seg- of GABAergic neurons and make symmetrical
ments of granule cells, and therefore function as synapses, it is likely that they provide an addi-
inhibitory interneurons. These cells are not neuro- tional means of inhibitory control of granule cell
chemically homogeneous, however, since subsets output.
appear to colocalize distinct categories of other
neuroactive substances (Ribak, 1992).
Neurons of the polymorphic cell layer

Neurons of the molecular layer Besides the mossy cell, there are a number of fusi-
form cells in the polymorphic layer. The main
The molecular layer is occupied primarily by difference between the fusiform cell types is
dendrites of the granule cells, pyramidal basket whether they have spines or not and the charac-
cells and polymorphic layer cells, as well as axons teristic shapes and sizes of the spines. One type,
and terminal axonal arbors from the entorhinal the long-spined multipolar cell first described by
cortex and other sources. At least two neuron Amaral (1978) has recently been called the HIPP
types are also present in the molecular layer. The cell (hilar perforant path-associated cell) (Fig. 10;
first is located deep in the molecular layer, has a Han et al., 1993). The conspicuous feature of this
multipolar or triangular cell body and gives rise to cell is the distribution of copious, long and often
an axon that produces a substantial terminal branched spines over its cell body and dendrites.
plexus largely limited to the outer two thirds of Intracellular staining techniques demonstrate that
the molecular layer. This neuron has aspiny dend- these cells have axons that ascend into the outer
rites that remain mainly within the molecular layer two-thirds of the molecular layer (i.e., the perfo-
and has been called the MOPP cell (molecular rant path zone) and terminate with symmetrical
layer perforant path-associated cell). This termi- and presumably inhibitory synapses on the dend-
nology was proposed by Han et al. (1993) to bring rites of granule cells. An amazing feature of these
some order to naming interneurons in the hippo- neurons is that their axonal plexus can extend for
campal formation. The lettering system refers to as much as 3.5 mm along the septotemporal axis of
16

Fig. 10. The long-spined cell. A montage (A) and line drawing (B) of a long-spined cell in the polymorphic layer. The extremely long
spines that characterize this cell type are marked by arrowheads, and can be proximal as well as distal to the cell body, although in this
example they were primarily located along distal dendrites. The axon of this cell collateralized in the molecular layer. Some of the long-
spined cells correspond to the GABAergic interneurons that colocalize somatostatin, so-called HIPP cells. Calibration bar ¼ 25 mm
(A), 50 mm (B).

the dentate gyrus (the entire length of the dentate Bakst et al., 1986; Freund and Buzsaki, 1996;
gyrus in the rat is only 10 mm) and may generate Sik et al., 1997; Boyett and Buckmaster, 2001).
as many as 100,000 synaptic terminals. Since in- The somatostatin-positive cells all colocalize with
hibitory interneurons typically have aspiny den- GABA, and are the source of the somatostatin
drites and relatively local axonal plexuses, this immunoreactive fibers and terminals in the outer
long spined multipolar/HIPP cell is a very atypical two-thirds of the molecular layer. This system of
interneuron! At least some of these HIPP cells fibers, which forms contacts on the distal dendrites
appear to correspond to the somatostatin/GABA of the granule cells, provides a third means for in-
cells that give rise to the somatostatin innervation hibitory control of granule cell activity, in addition
of the outer portion of the molecular layer. to the GABAergic basket cell plexus and axo-ax-
Antibodies directed against the peptide som- onic terminals of the chandelier cells. Since elec-
atostatin have revealed that neurons such as the tron microscopic studies have demonstrated that
HIPP cells scattered throughout the polymorphic the somatostatin cells are contacted by mossy fiber
layer are immunoreactive for this peptide, and ac- terminals, the projection to the outer molecular
count for approximately 16% of the GABAergic layer thus constitutes a local feedback inhibitory
cells in the dentate gyrus (Morrison et al., 1982; circuit.
17

Fig. 11. Neurons of the polymorphic region. The summary figure reprinted from Amaral (1978) illustrates the diversity of cells types
within the dentate gyrus and proximal CA3 region of the rat. Based on Golgi staining and cameral lucida drawings of individual
neurons from multiple preparations. See original paper for description of cell types.

Interestingly, unlike the mossy cell associational in the polymorphic layer of the dentate gyrus
projection that terminates more heavily at distant whose axonal plexus have not yet been well
levels of the dentate gyrus, the GABA/somatosta- described. A summary of the neurons in the pol-
tin projection terminates most heavily around the ymorphic region, to some extent still incompletely
level of the cells of origin, and termination rapidly understood, is shown in Fig. 11.
decreases within approximately 1.5 mm septally
and temporally to the cells of origin. Thus, the
Extrinsic afferents
mossy cell projection and the somatostatin/GABA
cell projection have terminal fields that are spa-
Afferent projections
tially complementary in both radial and septotem-
poral axes. The distribution suggests that the two
Although this topic is covered in more details in
cell types mediate distal excitation and local inhi-
other chapters (see Chapters 3 and 4), we will
bition, respectively.
provide a brief general synopsis of the afferent
There are also multipolar or triangular cells in
projections of the dentate gyrus.
the polymorphic layer with thin, aspiny dendrites
that extend both within the hilus and within the
molecular layer. The axons of these HICAP cells Entorhinal cortex projection to the dentate gyrus
(hilar commissural-associational pathway related
cells) extend through the granule cell layer and The dentate gyrus receives its major input from the
branch profusely in the inner third of the molec- entorhinal cortex, via the so-called perforant path-
ular layer. There is a variety of other neuron types way (Ramón y Cajal, 1893). The projection to the
18

dentate gyrus arises mainly from cells located in axon arbor in the molecular layer (Tamamaki and
layer II of the entorhinal cortex (Fig. 1), although Nojyo, 1993).
a minor component of the projection also comes While it is often assumed that the perforant path
from layers V and VI (Steward and Scoville, 1976; fibers from the entorhinal cortex are the only hip-
Deller et al., 1996). The entorhinal terminals are pocampal input reaching the dentate gyrus, it is
strictly confined to the outer two-thirds of the now clear that at least minor projections also arise
molecular layer where they form asymmetric in the presubiculum and parasubiculum (Kohler,
synapses that account for nearly 85% of the total 1985). These fibers enter the molecular layer of the
axospinous terminations (Nafstad, 1967; Hjorth- dentate gyrus and ramify in a zone that is inter-
Simonsen and Jeune, 1972). These contacts occur spersed between the lateral and medial perforant
primarily on the dendritic spines of granule cells, path projections. The presubicular axons tend to
although a small number of perforant path fibers be thicker than those from the entorhinal cor-
also form asymmetric synapses on the shafts of tex and give rise to collaterals that take a radial
GABA-positive interneurons. course in the molecular layer. Virtually nothing is
The perforant pathway can be divided into two currently known about which cells these fibers in-
parts based on the region of origin, pattern of ter- nervate or what type of transmitter they use. Since
mination and appearance in histochemical and the presubiculum receives the only direct input
immunohistochemical preparations. In the rat, the from the anterior thalamic nucleus, these fibers
two divisions have been called the lateral and me- provide a potential link by which thalamic infor-
dial perforant paths because they originate from mation could reach the dentate gyrus.
the lateral and medial entorhinal areas, respec-
tively. Interestingly, while the cells of origin for
these projections look very similar and the light Basal forebrain inputs: projections from the septal
microscopic appearance of their projections is in- nuclei
distinguishable, they do demonstrate substantial
histochemical and chemical neuroanatomical The dentate gyrus receives relatively few inputs
differences which remain largely unexplored. from subcortical structures. The most robust is the
Perforant path fibers originating in the lateral ent- projection from the septal nuclei (Mosko et al.,
orhinal area terminate in the most superficial third 1973; Swanson, 1978; Amaral and Kurz, 1985).
of the molecular layer, whereas the fibers origi- The septal projection arises from cells of the me-
nating from the medial entorhinal area terminate dial septal nucleus and the nucleus of the diagonal
in the middle third of the molecular layer (see band of Broca. Septal fibers heavily innervate cells
Chapter 3 for more detail on the perforant path of the polymorphic layer, particularly in a narrow
projection). These terminal zones are readily dis- region just subjacent to the granule cell layer.
tinguished by the classical Timm’s stain method Septal fibers are lightly distributed throughout the
for visualization of heavy metals which demon- molecular layer.
strates dense staining in the outer third of the mo- A major portion of the fibers of the septal pro-
lecular layer, a near absence of staining in the jection to the dentate gyrus are cholinergic. Thirty
middle third and a dark staining in the inner third to fifty percent of the cells in the medial septal
that is associated with the commissural/associat- nucleus and 50–75% of the cells in the nucleus of
ional connection (Fig. 2). Projections from both the diagonal band that project to the hippocampal
areas of the entorhinal cortex innervate the entire formation are cholinergic. Many of the other sep-
transverse extent of the molecular layer. The thin tal cells that project to the dentate gyrus are GAB-
axon branches (0.1 mm) in the molecular layer of Aergic. The most interesting facet of this hetero-
the dentate gyrus show periodic varicosities with a geneous septal projection is that the cholinergic
thickness of 0.5–1.0 mm. Most of entorhinal cortex and GABAergic components target different cell
layer II spiny stellate cells project up to 2 mm in types. Fibers of the septal GABAergic projections
the septotemporal direction forming a sheet-like terminate preferentially on other GABAergic
19

nonpyramidal cells, such as the basket pyramidal The dentate gyrus receives a minor and diffusely
cells of the dentate gyrus, and they form symmet- distributed dopaminergic projection that arises
rical, presumably inhibitory contacts. The heaviest mainly from cells located in the ventral tegmental
GABAergic septal termination is on interneurons area. The dopaminergic fibers terminate mainly in
located in the polymorphic layer. The cholinergic the polymorphic layer.
septal projection to the dentate gyrus, in contrast, The serotonergic projection that originates from
terminates mainly on granule cells, making asym- median and dorsal divisions of the raphe nuclei
metric, presumably excitatory contacts on dendri- also terminates most heavily in the polymorphic
tic spines, chiefly in the inner third of the layer in an immediately subgranular portion of the
molecular layer. Only 5–10% of the septal layer (Conrad et al., 1974; Moore and Halaris,
cholinergic synapses are on dentate interneurons. 1975; Köhler and Steinbusch, 1982; Vertes et al.,
The large mossy cells are innervated by cholinergic 1999). A number of GABAergic interneurons
fibers (Lübke et al., 1997). appear to be preferentially innervated by the se-
rotonergic fibers. The targets are often the pyram-
idal basket cells. Fusiform neurons in the region,
Supramammillary and other hypothalamic inputs particularly those that are stained for the calcium
binding protein calbindin, are also very heavily
The major hypothalamic projection to the dentate innervated. As with the cholinergic projection
gyrus arises from a population of large cells in the from the septum, many of the cells in the raphe
supramammillary area that caps and partially sur- nuclei that project to the hippocampal formation
rounds the medial mammillary nuclei ( Wyss et al., appear to be nonserotonergic, but their transmitter
1979; Dent et al., 1983; Vertes, 1993; Magloczky et is not known.
al., 1994). The supramammillary projection mainly
terminates in a narrow zone located just superficial
to the granule cell layer and lightly in the poly- Conclusions
morphic layer or the remaining portion of the
molecular layer. The vast majority of the sup- In the remainder of this book, the reader will
ramammillary fibers terminate on the proximal receive far greater detail on many of the neuroan-
dendrites of granule cells. This projection is exci- atomical features of the dentate gyrus that were
tatory and is likely using glutamate as a primary only briefly touched on in this chapter. Every
neurotransmitter (Kiss et al., 2000). Most, but not effort will also be made to correlate the neuro-
all, of the glutamatergic supramammillary neurons anatomy of the dentate gyrus with both its elect-
that project to the dentate gyrus also colocalize rophysiological and functional attributes. We will
calretinin; some of these cells also colocalize sub- close with a short reflection on the position of
stance P (Borhegyi and Leranth, 1997). the dentate gyrus both within the hippocampal
formation and within the brain at large.
It has been fashionable at times to highlight the
Brainstem inputs simple neuroanatomy of the dentate gyrus and to
claim that it provides a heuristic for studying and
The dentate gyrus receives a particularly promi- understanding the much more complicated neo-
nent noradrenergic input from the nucleus locus cortex. This is probably a misguided strategy.
coeruleus (Pickel et al., 1974; Swanson and Hart- There are numerous features of the dentate gyrus
man, 1975; Loughlin et al., 1986). The nor- that make it absolutely unique from a neuroana-
adrenergic fibers terminate mainly in the tomical point of view and thus presumably from a
polymorphic layer of the dentate gyrus and ex- functional point of view as well. The largely uni-
tend into the stratum lucidum of CA3, as if pref- directional nature of its inputs and outputs is one
erentially terminating in the zones occupied by distinctive feature. The distinctive mossy fibers
mossy fibers. that give rise to massive synaptic endings that have
20

as many as 40 active sites onto postsynaptic neu- Amaral, D.G. and Witter, M.P. (1989) The three dimensional
rons is another. Our view is that the distinctive organization of the hippocampal formation: a review of
anatomical data. Neuroscience, 31: 571–591.
neuroanatomy of the dentate gyrus has been
Bakst, I., Avendaño, C., Morrison, J.H. and Amaral, D.G.
sculpted by evolution to play a precise and unique (1986) An experimental analysis of the origins of the
role in the hippocampal information processing somatostatin immunoreactive fibers in the dentate gyrus of
that ultimately leads to the production of declar- the rat. J. Neurosci., 6: 1452–1462.
ative memories. Whatever computations the dent- Blackstad, T.W., Brink, K., Hem, J. and Jeune, B. (1970) Dis-
ate gyrus executes will undoubtedly be different tribution of hippocampal mossy fibers in the rat. An exper-
imental study with silver impregnation methods. J. Comp.
from, and will work in concert with, the compu- Neurol., 138: 433–450.
tations carried out by field CA3 and other portions Borhegyi, Z. and Leranth, C. (1997) Distinct substance P- and
of the hippocampal formation. The challenge for calretinin-containing projections from the supramammillary
the future will be to understand enough about the area to the hippocampus in rats — a species difference be-
tween rats and monkeys. Exp. Brain Res., 115: 369–374.
input/output patterns of the dentate gyrus to iden-
Boyett, J.M. and Buckmaster, P.S. (2001) Somatostatin-
tify the specific contribution it makes to memory immunoreactive interneurons contribute to lateral inhibitory
formation. circuits in the dentate gyrus of control and epileptic rats.
There are other unique features of the dentate Hippocampus, 11: 418–422.
gyrus. Why, for example, is there the conspicuous Buckmaster, P.S. and Amaral, D.G. (2001) Intracellular re-
production during adult life of new granule cells cording and labeling of mossy cells and proximal CA3 py-
ramidal cells in macaque monkeys. J. Comp. Neurol., 430:
and how do these support the unique function of 264–281.
the dentate gyrus? Does the formation of these Buckmaster, P.S., Strowbridge, B.W., Kunkel, D.D., Schmiege,
new neurons have significance for the broader D.L. and Schwartzkroin, P.A. (1992) Mossy cell axonal pro-
questions of stem cell recovery of function or does jections to the dentate gyrus molecular layer in the rat hip-
it reflect again the unique attributes of one partic- pocampal slice. Hippocampus, 2: 349–362.
Buckmaster, P.S., Wenzel, H.J., Kunkel, D.D. and Schwa-
ular brain region? Despite the ‘‘simplicity’’ of the rtzkroin, P.A. (1996) Axon arbors and synaptic connections
dentate gyrus, it will undoubtedly remain the topic of hippocampal mossy cells in the rat in vivo. J. Comp.
of extensive research for decades to come. And, Neurol., 366: 271–292.
the chapters that constitute the remainder of this Claiborne, B.J., Amaral, D.G. and Cowan, W.M. (1986) A light
and electron microscopic analysis of the mossy fibers of the
book will provide the reader both with a valuable
rat dentate gyrus. J. Comp. Neurol., 246: 435–458.
summary of the state of dentate gyrus science and Claiborne, B.J., Amaral, D.G. and Cowan, W.M. (1990) A
guidance for research into the future. quantitative three-dimensional analysis of granule cell
dendrites in the rat dentate gyrus. J. Comp. Neurol., 302:
206–219.
Conrad, L.C.A., Leonard, C.M. and Pfaff, D.W. (1974) Con-
nections of the median and dorsal raphe nuclei in the rat: an
References autoradiographic and degeneration study. J. Comp. Neurol.,
156: 179–206.
Acsady, L., Kamondi, A., Sik, A., Freund, T. and Buzsaki, G. Deller, T., Martinez, A., Nitsch, R. and Frotscher, M. (1996) A
(1998) GABAergic cells are the major postsynaptic targets of novel entorhinal projection to the rat dentate gyrus — direct
mossy fibers in the rat hippocampus. J. Neurosci., 18: innervation of proximal dendrites and cell bodies of granule
3386–3403. cells and gabaergic neurons. J. Neurosci., 16: 3322–3333.
Amaral, D.G. and Kurz, J. (1985) An analysis of the origins of Dent, J.A., Galvin, N.J., Stanfield, B.B. and Cowan, W.M.
the cholinergic and noncholinergic septal projections to the (1983) The mode of termination of the hypothalamic projec-
hippocampal formation of the rat. J. Comp. Neurol., 240: tion to the dentate gyrus: An EM autoradiographic study.
37–59. Brain Res., 258: 1–10.
Amaral, D.G. (1978) A Golgi study of cell types in the hilar Desmond, N.L. and Levy, W.B. (1985) Granule cell dendritic
region of the hippocampus in the rat. J. Comp. Neurol., 182: spine density in the rat hippocampus varies with spine shape
851–914. and location. Neurosci. Lett., 54: 219–224.
Amaral, D.G. and Lavenex, P. (2007) Hippocampal neuro- Freund, T.F. and Buzsaki, G. (1996) Interneurons of the hip-
anatomy. In: Andersen P., Morris R., Amaral D., Bliss T. pocampus [Review]. Hippocampus, 6: 347–470.
and O’Keefe J. (Eds.), The Hippocampus Book. Oxford Frotscher, M., Seress, L., Schwerdtfeger, W.K. and Buhl, E.
University Press, New York, p. 872. (1991) The mossy cells of the fascia dentata: a comparative
21

study of their fine structure and synaptic connections in ro- Moore, R.Y. and Halaris, A.E. (1975) Hippocampal innerva-
dents and primates. J. Comp. Neurol., 312: 145–163. tion by serotonin neurons of the midbrain raphe in the rat. J.
Gaarskjaer, F.B. (1978a) Organization of the mossy fiber sys- Comp. Neurol., 164: 171–184.
tem of the rat studied in extended hippocampi. II. Experi- Morrison, J.H., Benoit, R., Magistretti, P.J., Ling, N. and
mental analysis of fiber distribution with silver impregnation Bloom, F.E. (1982) Immunohistochemical distribution of
methods. J. Comp. Neurol., 178: 73–88. prosomatostatin-related peptides in hippocampus. Neurosci.
Gaarskjaer, F.B. (1978b) Organization of the mossy fiber sys- Lett., 34: 137–142.
tem of the rat studied in extended hippocampi. I. Terminal Mosko, S., Lynch, G. and Cotman, C.W. (1973) The distribu-
area related to number of granule and pyramidal cells. J. tion of septal projections to the hippocampus of the rat. J.
Comp. Neurol., 178: 49–72. Comp. Neurol., 152: 163–174.
Gaarskjaer, F.B. (1981) The hippocampal mossy fiber system of Nafstad, P.H.J. (1967) An electron microscope study on the
the rat studied with retrograde tracing techniques. Correla- termination of the perforant path fibres in the hippocampus
tion between topographic organization and neurogenetic and the fascia dentata. Zeitsch. Zellforsch. Mikrosk. Anat.,
gradients. J. Comp. Neurol., 203: 717–735. 76: 532–542.
Han, Z.S., Buhl, E.H., Lorinczi, Z. and Somogyi, P. (1993) A Pickel, V.M., Segal, M. and Bloom, F.E. (1974) A radioauto-
high degree of spatial selectivity in the axonal and dendritic graphic study of the efferent pathways of the nucleus locus
domains of physiologically identified local-circuit neurons in coeruleus. J. Comp. Neurol., 155: 15–42.
the dentate gyrus of the rat hippocampus. Eur. J. Neurosci., Ramón y Cajal, S. (1893) Estructura del asta de Ammon y
5: 395–410. fascia dentata. Ann. Soc. Esp. Hist. Nat., 22.
Hjorth-Simonsen, A. and Jeune, B. (1972) Origin and Rapp, P.R. and Gallagher, M. (1996) Preserved neuron number
termination of the hippocampal perforant path in the rat in the hippocampus of aged rats with spatial learning deficits.
studied by silver impregnation. J. Comp. Neurol., 144: Proc. Natl. Acad. Sci. U.S.A., 93: 9926–9930.
215–232. Ribak, C.E. (1992) Local circuitry of GABAergic basket cells in
Kiss, J., Csaki, A., Bokor, H., Shanabrough, M. and Leranth, the dentate gyrus. Epilepsy Res. Suppl., 7: 29–47.
C. (2000) The supramammillo-hippocampal and supramam- Ribak, C.E. and Seress, L. (1983) Five types of basket cell in the
millo-septal glutamatergic/aspartatergic projections in the hippocampal dentate gyrus: a combined Golgi and electron
rat: a combined. Neuroscience, 97: 657–669. microscopic study. J. Neurocytol., 12: 577–597.
Kohler, C. (1985) Intrinsic projections of the retrohippocampal Ribak, C.E., Seress, L. and Amaral, D.G. (1985) The
region in the rat brain. I. The subicular complex. J. Comp. development, ultrastructure and synaptic connections of
Neurol., 236: 504–522. the mossy cells of the dentate gyrus. J. Neurocytol., 14:
Köhler, C. and Steinbusch, H. (1982) Identification of serotonin 835–857.
and non-serotonin-containing neurons of the mid-brain rap- Ribak, C.E., Seress, L., Peterson, G.M., Seroogy, K.B., Fallon,
he projecting to the entorhinal area and the hippocampal J.H. and Schmued, L.C. (1986) A GABAergic inhibitory
formation. A combined immunohistochemical and fluores- component within the hippocampal commissural pathway. J.
cent retrograde tracing study in the rat brain. Neuroscience, Neurosci., 6: 3492–3498.
7: 951–975. Ribak, C.E., Vaughn, J.E. and Saito, K. (1978) Immunocyto-
Laurberg, S. and Sorensen, K.E. (1981) Associational and chemical localization of glutamic acid decarboxylase in neu-
commissural collaterals of neurons in the hippocampal for- ronal somata following colchicine inhibition of axonal
mation (hilus fasciae dentate and subfield CA3). Brain Res., transport. Brain Res., 140: 315–332.
212: 287–300. Scharfman, H.E. (1994) Evidence from simultaneous intracel-
Lorente, de Nó R. (1934) Studies on the structure of the cer- lular recordings in rat hippocampal slices that area CA3
ebral cortex. II. Continuation of the study of the ammonic pyramidal cells innervate dentate hilar mossy cells. J. Ne-
system. J. Psychol. Neurol., 46: 113–177. urophysiol., 72: 2167–2180.
Loughlin, S.E., Foote, S.L. and Bloom, F.E. (1986) Efferent Scharfman, H.E. (1995) Electrophysiological evidence that
projections of nucleus locus coeruleus: topographic organi- dentate hilar mossy cells are excitatory and innervate
zation of cells of origin demonstrated by three-dimensional both granule cells and interneurons. J. Neurophysiol., 74:
reconstruction. Neuroscience, 18: 291–306. 179–194.
Lübke, J., Deller, T. and Frotscher, M. (1997) Septal innerva- Seress, L. and Mrzljak, L. (1987) Basal dendrites of granule
tion of mossy cells in the hilus of the rat dentate gyrus — an cells are normal features of the fetal and adult dentate gyrus
anterograde tracing and intracellular labeling study. Exp. of both monkey and human hippocampal formations. Brain
Brain Res., 114: 423–432. Res., 405: 169–174.
Magloczky, Z., Acsady, L. and Freund, T.F. (1994) Principal Seress, L. and Pokorny, J. (1981) Structure of the granular layer
cells are the postsynaptic targets of supramammillary affer- of the rat dentate gyrus. A light microscopic and Golgi study.
ents in the hippocampus of the rat. Hippocampus, 4: J. Anat., 133: 181–195.
322–334. Sik, A., Penttonen, M. and Buzsaki, G. (1997) Interneurons in
McLardy, T. and Kilmer, W.L. (1970) Hippocampal circuitry. the hippocampal dentate gyrus — an in vivo intracellular
Am. Psychol., 25: 563–566. study. Eur. J. Neurosci., 9: 573–588.
22

Soriano, E. and Frotscher, M. (1989) A GABAergic axo-axonic association pathways in the rat. J. Comp. Neurol., 181:
cell in the fascia dentata controls the main excitatory hippo- 681–716.
campal pathway. Brain Res., 503: 170–174. Tamamaki, N. and Nojyo, Y. (1993) Projection of the
Soriano, E. and Frotscher, M. (1994) Mossy cells of the rat entorhinal layer-II neurons in the rat as revealed by intra-
fascia dentata are glutamate-immunoreactive. Hippocampus, cellular pressure-injection of neurobiotin. Hippocampus, 3:
4: 65–69. 471–480.
Steward, O. and Scoville, S.A. (1976) Cells of origin of ent- Vertes, R.P. (1993) PHA-L analysis of projections from the
orhinal cortical afferents to the hippocampus and fascia den- supramammillary nucleus in the rat. J. Comp. Neurol., 326:
tata of the rat. J. Comp. Neurol., 169: 347–370. 595–622.
Struble, R.G., Desmond, N.L. and Levy, W.B. (1978) Ana- Vertes, R.P., Fortin, W.J. and Crane, A.M. (1999) Projections
tomical evidence for interlamellar inhibition in the fascia of the median raphe nucleus in the rat [Review]. J. Comp.
dentata. Brain Res., 152: 580–585. Neurol., 407: 555–582.
Swanson, L.W. (1978) The anatomical organization of septo- Walker, M.C., Ruiz, A. and Kullmann, D.M. (2002) Do mossy
hippocampal projections. In: Gray A. (Ed.), Functions of the fibers release GABA? Epilepsia, 43: 196–202.
Septo-Hippocampal System. Elsevier North Holland, Am- West, M.J., Slomianka, L. and Gundersen, H.J.G. (1991) Un-
sterdam, pp. 25–48. biased stereological estimation of the total number of neu-
Swanson, L.W. and Hartman, B.K. (1975) The central ad- rons in the subdivisions of the rat hippocampus using the
renergic system. An immunofluorescence study of the loca- optical fractionator. Anat. Rec., 231: 482–497.
tion of cell bodies and their efferent connections in the rat Wyss, J.M., Swanson, L.W. and Cowan, W.M. (1979) Evi-
utilizing dopamine-beta-hydroxylase as a marker. J. Comp. dence for an input to the molecular layer and the stra-
Neurol., 163: 467–506. tum granulosum of the dentate gyrus from the
Swanson, L.W., Wyss, J.M. and Cowan, W.M. (1978) An au- supramammillary region of the hypothalamus. Anat. Em-
toradiographic study of the organization of intrahippocampal bryol., 156: 165–176.
Plate 1.1. The rat hippocampal formation. (A) Nissl-stained horizontal section through the hippocampal formation of the rat. The
major fields are indicated. Projections (1) originate from layer II of the entorhinal cortex (EC) and terminate in the molecular layer of
the dentate gyrus (DG) and in the stratum lacunosum-moleculare of the CA3 field of the hippocampus. An additional component of
the perforant path originates in layer III and terminates in the CA1 field of the hippocampus and the subiculum. Granule cells
of the DG give rise to the mossy fibers (2) that terminate both within the polymorphic layer of the DG and within stratum lucidum of
the CA3 field of the hippocampus. The CA3 field, in turn, gives rise to the Schaffer collaterals (3) that innervate the CA1 field of the
hippocampus. Pyramidal cells in CA1 project to the subiculum and to the deep layers of the EC. The subiculum also gives rise to
projections to the deep layers of the EC. (B) Line drawing to illustrate the major regional and laminar organization of the DG. The DG
is divided into a molecular layer (ml) a granule cell layer (gcl) and a polymorphic layer (pl). The molecular layer is divided into three
sublayers based on the laminar organization of inputs. The hippocampus is divided into CA3, CA2 and CA1 subfields. Within CA3, a
number of layers are defined. The main cell layer is the pyramidal cell layer (pcl). Deep to the pyramidal cell layer is the stratum oriens
(so); deep to this is the white matter of the alveus (al). Superficial to the pyramidal cell layer is stratum lucidum (sl), stratum radiatum
(sr) and stratum lacunosum-moleculare (sl-m). Fields CA2 and CA1 have the same layers as CA3 except for stratum lucidum. The
remainder of the hippocampal formation is made up of the subiculum (Sub), presubiculum (Pre), parasubiculum (Para) and EC. Layers
of these latter structures are indicated with roman numerals. Additional abbreviations: ab, angular bundle; fi, fimbria; hf, hippocampal
fissure. (C) Schematic illustration of DG and hippocampus to illustrate position of suprapyramidal blade, infrapyramidal blade and
crest of the DG. This model is used in subsequent illustrations to demonstrate the major cell types and connections of the DG. (For
B/W version, see page 4 in the volume.)
Plate 1.3. Horizontal sections through the rat hippocampal formation. This figure illustrates a more dorsally situated (A) and a more
ventrally situated (B) horizontal section through the rat hippocampal formation. The approximate level of the section is illustrated on a
3D reconstruction of a magnetic resonance image series of the rat brain. Subtle differences in the cytoarchitectonic organization are
seen throughout the hippocampal formation. The dentate gyrus, takes on a more ‘‘V’’ shape dorsally and a more ‘‘U’’ shape ventrally.
Calibration bar ¼ 250 mm. (For B/W version, see page 7 in the volume.)

Plate 1.4. The dentate granule cell. The characteristic features of the dentate granule cell are illustrated, including its axonal arbor. A
collateral plexus gives rise to numerous (200) typical synapses on cells located within the polymorphic layer. Most of these synapses
are onto the dendrites of inhibitory interneurons. Some of the large mossy fiber expansions are also distributed in the polymorphic
layer. Many of these terminate on the proximal dendrites of mossy cells. The mossy fiber axons ultimately enter the CA3 field where
they travel through the full transverse extent of the field. On their course, they terminate with mossy fiber expansions on a small
number (15–20) of CA3 pyramidal cells. Additional abbreviations: gc, granule cell; pc, pyramidal cell. (For B/W version, see page 8 in
the volume.)
Plate 1.6. The pyramidal basket cell. The cell body of the pyramidal basket cell is located at the interface between the granule cell layer
and the polymorphic cell layer. The axon (arrow) emerges from the apical dendrite. Collaterals of this axon form a curtain of terminals
that synapse with the granule cell bodies. Additional abbreviations: pbc, pyramidal basket cell. (For B/W version, see page 12 in the
volume.)

Plate 1.8. The mossy cell. A line drawing of a mossy cell (mc) in the polymorphic layer. The axon (arrow) develops a plexus within the
polymorphic layer, and also an ipsilateral projection to the inner molecular layer, known as the associational pathway. The main axon
also projects contralaterally to the inner molecular layer, forming the commissural pathway. The ipsilateral projection increases in
density with distance from the cell body of origin. (For B/W version, see page 13 in the volume.)
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 3

The perforant path: projections from the entorhinal


cortex to the dentate gyrus

Menno P. Witter1,2,

1
Institute for Clinical and Experimental Neurosciences, Department of Anatomy & Neurosciences, VU University Medical
Center, MF-G102C, P.O. Box 7057, 1007 MB Amsterdam, The Netherlands
2
Centre for the Biology of Memory, Norwegian University of Science and Technology, Trondheim, Norway

Abstract: This paper provides a comprehensive description of the organization of projections from the
entorhinal cortex to the dentate gyrus, which together with projections to other subfields of the hippo-
campal formation form the so-called perforant pathway. To this end, data that are primarily from an-
atomical studies in the rat will be summarized, complimented with comparative data from other species.
The analysis of the organization of any of the connections of the hippocampus, including that of the
entorhinal cortex to the dentate gyrus, is severely hampered because of the complex three-dimensional
shape of the hippocampus. In particular in rodents, but to a lesser extent also in primates, all traditional
planes of sectioning will result in sections that at some point or another do not cut through the hippo-
campus at an angle that is perpendicular to its long axis. To amend this, we will describe own unpublished
tracing data obtained in the rat with the use of the so-called extended preparation. A number of issues will
be addressed. First, data will be summarized which will clarify the laminar origin of the perforant pathway
within the entorhinal cortex. Second, we will discuss whether or not a radial organization, along the
proximo-distal dendritic axis of granule cells, characterizes the entorhinal-dentate projection. Third, we will
discuss whether this projection is governed by any transverse organization, and fourth, we will focus on the
organization along the longitudinal axis. Finally, the synaptic organization and the contralateral ent-
orhinal-dentate projection will be described briefly. Taken together, the available data suggest that the
projection from the entorhinal cortex to the dentate gyrus is a fairly well conserved connection, present in
all species studied, exhibiting a grossly similar organization.

Keywords: hippocampus; anatomy; parahippocampal region; topographical organization; subiculum

Introduction dentate gyrus with its major cortical input. Sec-


ond, the perforant pathway is among the path-
A description of the projection from the ent- ways in the brain that have been most actively
orhinal cortex to the dentate gyrus is an indispen- investigated, beginning with the earliest pioneers
sable part of any work on the dentate gyrus for at in neuroscience who described its remarkable
least two reasons. First, this pathway, generally structure and organization. These studies were
referred to as the perforant pathway, provides the followed by others, which clarified the central im-
portance of the perforant pathway to hippocam-
Corresponding author. Tel.:+31 20 4448048; pal function and plasticity, studies which continue
Fax:+31 20 4448054; E-mail: mp.witter@vumc.nl to this day.

DOI: 10.1016/S0079-6123(07)63003-9 43
44

The elaborate Golgi studies of Ramón y Cajal use of tools to circumvent a major problem with
(1911) and Lorente de Nó (1933), first demonstrated hippocampal studies in rodents, the fact that the
that the entorhinal cortex is the origin of an im- hippocampal formation and the entorhinal cortex
mensely strong projection to the dentate gyrus. are curved structures. Cutting through such curved
These observations were subsequently corroborated structures in any plane will result in sections that
and extended in a seemingly continuous stream of at some point or another do not cut through the
tracing studies using axonal degeneration techniques hippocampus and entorhinal cortex at an angle
(Blackstad, 1958; Hjorth-Simonsen, 1972; Hjorth- that is perpendicular to the respective long axes.
Simonsen and Jeune, 1972), followed by transport Initially described by Gaarskjaer (1978), and
of radioactively labeled amino acids and horse- advocated by Ishizuka (2001), and likewise by
radish peroxidase (Van Hoesen and Pandya, 1975; Amaral and Witter (1989), this problem can
Steward, 1976; Wyss, 1981; Witter and Groenewe- be amended by using the so-called extended prep-
gen, 1984; Witter, 1989; Witter et al., 1989b) and aration, and in a number of studies this approach
finally the more recently introduced sensitive tracing has resulted in improved understanding of the
with lectines and dextran-amines (Köhler, 1985; connectional organization of the system (Amaral
Witter, 1989; Tamamaki and Nojyo, 1993; Deller and Witter, 1989; Ishizuka et al., 1990; Ishizuka,
et al., 1996; Deller, 1998). In the same period, 2001).
a number of retrograde tracing studies, again using Many of the disputes about functions of the
a variety of different tracers, have further contrib- entorhinal-dentate projection, and entorhinal-hip-
uted to our current understanding of the entorhinal- pocampal connectivity more generally, are likely
dentate projection (Ruth et al., 1982, 1988; Witter to be related to this spatial problem, at least in
et al., 1989b; Dolorfo and Amaral, 1998). Although part. A proper understanding of the complex
the dentate gyrus is the ‘‘traditional’’ target of the three-dimensional organization of the system is
entorhinal-hippocampal fibers, there is ample evi- essential to evaluate data from functional studies
dence that the entorhinal cortex also projects to the which suggest heterogeneity, as has been suggested
hippocampal fields CA1–CA3, and to the subiculum for the dorsal versus the ventral hippocampus, for
(Steward, 1976; Steward and Scoville, 1976; Witter example (Witter et al., 1989a; Moser et al., 1993).
et al., 1989a; Desmond et al., 1994; Naber et al., Another, more recent example deals with the
2001; Baks-Te-Bulte et al., 2005). rather confusing data addressing the functional
In all species, the hippocampus and entorhinal relevance of the entorhinal cortex. Here, again,
cortex show a complex three-dimensional orienta- studies have benefited significantly from taking
tion and relationship. In rats in particular, this has into account the complex three-dimensional or-
lead to a general tendency to restrict hippocampal ganization of cortico-entorhinal-hippocampal con-
and entorhinal studies to those parts that are most nections (Fyhn et al., 2004; Hafting et al., 2005;
easily accessible, such as dorsal hippocampus and Hargreaves et al., 2005; Steffenach et al., 2005;
central parts of entorhinal cortex. Initially this re- Sargolini et al., 2006). Over the years, one of our
sulted in rather coarse descriptions of the overall aims has been to further our understanding of the
functionality and connectivity of entorhinal-dent- system through systematic descriptions of its com-
ate relationships, and it was only after analyses plex anatomical organization. For the present pa-
started to include the full extent of both entorhinal per, the focus will be on the entorhinal-dentate
cortex and dentate gyrus that we became aware of projection; in a recently published parallel paper,
the complex topographical organization of this the focus was on entorhinal projections to the
connection. It is now more widely appreciated that subiculum (Witter, 2006). The data support a cur-
the hippocampus and entorhinal cortex are not as rently prevailing view that there are two differen-
homogeneous in their organization as initially tially organized components within the entorhinal-
conceived, (cf. Witter and Groenewegen, 1990). dentate projection, originating from the lateral
A major reason for this new perspective was the and medial entorhinal cortex respectively.
45

Nomenclature portion of the granule cell layer will be called the


exposed blade (which is synonymous with infra-
Dentate gyrus pyramidal, ventral, outer, or free blade/limb). The
region of the ‘‘V’’ or ‘‘U’’ that unites the two
In all species, the dentate gyrus represents part of blades will be referred to as the crest, also called
the cortex situated closest to the free rim of the apex or vertex.
original cortical anlage, but the final position and Aside from the longitudinal or dorsoventral
overall structural organization of the dentate gyrus axis, a description of the organization of the dent-
and of the hippocampus as a whole may be quite ate gyrus is strongly facilitated by defining two
different among species (Stephan, 1975; Voogd additional axes, a transverse and a radial axis. The
et al., 1998). To give a striking example, one just transverse axis is oriented perpendicular to the
has to compare the position of the hippocampus in long axis, similar to the term ‘lamellar axis’ be-
the rat with that in primates. In the rat the struc- cause a section through the transverse axis would
ture extends from a dorsomedial position, in close reveal the classic lamellar arrangement of the hip-
proximity to the most caudodorsal part of the pocampus. The radial axis refers to an axis that is
septal complex to a ventrolateral position in close essentially perpendicular to the transverse axis. It
proximity to the most caudomedial parts of the can be defined as a perpendicular to the granule
amygdaloid complex. In contrast, in primates, this cell layer, extending from molecular layer through
same structure starts from a dorsocaudal position, the granule cell layer to the hilus. It therefore rep-
close to the splenium of the corpus callosum, ex- resents the orientation of the granule cell dendrites
tending ventrorostrally all the way again to appo- from their most distal apical extent to their origin
sition in close proximity to the amygdaloid at the granule cell soma. For the present descrip-
complex. Moreover, in the rat, the dorsal portion tion we will refer to the three layers of the dentate
of the dentate gyrus, plus parts of the hippocam- gyrus as molecular layer, granule layer, and the
pus proper, seem to fold backwards, continuing hilus of the dentate gyrus as the third, inner, or
into a structure generally referred to as the gyrus polymorph layer. The molecular layer will be
fasciolaris (fasciolaris cinereum), whereas in pri- further subdivided into inner, middle, and outer
mates such a folding seems to be the main char- one-thirds.
acteristic of the most ventral and anterior part,
referred to as genu and uncal portions, spatially
associated with the amygdaloid complex. A second Entorhinal cortex
example refers to the overall connotation that the
principal cell layer of dentate gyrus contains gran- In view of the classical notion that the entorhinal
ule cells, i.e. cells without basal dendrites but with cortex is the source of the projection to the dentate
an extensive apical tuft. In contrast, in the human gyrus (Ramón y Cajal, 1911), we and others have
and monkey dentate, a sizeable proportion, in hu- used this to delineate the entorhinal cortex of the
mans up to 30%, of the granule cells do have basal rat and the monkey (Insausti et al., 1997). The
dendrites (Seress and Mrzljak, 1987; Lim et al., entorhinal cortex comprises six layers, including
1997). the two cell-sparse layers I and IV, respectively
Irrespective of such differences, it is useful to called the molecular layer and lamina dissecans.
develop a coherent and generally applicable no- There are several reasons to use this nomenclature
menclature, for example to be able to distinguish for entorhinal cortical layers instead of the initially
one portion of the granule cell layer from another. proposed terms defined by Lorente de Nó (1933),
For this paper, I will call the portion of the granule who considered the lamina dissecans as the deep,
cell layer that is adjacent to CA1, the enclosed sparsely populated part of layer III, and used the
blade (which is synonymous with suprapyramidal, designation IV for the layer of large pyramidal
dorsal, or inner blade/limb), and the opposite cells which is layer Va in the currently employed
46

terminology (Witter et al., 1989a; Witter and Fiber pathways


Amaral, 2004).
In general, the entorhinal cortex can be subdi- After leaving the entorhinal cortex, perforant path
vided into two components generally referred to as fibers enter the underlying white matter and the an-
lateral and medial entorhinal cortex or Brodman’s gular bundle. They then traverse the pyramidal cell
areas 28a and 28b respectively (for a more detailed layer of the subiculum and cross the hippocampal
description and comparison of different nomen- fissure to enter the dentate gyrus, or distribute to the
clatures used in the rat and in different species the molecular layer of the subiculum and the hippocam-
reader is referred to a number of reviews (Witter pus. The entorhinal cortex fibers also take alternative
et al., 1989a). In the rat, and likewise in the mouse, routes such as projecting through the alveus before
the entorhinal cortex has been further subdivided entering the hippocampus, or by traversing the mo-
into dorsolateral (DLE), dorsal-intermediate lecular layers of the entorhinal cortex, pre and par-
(DIE), ventral-intermediate (VIE), caudal (CE), asubiculum. The general orientation of entorhinal
and medial (ME) subdivisions (Insausti et al., fibers within hippocampal subfields is parallel to the
1997; van Groen et al., 2003). Initially, we included principal cell layers, and it has been described that
a sixth domain called amygdalo-entorhinal cortex entorhinal fibers in the molecular layer of the CA-
(AE) as part of the entorhinal cortex because our fields continue around the tip of the enclosed blade
own tracing data indicated that this region also of the dentate gyrus to enter the dentate molecular
contributed to the projections to the dentate gyrus layer (Ramón y Cajal, 1911; Witter, 1989).
(Insausti et al., 1997). In a recent and more de-
tailed analysis of the efferent connectivity of this
region, it was shown that it is quite unlikely that Layers of origin in the entorhinal cortex
this so-called area AE indeed contributes projec-
tions to the dentate gyrus (Kemppainen et al., In rat, mouse, or monkey, the ipsilateral projection
2002; Majak and Pitkanen, 2003). Therefore, we to the dentate gyrus appears to arise mainly, if not
exclude this region in the present description. exclusively from layer II of the entorhinal cortex
In monkeys, humans, and in other species in (Steward and Scoville, 1976; Schwartz and
which the entorhinal cortex was described, such as Coleman, 1981; Ruth et al., 1982, 1988; Witter
cat, dog, and guinea pig (Krettek and Price, 1977; et al., 1989b; van Groen et al., 2003; Chrobak and
Amaral et al., 1987; Witter et al., 1989a; Insausti et Amaral, 2006). In humans, fetal material indicates a
al., 1995; Uva et al., 2004; Woznicka et al., 2006) similar origin (Hevner and Kinney, 1996). It is of
comparable partitioning schemes into multiple interest to note that in the brain of Alzheimer pa-
subdivisions of the entorhinal cortex have been tients, the entorhinal cortex layer II neurons are
proposed. However, in those species there is also a among the ones preferentially implicated in the dis-
tendency to consider the entorhinal cortex as com- ease, such that up to 50% of those neurons appar-
posed of two primary components, most likely re- ently disappear. This rather selective loss of layer II
flecting functional differences. neurons has been associated with changes in the
For the present description we will therefore outer part of the molecular layer, such as a decrease
maintain a subdivision of the entorhinal cortex in free glutamate and a decrease in markers asso-
into two main subdivisions, the lateral (LEC) and ciated with the perforant path, suggesting that in
medial (MEC) entorhinal cortices respectively. humans also the origin, and obviously termination,
Studies of the perforant pathway projection to of the entorhinal-dentate projection is similar to that
the dentate gyrus are most numerous and most reported in other species (Hyman et al., 1986, 1987,
detailed in the rat. Therefore, the descriptions 1988; Morys et al., 1994). There is evidence that a
will be initially provided for material gathered minor component of the projection to the dentate
from experiments in rats. When relevant, data gyrus also comes from the deep layers (IV–VI) of
from other species will be added in a comparative the entorhinal cortex (Köhler, 1985; Witter and
fashion. Amaral, 1991; van Groen et al., 2003). In rat and
47

monkey, neurons in layer II not only project to of the entorhinal cortex, aside from overall cyto-
the dentate gyrus but also to CA3 (Steward and architectonic differences, was based on observa-
Scoville, 1976; Witter and Amaral, 1991). Intra- tions that fibers originating in the LEC terminate
cellular tracing in the rat revealed that a single layer in the outer one-third of the molecular layer and
II cell may innervate not only the dentate gyrus and fibers from the MEC terminate in the middle one-
CA3 but also the subiculum (Tamamaki and Nojyo, third of the molecular layer (Hjorth-Simonsen and
1993), but whether this holds true for the monkey as Jeune, 1972; Hjorth-Simonsen, 1972). This initial
well is currently unknown. In contrast, in the idea was strengthened further by the striking la-
mouse, at least in one of the strains (C57BL/6J), minar staining pattern that was revealed by Timm
layer II cells appear to project only to the dentate stain for heavy metals, (Haug, 1976; Stanfield and
gyrus and does not extend collaterals to other hip- Cowan, 1979) and by immunocytochemistry using
pocampal subfields (van Groen et al., 2003). Al- antibodies against enkephalin and cholecystokinin
though this appears a striking species difference, (Fredens et al., 1984).
emphasized by several authors, it is currently not Although most consider a non-overlapping dis-
known whether this is a general phenomenon in all tribution in the dentate of LEC and MEC fibers, it
mouse strains or typical for this one strain. More- is important to note that there may be a contin-
over, our own material has indicated that in the rat uum in these fiber systems. This was originally
layer III cells contribute to the projection to CA3. suggested on the basis of experiments with antero-
As illustrated in Fig. 1A, a ventrally positioned in- grade transport of tritiated amino acids by Steward
jection in layers II and III of CE (case 88246) results (1976), who suggested that the perforant path
in strong labeling of the dentate gyrus, all CA fields should actually be subdivided into lateral, interme-
and the subiculum. In contrast, a similarly posi- diate, and medial components. The intermediate
tioned injection confined to layer III (case 89339) pathway was thought to originate in the inter-
does not result in any labeling in the dentate gyrus, mediate entorhinal area, an area most likely
as expected on the basis of the selective origin of the including VIE and ME, and to terminate in the
dentate component of the perforant path in layer II. molecular layer between the projections from LEC
However, in this case we did find some labeling in and MEC. In an elegant series of anterograde
CA3, as well as strong labeling in CA1 and the tracing experiments using small injections of triti-
subiculum (Fig. 1B, D). Likewise, this can be ob- ated amino acids in rats, Wyss (1981) reported that
served in case of an injection in layer III of VIE there is a striking radial gradient in the terminal
(case 88178; Fig. 1C, E). distribution of entorhinal fibers depending on the
On the basis of these data, it thus seems safe to position of the source in the entorhinal cortex.
conclude that in all species, layer II is the pre- Lateral portions of LEC project close to the pial
dominant if not exclusive source of the entorhinal surface and successively more medial portions
projections to the outer two-thirds of the molec- project closer to the granule cells. However, the
ular layer of the ipsilateral dentate gyrus (cf. data from this study also supported an intermedi-
Kunzle, 2002). In addition, these data indicate that ate perforant pathway, distributing to the border
in the rat, the projection to CA3 originates pre- region between the typical lateral and medial per-
dominantly in layer II with a minor component forant path terminal zones in the outer and middle
arising from layer III. The layer III projection to one-thirds of the molecular layer, respectively. A
area CA3 is apparently more prominent in mice. comparable conclusion was reached on the basis of
a study using very small injections of the ante-
rogradely transported tracer DiI in portions of the
Radial or layered terminal organization in the entorhinal cortex close to the rhinal fissure, i.e.
dentate gyrus projecting to the dorsal part of the dentate gyrus
(Tamamaki, 1997). When looking at a summary
One of the more convincing arguments to differ- figure taken from this particular study (Fig. 2),
entiate between the lateral and medial subdivisions two conclusions are apparent. First, the radial
48

Fig. 1. Layer II and to a lesser extent layer III contribute to the projections to CA3. Sections taken from unfolded preparations
(Gaarskjaer, 1978), stained for the presence of BDA and counterstained for Nissl-substances with cresyl violet. (A) Section from a case
with an injection centered in layers II and III of MEC (Fig. 1, case 89246). Strong labeling is present in the dentate gyrus, as well as in
CA3 (and CA1 and subiculum). (B) Section from a case with an injection of which the location corresponds to that illustrated in A, but
no tracer has been taken up by layer II neurons (case 88339). Note the absence of labeling in the dentate gyrus, strong labeling in CA1
and subiculum but also weak labeling in CA3 (boxed area shown in D). (C) Section from a case with an injection in layer III of LEC
(Fig. 2, case 88178). Similar to the case shown in B, labeling is present in CA1 and subiculum, absent in dentate gyrus, but weak
labeling can be seen in CA3 (boxed area shown in E). (D) Higher magnification of boxed area marked in B, showing the light terminal
plexus in CA3. (E) Higher magnification of boxed area marked in C, showing the light terminal plexus in CA3. Scale bar in A equals
1 mm; also holds for B and C. Scale bars in D and E equal 100 mm.

extent of the labeled terminal field in the dentate cells in layer II of the medial entorhinal cortex
gyrus almost never covers one-third of the molec- which projected to the dorsal part of the dentate
ular layer: one-fifth to one-sixth would be a more gyrus revealed yet another pattern, where the axon
appropriate description. Second, the terminal is confined to the middle one-third in the enclosed
fields show a gradual transition from a very distal blade but extends into the outer one-third of the
to a more proximal position along the apical exposed blade, i.e. the terminal zone becomes
dendrites of dentate granule cells. These observa- thicker in the exposed blade (Tamamaki and
tions led to the proposition that the entorhinal- Nojyo, 1993). This finding was supported by the
dentate projection is a continuum rather than results of anterograde tracing using DiI. Injection
subdivided into two or three discrete components. of DiI into lateral entorhinal cortex led to a thick
Interestingly, analysis of intracellularly labeled band of terminal labeling in the enclosed blade but
49

Fig. 2. Confocal microphotographs showing anterograde labeling in the enclosed blade of the dentate gyrus following a series of
anterograde tracer injections (DiI) along the rostro-caudal extent of the rhinal fissure. A is taken from the most anterior injection,
which is in LEC, F is taken from the most posterior position, which is in MEC. Cases D and E are most likely along the border between
LEC and MEC, although this has not been assessed in the original paper. Arrows demarcate laminar boundaries. Scale bar equals
100 mm. Adapted with permission from Tamamaki (1997).

comparatively thin labeling in the exposed blade. of their corresponding one-third of the molecular
Following DiI injections in the medial entorhinal layer.
cortex, the reverse pattern was observed. Our own What remains to be discussed is whether or not
data, based on visualizing the projections with the perforant path is organized as a continuum, as
PHA-L and BDA (Witter, 1989, 1990), to a large proposed by Tamamaki (1997), whether it com-
extent corroborates the observations reported by prises two non-overlapping systems — lateral and
Tamamaki and Wyss (Wyss, 1981; Tamamaki, medial, or whether there are three components —
1997). Whereas large injections generally label the lateral, intermediate and medial. To address this
full width of the outer or middle one-third of the point, we compared the radial distribution of an-
dentate molecular layer, small injections never re- terogradely labeled fibers following two closely
sult in consistent labeling throughout the entire spaced injections, centered around the border be-
outer or middle one-third. In these cases, the labe- tween VIE and ME in the rat (Figs. 3 and 4). The
led domain is closer to one-fifth or one-sixth of the injection in ME (case 88246), which is part of
entire width of the molecular layer, similar to the MEC, clearly produces a narrow terminal field in a
observations of Tamamaki (1997). There is a weak narrow distal part of the middle one-third of the
topographical arrangement emerging from these molecular layer (Fig. 3A). In contrast, a spatially
experiments. Injections in the lateral part of LEA close injection, but in VIE (case 88334), which is
or dorsolateral part of MEA more densely target part of LEC, labels a narrow terminal field in a
the more distal portions of the dendrite in their proximal part of the outer one-third of the molec-
respective terminal zone, whereas more ventrally ular layer (Fig. 3B). One can easily imagine that a
and medially positioned injections tend to distrib- slightly larger injection around the border region
ute fibers preferentially to more proximal portions between VIE and ME would result in a terminal
50

Fig. 3. Radial distribution of the entorhinal-dentate projections. (A) High power photomicrograph of a small portion of the enclosed
blade of the dentate gyrus, taken from a case with an injection in MEC. Indicated is the border between the molecular layer and
lacunosum-moleculare of CA1 (white line). Arrows indicate the measures of the width of the molecular layer (a), expressed as 100%
(see also the Y-axis in D), the position of the outer border or the terminal field (b) and that of the inner border (c), as measured from
the outer border of the molecular layer (see also C). (B) High power photomicrograph of a small portion of the enclosed blade of the
dentate gyrus, taken from a case with an injection in LEC. The picture is size-matched and merged with A to illustrate the striking
difference in position of the two terminal fields (see also D). (C) Section with terminal labeling in the dentate gyrus, illustrating the
measuring protocol applied in this study. The center of the line has been derived, and equally distanced lines, perpendicular to this
center-axis, have been generated to measure the position of the terminal field as indicated in A. This procedure is carried out in all
sections with a labeled terminal field. (D) Graphic representation of four representative cases with terminal fields confined to either the
outer 35% (cases 88183R and 88334; see also Fig. 2), or the middle portion between 35 and 66.5%. Exposed and enclosed blades have
been analyzed separately in view of the overall fluctuations in density and position of the terminal fields between the two blades. Note
that two closely positioned injections on either side of the border between LEC and MEC (cases 88334 and 89246 respectively) show
completely non-overlapping, though adjacent, terminal fields. Scale bar in C, D equals 100 mm.
51

Fig. 4. Longitudinal and transverse distribution of the entorhinal-dentate projections. (A) Unfolded representations of the dorso-
ventral extent of the dentate gyrus and terminal labeling resulting from injections in LEC (upper row) and MEC (lower row). (B)
Unfolded representation of the entorhinal cortex with injections in lateral and medial subdivisions. Case 88178 (light grey) is an
injection in layer III, illustrated in Fig. 1C. Abbreviations: AE, amygdalo-entorhinal area; LEC, lateral entorhinal cortex; MEC,
medial entorhinal cortex; PaS, parasubiculum.
52

pattern that is best described as an intermediate was present throughout the extent of the outer
pathway (see for example Fig. 2D, E). A problem two-thirds of the molecular layer. Subtle indica-
with all such types of analyses is that the data are tions for differences between lateral and medial
taken from a single level of a highly divergent perforant pathways were observed however. Ros-
projection. Note that most studies indicated differ- tral parts of EC, most likely homologous to LEC
ences between exposed and enclosed blades (see of the rat, project most densely to the outer one-
above) or between different longitudinal levels. third, whereas projections originating in the most
Therefore, we aimed to assess our terminal distri- caudal parts of the monkey EC, most likely
butions on the basis of all available data collected an area comparable to MEC of the rat, showed
throughout the entire terminal domain. As illus- a preference for the middle one-third. Interest-
trated schematically in Fig. 3C, we calculated the ingly, also the typical differentiation between the
centerline of the terminal field for multiple posi- terminal zones of the two perforant path compo-
tions and subsequently measured the radial thick- nents, as seen with the Timm stain, is largely ab-
ness of the molecular layer, taken perpendicular to sent in the monkey dentate, with the exception of
this centerline (measure a). We also measured the the anterior genu and uncal portion (Witter and
outer (measure b) and inner border (measure c) of Amaral, 1991).
the terminal field at all these radial points through- Taken together, the data do not allow for a
out all sections of the extended dentate gyrus that conclusive statement whether or not the radial or-
showed anterogradely labeled perforant path fib- ganization of the entorhinal-dentate projections
ers (Figs. 4 and 5). For both illustrated experi- supports the differentiation into two discrete lat-
ments, as well for all other experiments illustrated eral and medial components. One question is
throughout this paper, we collected such a popu- whether such distinctions influence function. We
lation of outer and inner border measures. We do know that lateral and medial entorhinal cortex
subsequently established that the outer border most likely convey different types of information
measures for the medial perforant pathway where to the dentate molecular layer, and the most strik-
significantly different from the inner border meas- ing difference seem to relate to spatial modulation,
ures for the lateral perforant pathway (Fig. 3D; with the cells in the medial entorhinal cortex being
Po0.05). more strongly spatially modulated relative to those
At this point, one may wonder what we know in the lateral entorhinal cortex (Hafting et al.,
about this issue in other species. In the mouse, no 2005; Hargreaves et al., 2005). We also know that
detailed studies with respect to the radial differ- these pathways are differently modulated and that
ences in the distribution of the entorhinal projec- the overall postsynaptic effects are different (see
tions to the dentate are available. The only the section on synaptic organization below). More
detailed account in mice seems to support a sep- important, however, is the conclusion that in
aration between fibers originating in LEC and all species, information from functionally differ-
MEC (van Groen et al., 2003). Irrespective of their ent entorhinal domains converges onto a single
origin in either DLE, DIE or VIE, LEC fibers ter- population of dentate granule cells (see section on
minate throughout the extent of the outer one- synaptic organization below).
third of the molecular layer. Fibers arising from
MEC preferentially terminate throughout the mid-
dle one-third, although in this pathway, thinner Transverse organization in the dentate gyrus
terminal zones have been noted, similar to what
has been observed in the rat. In contrast, the sit- There are conflicting papers on the transverse dis-
uation in the macaque monkey is in support of a tribution of the perforant path projection. In ear-
more continuous organization of the entorhinal- lier studies no differences were reported (Hjorth-
dentate projection (Witter et al., 1989b; Witter and Simonsen, 1971, 1972; Hjorth-Simonsen and
Amaral, 1991). Irrespective of the origin in EC, at Jeune, 1972; Steward, 1976). Wyss (1981) reported
all levels of the dentate gyrus innervated, labeling that the lateral perforant pathway preferentially
53

Fig. 5. Representative examples of longitudinal, transverse, and radial terminal distributions of the entorhinal-dentate projections.
(A–C) Injections in LEC. (D–F) Injections in MEC. For details see text.
54

projects to the enclosed blade of the dentate gyrus, the origin of the projection, when labeling is
whereas the medial component either does not present in the dentate molecular layer, it appears
show a preference or predominantly targets the to be distributed equally along both blades (Witter
exposed blade. It should be stressed, however, that et al., 1989b; van Groen et al., 2003).
dimensions of injection sites may lead to different In conclusion, the entorhinal-dentate projection
results, similar to what was described with respect appears to distribute along the transverse axis of
to the radial organization. The findings of Wyss the dentate gyrus in a homogeneous fashion. Al-
were supported by the observation that the termi- though a subtle topographical arrangement cannot
nal zone of LEC fibers is wider in the enclosed be excluded, and individual cells may differ in their
blade, covering almost the entire outer one-third particular target projection, there is no obvious
of the enclosed blade, whereas the terminal field in pattern. It is worth to emphasize though that con-
the exposed blade is much thinner (Tamamaki, vergent evidence points to functional differences
1997). For the MEC, the reverse appears to be the between the enclosed and exposed blades of the
case. Importantly, individual neurons show exten- dentate gyrus. For example in guinea pigs, cells in
sive variation in their target projection, demon- the enclosed blade as well as the associated hilar
strated by the axon arbors of layer II cells that mossy cells show morphological sex differences
were intracellularly labeled in vivo. An individual whereas such differences have not been observed
layer II cell may send axon collaterals along the in the exposed blade and associated hilar mossy
entire transverse extent of the dentate gyrus, or cells (Bartesaghi et al., 2003; Guidi et al., 2006).
project preferentially to either one of the two Differences between the two blades have further
blades. These data are important because they been reported with respect to sensitivity for hypo-
suggest that the EC projections as a whole may xia (Hara et al., 1990), neurogenesis (Choi et al.,
provide homogeneous innervation of the dentate, 2003), efficacy in activating hippocampal circuits
but individual components could have specific and (Scharfman et al., 2002), and loss of granule cells
selective functional influence. Indeed, using the following adrenalectomy (Jaarsma et al., 1992);
extended preparations and subsequent unfolding, (see also Chawla et al., this volume).
we noticed no systematic organization in a total of
21 experiments. In most cases, when in the core of
the projection, labeling was present in both blades, Longitudinal organization of entorhinal-dentate
although it was unequal in density in some in- projections
stances (Figs. 4 and 5). In case of injections in
LEC, we generally observed, in the dorsal and Entorhinal projections show a striking organiza-
ventral domains of the terminal field, a clear pref- tion along the longitudinal axis of the dentate
erence for the enclosed blade, whereas even in the gyrus. Originally described in the cat (Witter and
core, labeling was denser in the enclosed blade Groenewegen, 1984), it was later discovered to be
(Fig. 5A–C). In particular, more laterally placed a general governing principle in all species studied
injections in LEC (cases 90170, 881813R, 89156) (Ruth et al., 1982, 1988; Witter et al., 1989b;
showed extensive labeling in the dorsal part of the Dolorfo and Amaral, 1998; van Groen et al.,
enclosed blade, exclusively (Fig. 4, top row), in line 2003). In the rat, cells located laterally in the ent-
with the observations published by Wyss (1981). orhinal cortex project to dorsal parts of the dent-
Regarding the projections originating from MEC ate gyrus while cells located progressively more
(Fig. 4, lower row; Fig. 5D–F) no striking prefer- medially project to more ventral levels of the dent-
ence for either blade was apparent although differ- ate gyrus (Fig. 4). This organization leads to a
ences in density do occur between experiments pattern of connections such that the dorsal part of
(compare Fig. 5D, E and F) as well as within ex- the dentate gyrus receive inputs from lateral parts
periments (Fig. 5E). of the LEC and lateral and caudal parts of MEC,
In the mouse and the monkey, no transverse whereas the ventral portions of the dentate gyrus
organization has been described. Irrespective of receive input from more medial portions of both
55

LEC and MEC (Ruth et al., 1982, 1988; Witter labeling in the intermediate levels of the hippo-
et al., 1989a, 2000). This topographical organization campus, the longitudinal extent is generally larger
has been convincingly demonstrated in a series of (Fig. 4, cases 88183R, 89156, 89247, and 89246).
retrograde tracing experiments, in which discrete The longitudinal organization of the entorhinal-
injections in dorsal, mid-dorsoventral, and ventral dentate projection in the mouse, cat, and monkey
levels of the dentate gyrus resulted in labeled pop- is strikingly comparable. In these species again,
ulations of entorhinal neurons, in both LEC and lateral and caudal parts of the entorhinal cortex,
MEC, with a different lateral to medial position encompassing a band that parallels the lateral
(Dolorfo and Amaral, 1998). Although the do- border between entorhinal and perirhinal/parahip-
mains of the entorhinal cells projecting to these pocampal cortex and thus including parts of both
different longitudinal levels do not show much the lateral and medial entorhinal cortex preferen-
overlap, one should take into account that a single tially project to the dorsal (posterior) dentate
entorhinal layer II neuron, as shown with intra- gyrus. Progressively more medial portions of EC,
cellular filling, may distribute an axon along as again encompassing portions of both LEC and
much as 2 mm (20–25%) of the dorsoventral ex- MEC project to increasingly more ventral (ante-
tent (Tamamaki and Nojyo, 1993) and that re- rior) portions of the dentate gyrus (Witter et al.,
stricted injections in EC in almost all cases result 1989a, b; van Groen et al., 2003).
in labeling extending over at least one-half to up to We may thus conclude that of all the principles
two-thirds of the long axis (Witter et al., 1989a; governing the organization of the perforant path-
Tamamaki, 1997), such that overlapping terminal way projection to the dentate gyrus, the longitu-
zones in the dentate gyrus of these populations of dinal one is most reliable throughout the animal
layer II neurons are likely to exist. This widespread kingdom. Whether this also holds true for the hu-
distribution along the long axis demonstrated by man is currently unknown but it seems plausible
anterograde tracing seems at odds with the rather given the number of studies supporting functional
restrictive labeling in the entorhinal cortex result- differences along the long axis that would be likely
ing from retrograde tracing methods. One may to reflect this longitudinal topography. It is clear,
therefore pose the question whether this longitu- however, that further research is needed to fully
dinal topographical organization actually holds understand and assess the potential functional im-
true (Tamamaki, 1997). However, it is essential to plications of these findings. Complicating factors
take into account the effect of injection size, which are the strong longitudinal associational pathways
quite often is not easily determined in anatomical present within several of the hippocampal subfields
tracing experiments. Most data support the idea (Witter and Amaral, 2004). Since we do not yet
that small foci of entorhinal layer II cells adhere to appreciate the functional relevance of those strong
the overall 25% longitudinal axonal distribution, longitudinal associational connections, functional
as described on the basis of single cell axonal dis- differences along the long axis may be hard to as-
tributions. sess.
The unfolded approach unmasks another fea-
ture of this projection that has not been appreci-
ated before. Irrespective of the size of the injection, Contralateral projections
the position of the origin in the entorhinal cortex,
and thus the overall terminal distribution along In the rat, the entorhinal cortex projects to the
the longitudinal axis, appears critical. In case of an ipsilateral dentate gyrus, and also gives rise to a
injection in the entorhinal cortex, which strongly crossed projection to the contralateral dentate
labels the extreme dorsal or ventral tips of the gyrus, as well as the contralateral CA3 and CA1
dentate gyrus (Fig. 4, cases 90170, 88334, 89361, subfields, and the subiculum. The crossed ent-
and 88400), the terminal fields do not extend over orhinal projection is most prominent to the more
more than about 20–25% of the long axis. In con- dorsal portions of the hippocampal subfields and
trast, following injections that produce strong rapidly diminishes in density at more temporal
56

levels (Goldowitz et al., 1975; Steward, 1976). known to distribute local axonal plexi to the per-
With respect to the laminar origin of the crossed forant path terminal zone (Bakst et al., 1985, 1986;
projections, Steward and Scoville (Steward and Halasy and Somogyi, 1993; Boyett and Buckmaster,
Scoville, 1976) reported that it matches the ipsi- 2001). It is most likely that the so-called molecular
lateral origin. The crossed dentate projection, that layer perforant path-associated cells (MOPP) as well
thus originates from layer II cells, mainly takes the as the hilar commissural-associational pathway re-
more common perforant path trajectory, i.e. cross- lated cells (HICAP)(Han et al., 1993) are also
ing the midline through the ventral hippocampal among the postsynaptic targets of entorhinal axons,
commissure. In the mouse, almost no crossed pro- but this remains to be established. Note that in
jection to the dentate gyrus have been observed, in monkeys, NPY-positive cells in the hilus do not
contrast to the rat (van Groen et al., 2003). In extend dendrites into the molecular layer (Nitsch
rabbit and cat, a distinct projection to the cont- and Leranth, 1991). Whether this implies a species
ralateral DG, mainly limited to the dorsal part, difference in connectivity or in the expression of
has been described, similar to that in the rat certain proteins is not clear.
(Goldowitz et al., 1975; Hjorth-Simonsen and In addition to the main innervation arising from
Zimmer, 1975; Steward and Scoville, 1976; Wyss, layer II cells in the entorhinal cortex, a projection
1981). In the monkey, this projection is very mod- originating from deep layers has been described
est, limited to the uncal/genu portion only (Witter above (Köhler, 1985). This projection preferen-
and Amaral, 1991). tially distributes to the inner portion of the mo-
lecular layer, the granule cell layer, as well as the
subgranular zone, where it establishes asymmetri-
Synaptic organization of the entorhinal-dentate cal synapses onto granule cell dendrites as well as
projection on their somata and onto spine-free dendrites in
the subgranular zone. The latter most likely rep-
In the molecular layer of the dentate gyrus in the resent dendrites of interneurons (Deller et al.,
rat, the terminals of the perforant path fibers in the 1996).
outer two-thirds of the molecular layer make up at The observation that most of the entorhinal
least 85% of the total synaptic population (Nafstad, synapses in the dentate gyrus are asymmetric
1967; Matthews et al., 1976). Entorhinal fibers form strongly suggests that the pathway is largely exci-
mainly, if not exclusively, asymmetric synapses tatory, most likely using glutamate as primary
(Nafstad, 1967; Matthews et al., 1976; Deller and transmitter (White et al., 1977). Although not
Leranth, 1990; Leranth et al., 1990). These occur studied in detail, prominent differences in these
most frequently on the dendritic spines of dentate respects between the lateral and medial perforant
granule cells, although a small proportion of perfo- pathway are unlikely (Nafstad, 1967; Matthews et
rant path fibers terminate on non-spiny dendrites of al., 1976), similar to observations recently reported
presumed interneurons. These include parvalbumin/ for the lateral and medial entorhinal projections to
GABA-immunoreactive (Zipp et al., 1989), som- the subiculum (Baks-Te-Bulte et al., 2005). How-
atostatin-positive, and NPY-positive neurons with ever, regarding the chemical characteristics, differ-
cell bodies located in the hilus and apical dendrites ences between the two pathways have been
which extend into the outer portions of the molec- described. Terminals of the lateral perforant path-
ular layer (Scharfman, 1991; Soriano and Frotscher, way are also enkephalin immunoreactive, whereas
1993). Since at least half of the NPY-positive neu- those of the medial pathway are immunoreactive
rons also stain for somatostatin (Swanson and for CCK and dynorphin (Fredens et al., 1984; van
Köhler, 1986) and some co-localization of NPY and Abeelen, 1989). Medial perforant path fibers are
parvalbumin has been reported (Deller and Leranth, immunoreactive for the metabotropic glutamate
1990), it cannot be excluded that what may seem receptor, mGLUR 2/3, whereas the lateral fibers
as three different population of interneuronal tar- are not. There is also convincing evidence that
gets, may in fact be the population of interneurons neurons in LEC and MEC that project to the
57

dentate gyrus are markedly different with respect suggest that they play a minor role in the ent-
to their electrophysiological properties (van Der orhinal input to the dentate gyrus compared to the
Linden and Lopes da Silva, 1998; Wang and Lam- primary projection from layer II neurons.
bert, 2003). Surprisingly, there is, as yet, no dis- Although there is converging evidence that the
tinctive marker at the level of the entorhinal cortex entorhinal-dentate projection can be subdivided
for cells that give rise to the lateral and medial into two, functionally different systems, the precise
perforant path projections. spatial relationship between those two systems and
In view of the extensive focus on whether or not the radial terminal distribution in the molecular
the two pathways predominantly target different layer of the dentate gyrus is not entirely clear at
segments of granule cell dendrites (see section on the present time. Although generally accepted in
radial organization above), there has been surpris- rodents, and quite often used as a model system to
ingly less interest in the question whether or not it understand mechanisms for lamina-specific axon
is a general feature for lateral and medial perforant outgrowth (see chapter by M. Frotscher in this
path fibers to converge onto a single granule cell. volume), this radial differentiation is less apparent
Although this seems likely in view of the overall in the monkey. Taken together with the other spe-
‘‘en passant’’ type of termination of both pathways cies differences in the two pathways discussed
and the fact that both indeed distribute along the above, the question can be raised whether these
entire transverse extend of the dentate gyrus, thus species differences in the topographic organization
increasing the likelihood that each dentate granule of the perforant path impart functional differences
cells receives convergent input from multiple ent- to distinct species. Physiological differences would
orhinal layer II cells, there is no strong anatomical be expected if the organization of the projections is
data to support the idea of convergence. In view of not the same, given the evidence that the lateral
the surprising observation of an apparent lack of and medial pathways differentially affect dentate
convergence in similarly organized pathways in the activity upon stimulation. Moreover, the effects of
subiculum (Cappaert et al., 2005), this issue of either pathway on dentate activity can be manip-
convergence at the single cell level remains to be ulated with a number of pharmacological tools
resolved. that appear selective for only one or the other (see
chapter by Bramham in this volume).
A critical question that seems yet to be ad-
Summary and functional comments dressed is whether or not the position of a single
entorhinal synapse along the proximo-distal extent
The main origin of the projection from the ent- of the apical dendrite of a granule cells is a nec-
orhinal cortex to the dentate gyrus is throughout essary element to explain the function of that in-
layer II cells in EC. Entorhinal axons of layer II put. Furthermore, it is not clear whether fibers
cells have their terminals predominantly, if not from LEC and MEC must target individual gran-
exclusively, in the outer two-thirds of the molec- ule cells in specific convergent patterns in order to
ular layer of the dentate gyrus. An additional pro- transfer information from entorhinal cortex nor-
jection arises from cells in deeper layers of the mally. In view of the primate organization, this
entorhinal cortex, mainly deep V and VI. In the seems an unlikely proposition. Surprisingly, there
rat, and likely in the mouse as well, these deeply appears no solid anatomical evidence for the gen-
located cells appear to be the source of most if not erally accepted view that all granule cells receive
all of the perforant path fibers that distribute out- convergent inputs from both pathways.
side the ‘‘traditional’’ terminal zone in the outer No clear conclusions can be derived with respect
molecular layer, showing a more dispersed termi- to a transverse organization of this projection. From
nal distribution in the hilar region, inner molecular this overview, the generally accepted view that there
and granular layers of the DG. However, these is no transverse organization appears acceptable. In
minor projections have received little emphasis, contrast, the entorhinal-dentate system shows a
and remain largely unexplored, and most would striking organization along the longitudinal axis of
58

the dentate gyrus such that lateral and caudal por- topographical issues described here, we have to un-
tions of the entorhinal cortex project preferentially derstand the unique, and most likely complemen-
to the dorsal (posterior in primates) part of the tary, contributions of each subfield of the
dentate gyrus, whereas more medial and anterior hippocampus and the entorhinal cortex, as well as
portions of the entorhinal cortex projects preferen- the roles of the longitudinal associational connec-
tially to more ventral (anterior in primates) portions tions.
of the dentate gyrus. It is of interest that the pro-
jections to the dorsal (posterior) part generally show
a much more dispersed terminal distribution, up to Acknowledgments
almost 60% of the entire length, whereas those to
the ventral (anterior) dentate appear more restricted This paper is based on large quantities of work
to maximally 30% of the long axis. This appears to carried out by many colleagues over the years. I am
be in correspondence to the differential longitudinal greatly indebted to them and their well-prepared
spread of the intrinsic CA3 associative system. This publications on this subject. Moreover, for the ex-
clear cut organization along the long axis has trig- perimental data described, I owe my former tech-
gered numerous experiments trying to relate differ- nician, Mrs. B. Jorritsma-Byham, who performed
ences in inputs to the different lateral-to-medial most of the analyses, assisted by a sizable group of
bands in entorhinal cortex to a functional differen- under-graduate students. Dr. G. Doctor developed
tiation along the long axis of the hippocampus, as the statistical routines and performed the measure-
initially proposed (Witter et al., 1989a). It has been ments on the radial position of terminal fields.
established in rats that dorsal hippocampal lesions Finally, I want to thank Helen Scharfman for her
selectively interfere with certain forms of spatial valuable discussions and contributions to the man-
learning and memory (Moser and Moser, 1998). uscript. This work has been supported in part by
Likewise, lesions of the related entorhinal input grants from HFSPO and the Dutch Organization
zone produces spatial deficits (Steffenach et al., for Scientific Research (NWO).
2005). In contrast, ventral hippocampal lesions do
not result in clear spatial deficits but lead to more
motivational deficits (Kjelstrup et al., 2002) and References
again comparable behavioral effects have been re-
van Abeelen, J.H. (1989) Genetic control of hippocampal
ported to result from lesions of the corresponding cholinergic and dynorphinergic mechanisms regulating nov-
entorhinal input zone (Steffenach et al., 2005). In elty-induced exploratory behavior in house mice. Experientia,
non-human primates this issue has not been ad- 45(9): 839–845.
dressed in any detail, but human functional imaging Amaral, D.G., Insausti, R. and Cowan, W.M. (1987) The ent-
experiments indeed did report functional differences orhinal cortex of the monkey. I. Cytoarchitectonic organiza-
tion. J. Comp. Neurol., 264(3): 326–355.
along the longitudinal axis. However, these exper- Amaral, D.G. and Witter, M.P. (1989) The three-dimensional
imental data mainly support differences in encoding organization of the hippocampal formation: a review of an-
and retrieval processes and as such are not easily atomical data. Neuroscience, 31(3): 571–591.
comparable to the rodent data. In order to reconcile Bakst, I., Avendano, C., Morrison, J.H. and Amaral, D.G.
this current apparent mismatch between animal and (1986) An experimental analysis of the origins of somatosta-
tin-like immunoreactivity in the dentate gyrus of the rat. J.
human data, we most likely have to take the strong Neurosci., 6(5): 1452–1462.
intrinsic wiring of both the hippocampus and the Bakst, I., Morrison, J.H. and Amaral, D.G. (1985) The distri-
entorhinal cortex into account (Witter and Amaral, bution of somatostatin-like immunoreactivity in the monkey
2004; Witter and Moser, 2006). In particular the hippocampal formation. J. Comp. Neurol., 236(4): 423–442.
Baks-Te-Bulte, L., Wouterlood, F.G., Vinkenoog, M. and Wit-
intrinsic interactions of the hippocampus, both
ter, M.P. (2005) Entorhinal projections terminate onto prin-
along the transverse and the long axis may compli- cipal neurons and interneurons in the subiculum: a
cate our understanding of this issue. In order to be quantitative electron microscopical analysis in the rat.
able to value the relevance of the organizational and Neuroscience, 136(3): 729–739.
59

Bartesaghi, R., Guidi, S., Severi, S., Contestabile, A. and Ciani, related to number of granule and pyramidal cells. J. Comp.
E. (2003) Sex differences in the hippocampal dentate gyrus of Neurol., 178(1): 49–72.
the guinea-pig before puberty. Neuroscience, 121(2): 327–339. Goldowitz, D., White, W.F., Steward, O., Lynch, G. and Cot-
Blackstad, T.W. (1958) On the termination of some afferents to man, C. (1975) Anatomical evidence for a projection from the
the hippocampus and fascia dentata. An experimental study entorhinal cortex to the contralateral dentate gyrus of the rat.
in the rat. Acta Anat., 35: 202–214. Exp. Neurol., 47(3): 433–441.
Boyett, J.M. and Buckmaster, P.S. (2001) Somatostatin- van Groen, T., Miettinen, P. and Kadish, I. (2003) The ent-
immunoreactive interneurons contribute to lateral inhibitory orhinal cortex of the mouse: organization of the projection to
circuits in the dentate gyrus of control and epileptic rats. the hippocampal formation. Hippocampus, 13(1): 133–149.
Hippocampus, 11(4): 418–422. Guidi, S., Severi, S., Ciani, E. and Bartesaghi, R. (2006) Sex
Cappaert, N.L.M., Wadman, W.J. and Witter, M.P. (2005) differences in the hilar mossy cells of the guinea-pig before
Spatiotemporal analyses of interactions between entorhinal puberty. Neuroscience, 139(2): 565–576.
and CA1 projections to the subiculum of the rat. Soc. Ne- Hafting, T., Fyhn, M., Molden, S., Moser, M.B. and Moser,
urosci. Abstr., 73.13. E.I. (2005) Microstructure of a spatial map in the entorhinal
Choi, Y.S., Lee, M.Y., Sung, K.W., Jeong, S.W., Choi, J.S., cortex. Nature, 436(7052): 801–806.
Park, H.J., Kim, O.N., Lee, S.B. and Kim, S.Y. (2003) Re- Halasy, K. and Somogyi, P. (1993) Subdivisions in the multiple
gional differences in enhanced neurogenesis in the dentate GABAergic innervation of granule cells in the dentate gyrus
gyrus of adult rats after transient forebrain ischemia. Mol. of the rat hippocampus. Eur. J. Neurosci., 5(5): 411–429.
Cells, 16(2): 232–238. Han, Z.S., Buhl, E.H., Lorinczi, Z. and Somogyi, P. (1993) A
Chrobak, J.J. and Amaral, D.G. (2006) The entorhinal cortex high degree of spatial selectivity in the axonal and dendritic
of the monkey. VII. Intrinsic connections. J. Comp. Neurol., domains of physiologically identified local-circuit neurons in
500(4): 612–633. the dentate gyrus of the rat hippocampus. Eur. J. Neurosci.,
Deller, T. (1998) The anatomical organization of the rat fascia 5(5): 395–410.
dentate — new aspects of laminar organization as revealed by Hara, H., Onodera, H., Kogure, K. and Akaike, N. (1990) The
anterograde tracing with Phaseolus vulgaris-leucoagglutinin regional difference of neuronal susceptibility in the dentate
(Pha-L). Anat. Embryol., 197(2): 89–103. gyrus to hypoxia. Neurosci. Lett., 115(2–3): 189–194.
Deller, T. and Leranth, C. (1990) Synaptic connections of ne- Hargreaves, E.L., Rao, G., Lee, I. and Knierim, J.J. (2005)
uropeptide Y (NPY) immunoreactive neurons in the hilar Major dissociation between medial and lateral entorhinal in-
area of the rat hippocampus. J. Comp. Neurol., 300(3): put to dorsal hippocampus. Science, 308(5729): 1792–1794.
433–447. Haug, F.M. (1976) Sulphide silver pattern and cytoarchitec-
Deller, T., Martinez, A., Nitsch, R. and Frotscher, M. (1996) A tonics of parahippocampal areas in the rat. Special reference
novel entorhinal projection to the rat dentate gyrus: direct to the subdivision of area entorhinalis (area 28) and its de-
innervation of proximal dendrites and cell bodies of granule marcation from the pyriform cortex. Adv. Anat. Embryol.
cells and GABAergic neurons. J. Neurosci., 16(10): Cell Biol., 52(4): 3–73.
3322–3333. Hevner, R.F. and Kinney, H.C. (1996) Reciprocal entorhinal-
van Der Linden, S. and Lopes da Silva, F.H. (1998) Compar- hippocampal connections established by human fetal mid-
ison of the electrophysiology and morphology of layers III gestation. J. Comp. Neurol., 372(3): 384–394.
and II neurons of the rat medial entorhinal cortex in vitro. Hjorth-Simonsen, A. (1971) Hippocampal efferents to the ipsi-
Eur. J. Neurosci., 10(4): 1479–1489. lateral entorhinal area: an experimental study in the rat. J.
Desmond, N.L., Scott, C.A., Jane, J.A. and Levy, W.B. (1994) Comp. Neurol., 142(4): 417–437.
Ultrastructural identification of entorhinal cortical synapses Hjorth-Simonsen, A. (1972) Projection of the lateral part of the
in CA1 stratum lacunosum-moleculare of the rat. Hippo- entorhinal area to the hippocampus and fascia dentata. J.
campus, 4(5): 594–600. Comp. Neurol., 146(2): 219–232.
Dolorfo, C.L. and Amaral, D.G. (1998) Entorhinal cortex of Hjorth-Simonsen, A. and Jeune, B. (1972) Origin and termina-
the rat: topographic organization of the cells of origin of the tion of the hippocampal perforant path in the rat studied by
perforant path projection to the dentate gyrus. J. Comp. silver impregnation. J. Comp. Neurol., 144(2): 215–232.
Neurol., 398(1): 25–48. Hjorth-Simonsen, A. and Zimmer, J. (1975) Crossed pathways
Fredens, K., Stengaard-Pedersen, K. and Larsson, L.I. (1984) from the entorhinal area to the fascia dentata. I. Normal in
Localization of enkephalin and cholecystokinin immunore- rabbits. J. Comp. Neurol., 161(1): 57–70.
activities in the perforant path terminal fields of the rat hip- Hyman, B.T., Kromer, L.J. and Van Hoesen, G.W. (1988) A
pocampal formation. Brain Res., 304(2): 255–263. direct demonstration of the perforant pathway terminal zone
Fyhn, M., Molden, S., Witter, M.P., Moser, E.I. and Moser, in Alzheimer’s disease using the monoclonal antibody Alz-50.
M.B. (2004) Spatial representation in the entorhinal cortex. Brain Res., 450(1–2): 392–397.
Science, 305(5688): 1258–1264. Hyman, B.T., Van Hoesen, G.W. and Damasio, A.R. (1987)
Gaarskjaer, F.B. (1978) Organization of the mossy fiber system Alzheimer’s disease: glutamate depletion in the hippocampal
of the rat studied in extended hippocampi. I. Terminal area perforant pathway zone. Ann. Neurol., 22(1): 37–40.
60

Hyman, B.T., Van Hoesen, G.W., Kromer, L.J. and Damasio, Morys, J., Sadowski, M., Barcikowska, M., Maciejewska, B.
A.R. (1986) Perforant pathway changes and the memory and Narkiewicz, O. (1994) The second layer neurones of
impairment of Alzheimer’s disease. Ann. Neurol., 20(4): the entorhinal cortex and the perforant path in physiologi-
472–481. cal ageing and Alzheimer’s disease. Acta Neurobiol. Exp.
Insausti, R., Herrero, M.T. and Witter, M.P. (1997) Entorhinal (Wars.), 54(1): 47–53.
cortex of the rat: cytoarchitectonic subdivisions and the or- Moser, E., Moser, M.B. and Andersen, P. (1993) Spatial learn-
igin and distribution of cortical efferents. Hippocampus, 7(2): ing impairment parallels the magnitude of dorsal hippocam-
146–183. pal lesions, but is hardly present following ventral lesions. J.
Insausti, R., Tunon, T., Sobreviela, T., Insausti, A.M. and Neurosci., 13(9): 3916–3925.
Gonzalo, L.M. (1995) The human entorhinal cortex: a cyto- Moser, M.B. and Moser, E.I. (1998) Functional differentiation
architectonic analysis. J. Comp. Neurol., 355(2): 171–198. in the hippocampus. Hippocampus, 8(6): 608–619.
Ishizuka, N. (2001) Laminar organization of the pyramidal cell Naber, P.A., Lopes da Silva, F.H. and Witter, M.P. (2001)
layer of the subiculum in the rat. J. Comp. Neurol., 435(1): Reciprocal connections between the entorhinal cortex and
89–110. hippocampal fields CA1 and the subiculum are in register
Ishizuka, N., Weber, J. and Amaral, D.G. (1990) Organization with the projections from CA1 to the subiculum. Hippocam-
of intrahippocampal projections originating from CA3 py- pus, 11(2): 99–104.
ramidal cells in the rat. J. Comp. Neurol., 295(4): 580–623. Nafstad, P.H. (1967) An electron microscope study on the ter-
Jaarsma, D., Postema, F. and Korf, J. (1992) Time course and mination of the perforant path fibres in the hippocampus and
distribution of neuronal degeneration in the dentate gyrus the fascia dentata. Z. Zellforsch. Mikrosk. Anat., 76(4):
of rat after adrenalectomy: A silver impregnation study. 532–542.
Hippocampus, 2(2): 143–150. Nitsch, R. and Leranth, C. (1991) Neuropeptide Y (NPY)-
Kemppainen, S., Jolkkonen, E. and Pitkanen, A. (2002) Pro- immunoreactive neurons in the primate fascia dentata; occa-
jections from the posterior cortical nucleus of the amygdala sional coexistence with calcium-binding proteins: a light
to the hippocampal formation and parahippocampal region and electron microscopic study. J. Comp. Neurol., 309(4):
in rat. Hippocampus, 12(6): 735–755. 430–444.
Kjelstrup, K.G., Tuvnes, F.A., Steffenach, H.A., Murison, R., Ramón y Cajal, S. (1911) Histologie du SystemeNerveux de
Moser, E.I. and Moser, M.B. (2002) Reduced fear expression l’Homme et des Vertebres. Maloine, Paris.
after lesions of the ventral hippocampus. Proc. Natl. Acad. Ruth, R.E., Collier, T.J. and Routtenberg, A. (1982) Topog-
Sci. U.S.A., 99(16): 10825–10830. raphy between the entorhinal cortex and the dentate septo-
Köhler, C. (1985) A projection from the deep layers of the temporal axis in rats. I. Medial and intermediate entorhinal
entorhinal area to the hippocampal formation in the rat projecting cells. J. Comp. Neurol., 209(1): 69–78.
brain. Neurosci. Lett., 56(1): 13–19. Ruth, R.E., Collier, T.J. and Routtenberg, A. (1988) Topo-
Krettek, J.E. and Price, J.L. (1977) Projections from the am- graphical relationship between the entorhinal cortex and the
ygdaloid complex and adjacent olfactory structures to the septotemporal axis of the dentate gyrus in rats. II. Cells
entorhinal cortex and to the subiculum in the rat and cat. J. projecting from lateral entorhinal subdivisions. J. Comp.
Comp. Neurol., 172(4): 723–752. Neurol., 270(4): 506–516.
Kunzle, H. (2002) Distribution of perihippocampo-hippocampal Sargolini, F., Fyhn, M., Hafting, T., McNaughton, B.L., Wit-
projection neurons in the lesser hedgehog tenrec. Neurosci. ter, M.P., Moser, M.B. and Moser, E.I. (2006) Conjunctive
Res., 44(4): 405–419. representation of position, direction, and velocity in ent-
Leranth, C., Malcolm, A.J. and Frotscher, M. (1990) Afferent orhinal cortex. Science, 312(5774): 758–762.
and efferent synaptic connections of somatostatin-immuno- Scharfman, H.E. (1991) Dentate hilar cells with dendrites in the
reactive neurons in the rat fascia dentata. J. Comp. Neurol., molecular layer have lower thresholds for synaptic activation
295(1): 111–122. by perforant path than granule cells. J. Neurosci., 11(6):
Lim, C., Blume, H.W., Madsen, J.R. and Saper, C.B. (1997) 1660–1673.
Connections of the hippocampal formation in humans. I. The Scharfman, H.E., Sollas, A.L., Smith, K.L., Jackson, M.B. and
mossy fiber pathway. J. Comp. Neurol., 385(3): 325–351. Goodman, J.H. (2002) Structural and functional asymmetry
Lorente de Nó, R. (1933) Studies on the structure of the cer- in the normal and epileptic rat dentate gyrus. J. Comp.
ebral cortex. J. Psychol. Neurol., 45(6): 26–438. Neurol., 454(4): 424–439.
Majak, K. and Pitkanen, A. (2003) Projections from the per- Schwartz, S.P. and Coleman, P.D. (1981) Neurons of origin of
iamygdaloid cortex to the amygdaloid complex, the hippo- the perforant path. Exp. Neurol., 74(1): 305–312.
campal formation, and the parahippocampal region: a PHA- Seress, L. and Mrzljak, L. (1987) Basal dendrites of granule
L study in the rat. Hippocampus, 13(8): 922–942. cells are normal features of the fetal and adult dentate gyrus
Matthews, D.A., Cotman, C. and Lynch, G. (1976) An electron of both monkey and human hippocampal formations. Brain
microscopic study of lesion-induced synaptogenesis in the Res., 405(1): 169–174.
dentate gyrus of the adult rat. I. Magnitude and time course Soriano, E. and Frotscher, M. (1993) GABAergic innervation
of degeneration. Brain Res., 115(1): 1–21. of the rat fascia dentata: a novel type of interneuron in the
61

granule cell layer with extensive axonal arborization in the Witter, M.P. (1989) Connectivity of the rat hippocampus. In:
molecular layer. J. Comp. Neurol., 334(3): 385–396. Chan-Palay V. and Köhler C. (Eds.), The Hippocampus —
Stanfield, B.B. and Cowan, W.M. (1979) The morphology of New Vistas. Allen R. Liss, New York, pp. 53–69.
the hippocampus and dentate gyrus in normal and reeler Witter, M.P. (1990) Organization of the entorhinal projections
mice. J. Comp. Neurol., 185(3): 393–422. to the hippocampal formation of the rat. Soc. Neurosci.
Steffenach, H.A., Witter, M., Moser, M.B. and Moser, E.I. Abstr., 16: 124.
(2005) Spatial memory in the rat requires the dorsolateral Witter, M.P. (2006) Connections of the subiculum of the rat:
band of the entorhinal cortex. Neuron, 45(2): 301–313. topography in relation to columnar and laminar organiza-
Stephan, H. (1975) Allocortex. In: Borgman W. (Ed.), Handbuch tion. Behav. Brain Res., 174(2): 251–264.
der Mikroskopischen Anatomie des Menschen. Springer- Witter, M.P. and Amaral, D.G. (1991) Entorhinal cortex of the
Verlag, Berlin. monkey. V. Projections to the dentate gyrus, hippocampus,
Steward, O. (1976) Topographic organization of the projections and subicular complex. J. Comp. Neurol., 307(3): 437–459.
from the entorhinal area to the hippocampal formation of the Witter, M.P. and Amaral, D.G. (2004) Hippocampal forma-
rat. J. Comp. Neurol., 167(3): 285–314. tion. In: Paxinos G. (Ed.), The Rat Nervous System (3rd
Steward, O. and Scoville, S.A. (1976) Cells of origin of ent- Ed.). Elsevier Academic Press, San Diego, CA, pp. 635–704.
orhinal cortical afferents to the hippocampus and fascia den- Witter, M.P. and Groenewegen, H.J. (1984) Laminar origin and
tata of the rat. J. Comp. Neurol., 169(3): 347–370. septotemporal distribution of entorhinal and perirhinal pro-
Swanson, L.W. and Köhler, C. (1986) Anatomical evidence for jections to the hippocampus in the cat. J. Comp. Neurol.,
direct projections from the entorhinal area to the entire cor- 224(3): 371–385.
tical mantle in the rat. J. Neurosci., 6(10): 3010–3023. Witter, M.P. and Groenewegen, H.J. (1990) The subiculum:
Tamamaki, N. (1997) Organization of the entorhinal projection cytoarchitectonically a simple structure, but hodologically
to the rat dentate gyrus revealed by Dil anterograde labeling. complex. Prog. Brain Res., 83: 47–58.
Exp. Brain Res., 116(2): 250–258. Witter, M.P., Groenewegen, H.J., Lopes da Silva, F.H. and
Tamamaki, N. and Nojyo, Y. (1993) Projection of the entorhinal Lohman, A.H. (1989a) Functional organization of the ex-
layer II neurons in the rat as revealed by intracellular pressure- trinsic and intrinsic circuitry of the parahippocampal region.
injection of neurobiotin. Hippocampus, 3(4): 471–480. Prog. Neurobiol., 33(3): 161–253.
Uva, L., Gruschke, S., Biella, G., de Curtis, M. and Witter, Witter, M.P. and Moser, E.I. (2006) Spatial representation and
M.P. (2004) Cytoarchitectonic characterization of the para- the architecture of the entorhinal cortex. Trends Neurosci.,
hippocampal region of the guinea pig. J. Comp. Neurol., 29(12): 671–678.
474(2): 289–303. Witter, M.P., Van Hoesen, G.W. and Amaral, D.G. (1989b)
Van Hoesen, G.W. and Pandya, D.N. (1975) Some connections Topographical organization of the entorhinal projection to
of the entorhinal (area 28) and perirhinal (area 35) cortices of the dentate gyrus of the monkey. J. Neurosci., 9(1): 216–228.
the rhesus monkey. III. Efferent connections. Brain Res., Witter, M.P., Wouterlood, F.G., Naber, P.A. and van Haeften,
95(1): 39–59. T. (2000) Anatomical organization of the parahippocampal-
Voogd, J., Nieuwenhuis, R., van Dongen, P.A.M. and Ten hippocampal network. Ann. N.Y. Acad. Sci., 911: 1–24.
Donkelaar, H.J. (1998) Mammals. In: Dubbeldam J.L., van Woznicka, A., Malinowska, M. and Kosmal, A. (2006) Cyto-
Dongen P.A.M. and Voogd J. (Eds.), The Central Nervous architectonic organization of the entorhinal cortex of the ca-
System of Vertebrates, Vol. 3. Springer-Verlag, Berlin, nine brain. Brain Res. Rev., 52(2): 346–367.
pp. 1637–2098. Wyss, J.M. (1981) An autoradiographic study of the efferent
Wang, X. and Lambert, N.A. (2003) Membrane properties of connections of the entorhinal cortex in the rat. J. Comp.
identified lateral and medial perforant pathway projection Neurol., 199(4): 495–512.
neurons. Neuroscience, 117(2): 485–492. Zipp, F., Nitsch, R., Soriano, E. and Frotscher, M. (1989)
White, W.F., Nadler, J.V., Hamberger, A., Cotman, C.W. and Entorhinal fibers form synaptic contacts on parvalbumin-
Cummins, J.T. (1977) Glutamate as transmitter of hippo- immunoreactive neurons in the rat fascia dentata. Brain Res.,
campal perforant path. Nature, 270(5635): 356–357. 495(1): 161–166.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 4

Extrinsic afferent systems to the dentate gyrus

Csaba Leranth1,2, and Tibor Hajszan1,3

1
Department of Obstetrics, Gynecology, and Reproductive Sciences, Yale University School of Medicine,
333 Cedar Street, FMB 312, New Haven, CT 06520, USA
2
Department of Neurobiology, Yale University School of Medicine, New Haven, CT 06510, USA
3
Department of Biophysics, Biological Research Center, Hungarian Academy of Sciences, H-6726 Szeged, Hungary

Abstract: The dentate gyrus is the first stage of the intrahippocampal, excitatory, trisynaptic loop, and a
primary target of the majority of entorhinal afferents that terminate in a laminar fashion on granule cell
dendrites and carry sensory information of multiple modalities about the external world. The electric
activity of the trisynaptic pathway is controlled mainly by different types of local, GABAergic interneu-
rons, and subcortical and commissural afferents. In this chapter we will outline the origin and postsynaptic
targets in the dentate gyrus of chemically identified subcortical inputs. These systems are afferents orig-
inating from the medial septum/diagonal band of Broca GABAergic and cholinergic neurons, neuro-
chemically distinct types of neurons located in the supramammillary area, serotonergic fibers from the
median raphe, noradrenergic afferents from the pontine nucleus, locus ceruleus, dopamine axons origi-
nating in the ventral tegmental area, and the commissural projection system. Because of the physiological
implications, these afferents are discussed in the context of the glutamatergic innervation of the dentate
gyrus. One common feature of the extrinsic dentate afferent systems is that they originate from a relatively
small number of neurons. However, the majority of these afferents are able to exert a powerful control over
the electrical activity of the hippocampus. This strong influence is due to the fact that the majority of the
extrinsic afferents terminate on a relatively small, but specific, populations of neurons that are able to
control large areas of the hippocampal formation.

Keywords: septum; supramammillary area; median raphe; commissural connection; acetylcholine;


dopamine; noradrenaline; serotonin

Introduction and anatomical overview Even if only a fraction of these connections are
active at the same time, a network with virtually
It has been known for more than 100 years that endless numbers of possible active patterns emerge
neurons present in any cortical area are far from (Freund and Buzsaki, 1996). Thus, even if we will
being uniform regarding their morphology and in- be able to understand the complete molecular and
trinsic and extrinsic connections (Ramon y Cajal, biophysical properties of a single hippocampal
1893, 1911). This suggests that they are able to neuron, in order to understand the function of the
interact with each other in a complex fashion. hippocampal formation, including the dentate
gyrus, as a whole, it is indispensable to elucidate
Corresponding author. Tel.: +1 203 785 4748; both the intrinsic and extrinsic connections of
E-mail: csaba.leranth@yale.edu these cells.

DOI: 10.1016/S0079-6123(07)63004-0 63
64

The major cell types and the majority of intrin- laminar fashion on granule cell dendrites
sic and extrinsic connections of the hippocampal (see Chapter 1) and carry sensory information of
formation, which comprises the cornu ammonis, multiple modalities about the external world. The
divided into three subfields, CA1–CA3 (Lorente de principal cells of the dentate gyrus are the granule
No, 1934) and the dentate gyrus (fascia dentata cells, 1 million in rats and 5 million in non-
and hilus), have been well known since the studies human primates (Claiborne et al., 1986; Seress,
of Ramon y Cajal (1893) and Lorente de No 1988), and the mossy cells (Amaral, 1978). The
(1934) and have been reviewed extensively (e.g., small cell bodies of the granule cells (8–12 mm)
Amaral and Witter, 1989, 1995; Lopes de Silva form the granule cell layer. Granule cells have two
et al., 1990). It should be noted, however, that the main, radially oriented, spiny dendrites emitting
term hippocampal formation sometimes is used several fine branches, which reach the hippocam-
to include the subicular complex and entorhinal pal fissure. The axons of the granule cells, the
cortex (Amaral and Witter, 1995). mossy fiber axons, originate from the opposite
The three dimensional position of the hippo- pole of the soma relative to the dendrites, and
campal formation in the brain is rather complex. enter the dentate hilus, where they give rise to
In the majority of species, the hippocampal for- several collaterals (Claiborne et al., 1986). Recur-
mation is an elongated structure with its long axis rent collaterals periodically enter the granule cell
extending in a C-shaped fashion from the septal layer, climb along the cell bodies and dendrites of
nuclei of the basal forebrain rostrodorsally, over presumed basket cells, and form synapses (Ribak
and behind the diencephalon, to the incipient and Peterson, 1991). The main axonal projection
temporal lobe caudoventrally (see more details in of the granule cells leaves the hilar region and
Amaral and Witter, 1995). courses through the stratum lucidum of the CA3
Although there is still a debate about the classi- subfield, where it forms the giant mossy terminals
fication of hippocampal neurons, the simple synapsing on the proximal dendrites of pyramidal
terms of principal and non-principal neurons are cells.
accepted (Freund and Buzsaki, 1996). The hippo- The dentate hilus is located subjacent to the
campal principal neurons are the pyramidal and granule cell layer and extends to the border of the
granule cells located in the CA1–CA3 subfields of dendritic layer of CA3 that is interposed between
the ammon’s horn and dentate gyrus, respectively. the upper (suprapyramidal) and lower (infrapy-
In addition, the mossy cells of the dentate gyrus can ramidal) blades of the dentate gyrus. The principal
be referred to as principal cells, as discussed further and most numerous cell type in the hilus is the
below. The granule cells and pyramidal neurons, mossy cell. These neurons are characterized by
with their intrahippocampal connections, are the their densely spiny dendrites and several thorny
major components of the so-called trisynaptic cir- excrescences on both the cell body and proximal
cuit of the hippocampus, which is unidirectional dendritic shafts and their dendrites are mostly
and glutamatergic. Layer 2 pyramidal cells of the confined to the hilus (Amaral, 1978). However,
entorhinal cortex project to granule cells (and single mossy dendrites were observed penetrating
to some extent, to CA3 pyramidal cells) via the the stratum moleculare (Scharfman, 1991; Scharf-
perforant pathway. Granule cells, via their mossy man, 1995; Soltesz and Mody, 1994). Axons of
fiber projection, organized in a laminar fashion, mossy cells innervate the inner third of the dentate
terminate on CA3 pyramidal neurons, which send molecular layer of both the ipsi- and contralateral
their Schaffer collaterals to the CA1 pyramidal dentate gyrus and also have collaterals in the hilus
cells. CA1 pyramidal neurons, in turn, project back (Amaral, 1978; Laurberg and Sorensen, 1981;
to the entorhinal cortex, via the subiculum. Ribak et al., 1985; Buckmaster et al., 1996). These
As mentioned above, the dentate gyrus is the intrahilar collaterals of mossy cells terminate on
first stage of the intrahippocampal, excitatory, unidentified dendrites in the hilus and on dend-
trisynaptic loop, being the primary target of the rites of hilar interneurons (Frotscher and Zimmer,
majority of entorhinal afferents that terminate in a 1983a, b; Frotscher et al., 1984; Ribak et al., 1985;
65

Buckmaster et al., 1996). Because both the CA3 of GABAergic, parvalbumin (PA)-containing
pyramidal cells and mossy axons form asymmetric, (Freund, 1989) neurons located in the midline of
glutamatergic, excitatory synapses, many authors the MSDB (Kiss et al., 1990). The type 2 axons are
consider the mossy cells as modified CA3 pyram- GABA-negative (Freund and Antal, 1988), but are
idal neurons (Soriano and Frotscher, 1994; Scharf- immunoreactive for ChAT, the synthesizing en-
man, 1995, 1999). Therefore, mossy cells are zyme of acetylcholine (Frotscher and Leranth,
considered as excitatory principal cells that pro- 1985, 1986; Leranth and Frotscher, 1987). The
vide long-range ipsilateral and commissural pro- parent neurons of these axons are also in the
jections into the dentate gyrus (Amaral, 1978). MSDB and are positioned in a way so that they
The electric activity of the aforementioned ex- surround the septo-hippocampal GABAergic cells
citatory signal loop is controlled mainly by differ- (Kiss et al., 1990).
ent types of local, GABAergic interneurons (for
review, see Freund and Buzsaki, 1996) and sub-
cortical and commissural afferents. In this chapter MSDB cholinergic innervation of the dentate gyrus
we will outline our recent knowledge regarding the
origin and postsynaptic targets in the dentate Basic and clinical studies have long recognized the
gyrus of chemically identified subcortical inputs, importance of cholinergic mechanisms in cognitive
including afferents originating from the medial function (Givens and Sarter, 1997), and drugs
septum/diagonal band of Broca (MSDB) GAB- which increase synaptic acetylcholine levels are
Aergic and cholinergic neurons, the dentate pro- currently the most common for the treatment of
jection of neurochemically different types of cognitive deficits associated with disorders such as
neurons located in the supramammillary area Alzheimer’s disease, albeit, with limited effective-
(SUM), serotonergic fibers from the median rap- ness (Benzi and Moretti, 1998). The septo-hippo-
he (MR), noradrenergic afferents from the pontine campal pathway (SH), which originates in the
nucleus, locus ceruleus, dopamine axons originat- MSDB and shows progressive degeneration in
ing in the ventral tegmental area, and the com- Alzheimer’s disease (Whitehouse et al., 1982), has
missural system. specifically been implicated in cognitive function.
Lesions of the fimbria-fornix, which carry SH
cholinergic (Lewis and Shute, 1967) fibers to the
Septo-hippocampal connections hippocampal formation, including the dentate
gyrus, interfere both with learning and memory
In the 1970s, it had been shown that anterograd- tasks and with generation of the theta rhythm in
ely-transported radiolabeled amino acids, injected rats (Brito and Brito, 1990). These deficits can be
into the MSDB, could be detected in hippocampal attenuated by grafting acetylcholine-producing
neurons (Rose and Schubert, 1977). Studies in the cells to the hippocampus (Dunnett et al., 1982;
seventies and eighties, using the combination of Dickinson-Anson et al., 1998).
retrograde tracer technique and immunostaining Electrophysiological experiments have demon-
for choline acetyltransferase (ChAT) and gluta- strated that excitatory, inhibitory, and disinhibi-
mate decarboxylase (GAD) showed that there tory responses can be recorded as a result of the
were two major populations of MSDB neurons iontophoretic application of acetylcholine in the
projecting to the hippocampus (for review, see dentate gyrus (Wheal and Miller, 1980; Fricke and
Leranth and Frotscher, 1989). Later, experiments Prince, 1984). It is not clear, however, from these
using anterograde labeling with the lectin PHAL experiments whether acetylcholine exerts these
revealed two types of septohippocampal fibers. effects at the level of the primary dendrites,
One has large boutons that occur in clusters, directly onto the perikaryon of principal cells or
whereas the other has small boutons and arborizes indirectly, via interneurons (Hounsgaard, 1978).
more diffusely (Nyakas et al., 1987). Type 1 axons Histochemical staining procedures to visualize
are immunoreactive for GABA and are processes putative cholinergic neurons and fibers applying
66

acetylcholinesterase (ACHE) reaction were first At the electron microscopic level, the vast
used in the dentate gyrus in the early 1960’s majority of ChAT-positive synaptic boutons con-
(Storm-Mathisen and Blackstad, 1964). These au- tacts dendritic shafts and forms predominantly
thors described a laminar pattern of ACHE fiber symmetric synaptic contacts. In addition, some
staining in the dentate gyrus; the densest band oc- asymmetric synapses, 11% of the total number
curring at the interface between the granule cell of ChAT synapses (Clarke, 1985), could also be
layer and the molecular layer. Shute and Lewis observed on dendritic spines. There is a possibility
(1966) modified this ACHE histochemical tech- that axons forming symmetric and asymmetric
nique to permit examination of the dentate gyrus, contacts originate from different population of
stained for ACHE, using the electron microscope. cholinergic neurons; intrinsic and extrinsic. How-
Their study revealed several neurons histochemi- ever, it seems unlikely since, lesions of the septo-
cally stained for ACHE, numerous axonal fibers hippocampal pathway cause an almost complete
and some synaptic boutons in contact with pre- removal of cholinergic markers in the dentate
dominantly dendritic shafts. Biochemical measure- gyrus (Mellgren and Srebro, 1973).
ments of the levels of ChAT in the dentate gyrus In the dentate gyrus, the majority of the postsy-
also indicate a supragranular band of intense naptic targets of cholinergic boutons are the gran-
cholinergic expression (Fonnum, 1970). ule cells and the ChAT-immunoreactive boutons
The first detailed light and electron microscopic form both axodendritic and axospinous synapses
studies using immunostaining for ChAT on the with these neurons. All of the axospinous synapses
cholinergic innervation of the dentate gyrus were observed are asymmetric (Fig. 1). In addition,
performed in the middle 1980s (Frotscher and 5–10% of all postsynaptic elements of choliner-
Leranth, 1985, 1986; Wainer et al., 1984; Clarke, gic axons show ultrastructural features of inter-
1985; Leranth and Frotscher, 1987). The light mi- neurons. A combined electron microscopic, Golgi
croscopic analyses revealed a dense plexus of impregnation and ChAT immunostaining study
ChAT-immunoreactive fibers in the dentate gyrus has shown symmetric synaptic contacts between
forming a supragranular band at the interface be- ChAT-containing axon terminals and non-spiny,
tween the granule cell layer and molecular layer. smooth dendritic shafts, characteristic for inter-
Very little staining was in the granule cell layer and neurons (Frotscher and Leranth, 1986; Fig. 2).
only a few fibers were located in the hilar region. This 5–10% appears to correspond to the propor-
In addition, ChAT-immunoreactive neurons could tion of interneurons in the neuropil, suggesting
also be observed in the dentate hilus (Clarke, 1985; that the interneurons are contacted in a ‘‘quasi-
Frotscher et al., 1986). These ChAT-positive cells random’’ fashion. Direct evidence that interneu-
are rare and non-principal neurons. They are rel- rons are indeed innervated by cholinergic fibers is
atively small with round or ovoid perikarya, which provided by an electron microscopic double-
give rise to thin spine-free dendrites and are very immunostaining study that showed ChAT-posi-
similar to ChAT-immunoreactive cells in the neo- tive axon terminals forming symmetric synaptic
cortex of the same animals but were quite different contacts with hilar GAD (glutamic acid dec-
from cholinergic neurons in the basal forebrain, arboxylase)- and SOM (somatostatin)-containing
medial septal nucleus, and neostriatum, which neurons (Leranth and Frotscher, 1987).
were larger and more intensely immunostained. In addition to the cholinergic innervation of
Electron-microscopic analysis of ChAT-contain- dentate granule cells and interneurons, MSDB
ing cells in the hippocampus and fascia dentata cholinergic cells innervate mossy cells densely. A
revealed that their afferent synaptic contacts are correlated light and electron microscopic double-
mainly the asymmetric type, and are located on immunostaining study has demonstrated numerous
their cell bodies and smooth proximal dendrites. axosomatic synaptic contacts between ChAT-
The nuclei of the immunoreactive cells exhibited containing axon terminals and calcitonin gene-
deep indentations, which are a characteristic of related peptide (CGRP)-immunoreactive neurons
non-pyramidal neurons (Frotscher et al., 1986). in the dentate hilar area (Deller et al., 1999). CGRP
67

Fig. 1. Low (Panel a) and high power (Panel b) electron micrographs (received from Dr. Michael Frotscher) show the result of a
combined Golgi impregnation and ChAT immunostaining experiment. On Panel a, Golgi-impregnated (gold-toned) spiny granule cell
dendrites are seen. Arrow on the same panel points at a ChAT immunoreactive bouton contacting the spine head of the dendrite (D).
Panel b shows the asymmetric synaptic contact (arrow) between the two profiles. Bar scales ¼ 1 mm.

of these neurons could modulate granule cell excit-


ability throughout large portions of the dentate
gyrus.

MSDB GABAergic innervation of the dentate gyrus

It has been shown that when the perforant path is


stimulated with a brief test pulse, it evokes exci-
tatory postsynaptic potentials (EPSPs) in the gran-
ule cell dendrites, which could be recorded
collectively in the dentate hilus as a positive-going
Fig. 2. Electron micrograph taken from the molecular layer of field potential. At greater stimulus intensities, a
the dentate gyrus immunostained for ChAT. A ChAT-immuno- negative-going component is superimposed on the
reactive axon terminal forms asymmetric synaptic contact with positive-going field potential, as the result of
a non-spiny dendritic shaft (D). The lack of dendritic spines the synchronous activation of many granule cells.
indicates that this dendrite is the process of an interneuron.
This component is referred to as a population
Asterisks label axon terminals forming asymmetric synapses
with the same dendrite. Bar scale ¼ 1 mm. spike (PS). When a conditioning pulse is applied to
the medial septum just prior to a perforant path
is an accepted marker for mossy cells (Freund et al., test pulse, the EPSP remains unchanged, but the
1997). Since mossy cells project for a long distance PS is greater than that expected from the sum of
along the longitudinal axis of the hippocampus the potentials evoked by either stimulus alone
(Ribak et al., 1985), the cholinergic innervation (Alvarez-Leefmans and Gardner-Medwin, 1975;
68

Fantie and Goddard, 1982). Thus, the septohip- facilitated the granule cell population spike evoked
pocampal input appears to facilitate the discharge by perforant path stimulation, and infusion into
of granule cells. An interesting and unexpected the dentate hilus of a GABAA receptor antagonist,
finding of these studies was that stimulation of the picrotoxin, blocked the facilitation. This observa-
medial septum alone, at a site that produced the tion suggests that the facilitatory effect of MSDB
facilitation of the PS in the dentate gyrus, did not stimulation is mediated through an inhibitory
necessarily evoke a field potential in the dentate connection from the MSDB onto inhibitory inter-
gyrus. This finding, together with the failure of neurons in the dentate gyrus, and that this con-
medial septal stimulation to affect the perforant nection may utilize the neurotransmitter GABA.
path-evoked field potential, upon which the PS is Indeed, it has been demonstrated in a series
superimposed suggests that the facilitation does of concomitant morphological studies that SH (in-
not result from summation of excitatory input to cluding the septo-dentate) GABAergic terminals
the granule cells. Two mechanisms that have been always terminate on GABAergic interneurons, in
proposed by Buzsaki (1984) to account for these the rat (Freund and Antal, 1988; Gulyas et al.,
effects are: (1) medial septal input acts directly on 1990) and monkey hippocampus (Gulyas et al.,
granule cells to facilitate activation (Robinson and 1991). Experiments using double immunostaining
Racine, 1984), and (2) the medial septum acts in- for PHAL and markers for different subsets of in-
directly, via the inhibitory interneurons responsi- terneurons have revealed that all examined sub-
ble for feed-forward and feed-back inhibition populations of hippocampal GABAergic
(Kandel et al., 1961; Anderson and Eccles, 1962; interneurons, including those co-expressing parv-
Mosko et al., 1973; Buzsaki and Eidelburg, 1982). albumin, calbindin, SOM, neuropeptide Y, chole-
Anatomical studies of the septo-dentate termi- cystokinin, and vasoactive intestinal polypeptide,
nation patterns show that the innervation of the receive input from GABAergic septohippocampal
supragranular and molecular layers are consider- afferents (Freund and Antal, 1988; Gulyas et al.,
ably less dense than the innervation of the dentate 1990; Miettinen and Freund, 1992a, b; Acsady
hilus and subgranular layer (Lynch et al., 1978; et al., 1993). The typical innervation pattern is
Chandler and Crutcher, 1983). Although MSDB multiple, ‘‘climbing fiber-like’’ contacts on the
fibers in the supragranular and molecular layers soma, and proximal and distal dendrites of the
are likely to terminate on granule cells, fibers in the postsynaptic interneurons (Fig. 3). All these con-
subgranular and hilar region appear to terminate tacts have proved to be symmetrical synapses. The
on cells having the morphological characteristics area where most of the interneurons appear to re-
of interneurons (Rose and Schubert, 1977; ceive septal GABAergic input is the CA3 subfield,
Chandler and Crutcher, 1983). If SH fibers form particularly strata oriens and pyramidale (Rose
excitatory connections onto inhibitory interneu- and Schubert, 1977; Nyakas et al., 1987; Freund
rons, their activation might synchronize granule and Antal, 1988; Gaykema et al., 1990). The dent-
cell responses, resulting in a larger population ate hilus is also heavily innervated, but in the CA1
spike from a subsequent perforant path input region, a relatively smaller proportion of the inter-
(Buzsaki, 1984). Alternatively, if they form inhib- neurons receive multiple synaptic inputs. No sys-
itory connections onto the interneurons, septal ac- tematic differences have been observed in the
tivation might reduce both tonic and feed-forward termination pattern of septal afferents between
inhibition of the granule cells, allowing a larger the dorsal and ventral hippocampus.
number of these neurons to be activated by the
perforant path input, again resulting in a larger
population spike. To determine the validity of Interactions between the septohippocampal
these hypotheses, Bilkey and Goddard (1985) per- cholinergic and GABAergic systems
formed a very elegant study. They have shown that
a conditioning pulse to the medial septum, al- In order to better understand the function of the
though eliciting no field potential of its own, septohippocampal projection system that contains
69

acetylcholine (ACh) release (Monmaur and


Breton, 1991; Givens and Olton, 1994, 1995;
Apartis et al., 1998; Bland and Oddie, 1998;
Dickinson-Anson et al., 1998). As such, it has
been presumed that the memory-enhancing effects
of intraseptal administration of muscarinic ago-
nists occurs due to increased firing of MSDB
cholinergic neurons (Markowska et al., 1995;
Givens and Sarter, 1997). These assumptions have
been based on earlier studies that reported a mu-
scarinic receptor-mediated increase in firing of
MSDB neurons (Dutar et al., 1983; Lamour et al.,
1984), which lead to the hypothesis that Ach, via
muscarinic receptors, has a positive feed-back
effect on MSDB cholinergic neurons (Dutar
et al., 1983; Lamour et al., 1984; see review in,
Fig. 3. Light micrograph (kindly provided by Dr. Attila Wu et al., 2004). However, a recent combined
Gulyas) shows the result of a combined anterograde tracing electrophysiological and morphological study
and calretinin immunostaining study, in the dentate gyrus. The
anterograde tracer, biotinylated dextran amine (BDA) was in-
(Wu et al., 2004), using a novel fluorescent labe-
jected into the medial septum diagonal band. Large boutons of ling technique to selectively visualize live septo-
BDA-containing axons form multiple, basket-like, putative hippocampal cholinergic neurons has demonstrated
synaptic contacts with the soma of calretinin-immunoreactive that administration of muscarinic agonists to the
(labeled with a brown diaminobenzidine reaction product) neu- MSDB do not excite septo-hippocampal choliner-
rons. One of these cells (arrows) located in the supragranular
layer (SgL) the other is at the border between the granule cell
gic neurons, instead they inhibit a subpopula-
layer (GcL) and dentate hilar area (H). Bar scale ¼ 50 mm. tion of them. In contrast, septo-hippocampal
(See Color Plate 4.3 in color plate section.) GABAergic neurons (visualized by retrograde
tracing and parvalbumin immunostaining tech-
niques) are profoundly excited by muscarine
a cholinergic and a GABAergic component, the administration into the MSDB. Thus the cogni-
anatomical and functional interaction between tion enhancing effects of muscarinic drugs in the
these two systems must be briefly discussed, at MSDB cannot be attributed to an increase in hip-
the level of the MSDB. pocampal ACh release. Instead, disinhibitory
It has to be noted that intrinsic cholinergic mechanisms (Freund and Antal, 1988), due to in-
mechanisms operating within the MSDB are also creased impulse flow in the septo-hippocampal
critical for learning and memory. Infusion of GABAergic pathway, may underlie the cognition-
muscarinic agonists directly into the MSDB elicit enhancing effects of muscarinic agonists. In sup-
continuous hippocampal theta rhythm (Monmaur port of this view is the morphological observation
and Breton, 1991; Lawson and Bland, 1993) and that axon collaterals of the septo-hippocampal
facilitate learning and memory-related behaviors cholinergic neurons heavily innervate MSDB se-
both in young (Givens and Olton, 1990) and aged pto-hippocampal GABAergic neurons (Leranth
rats (Markowska et al., 1995). Also, intraseptal and Frotscher, 1989; Fig. 4). These connections
infusions of muscarinic agonists can alleviate could be involved in the aforementioned process.
systemic scopolamine-induced amnesia, suggesting
that the MSDB is a critical locus for the mnemonic
effects of muscarinic drugs (Givens and Olton, Supramamillo-dentate connections
1995). In general, it is assumed that improvements
in MSDB-related learning and memory tasks It has been shown that prestimulation of sup-
occur as a result of an increase in hippocampal ramammillary (SUM) neurons significantly
70

hippocampus in the rat (Segal and Landis, 1974).


Later, Amaral and Cowan (1980) showed that
horseradish peroxidase (HRP) injected into the
monkey hippocampus also results in labeled cells
in the SUM. Furthermore, it was noted that more
cells ipsilateral to the site of injection were labeled
than cells in the contralateral SUM and a large
number of SUM efferents terminate in the dentate
gyrus since, local application of the retrograde
tracer, Evans blue to the upper blade of the dent-
ate gyrus produced labeled cell bodies in the SUM
(Harley et al., 1983). Labeling was observed
throughout the rostrocaudal aspect of the SUM.
Therefore, Amaral and Cowan (1980) and Harley
et al. (1983) postulated that afferents from the
SUM to the hippocampus involve at least as many
cells in the SUM as the septal neurons, which give
rise to the septohippocampal projection. The spe-
cific pathway containing SUM efferents to the
hippocampus remains debatable. The medial fore-
brain bundle (Haglund et al., 1984) and fornix
(Veazey et al., 1982) have been suggested as likely
candidates. More specifically, Veazey et al. (1982)
have postulated that the medial forebrain bundle
carries SUM efferents to septal nuclei, whereas the
fornix carries projection from the SUM to the
hippocampus. At the level of dorsal hippocampus,
labeled fibers have been described in the fimbria
Fig. 4. Electron micrograph shows the result of a double
immunostaining experiment for choline acetyltransferase (Haglund et al., 1984) or the subcallosal fornix
(ChAT) and glutamic acid decarboxylase (GAD), in the rat (Wiss et al., 1979).
medial septum diagonal band of Broca. Immunoreactivity for Regarding the neurochemical nature of the
ChAT and GAD was visualized with two contrasting immuno- SUM-hippocampal pathway, studies were able to
marker, diaminobenzidine reaction and ferritin labeling, re-
demonstrate that calretinin-immunoreactive axon
spectively. A ChAT-immunoreactive bouton forms asymmetric
synaptic contact (arrowheads) with a GAD immunoreactive terminals in the inner molecular layer of the dent-
dendrite. Bar scale ¼ 1 mm. ate gyrus (and in the pyramidal layer of CA2)
originate in the SUM. Colocalization studies pro-
enhances perforant path-elicited population spikes vided evidence that these large projecting neurons
in the fascia dentata (Mizumori et al., 1989), an contained both calretinin and substance P (SP) but
effect that could be mimicked by glutamate injec- lacked GABA (Nitsch and Leranth, 1993). This
tions to the lateral supramammillary area (Carre observation indicates that the nature of the hypo-
and Harley, 1991). Moreover, Dahl and Winson thalamo-hippocampal projection to the dentate is
(1986) speculated about a neuronal control mech- likely to be excitatory. In fact, SP-containing, long
anism, a ‘‘gate,’’ mediated by supramammillary projection systems have been shown to exert ex-
afferents that would facilitate information flow in citatory actions (Nicoll et al., 1980). Conversely, in
the rat dentate gyrus, in a behavior-dependent the rat, the functional properties of the hypotha-
manner. lamo-hippocampal afferent system have been re-
Since the mid-1970s, it has been known that ported to be at least partially inhibitory (Segal,
a projection exists between the SUM and 1979). The study by Segal (1979) raises a question
71

of how can the physiological action on hippocam-


pal principal neurons of a putatively excitatory
pathway become inhibitory. A possible explana-
tion for this is that the supramammillo-hippocam-
pal afferents, in addition to terminating on
principal neurons (Magloczky et al., 1994), inner-
vate hippocampal GABAergic interneurons.
Indeed, in a study dealing with the innervation
of the primate hippocampal formation, thick and
varicose SP-immunoreactive axons, forming bas-
ket-like structures were identified adjacent to the
granule cell layer and in the hilar area of the dent-
ate gyrus and in the molecular layer of the middle
portion of CA3 (Nitsch and Leranth, 1994). The
location of these basket-like structures outside the
principal cell layers indicates that they contact
hippocampal non-principal neurons. Furthermore,
these SP-containing axons disappear after fimbria-
fornix transection, indicating that they are of
extrinsic origin (Nitsch and Leranth, 1994). A
subsequent study on non-human primates (Lera-
nth and Nitsch, 1994), applying correlated light
and electron microscopic immunocytochemical,
double-labeling technique for SP and parvalbu-
min and SP and calbindin and subsequent post-
embedding GABA-immunostaining revealed that
Fig. 5. Light micrograph taken from a double immunostained
this supramammillo-hippocampal afferent system vibratome section of the monkey dentate gyrus. Immunoreac-
establishes multiple, exclusively asymmetric tivity for substance P was labeled with a dark-blue Ni-di-
synapses with three specific subpopulations of aminobenzidine reaction, while immunostaining for
non-pyramidal cells: (1) a small portion of parv- parvalbumin was visualized by a brown diaminobenzidine re-
albumin-containing basket cells located in or action. The soma and dendrites of the parvalbumin-containing
cell embedded into the granule cell layer (GcL) is contacted by
adjacent to the granule cell layer of the dentate several substance P-immunoreactive axon terminals (arrows).
gyrus (Fig. 5), which, therefore, inhibit only a Bar scale ¼ 10 mm. (See Color Plate 4.5 in color plate section.)
subpopulation of granule cells; (2) some of the
calbindin-immunoreactive neurons located in the will transfer excitatory signals differently than
hilar area and in the granule cell layer (Fig. 6); and those that are controlled by a feed-forward inhib-
(3) calbindin-positive cells occurring exclusively in itory mechanism initiated by these fibers (see
the stratum moleculare of the middle portion of Fig. 10, in Leranth and Nitsch, 1994).
the CA3 subfield. Postembedding immunostaining
for GABA revealed that the aforementioned
calbindin-containing cells in area CA3 are Catecholaminergic brainstem-dentate connections
GABAergic inhibitory neurons. The results of this
study indicate that supramammillary afferents, a Serotonergic afferents
portion of which is glutamatergic (Kiss et al.,
2000) can effectively filter the information flow at The serotonergic raphe-hippocampal pathway has
different points along the trisynaptic circuit in the a powerful effect on hippocampal electric activity,
monkey hippocampal formation. Dentate granule depression-associated synaptic plasticity, and
cells, which are only stimulated by SUM afferents, cognitive behavior. Electrophysiological and
72

Fig. 6. Light (Panel a) and electron micrographs (Panels b, c, d) depicted from the monkey dentate gyrus demonstrate the result of a
correlated light and electron microscopic double immunostaining for substance P (labeled by a dark-blue Ni-diaminobenzidine
reaction) and calbindin (brown diaminobenzidine chromogen). Panel a shows a calbindin-immunoreactive neuron embedded into the
granule cell layer forming putative synaptic contacts (A–D) with substance P-containing axon terminals. Electron microscopic analysis
of ultrathin sections cut from the same area shows that boutons A and D form robust asymmetric synaptic contacts (arrowheads on
panels c and d) with the soma of this calbindin-containing cell (CB on panel b). Gc-granule cell. Bar scales ¼ panel a, 10 mm; panels
b–c, 1 mm.
73

pharmacological studies have shown that the typ- by axons originating in the median raphe. On
ical effect of serotonin on hippocampal neurons is the other hand, calbindin- and, to a smaller
a hyperpolarization evoked by an increase in K+ extent, cholecystokinin-containing interneurons
conductance, but depolarization and reduction of are targets for both pathways. In some cases
afterhyperpolarization were also reported (for the same individual calbindin- or cholecystokinin-
review, see Freund et al., 1990). containing neurons received multiple contacts
The serotonergic innervation of the hippocam- from afferents of both MSDB and median raphe
pus originates largely from the MR and a less origin. Thus, these observations indicate that
numerous projection arises from the dorsal raphe different subcortical nuclei modulate largely differ-
(for review, see Tork, 1990) and includes two types ent inhibitory circuits. However, considering
of fibers (Kosofsky and Molliver, 1987). In the the occasional convergence of the two subcortical
dentate gyrus, the most numerous are the thin ax- nuclei not only onto the same type, but also onto
ons with small evenly distributed varicosities that the same individual interneurons, the authors pro-
are present in the subgranular area. The other fiber posed that a particular inhibitory function, most
type has larger boutons that form clusters along probably feed-forward inhibition in the distal
the hilar border of the stratum granulosum and dendritic region, is under the control of both
contact secondary dendritic branches of two pathways.
cell types: GABAergic, pyramidal-shaped neurons It has to be noted that the MR serotonin system
located subjacent to the granule cell layer, and could also effect the hippocampus via an indirect
fusiform cells in the hilus. Both types of these route. The MR serotonergic neurons heavily in-
GABAergic cells seem to contain calbindin, but nervate the MSDB (Leranth and Vertes, 1999) and
none of them are immunoreactive for parvalbumin exert a robust stimulatory effect on parvalbumin-
(Freund et al., 1990; Halasy et al., 1992). containing septohippocampal GABAergic neurons
Based on these morphological data, it has been (Alreja, 1996). These GABAergic septal cells, in
suggested that the two types of serotonergic axons turn, selectively innervate hippocampal basket and
have different mechanisms of action in the hippo- chandelier cells (Freund and Antal, 1988), which
campus. Axons with small varicosities (originating are known to have powerful inhibitory effects on
in the dorsal raphe; Kosofsky and Molliver, 1987) the output sector (soma and axon hillock) of prin-
release serotonin at non-synaptic sites, diffusely, cipal neurons (Freund and Antal, 1988). Thus, se-
and target cells having 5-HT1-2 receptors, to exert rotonergic stimulation of the septohippocampal
a slow, tonic, G-protein-mediated action. The GABA system results in a disinhibition of princi-
other type of serotonergic axons with large pal cells. The situation is more complex, because a
boutons (originating in the median raphe; Kosof- population of MR serotonergic neurons project to
sky and Molliver, 1987) always form synaptic both the hippocampus and MSDB (e.g., McKenna
contacts with GABAergic interneurons that have and Vertes, 2001).
5-HT3 receptors (Halasy et al., 1992). Thus, stim-
ulation of median raphe serotonin neurons could
result in fast excitation of these GABAergic cells Noradrenergic and dopaminergic afferents to the
and an enhanced GABAA receptor-mediated inhi- dentate gyrus
bition of granule cells, a mechanism described in
the CA1 area (Roppert and Guy, 1991). Noradrenergic fibers in the hippocampus originate
An important question is whether there is from the locus ceruleus. Noradrenaline in the
convergence of the MSDB and MR hippocampal dentate gyrus promotes and permits long-term
inputs on the same hippocampal interneurons. perforant path potentiation. Furthermore, phasic
A very elegant morphological study of Miettinen locus ceruleus activation produces a delayed pro-
and Freund (1992a) has demonstrated that parv- tein synthesis-dependent long-term potentiation of
albumin-containing interneurons are innervated synaptic plasticity, suggesting a selective role in
by MSDB GABAergic afferents, but are avoided long-term memory and increases in the perforant
74

path-evoked population spike (see Chapter 10). which are likely to reside within the hippocampus.
Potentiation of the perforant path population Most of our knowledge derives from receptor
spike by noradrenaline could be the result of in- studies, characterizing the role of D1–D5 recep-
creased granule cell excitability or reduced inhibi- tors. The results of these studies indicate that in
tion from interneurons of granule cells, or both. the dentate gyrus, via the aforementioned
However, unit recording in the dentate gyrus dopamine receptor subtypes, dopamine is in-
shows that exogenous noradrenaline inhibits gran- volved in depotentiation and serves to maintain
ule cells and excites presumed inhibitory interneu- synaptic facilitation in recently potentiated path-
rons (Brown et al., 2005). On the other hand, ways. The net effect helps consolidate information
activation of a2- or b-adrenoceptors excites both storage (e.g., in Manahan-Voughan and Kulla,
the interneurons and granule cells (Brown et al., 2003).
2005). Increased inhibitory interneuron activity
and inhibition of granule cells is inconsistent with
the enhanced granule cell responsiveness and en- Commissural connections of the dentate gyrus
hanced plasticity in the dentate gyrus reported
with population recording. Thus, a critical ques- The majority of commissural fibers occupy the
tion concerns the action of noradrenaline on the inner molecular layer of the dentate gyrus (e.g.,
dentate gyrus interneurons. Blackstad, 1956). However, some commissural fib-
In the hippocampal formation, noradrenergic ers were described that do not follow the ‘‘classi-
innervation is particularly dense in areas receiving cal’’ pattern of fiber lamination and terminate in
mossy fiber inputs, including the hilus of the dent- the outer molecular layer (Deller et al., 1995,
ate gyrus and stratum lucidum of the CA3 (e.g., 1996a, b; Deller, 1998). The main postsynaptic
Moudy et al., 1993). In these two areas, the ma- targets of the commissural projection are the gran-
jority of noradrenergic varicosities do not make ule cell dendrites (e.g., Frotscher and Zimmer,
conventional synaptic contacts. However, those 1983b; Seress and Ribak 1984), but they also
that form synapses terminate on dendritic shafts innervate interneurons (Frotscher and Zimmer,
and somata of GABAergic interneurons forming 1983a; Seress and Ribak, 1984), which were shown
symmetric membrane specializations (Frotscher to contain GAD (Frotscher et al., 1984). These
and Leranth, 1988; Milner and Bacon, 1989). It interneurons do not represent a homogeneous cell
should be noted, however, that asymmetric synap- population since, they could be distinguished by
tic contacts between tyrosine hydroxylase-contain- their different neuropeptide content, e.g., vasoac-
ing (presumably noradrenergic) axon terminals tive intestinal polypeptide (Leranth and Frotscher,
and dendritic spines were also observed, in the 1983), neuropeptide Y (Deller and Leranth, 1990),
stratum lucidum of CA3 (Frotscher and Leranth, and parvalbumin (Deller et al., 1994). This ar-
1988). Symmetric (presumably inhibitory) rangement, that commissural fibers terminate on
synapses support earlier physiological studies sug- both principal and inhibitory interneurons in the
gesting that noradrenaline disinhibits hippocampal dentate gyrus, represents the anatomical basis of
pyramidal neurons by decreasing the excitability the feed-forward inhibitory mechanism elicited by
of GABAergic interneurons (e.g., Madison and stimulation of the commissural system (Buzsaki
Nicoll, 1988). and Eidelburg, 1981; Buzsaki, 1984).
In contrast to the very dense noradrenergic in- Retrograde tracing experiments revealed that
nervation of the dentate gyrus, this structure re- the majority of the cells of origin of the commis-
ceives only a minor and diffusely distributed sural system located in the hilus of the fascia den-
dopaminergic projection that arises mainly from tata (e.g., Hjorth-Simonsen and Laurberg, 1977;
the ventral tegmental area (Swanson, 1982). In Berger et al., 1980) and several groups of neurons
spite of the well-established effects of dopamine in have been identified that contribute to this fiber
hippocampus-related mnemonic functions, little is system. The majority of commissural fibers are
known about the mechanisms of these effects, axon collaterals of the glutamate-containing
75

mossy cells that are also considered as associatio- Glutamatergic innervation of the dentate gyrus
nal/commissural neurons (Ribak et al., 1985;
Scharfman and Schwartzkroin, 1988; Scharfman, During the last several decades, visualizing neu-
1992, 1995; Soriano and Frotscher, 1994). The rons that utilize glutamate as a neurotransmitter
second major population of cells projecting to the has proven to be a difficulty due to the lack of
contralateral hippocampus includes different types specific glutamatergic markers. As a result, one of
of GABAergic interneurons. The first hint of a the most influential and significant neuronal
possible contribution of GABAergic interneurons systems has remained grossly under investigated.
to the commissural projection was published by At the turn of the millennium, two independent
Seress and Ribak (1983). They showed that in the research groups discovered that a protein that
hilar area, more than 60% of neurons were GAD has previously been suggested to mediate the Na-
positive, whereas 80% of hilar neurons project dependent uptake of inorganic phosphate across
commissurally, suggesting that at least some of the plasma membrane, also transports glutamate
the hilar projection must be GABAergic. Later, into synaptic vesicles (Bellocchio et al., 2000;
this proposition was verified by experiments Takamori et al., 2000). This protein, originally
using combination of retrograde tracing and called brain-specific Na+-dependent inorganic
GAD immunochemistry, as well as anterograde phosphate contransporter (BNPI) (Ni et al.,
labeling and degeneration (Ribak et al., 1986). 1994), became the first of what are now called ve-
The GABAergic commissural projection is not sicular glutamate transporters (VGLUT1), and
homologous. Different populations of commissu- revolutionized research about the central glut-
rally projecting GABAergic cells co-express SOM amatergic system. Shortly after this groundbreak-
(Zimmer et al., 1983; Leranth and Frotscher, 1987), ing discovery, a second (VGLUT2) and a third
neuropeptide Y (Deller and Leranth, 1990), parv- (VGLUT3) vesicular glutamate transporter has
albumin (Goodman and Sloviter, 1992), or other also been identified and cloned (Aihara et al.,
markers (Leranth and Frotscher, 1987; Sloviter 2000; Bai et al., 2001; Fremeau et al., 2002; Gras
and Nilaver, 1987). However, not all neurons of a et al., 2002; Schafer et al., 2002; Takamori et al.,
given cell type have a commissural collateral. 2002). Since then, a vast amount of data has been
While most, if not all mossy cells appear to project collected that has led to a better understanding of
bilaterally (e.g., Frotscher et al., 1991; Scharfman, the brain glutamatergic circuitry, including that in
1992), only 4–5% of the SOM- (Zimmer et al., the dentate gyrus.
1983) and only 2% of the NPY-containing
neurons (Deller and Leranth, 1990) project to the
contralateral dentate gyrus. In addition to the BNPI/VGLUT1
aforementioned neurons, commissurally projecting
CA3c pyramidal neurons were also described Distribution of BNPI/VGLUT1 containing fibers
(Gottlieb and Cowan, 1973; Laurberg, 1979; and terminals in the rat dentate gyrus has been
Voneida et al., 1981). These CA3c pyramids investigated by several groups (Bellocchio et al.,
appear to have commissural collaterals, which 1998; Fremeau et al., 2001; Kaneko and Fujiyama,
arborize in the contralateral hilus, similar to their 2002; Kaneko et al., 2002). VGLUT1 fiber density
ipsilateral collaterals (Ishizuka et al., 1990; is moderate to intense in most regions of the dent-
Li et al., 1994). ate gyrus except the granule cell layer (Kaneko
Based on the existence of a GABAergic com- et al., 2002). A more detailed description has come
missural projection, Freund and Buzsaki (1996) from Bellocchio and colleagues (Bellocchio et al.,
have suggested that there is a component of direct 1998). The outer two-thirds of the molecular layer
inhibition in the feed-forward inhibitory response label more strongly for BNPI than does the inner
evoked in the dentate gyrus by commissural stim- one-third. Further examination of the CA3 region
ulation (Buzsaki and Eidelburg, 1981; Buzsaki, under higher magnification revealed coarse gran-
1984). ular labeling at the periphery of the pyramidal cell
76

layer, strongly suggestive of mossy fiber synapses. phosphate cotransporter (DNPI), having similar
Electron microscopic, immunoperoxidase labeling functions as BNPI (Hisano et al., 2000). In a
in stratum lucidum of CA3 showed prominent re- rather comprehensive study, VGLUT2-positive
action product in large axon terminals having the fibers and terminals in the rat hippocampus have
morphological characteristics of mossy fiber bou- been characterized, and their possible sources of
tons. In the hilar area of the dentate gyrus, where origin defined (Halasy et al., 2004). The highest
mossy fiber collaterals also terminate, a few large density of VGLUT2-immunoreactive boutons has
terminals similar to those in the CA3 region been observed in the inner molecular layer, while
contain BNPI immunoreactivity. Many smaller only scattered VGLUT2 positive fibers have been
terminals in the dentate gyrus that form asymmet- detected in the other areas of the dentate gyrus,
ric (Gray type I) synapses with dendritic spines such as the stratum moleculare and the hilus. In
are also labeled for BNPI. Symmetric synapses do general, VGLUT2-immunoreactive boutons estab-
not contain detectable BNPI, supporting a specific lish exclusively asymmetric synapses. Analysis of
role for the protein in excitatory transmission. In VGLUT2-positive synaptic boutons revealed that
addition, many terminals forming asymmetric the majority of postsynaptic targets in the dentate
synapses do not contain detectable BNPI, indicat- gyrus are dendritic spines, followed by dendritic
ing expression only in a subset of excitatory shafts and granule cell somata (Halasy et al.,
synapses (Bellocchio et al., 1998). 2004). These findings are consistent with the re-
The distribution of BNPI containing boutons in sults of previous studies in the rat (Fremeau et al.,
the stratum moleculare suggests that they repre- 2001; Kaneko and Fujiyama, 2002; Kaneko et al.,
sent mainly perforant path inputs from the ent- 2002; Varoqui et al., 2002).
orhinal cortex. Indeed, hybridization with the Because only low levels of VGLUT2 mRNA
BNPI probe resulted in a strong hybridization sig- expression have been observed within the hippo-
nal in neuron-enriched regions of the entorhinal campus of rats (Hisano et al., 2000; Fremeau et al.,
cortex (Ni et al., 1995). On the other hand, the 2001; Herzog et al., 2001), and the granule cell
BNPI positivity of mossy fiber-like terminals in the layer of the dentate gyrus shows no signal for
hilus and CA3 suggests that they originate from DNPI by in situ hybridization (Fremeau et al.,
granule cells. Similar to cell bodies elsewhere in the 2001), dentate VGLUT2 positive boutons are
brain, the somata of dentate gyrus granule likely to be of extrahippocampal origin. It has to
cells show no detectable BNPI immunoreactivity be noted, however, that VGLUT2 mRNA in mice
(Bellocchio et al., 1998). By contrast, a strong is concentrated in pyramidal neurons of the hip-
BNPI hybridization signal has been observed in pocampus (Bai et al., 2001), indicating the poten-
the pyramidal neurons of the hippocampus and tial for considerable species differences in the
granule cells of the dentate gyrus, suggesting that hippocampal expression of VGLUT2. In rats,
these cells synthesize the protein and then trans- Fremeau and colleagues suggest a hypothalamic
port it to their terminals (Ni et al., 1994, 1995; origin for the supragranular VGLUT2-positive
Fremeau et al., 2001; Herzog et al., 2001). In con- fiber cluster because cells in the hypothalamus that
clusion, BNPI/VGLUT1 appears to be the pri- project to this layer strongly express DNPI mRNA
mary glutamatergic marker of principal neurons, (Fremeau et al., 2001). In addition to the hypo-
including dentate granule cells (Varoqui et al., thalamus, VGLUT2 has been found in most septal
2002). neurons (Hajszan et al., 2004), and several of these
neurons are also positive for GABAergic or
cholinergic markers (Danik et al., 2005). How-
DNPI/VGLUT2 ever, deafferentation studies (fimbria-fornix tran-
section and entorhinal cortex ablation) led to no
Prior to its identification as VGLUT2 (Aihara significant differences in either the density or dis-
et al., 2000), this protein has been known as differ- tribution pattern of VGLUT2-positive boutons in
entiation-associated Na+-dependent inorganic the dentate gyrus (Halasy et al., 2004), suggesting
77

that the majority of dental VGLUT2 boutons are Regarding the source of dentate VGLUT3 fib-
of intrahippocampal origin. Indeed, light micro- ers, one possibility is an intrahippocampal origin.
scopic observation of the hippocampus of col- Although moderate levels of VGLUT3 mRNA
chicine-treated rats revealed a large number of expression have been observed in the principal
VGLUT2-immunoreactive cell bodies. In the dent- cells of pyramidal and dentate granule cell layers
ate gyrus, the hilar mossy cells represent the most in the rat (Fremeau et al., 2002), the distribution
heavily labeled population, while a small propor- of VGLUT3-immunoreactive fibers resembles
tion of granule cells in the subgranular zone are that observed with markers of GABA terminals.
also VGLUT2-positive (Halasy et al., 2004). It is Indeed, the VGLUT3 gene is expressed in scat-
well known that mossy cells in the dentate gyrus tered interneurons in the hilus of the dentate gyrus
project to the inner molecular layer (Frotscher (Fremeau et al., 2002; Gras et al., 2002; Herzog et
et al., 1991). Thus, VGLUT2-containing hilar al., 2004). A similar distribution pattern of
mossy cells may be the source of the VGLUT2 VGLUT3 mRNA has been demonstrated in mice
fiber network in the dentate gyrus. (Schafer et al., 2002). Immunoreactivity for
VGLUT3 is undetectable in pyramidal and dent-
ate granule cells (Fremeau et al., 2002; Somogyi
et al., 2004). More importantly, a partial colocal-
VGLUT3 ization of VGLUT3 and the GABA marker glut-
amic acid decarboxylase (GAD) has been observed
VGLUT3 is the only known vesicular glutamate in the hilar interneurons of the dentate gyrus by
transporter that began its ‘‘career’’ without previ- means of both in situ hybridization (Herzog et al.,
ous history as inorganic phosphate cotransporter. 2004) and immunohistochemistry (Fremeau et al.,
It has been identified in a direct search for addi- 2002). In addition, colocalization of VGLUT3
tional VGLUT molecules, and cloned (Fremeau with vesicular inhibitory amino acid transporter,
et al., 2002; Schafer et al., 2002; Takamori et al., a marker of GABAergic neurons and boutons
2002). VGLUT3-immunoreactive terminals sur- revealed numerous double-labeled nerve endings
round the granule cell layer of the rat dentate in the perisomatic terminals that contact the soma
gyrus. VGLUT3 labeling is intense in the molec- of granule cells (Herzog et al., 2004). More spe-
ular layer, corresponding to the proximal part of cifically, Somogyi and colleagues (Somogyi et al.,
the granule cell dendrites, and along the hilar bor- 2004) have shown that all VGLUT3-positive so-
der with the granule cell layer, corresponding to mata are immunoreactive for cholecystokinin, a
the zone where granule cell axons emerge. A dense marker of a subpopulation of basket cells, and
VGLUT3 network also surrounds the soma of none express markers for other interneuron types
granule cells (Fremeau et al., 2002; Gras et al., in the hippocampus. Boutons expressing
2002; Herzog et al., 2004). A similar distribution VGLUT3, cholecystokinin and GAD are most
pattern has been observed in mice (Schafer et al., abundant in the cell layers of the hippocampus
2002). In rats, immunoparticles for VGLUT3 ac- (Somogyi et al., 2004).
cumulate over vesicle clusters in terminals making Another source of dentate VGLUT3 fibers may
classical asymmetrical synapses in the hippocam- be the subcortical neurons, such as the septohip-
pus (Gras et al., 2002). However, a large number pocampal neurons, the mesolimbic dopaminergic,
of VGLUT3-positive terminals also form symmet- and raphe serotonergic cells. However, the data
rical synapses (Fremeau et al., 2002; Gras et al., obtained so far are controversial. In the rat, the
2002). Further, the labeled terminals make contact substantia nigra pars compacta and ventral teg-
with the shaft of proximal dendrites, a location mental area contain moderate levels of VGLUT3
characteristic of inhibitory synapses. However, mRNA, suggesting expression by dopamine neu-
only a subset of symmetric synapses in the hippo- rons (Fremeau et al., 2002). By contrast, no
campus appears to stain for VGLUT3 (Fremeau VGLUT3 mRNA expression has been detected
et al., 2002). in the substantia nigra by Herzog and colleagues
78

(Herzog et al., 2004), and lack of VGLUT3 ex- small number of neurons. However, the majority of
pressing neurons has also been observed in the these afferents are able to exert a powerful control
septum and the vertical limb of the diagonal band over the electric activity of the hippocampus. In
in the same study. Furthermore, a very high den- some of the afferent systems, this efficacy is due to
sity of VGLUT3 mRNA has been found in the the fact that the majority of the extrinsic afferents
dorsal raphe and MR nuclei (Fremeau et al., 2002; terminate on a relatively small, but specific popu-
Herzog et al., 2004). All of the serotonin trans- lations of neurons. These target cells include differ-
porter-positive neurons also express VGLUT3 in ent types of GABAergic interneurons, which, in
the dorsal raphe and MR. Interestingly, numerous turn have the ability to control a large number of
neurons from the dorsal raphe express VGLUT3 principal cells (Buzsaki, 1984; Freund and Buzsaki,
but no serotonin transporter (Gras et al., 2002). 1996). For example, The MSDB GABAergic
However, double-staining for VGLUT3 and the neurons seem to innervate only the GABAergic
catecholamine marker tyrosine hydroxylase or the chandelier and basket cells that have a great influ-
plasma membrane serotonin transporter revealed ence (disinhibition) on the output sector, soma and
no colocalization in the axons of the dentate axon hillock, of granule cells (Freund and Antal,
gyrus (Fremeau et al., 2002). On the other hand, 1988). Serotonergic fibers originating in the MR,
Somogyi and colleagues (Somogyi et al., 2004) also selectively innervate a specific population of
have found a population of VGLUT3 boutons in calbindin-containing GABA cells that terminate
the hippocampus that are negative for GAD but on the dendritic shafts of principal cells. Thus, in
are labeled for vesicular monoamine transporter contrast to the disinhibitory function of the MSDB-
type 2 (VMAT2), plasmalemmal serotonin trans- hippocampal GABA system, activation of the
porter or serotonin, but no colocalization has been MR-hippocampal serotonergic pathway could re-
found in terminals containing VGLUT3 and ve- sult in inhibition of the input of principal neurons
sicular acetylcholine transporter. In mice, the (Freund et al., 1990). The termination pattern of
highest VGLUT3 mRNA expression levels and SUM afferents also seems to be very specific. The
highest density of VGLUT3 mRNA-expressing overwhelming majority of fibers originating in the
cells have been observed in the raphe nuclei. In SUM form robust asymmetric synaptic contacts
contrast to its mRNA, VGLUT3 immunoreactiv- with the primary dendrites of granule cells (Leranth
ity is undetectable in the cell bodies of raphe nuclei and Nitsch, 1994; Magloczky et al., 1994; Nitsch
in vivo (Schafer et al., 2002). In the hippocampus, and Leranth, 1996) and, in addition, terminate on
VGLUT3 immunoreactivity is absent from choli- specifically positioned GABA interneurons, in a
nergic synapses. Confocal double-immunofluores- way that they could effectively filter/contrast the
cence revealed the presence of VGLUT3 in a signal flow in the trisynaptic circuit (Leranth and
subpopulation of both thick and thin VMAT2- Nitsch, 1994).
positive varicose fibers in the hippocampus, as well The functional importance of these three major
as the absence of VGLUT3 from tyrosine hydro- extrahippocampal afferent systems is highlighted
xylase positive terminals. Thus, the hippocampal by recent observations related to the mnemonic
VGLUT3/VMAT2 system most likely stems from functions of the hippocampus that are associated
the dorsal raphe and MR nuclei and represents a with synaptoplastic effects of gonadal hormones.
novel neuronal projection subsystem encoding Local estrogen administration into the MSDB
both glutamatergic and serotonergic neurotrans- (Lâm and Leranth, 2003), MR (Prange-Kiel et al.,
mission in the mouse (Schafer et al., 2002). 2004), and SUM (Leranth and Shanabrough, 2001)
of ovariectomized rats all resulted in a robust in-
crease in the density of spine synapses in the hip-
Conclusion pocampus. In contrast, transection of the fimbria
fornix, which contains the hippocampal projection
One common feature of the extrinsic dentate affer- of these subcortical structures, prevents (Leranth
ent systems is that they originate from a relatively et al., 2000) the known effects (Gould et al., 1990).
79

Acknowledgments inorganic phosphate transporter. Science, 289(5481):


957–960.
Benzi, G. and Moretti, A. (1998) Is there a rationale for the
This work was supported by NIH grants
use of acetylcholinesterase inhibitors in the therapy of
MH060858 and NS042644 to C.L. and Alzheimer’s disease? Eur. J. Pharmacol., 346: 1–13.
MH074021 to T.H., as well as by a Hungarian Berger, T.W., Semple-Rowland, S. and Basset, J. (1980) Hip-
National Office for Research and Technology pocampal polymorph neurons are the cells of origin for ipsi-
grant RET-08/04. lateral association and commissural afferents to the dentate
gyrus. Brain Res., 215: 329–336.
Bilkey, D.K. and Goddard, G.V. (1985) Medial septal facili-
References tation of hippocampal granule cell activity is mediated by
inhibition of inhibitory interneurons. Brain Res., 361:
Acsady, L., Halasy, K. and Freund, T.F. (1993) Calretinin is 99–106.
present in non-pyramidal cells of the rat hippocampus–III. Blackstad, T.W. (1956) Commissural connections of the hip-
Their inputs from the median raphe and medial septal nuclei. pocampal region of the rat, with special reference to their
Neuroscience, 52: 829–841. mode of termination. J. Comp. Neurol., 105: 417–537.
Aihara, Y., Mashima, H., Onda, H., Hisano, S., Kasuya, H., Bland, B.H. and Oddie, S.D. (1998) Anatomical, electrophys-
Hori, T., Yamada, S., Tomura, H., Yamada, Y., Inoue, I., iological and pharmacological studies of ascending brainstem
Kojima, I. and Takeda, J. (2000) Molecular cloning of a hippocampal synchronizing pathways. Neurosci. Biobehav.
novel brain-type Na(+)-dependent inorganic phosphate co- Rev., 22: 259–273.
transporter. J. Neurochem., 74(6): 2622–2625. Brito, G.N. and Brito, L.S. (1990) Septohippocampal system
Alreja, M. (1996) Excitatory actions of serotonin on GAB- and the prelimbic sector of frontal cortex: a neuropsycho-
Aergic neurons of the medial septum diagonal band of Broca. logical battery analysis in the rat. Behav. Brain Res., 36:
Synapse, 22: 15–27. 127–146.
Alvarez-Leefmans, F.J. and Gardner-Medwin, A.R. (1975) In- Brown, R.A., Walling, S.G., Milway, J.S. and Harley, C.W.
fluences of septum on the hippocampal dental area which are (2005) Locus ceruleus activation suppresses feedforward in-
unaccompanied by field potentials. J. Physiol. (Lond.), 249: terneurons and reduces beta-gamma electroencephalogram
14–16. frequencies while it enhances theta frequencies in rat dentate
Amaral, D.G. (1978) A Golgi study of cell types in the hilar gyrus. J. Neurosci., 25: 1985–1991.
region of the hippocampus in the rat. J. Comp. Neurol., 182: Buckmaster, P.S., Wentzel, H.J., Kunkel, D.D. and Schwa-
851–914. rtzkroin, P.A. (1996) Axon arbors and synaptic connections
Amaral, D.G. and Cowan, W.M. (1980) Subcortical afferents of hippocampal mossy cells in the rat in vivo. J. Comp.
to the hippocampal formation in the monkey. J. Comp. Neurol., 366: 270–292.
Neurol., 189: 573–591. Buzsaki, G. (1984) Feed-forward inhibition in the hippocampal
Amaral, D.G. and Witter, M.P. (1989) The three-dimensional formation. Progr. Neurobiol., 22: 131–153.
organization of the hippocampal formation: a review of Buzsaki, G. and Eidelburg, E. (1981) Commissural projection
anatomical data. Neuroscience, 31: 571–591. to the dentate gyrus of the rat: evidence for feed-forward
Amaral, D.G. and Witter, M.P. (1995) Hippocampal forma- inhibition. Brain Res., 230: 346–350.
tion. In: Paxinos G. (Ed.), The rat nervous system (2nd ed.). Buzsaki, G. and Eidelburg, E. (1982) Direct afferent excitation
Academic Press, New York, pp. 443–494. and long-term potentiation of hippocampal interneurons. J.
Anderson, P. and Eccles, J.C. (1962) Inhibitory phasing of Neurophysiol., 48: 597–607.
neuronal discharge. Nature (London), 196: 645–647. Carre, G.P. and Harley, C.W. (1991) Population spike facili-
Apartis, E., Poindessous-Jazat, F.R., Lamour, Y.A. and Bass- tation in the dentate gyrus following glutamate administra-
ant, M.H. (1998) Loss of rhythmically bursting neurons in rat tion to the lateral supramammillary nucleus. Brain Res., 568:
medial septum following selective lesion of septohippocampal 307–310.
cholinergic system. J. Neurophysiol., 79: 1633–1642. Chandler, J.P. and Crutcher, K.A. (1983) The septohippocam-
Bai, L., Xu, H., Collins, J.F. and Ghishan, F.K. (2001) pal projection in the rat: an electron microscopic horseradish
Molecular and functional analysis of a novel neuronal vesic- peroxidase study. Neuroscience, 10: 685–696.
ular glutamate transporter. J. Biol. Chem., 276(39): Claiborne, B.J., Amaral, D.G. and Cowan, W.M. (1986) A light
36764–36769. and electron microscopic analysis of the mossy fibers of the
Bellocchio, E.E., Hu, H., Pohorille, A., Chan, J., Pickel, V.M. rat dentate gyrus. J. Comp. Neurol., 246: 435–458.
and Edwards, R.H. (1998) The localization of the brain-spe- Clarke, D.J. (1985) Cholinergic innervation of the rat dentate
cific inorganic phosphate transporter suggests a specific pre- gyrus: an immunocytochemical and electron microscopical
synaptic role in glutamatergic transmission. J. Neurosci., study. Brain Res., 360: 349–354.
18(21): 8648–8659. Dahl, D. and Winson, J. (1986) Influence of neurons of the
Bellocchio, E.E., Reimer, R.J., Fremeau, R.T.J. and Edwards, parafascicular region on neuronal transmission from perfo-
R.H. (2000) Uptake of glutamate into synaptic vesicles by an rant path through dentate gyrus. Brain Res., 377: 211–219.
80

Danik, M., Cassoly, E., Manseau, F., Sotty, F., Mouginot, D. Fremeau, R.T., Troyer, M.D., Pahner, I., Nygaard, G.O.,
and Williams, S. (2005) Frequent coexpression of the vesic- Tran, C.H., Reimer, R.J., Bellocchio, E.E., Fortin, D.,
ular glutamate transporter 1 and 2 genes, as well as coex- Storm-Mathisen, J. and Edwards, R.H. (2001) The expres-
pression with genes for choline acetyltransferase or glutamic sion of vesicular glutamate transporters defines two classes of
acid decarboxylase in neurons of rat brain. J. Neurosci. Res., excitatory synapse. Neuron, 31: 247–260.
81(4): 506–521. Freund, T.F. (1989) GABAergic septohippocampal neurons
Deller, T. (1998) The anatomical organization of the rat fascia contain parvalbumin. Brain Res., 478: 375–381.
dentata: new aspects of laminar organization as revealed by Freund, T.F. and Antal, M. (1988) GABA-containing neurons
anterograde tracing with phaseolus vulgaris-leucoagglutinin. in the septum control inhibitory interneurons in the hippo-
Anat. Embryol., 197: 89–103. campus. Nature, 336: 170–173.
Deller, T., Katona, I., Cozzari, C., Frotscher, M. and Freund, Freund, T.F. and Buzsaki, G. (1996) Interneurons of the hip-
T.F. (1999) Cholinergic innervation of mossy cells in the rat pocampus. Hippocampus, 6: 347–470.
fascia dentata. Hippocampus, 9: 314–320. Freund, T.F., Gulyas, A.I., Acsady, L., Gorcs, T. and Toth, K.
Deller, T. and Leranth, C. (1990) Synaptic connections of NPY- (1990) Serotonergic control of the hippocampus via local in-
immunoreactive neurons in the rat hilar area. J. Comp. hibitory interneurons. Proc. Natl. Acad. Sci. U.S.A., 87:
Neurol., 300: 433–447. 8501–8505.
Deller, T., Martinez, A., Nitsch, R. and Frotscher, M. (1996a) Freund, T.F., Hajos, N., Acsady, L., Gorcs, T.J. and Katona, I.
A novel entorhinal projection to the rat dentate gyrus: direct (1997) Mossy cells of the rat dentate gyrus are immunore-
innervation of proximal dendrites and granule cell bodies of active for calcitonin gene-related peptide (CGRP). Eur. J.
granule cells and GABAergic neurons. J. Neurosci., 16: Neurosci., 9: 1815–1830.
3322–3333. Fricke, R.A. and Prince, D.A. (1984) Electrophysiology of
Deller, T., Nitsch, R. and Frotscher, M. (1994) Associational dentate gyrus granule cells. J. Neurophysiol., 51: 195–209.
and commissural afferents of parvalbumin-immunoreactive Frotscher, M. and Leranth, C. (1985) Cholinergic innerva-
neurons in the rat hippocampus: a combined immunocyto- tion of the rat hippocampus as revealed by choline ace-
chemical and PHAL study. J. Comp. Neurol., 350: 612–622. tyltransferase immunocytochemistry: A combined light and
Deller, T., Nitsch, R. and Frotscher, M. (1995) Phaseolus vul- electron microscopic study. J. Comp. Neurol., 239:
garis- Leucoagglutinin (PHAL) tracing of commissural fibers 237–246.
to the rat fascia dentata: evidence for a previously unknown Frotscher, M. and Leranth, C. (1986) The cholinergic innerva-
commissural projection to the outer molecular layer. J. tion of the rat fascia dentata: identification of target struc-
Comp. Neurol., 352: 55–68. tures on granule cells by combining choline acetyltransferase
Deller, T., Nitsch, R. and Frotscher, M. (1996b) Heterogeneity immunocytochemistry and Golgi impregnation. J. Comp.
of the commissural projection to the rat dentate gyrus: a Neurol., 243: 58–70.
Phaseolus vulgaris- Leucoagglutinin tracing study. Neurosci- Frotscher, M. and Leranth, C. (1988) Catecholaminergic in-
ence, 75: 111–121. nervation of pyramidal and GABAergic nonpyramidal neu-
Dickinson-Anson, H., Aubert, I., Gage, F.H. and Fisher, L.J. rons in the rat hippocampus. Double label immunostaining
(1998) Hippocampal grafts of acetylcholine-producing cells with antibodies against tyrosine hydroxylase and glutamate
are sufficient to improve behavioral performance following a decarboxylase. Histochemistry, 88: 313–319.
unilateral fimbria-fornix lesion. Neuroscience, 84: 771–781. Frotscher, M., Leranth, C., Lubbers, K. and Oertel, W.H.
Dunnett, S.B., Low, W.C., Iversen, S.D., Stenevi, U. and (1984) Commissural afferents innervate glutamate dec-
Bjorklund, A. (1982) Septal transplants restore maze learning arboxylase immunoreactive nonpyramidal neurons in the
in rats with fornix-fimbria lesions. Brain Res., 251: 335–348. guinea pig hippocampus. Neurosci. Lett., 46: 137–143.
Dutar, P., Lamour, Y. and Jobert, A. (1983) Acetylcholine Frotscher, M., Schlander, M. and Leranth, C. (1986) Choli-
excites identified septo-hippocampal neurons in the rat. nergic neurons in the hippocampus. A combined light- and
Neurosci. Lett., 43: 43–47. electron-microscopic immunocytochemical study in the rat.
Fantie, B.D. and Goddard, G.V. (1982) Septal modulation of Cell Tissue Res., 246: 293–301.
the population spike in the fascia dentata produced by Frotscher, M., Seress, L., Schwerdtfeger, W.K. and Buhl, E.
perforant path stimulation in the rat. Brain Res., 252: (1991) The mossy cells of the fascia dentata: a comparative
227–237. study of their fine structure and synaptic connections in
Fonnum, F. (1970) Topographical and subcellular localization rodents and primates. J. Comp. Neurol., 312: 145–163.
of choline acetyltransferase in rat hippocampal formation. J. Frotscher, M. and Zimmer, J. (1983a) Commissural fibers ter-
Neurochem., 17: 1029–1037. minate on nonpyramidal neurons in the guinea pig hippo-
Fremeau, R.T., Burman, J., Qureshi, T., Tran, C.H., Proctor, campus-a combined Golgi/EM degeneration study. Brain
J., Johnson, J., Zhang, H., Sulzer, D., Copenhagen, D.R., Res., 265: 289–293.
Storm-Mathisen, J., Reimer, R.J., Chaudhry, F.A. and Ed- Frotscher, M. and Zimmer, J. (1983b) Lesion-induced
wards, R.H. (2002) The identification of vesicular glutamate mossy fibers to the inner molecular layer of the rat fascia
transporter 3 suggests novel modes of signaling by glutamate. dentata: identification of postsynaptic granule cells by the
Proc. Natl. Acad. Sci. U.S.A., 99: 14488–14493. Golgi-EM technique. J. Comp. Neurol., 336: 170–173.
81

Gaykema, R.P., Luiten, P.G., Nyakas, C. and Traber, J. (1990) transport study with the lectin PHA-L in the rat. J. Comp.
Cortical projection patterns of the medial septum-diagonal Neurol., 229: 171–185.
band complex. J. Comp. Neurol., 293: 103–124. Harley, C.W., Lacaille, J.-C. and Galway, M. (1983) Hypo-
Givens, B.S. and Olton, D.S. (1990) Cholinergic and thalamic afferents to the dorsal dentate gyrus contain ace-
GABAergic modulation of medial septal area: effect on tylcholinesterase. Brain Res., 270: 335–339.
working memory. Behav. Neurosci., 104: 849–855. Herzog, E., Bellenchi, G.C., Gras, C., Bernard, V., Ravassard,
Givens, B.S. and Olton, D.S. (1994) Local modulation of basal P., Bedet, C., Gasnier, B., Giros, B. and El Mestikawy, S.
forebrain: effects on working and reference memory. J. (2001) The existence of a second vesicular glutamate trans-
Neurosci., 14: 3578–3587. porter specifies subpopulations of glutamatergic neurons. J.
Givens, B.S. and Olton, D.S. (1995) Bidirectional modulation Neurosci., 21: 1–6.
of scopolamine-induced working memory impairments by Herzog, E., Gilchrist, J., Gras, C., Muzerelle, A., Ravassard, P.,
muscarinic activation of the medial septal area. Neurobiol. Giros, B., Gaspar, P. and El Mestikawy, S. (2004) Locali-
Learn. Mem., 63: 269–276. zation of VGLUT3, the vesicular glutamate transporter type
Givens, B. and Sarter, M. (1997) Modulation of cognitive 3, in the rat brain. Neuroscience, 123: 983–1002.
processes by transsynaptic activation of the basal forebrain. Hisano, S., Hoshi, K., Ikeda, Y., Maruyama, D., Kanemoto,
Behav. Brain Res., 84: 1–22. M., Ichijo, H., Kojima, I., Takeda, J. and Nogami, H. (2000)
Goodman, J.H. and Sloviter, R.S. (1992) Evidence for com- Regional expression of a gene encoding a neuron-specific
missurally projecting parvalbumin-immunoreactive basket Na(+)-dependent inorganic phosphate cotransporter
cells in the dentate gyrus of the rat. Hippocampus, 2: 13–22. (DNPI) in the rat forebrain. Brain Res. Mol. Brain Res.,
Gottlieb, D.I. and Cowan, W.M. (1973) Autoradiographic 83: 34–43.
studies of the commissural and ipsilateral association con- Hjorth-Simonsen, A. and Laurberg, S. (1977) Commissural
nections of the hippocampus and dentate gyrus of the rat. I. connections of the fascia dentata in the rat. J. Comp. Neurol.,
The commissural connections. J. Comp. Neurol., 149: 174: 591–606.
393–422. Hounsgaard, J. (1978) Presynaptic inhibitory action of acetyl-
Gould, E., Woolley, C.S., Frankfurt, M. and McEwen, B.S. choline in area CA1 of the hippocampus. Exp. Neurol., 62:
(1990) Gonadal steroids regulate dendritic spine density in 787–797.
hippocampal pyramidal cells in adulthood. J. Neurosci., 10: Ishizuka, N., Weber, J. and Amaral, D.G. (1990) Organization
1286–1291. of intrahippocampal projections originating from CA3 py-
Gras, C., Herzog, E., Bellenchi, G.C., Bernard, V., Ravassard, ramidal cells in the rat. J. Comp. Neurol., 295: 580–623.
P., Pohl, M., Gasnier, B., Giros, B. and El Mestikawy, S. Kandel, E.R., Spencer, W.A. and Brinley, F.J. (1961) Electro-
(2002) A third vesicular glutamate transporter expressed by physiology of hippocampal neurons. I. Sequential invasion
cholinergic and serotoninergic neurons. J. Neurosci., 22: and synaptic organization. J. Neurophysiol., 24: 225–242.
5442–5451. Kaneko, T. and Fujiyama, F. (2002) Complementary distribu-
Gulyas, A.I., Gorcs, T.J. and Freund, T.F. (1990) Innervation tion of vesicular glutamate transporters in the central nervous
of different peptide-containing neurons in the hippocampus system. Neurosci. Res., 42: 243–250.
by GABAergic septal afferents. Neuroscience, 37: 31–44. Kaneko, T., Fujiyama, F. and Hioki, H. (2002) Immunohisto-
Gulyas, A.I., Seress, L., Toth, K., Acsady, L., Antal, M. and chemical localization of candidates for vesicular glutamate
Freund, T.F. (1991) Septal GABAergic neurons innervate transporters in the rat brain. J. Comp. Neurol., 444(1):
inhibitory interneurons in the hippocampus of the macaque 39–62.
monkey. Neuroscience, 41: 381–390. Kiss, J., Csaki, A., Bokor, H., Shanabrough, M. and Leranth,
Haglund, L., Swanson, L.W. and Kohler, C. (1984) The pro- C. (2000) The supramammillo-hippocampal and supramam-
jection of the supramammillary nucleus to the hippocampal millo-septal glutamatergic/aspartatergic projection in the rat:
formation: an immunohistochemical and anterograde trans- a combined [3H]D-aspartate autoradiographic and immuno-
port study with the lectin PHA-L in the rat. J. Comp. histochemical study. Neuroscience, 97: 657–669.
Neurol., 229: 171–185. Kiss, J., Patel, A.J. and Freund, T.F. (1990) Distribution of
Hajszan, T., Alreja, M. and Leranth, C. (2004) Intrinsic vesic- septohippocampal neurons containing parvalbumin or cho-
ular glutamate transporter 2-immunoreactive input to septo- line acetyltransferase in the rat brain. J. Comp. Neurol., 298:
hippocampal parvalbumin-containing neurons: novel 362–372.
glutamatergic local circuit cells. Hippocampus, 14: 499–509. Kosofsky, B.E. and Molliver, M.E. (1987) The serotonergic
Halasy, K., Hajszan, T., Kovacs, E.G., Lam, T.T. and Leranth, innervation of cerebral cortex: different classes of axon ter-
C. (2004) Distribution and origin of vesicular glutamate minals arise from dorsal and medial raphe nuclei. Synapse, 1:
transporter 2-immunoreactive fibers in the rat hippocampus. 153–168.
Hippocampus, 14: 908–918. Lâm, T.T. and Leranth, C. (2003) Role of the medial septum
Haglund, L., Swanson, L.W. and Kohler, C. (1984) The pro- diagonal band of Broca cholinergic neurons in estrogen-
jection of the supramammillary nucleus to the hippocampal induced spine synapse formation on hippocampal CA1 py-
formation: an immunohistochemical and anterograde ramidal cells of female rats. Eur. J. Neurosci., 17: 1997–2005.
82

Lamour, Y., Dutar, P. and Jobert, A. (1984) Septo-hippocam- In: Elliot K. and Whelan J. (Eds.), Functions of the Hippo-
pal and other medial septum diagonal band neurons: elect- campal System. Elsevier, Amsterdam, pp. 5–24.
rophysiological and pharmacological properties. Brain Res., Madison, D.V. and Nicoll, R.A. (1988) Norepinephrine de-
309: 227–239. creases synaptic inhibition in the rat hippocampus. Brain
Laurberg, S. (1979) Commissural and intrinsic connections of Res., 442: 131–138.
the rat hippocampus. J. Comp. Neurol., 184: 685–708. Magloczky, Z., Acsady, L. and Freund, T.F. (1994) Principal
Laurberg, S. and Sorensen, K.E. (1981) Associational and cells are the postsynaptic targets of supramammillary affer-
commissural collaterals of neurons in the hippocampal for- ents in the hippocampus of the rat. Hippocampus, 4:
mation (hilus fasciae dentatae and subfield CA3). Brain Res., 322–334.
212: 287–300. Manahan-Voughan, D. and Kulla, A. (2003) Regulation of de-
Lawson, V.H. and Bland, B.H. (1993) The role of the septo- potentiation and long-term potentiation in the dentate gyrus
hippocampal pathway in the regulation of hippocampal field of freely moving rats by dopamine D2-like receptors. Cereb.
activity and behavior: analysis by the intraseptal microinfu- Cortex, 13: 123–135.
sion of carbachol, atropine, and procaine. Exp. Neurol., 120: Markowska, A.L., Olton, D.S. and Givens, B. (1995) Choli-
132–144. nergic manipulations in the medial septal area: age-related
Leranth, C. and Frotscher, M. (1983) Commissural afferents to effects on working memory and hippocampal electrophysi-
the rat hippocampus terminate on vasoactive intestinal po- ology. J. Neurosci., 15: 2063–2073.
lypeptide-like immunoreactive non-pyramidal neurons: An McKenna, J.T. and Vertes, R.P. (2001) Collateral projections
EM immunocytochemical degeneration study. Brain Res., from the median raphe nucleus to the medial septum and
276: 357–361. hippocampus. Brain Res. Bull., 54: 619–630.
Leranth, C. and Frotscher, M. (1987) Cholinergic innervation Mellgren, S.I. and Srebro, B. (1973) Changes in acetylcholine-
of hippocampal GAD- and somatostatin-immunoreactive sterase and distribution of degenerating fibres in the hippo-
commissural neurons. J. Comp. Neurol., 261: 33–47. campal region after septal lesion in the rat. Brain Res., 52:
Leranth, C. and Frotscher, M. (1989) The organization of the 19–35.
septal region in the rat brain: cholinergic-GABAergic inter- Miettinen, R. and Freund, T.F. (1992a) Convergence and seg-
connections and the termination of hippocampo-septal fibers. regation of septal and median raphe inputs onto different
J. Comp. Neurol., 289: 304–314. subsets of hippocampal inhibitory interneurons. Brain Res.,
Leranth, C. and Nitsch, R. (1994) Morphological evidence that 594: 263–272.
hypothalamic substance P-containing afferents are capable of Miettinen, R. and Freund, T.F. (1992b) Neuropeptide-Y-con-
filtering the signal flow in the monkey hippocampal forma- taining interneurons in the hippocampus receive synaptic in-
tion. J. Neurosci., 14: 4079–4094. put from median raphe and GABAergic septal afferents.
Leranth, C. and Shanabrough, M. (2001) Supramammillary Neuropeptides, 22: 185–193.
area mediates subcortical estrogenic action on hippocampal Milner, T.A. and Bacon, C.E. (1989) GABAergic neurons in the
synaptic plasticity. Exp. Neurol., 167: 445–450. rat hippocampal formation: ultrastructure and synaptic re-
Leranth, C., Shanabrough, M. and Horvath, T.L. (2000) Hor- lationships with catecholaminergic terminals. J. Neurosci.,
monal regulation of hippocampal spine synapse density in- 10: 3410–3427.
volves subcortical mediation. Neuroscience, 101: 349–356. Mizumori, S.J.Y., McNaughton, B.L. and Barnes, C.A. (1989)
Leranth, C. and Vertes, R.P. (1999) Median raphe serotonergic A comparison of supramammillary and medial septal influ-
innervation of medial septum/diagonal band of Broca ences on hippocampal field potentials and single-unit activity.
(MSDB) parvalbumin-containing neurons: possible involve- J. Neurophysiol., 61: 15–31.
ment of the MSDB in the desynchronization of hippocampal Monmaur, P. and Breton, P. (1991) Elicitation of hippocampal
EEG. J. Comp. Neurol., 410: 586–598. theta by intraseptal carbachol injection in freely moving rats.
Lewis, P.R. and Shute, C.C.D. (1967) The cholinergic limbic Brain Res., 544: 150–155.
system: projections to hippocampal formation, medial cortex Mosko, S., Lynch, G. and Cotman, C.W. (1973) The distribu-
nuclei of the ascending cholinergic reticular system and the tion of septal projections to the hippocampus in the rat.
subfornical organ and supra-optic crest. Brain, 90: 521–537. J. Comp. Neurol., 52: 163–174.
Li, X.-G., Somogyi, P., Ylinen, A. and Buzsaki, G. (1994) The Moudy, A.M., Kunkel, D.D. and Schwartzkroin, P.A. (1993)
hippocampal CA3 network: an in vivo intracellular labeling Development of dopamine-beta-hydroxylase-positive fber in-
study. J. Comp. Neurol., 339: 181–208. nervation of the rat hippocampus. Synapse, 15: 307–318.
Lopes de Silva, F.H., Witter, M.P., Boeijinga, P.H. and Ni, B., Rosteck, P.R.J., Nadi, N.S. and Paul, S.M. (1994)
Lohman, A.H.M. (1990) Anatomic organization and phys- Cloning and expression, of a cDNA encoding a brain-specific
iology of the limbic cortex. Physiol. Rev., 70: 453–511. Na(+)-dependent inorganic phosphate cotransporter. Proc.
Lorente de No, R. (1934) Studies on the structure of the cer- Natl. Acad. Sci. U.S.A., 91(12): 5607–5611.
ebral cortex II. Continuation of the study of the ammonic Ni, B., Wu, X., Yan, G.M., Wang, J. and Paul, S.M. (1995)
system. J. Psychol. Neurol., 46: 113–177. Regional expression and cellular localization of the Na(+)-
Lynch, G., Rose, G. and Gall, C. (1978) Anatomical and dependent inorganic phosphate cotransporter of rat brain.
functional aspects of the septo-hippocampal projections. J. Neurosci., 15(8): 5789–5799.
83

Nicoll, R.A., Schenker, C. and Leeman, S.E. (1980) Substance Scharfman, H.E. (1992) Differentiation of rat dentate neurons
P as a transmitter candidate. Ann. Rev. Neurosci., 2: by morphology and electrophysiology in hippocampal slices:
227–268. granule cells, spiny hilar cells and aspiny ‘fast-spiking’ cells.
Nitsch, R. and Leranth, C. (1993) Calretinin immunoreactivity In: Ribak, C.E., Gall, C.M. and Mody, I. (Eds.), The Dentate
in the monkey hippocampal formation. II: Intrinsic GAB- Gyrus and its Role in Seizures: Epilepsy Research Supple-
Aergic and hypothalamic non-GABAergic systems. An ex- ments. Elsevier, Amsterdam, Vol. 7, pp. 93–109.
perimental tracing and coexistence study. Neuroscience, 55: Scharfman, H.E. (1995) Electrophysiological evidence that
797–812. dentate hilar mossy cells are excitatory and innervate both
Nitsch, R. and Leranth, C. (1994) Substance P-containing granule cells and interneurons. J. Neurophysiol., 74: 179–194.
hypothalamic afferents to the monkey hippocampus. An Scharfman, H.E. (1999) The role of nonprincipal cells in dent-
immunocytochemical, tract tracing, and coexistence study. ate gyrus excitability and its relevance to animal models of
Exp. Brain Res., 101: 231–240. epilepsy and temporal lobe epilepsy. Adv. Neurol., 79:
Nitsch, R. and Leranth, C. (1996) GABAergic neurons in the 805–820.
rat dentate gyrus are innervated by subcortical calretinin- Scharfman, H.E. and Schwartzkroin, P.A. (1988) Electrophys-
containing afferents. J. Comp. Neurol., 364: 425–438. iology of morphologically identified mossy cells of the dent-
Nyakas, C., Luiten, P.G.M., Spencer, D.G. and Traber, J. ate hilus recorded in guinea pig hippocampal slices.
(1987) Detailed projection patterns of septa1 and diagonal J. Neurosci., 8: 3812–3821.
band efferents to the hippocampus in the rat with emphasis Segal, M. (1979) A potent inhibitory monosynaptic hypotha-
on innervation of CAI and dentate gyrus. Brain Res. Bull., lamo-hippocampal connection. Brain Res., 162: 137–141.
18: 533–545. Segal, M. and Landis, S. (1974) Afferents to the hippocampus
Prange-Kiel, J., Rune, G.M. and Leranth, C. (2004) Median of the rat studied with the method of retrograde transport of
raphe mediates estrogenic effects to the hippocampus in fe- horseradish peroxidase. Brain Res., 78: 1–15.
male rats. Eur. J. Neurosci., 19: 309–317. Seress, L. (1988) Interspecies comparison of the hippocampal
Ramon y Cajal, S. (1893) Estructura del asfa de Ammon y formation shows increased emphasis on the regio superior in
fascia dentata. Ann. SOC Esp. Hist. Nat., 22. the Ammon’s horn of the human brain. J. Hirnforsch., 29:
Ramon y Cajal, S. (1911) Histologie de systeme nerveux de 335–340.
I’Homme et des vertebres tomme II. Maloine, Paris. Seress, L. and Ribak, C.E. (1983) GABAergic cells in the dent-
Ribak, C.E. and Peterson, G.M. (1991) Intragranular mossy ate gyrus appear to be local circuit and projection neurons.
fibers in rats and gerbils form synapses with the somata Exp. Brain Res., 50: 173–182.
and proximal dendrites of basket cells in the dentate gyrus. Seress, L. and Ribak, C.E. (1984) Direct commissural connec-
Hippocampus, 1: 355–364. tions to basket cells of the hippocampal dentate gyrus: an-
Ribak, C.E., Seress, L. and Amaral, D.G. (1985) The atomical evidence for feed-forward inhibition. J. Neurocytol.,
development, ultrastructure and synaptic connections of 13: 215–225.
the mossy cells of the dentate gyrus. J. Neurocytol., 14: Shute, C.C.D. and Lewis, P.R. (1966) Electron microscopy of
835–857. cholinergic terminals and acetylcholinesterase-containing
Ribak, C.E., Seress, L., Peterson, G.M., Serooggy, K.B., Fall- neurons in the hippocampal formation of the rat.
on, J.H. and Schmued, L.C. (1986) A GABAergic inhibitory Z. Zellforsch. Mikrosk. Anat., 69: 334–343.
component within the hippocampal commissural pathway. J. Sloviter, R.S. and Nilaver, G. (1987) Immunocytochemical lo-
Neurosci., 6: 3492–3498. calization of GABA-, cholecystokinin, vasoactive intestinal
Robinson, G.B. and Racine, R.J. (1984) Interactions between polypeptide, and somatostatin-like immunoreactivity in the
septodentate and perforant-path inputs to the dentate gyrus. area dentata and hippocampus of the rat. J. Comp. Neurol.,
Soc. Neurosci. Abstr., 10: 79. 256: 42–60.
Roppert, N. and Guy, N. (1991) Serotonin facilitates Soltesz, I. and Mody, I. (1994) Patch-clamp recordings reveal
GABAergic transmission in the CA1 region of the rat hip- powerful GABAergic inhibition in dentate hilar neurons.
pocampus in vitro. J. Physiol. (London), 441: 121–136. J. Neurosci., 14: 2365–2376.
Rose, G. and Schubert, P. (1977) Release and transfer of Somogyi, J., Baude, A., Omori, Y., Shimizu, H., El Mestikawy,
[3H]adenosine derivatives in the cholinergic septa1 system. S., Fukaya, M., Shigemoto, R., Watanabe, M. and Somogyi,
Brain Res., 121: 353–357. P. (2004) GABAergic basket cells expressing cholecystokinin
Schafer, M.K., Varoqui, H., Defamie, N., Weihe, E. and Erick- contain vesicular glutamate transporter type 3 (VGLUT3) in
son, J.D. (2002) Molecular cloning and functional identifi- their synaptic terminals in hippocampus and isocortex of the
cation of mouse vesicular glutamate transporter 3 and its rat. Eur. J. Neurosci., 19: 552–569.
expression in subsets of novel excitatory neurons. J. Biol. Soriano, E. and Frotscher, M. (1994) Mossy cells of the rat
Chem., 277(52): 50734–50748. fascia dentata are glutamate-immunoreactive. Hippocampus,
Scharfman, H.E. (1991) Dentate hilar cells with dendrites in the 4: 65–70.
molecular layer have lower thresholds for synaptic activation Storm-Mathisen, J. and Blackstad, T.W. (1964) Cholinesterase
by perforant path than granule cells. J. Neurosci., 11: in the hippocampal region. Distribution and relation to
1660–1673. architectonics and afferent systems. Acta Anat., 56: 216–253.
84

Swanson, L.W. (1982) The projection of the ventral tegmental Wainer, B.H., Bolam, J.P., Freund, T.F., Henderson, Z., Tot-
area and adjacent regions: a combined fluorescent and ret- terdell, S. and Smith, A.D. (1984) Cholinergic synapses in the
rograde tracer and immunofluorescence study in the rat. rat brain: a correlated light and electron microscopic
Brain Res. Bull., 9: 321–353. immunohistochemical study employing a monoclonal anti-
Takamori, S., Malherbe, P., Broger, C. and Jahn, R. (2002) body against choline acetyltransferase. Brain Res., 308:
Molecular cloning and functional characterization of human 69–76.
vesicular glutamate transporter 3. EMBO Rep., 3: 798–803. Wheal, H.V. and Miller, J.J. (1980) Pharmacological identifi-
Takamori, S., Rhee, J.S., Rosenmund, C. and Jahn, R. (2000) cation of acetylcholine and glutamate excitatory systems in
Identification of a vesicular glutamate transporter that de- the dentate gyrus of the rat. Brain Res., 182: 145–155.
fines a glutamatergic phenotype in neurons. Nature, 407: Whitehouse, P.J., Price, D.L., Struble, R.G., Clark, A.W.,
189–194. Coyle, J.T. and Delon, M.R. (1982) Alzheimer’s disease
Tork, I. (1990) Anatomy of the serotonergic system. Ann. N.Y. and senile dementia: loss of neurons in the basal forebrain.
Acad. Sci. U.S.A., 600: 9–34. Science, 215: 1237–1239.
Varoqui, H., Schafer, M.K., Zhu, H., Weihe, E. and Erickson, Wiss, J.M., Swanson, L.W. and Cowan, W.M. (1979) Evidence
J.D. (2002) Identification of the differentiation-associated for an input to the molecular layer and the stratum granulo-
Na+/PI transporter as a novel vesicular glutamate trans- sum of the dentate gyrus from the supramammillary region
porter expressed in a distinct set of glutamatergic synapses. of the hypothalamus. Anat. Embryol., 156: 165–176.
J. Neurosci., 22: 142–155. Wu, M., Hajszan, T., Xu, C., Leranth, C. and Alreja, M. (2004)
Veazey, R.B., Amaral, D.G. and Cowan, W.M. (1982) The Group I metabotropic glutamate receptor activation pro-
morphology and connections of the posterior hypothalamus duces a direct excitation of identified septohippocampal
in the cynomolgus monkey (Macaca fascicularis). II. Efferent cholinergic neurons. J. Neurophysiol., 92: 1216–1225.
connections. J. Comp. Neurol., 207: 135–156. Zimmer, J., Laurberg, S. and Sunde, N. (1983) Neuroanatom-
Voneida, T.J., Vardaris, R.M., Fish, S.E. and Reiheld, C.T. ical aspects of normal and transplanted hippocampal tissue.
(1981) The origin of the hippocampal commissure in the rat. In: Seifert W. (Ed.), Neurobiology of the Hippocampus.
Anat. Rec., 201: 91–103. Academic Press, New York, pp. 39–64.
Plate 4.3. Light micrograph (kindly provided by Dr. Attila Gulyas) shows the result of a combined anterograde tracing and calretinin
immunostaining study, in the dentate gyrus. The anterograde tracer, biotinylated dextran amine (BDA) was injected into the medial
septum diagonal band. Large boutons of BDA-containing axons form multiple, basket-like, putative synaptic contacts with the soma
of calretinin-immunoreactive (labeled with a brown diaminobenzidine reaction product) neurons. One of these cells (arrows) located in
the supragranular layer (SgL) the other is at the border between the granule cell layer (GcL) and dentate hilar area (H). Bar
scale ¼ 50 mm. (For B/W version, see page 69 in the volume.)

Plate 4.5. Light micrograph taken from a double immunostained vibratome section of the monkey dentate gyrus. Immunoreactivity
for substance P was labeled with a dark-blue Ni-diaminobenzidine reaction, while immunostaining for parvalbumin was visualized by a
brown diaminobenzidine reaction. The soma and dendrites of the parvalbumin-containing cell embedded into the granule cell layer
(GcL) is contacted by several substance P-immunoreactive axon terminals (arrows). Bar scale ¼ 10 mm. (For B/W version, see page 71
in the volume.)
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 5

The dentate mossy fibers: structural organization,


development and plasticity

Morten Blaabjerg and Jens Zimmer

Anatomy and Neurobiology, Institute of Medical Biology, University of Southern Denmark, Winslowparken 21,
DK-5000 Odense C, Denmark

Abstract: Hippocampal mossy fibers are the axons of the dentate granule cells and project to hippocampal
CA3 pyramidal cells and mossy cells of the dentate hilus (CA4) as well as a number of interneurons in the
two areas. Besides their role in hippocampal function, studies of which are still evolving and taking
interesting turns, the mossy fibers display a number of unique features with regard to axonal projections,
terminal structures and synaptic contacts, development and variations among species and strains, as well as
to normal occurring and lesion-induced plasticity and neural transplantation. These features are the topic
of this review, which will use the mossy fiber system of the rat as basis and reference in its aim to provide an
up-to-date, yet historically based guide to students in the field.

Keywords: granule cells; hippocampus; CA3; CA1; dentate hilus; lesion-induced sprouting; neural
transplantation

Introduction to the dentate mossy fiber system collaterals of the MFs in the dentate hilus also
contact several types of neurons there, deep in the
The dentate mossy fibers (MFs) are generally hilus as well as immediately below the granule cell
known to be the second link in the classical trisy- layer. The fine, filopodial processes, originally
naptic pathway from the entorhinal cortex to the identified only to extend from the
hippocampal subfield CA1, including: (1) the ento- classical, large MF terminals, have recently been
rhinal perforant path projection to dentate granule shown to make separate and specific synaptic con-
cells; (2) the dentate MF projection to CA3 py- tacts with interneurons, both within the dentate
ramidal cells; and (3) the CA3 Schaffer collateral hilus and in CA3, thereby adding new functional
projection to CA1 (Andersen et al., 1969). In line dimensions to the dentate MF system and its role
with this, MFs arise from dentate granule cells, in hippocampal function (Frotscher et al., 1994,
located in the tightly packed cell layer of the fascia 2006; Acsády et al., 1998; Henze et al., 2000).
dentata or dentate gyrus, and project their main The dentate MFs were first described by Golgi
axons through the dentate hilus (CA4) and to the (1886) and Sala (1891), and were named mossy
proximal parts of the apical (and basal) dendrites fibers by Ramón y Cajal (1893, 1911), based on the
of CA3 pyramidal cells. The extensive network of similarity of their large, irregular terminals with
the moss-like terminal varicosities of the cere-
Corresponding author. Tel.: (+45) 6550 3801; bellar MFs, which he had described earlier. Addi-
Fax: (+45) 6550 6321; E-mail: jzimmer@health.sdu.dk tional contributions, based on Golgi-impregnated

DOI: 10.1016/S0079-6123(07)63005-2 85
86

material, were made by Schaffer (1882) and 1991; Frotscher et al., 1994; Acsády et al., 1998).
Koelliker (1896) and later Lorente de Nó (1934), One rather unique and widely used method for
a student of Ramón y Cajal, who added further ‘‘bulk visualization’’ of MF terminals, and thereby
details about the synaptic nature of contacts be- the MF terminal projection, is the histochemical
tween the giant MF terminals and the complex Timm sulfide silver method (Timm, 1958; Haug,
spines emerging from the proximal dendrites of 1967; Danscher, 1981), which takes advantage of
large pyramidal cells of regio inferior (CA3) and a very high concentration of chelatable zinc in the
the so-called mossy cells of the dentate hilus MF terminals (see Danscher et al., 1985). After
(CA4). precipitation by sulfide, the chelatable zinc can be
The dentate MFs display various differences in visualized by physical development and the stained
number, trajectory and termination among species elements analyzed by both light and electron mi-
and strains and individuals within a given species. croscopy (Danscher and Zimmer, 1978; Laurberg
Here we present a guide to the structural organ- and Zimmer, 1981). After Timm staining, the MF
ization of the MF system. For the sake of sim- projection and its individual terminals stand out
plicity, we will use the common Wistar and distinctly black, in contrast to the other unstained
Sprague-Dawley laboratory rats as the standard to lightly-stained (yellowish to brown) laminar-
to which other species are compared. specific terminal fields in the hippocampus and
fascia dentata (Figs. 2–4).

Visualization of dentate mossy fibers and terminals

The axonal projection and termination of dentate Trajectory and termination of mossy fibers in
MF have been visualized by several techniques, CA3 (regio inferior)
including classical Golgi-impregnation (Schaffer,
1882; Golgi, 1886; Sala, 1891; Ramón y Cajal, Within the hippocampal region, there are about
1893, 1911; Koelliker, 1896; Lorente de Nó, 1934; one million granule cells in the rat fascia dentata
Amaral, 1978, 1979; Duffy and Rakic, 1983; (West et al., 1991), about 10 million granule cells
Soriano et al., 1983); axonal tract tracing meth- in the domestic pig (Holm and West, 1994) and
ods, like silver staining of lesion-induced antero- about 15 million granule cells in man (West and
grade axonal degeneration (Blackstad et al., 1970; Gundersen, 1990). Each cell gives rise to a thin,
Gaarskjaer, 1978a) and axonal transport of extra- unmyelinated MF (diameter 0.1–0.7 mm) (Gaarsk-
and intracellularly injected tracers (Lynch et al., jaer, 1978a; Amaral, 1979; Claiborne et al., 1986;
1973; Swanson et al., 1978; Gaarskjaer, 1981; West Acsády et al., 1998), which projects into the sub-
et al., 1982; Claiborne et al., 1986; Acsády et al., jacent dentate hilus. During its passage through
1998; Vida and Frotscher, 2000); histochemical the hilus, each MF gives off a number of branch-
and immunocytochemical staining for proteins like ing collaterals, which contact various hilar and
calbindin (Lim et al., 1997; Keuker et al., 2003) immediately subgranularly positioned neurons,
and neuropeptides like enkephalin, dynorphin, whereafter the main axons, grouped in bundles
cholecystokinin (CCK), neuropeptide Y (NPY) without intervening glia (Fig. 1) (Blackstad and
and substance P (Fitzpatrick and Johnson, 1981; Kjaerheim, 1961), continue toward the pyramidal
Stengaard-Pedersen et al., 1983; Gall, 1984, 1988; cell layer of regio inferior (CA3). Within CA3, the
Zimmer et al., 1988a, b; Holm et al., 1992, 1993); main MFs do not branch or give off collaterals as
and ultrastructural analysis by transmission elec- they traject in a laminar fashion along the pyram-
tron microscopy alone (Fig. 1) (Blackstad and idal cell layer across the transverse axis of the
Kjaerheim, 1961; Amaral, 1979) or in combination hippocampus, before bending to take a more
with the other methods mentioned above (see longitudinal course in temporal direction when
Amaral, 1979; Frotscher and Zimmer, 1983; approaching CA1 (Gaarskjaer, 1978b, 1986;
Frotscher, 1985, 1989; Frotscher and Seress, Acsády et al., 1998).
87

Fig. 1. Electron microscopy (EM) (A, B) of large MF terminals (mft) from the suprapyramidal MF layer in CA3 of the adult rat
hippocampus. The large terminals and terminal profiles form asymmetric synaptic contacts with a number of dendritic spines (sp).
Slender extension from these (spinules, s) invade the terminals. Micrographs were kindly provided by M. Frotscher. Abbreviations: a,
unmyelinated MFs; d, dendrites; mft, MF terminal; s, small dendritis spine extensions, spinules, invading MF terminals; sp, dendritic
spines. Scale bars: 2 mm (A), 1 mm (B).

Supra- and infrapyramidal bands of hippocampal hippocampi, Gaarskjaer (1978b) found that the
mossy fibers and terminals suprapyramidal MF terminals in CA3c actually
were located within the rather dispersed pyramidal
As they approach the part of the hippocampal cell layer, and that only a few infrapyramidal MF
pyramidal cell layer is closest to the hilus (CA3c), terminals were present in the corresponding
the MFs form, in classical terms, a ‘‘suprapyram- area of CA3 basal dendrites. After horseradish
idal band’’, passing along the most proximal parts peroxidase-labeling of MFs following intra- and
of the apical dendrites of CA3 pyramidal cells, and extracellular injections in different parts of the
an ‘‘infrapyramidal band’’, contacting correspond- granule cell layer of acute hippocampal slices from
ing proximal parts of the basal dendrites young adult Sprague-Dawley rats, Claiborne et al.
(Figs. 2–4A). Due to its homogeneous transpar- (1986) found that MFs from granule cells located
ency in unstained sections, the suprapyramidal at different supra- to infrapyramidal locations
MF layer is also referred to as stratum lucidum along the transverse axis of the dentate granule cell
(Ganser, 1882). The infrapyramidal MF projection layer projected into MF bundles both above and
is especially subject to variation among species within the pyramidal cell layer in CA3c. MF
and strains (Fig. 2 and below). By studying the bundles within the deep part and below the CA3c
MF system in so-called ‘‘extended hippocampi’’ of pyramidal cell layer originated from granule cells
Wistar rats, i.e., after having stretched out the located in the infrapyramidal (free or outer) blade
longitudinal septotemporal bend of dissected of the dentate. Granule cells from different
88

Fig. 2. Timm stained hippocampal sections from rat (A), guinea pig (B), cat (C), European hedgehog (D), dog (E) and man (F),
illustrating the distribution of intensely stained, black MF terminals in the different species, as well as other terminal fields and general
organization. Note species specific characteristics, like the ‘‘end bulb’’ at the terminal part of the MF layer in guinea pig (black asterisk,
B) and the layering of the dentate hilus with MF terminal-free, intermediate or plexiforme sublayer in the same species (white asterisk,
B), and MF projections into CA1 in some cats (arrows, C) and European hedgehog (D). Except for the black MF terminal staining, the
neuropil of the human hippocampus is virtually unstained, due to the use of special Timm staining procedures for non-perfused tissue
(Danscher and Zimmer, 1978). Abbreviations: CA1, hippocampal regio superior or subfield CA1; CA3, hippocampal regio inferior or
subfield CA3; FD, fascia dentata (or denate gyrus); g, dentate granule cell layer; h, dentate hilus or CA4; m, dentate molecular layer;
mf, MF layer; Sub, subiculum. Scale bars: 500 mm (See Color Plate 5.2 in color plate section.)
89

Fig. 3. Timm stained mouse hippocampal sections, illustrating strain differences in MFs terminal projections between BALB7C (A)
and C57B mice (B) and Reeler mutants (C). BALB/c mice have no infrapyramidal MFs, but a short intrapyramidal bundle (black
asterisk, A), compared to long infrapyramidal bundles in C57B mice (black asterisk, B). In Reeler mutant mice, granule and pyramidal
cell layers are disorganized, resulting in a corresponding distribution of MF terminals in the confluent granule cell layer/hilar area and
CA3. However, MFs still avoid CA1, similar to normal mice. Abbreviations: g, dentate granule cell layer; h, dentate hilus (CA4); m,
dentate molecular layer. Scale bar: 500 mm (See Color Plate 5.3 in color plate section.)

locations along the transverse axis of the granule Vaughn et al. (1977), but might also be induced
cell layer all sent MFs to the distal parts of the by strain differences in thyroid hormone levels,
suprapyramidal MF layer in CA3a, adjacent to which affect the formation of dentate granule cells
CA2/CA1 (Lorente de Nó, 1934). Granule cells and the size of the MF projection (Lauder and
located near the tip of the suprapyramidal (hidden Mugnaini, 1980; Fredens, 1981; Lipp et al., 1984;
or inner) part of the granule cell layer projected Madeira et al., 1988), or differences in the ratio
directly into the suprapyramidal MF layer, distal between the number of granule cells and target
to the hilar border, after passing through the in- CA3 pyramidal cells (Gaarskjaer, 1978a). The
tervening part of startum radiatum, as also shown difference in size of the granule cell layer between
in Ramón y Cajal’s figure 478 from a Golgi-stained the BALB/cJ and C57 mice is seen by comparing
guinea pig hippocampus (Ramón y Cajal, 1911). Fig. 3A and B. Systemic application of the cell
Having considered the presence and composi- proliferation inhibitor methylazoxymethanol-
tion of supra-, intra- and infrapyramidal MF bun- acetate (MAM) to pregnant rats during the late
dles in primarily Wistar and Sprague-Dawley rats, fetal period, when the majority of hippocampal py-
it should be noted that considerable differences ramidal cells are formed, induces larger than normal
exist between various strains of rats (Dimitrieva infrapyramidal MF projections from the virtually
et al., 1993) and mice. BALB/cJ mice have no normal sized fascia dentata into a smaller than
infrapyramidal MF bundles, but a single intrapy- normal CA3. In contrast, infrapyramidal MFs are
ramidal bundle (Fig. 3A), whereas SM/J mice almost absent, when MAM is administered to early
have a distinct infrapyramidal, but no intrapy- postnatal pups during the time when most dentate
ramidal MF bundle (Barber et al., 1974). These granule cells develop. This is likely to be due to the
mouse strain differences have been linked to fact that MAM reduces the number of the granule
different patterns of pyramidal cell generation cells relative to the virtually normal numbers of
(SM/J ¼ inside-out; BALB/cJ ¼ outside-in) by CA3 pyramidal cells (unpublished observations).
90

Fig. 4. Timm stained MF terminals (black) in stratum lucidum of CA3 (mf, A) and dentate hilus (B) in semi thin plastic section from
adult rat. (A) In CA3, the densely stained MF terminals in the suprapyramidal MF layer (mf) are usually large and contact CA3
pyramidal cell dendrites, but there are more homogeneous, lighter stained MF terminals in similar locations. A small group of MF
terminals (imf) are present below the pyramidal cell layer (p). Lighter, Timm stained elements in stratum oriens (o) and stratum
radiatum (r) are terminals of the commissural/associational projections to these layers. (B) In the hilus, a large mossy cell body (mc) is
seen in contact with MF terminals, which are of variable size and also contacting other dendritic structures. Scale bars: 25 mm (A);
15 mm (B).

Both the width of the suprapyramidal layer and of MF terminal fields corresponded to a reduction
the extent of the infrapyramidal MF layer in CA3 in the ratio between the number of granule cells
vary along the longitudinal (septotemporal) axis and CA3 pyramidal cells moving from septal to
of the rat hippocampus (Gaarskjaer, 1978a, b). At more posterior and temporal levels (Gaarskjaer,
septal levels, MF terminals form an infrapyrami- 1978a). Toward the temporal end of the hippo-
dal layer, which extends along the deep side of the campus, the border between the suprapyramidal
CA3 pyramidal cell layer to the CA2-CA1 transi- MF and CA3 pyramidal cell layer becomes
tion as an almost equal counterpart to the supra- ‘‘fuzzy,’’ and a small portion of intrapyramidal
pyramidal MF layer. Many MF terminals are also MFs reappear distally (Gaarskjaer, 1978a, b).
located within the pyramidal cell layer itself, which
at these levels is wide and loosely packed. Moving Transverse and longitudinal trajectories of mossy
temporally, the intrapyramidal and infrapyramidal fibers in CA3
MF terminals disappear first from the distal
(CA3a) and intermediate (CA3b) parts of CA3. While projecting along the more proximal parts of
In parallel, the suprapyramidal terminal zone CA3 (CA3c and adjacent parts of CA3b), rat MFs
becomes relatively thin. The reduction in the size deviate only slightly in temporal direction relative
91

to the level of origin (Blackstad et al., 1970; subfield, named CA2, and characterized by a
Gaarskjaer 1978b, 1981; Claiborne et al., 1986; mixed content of small CA1-like pyramidal cells
Amaral and Witter, 1989; Tamamaki and Nojyo, and larger CA3-like pyramidal cells, which, how-
1991), remaining within a septotemporal span or ever, were devoid of complex spines and MF input
lamina of 1.1 mm (Acsády et al., 1998) or less (see also Tamamaki et al., 1988; Barthesaghi and
(Gaarskjaer, 1981). As the MFs enter more distal Ravasi, 1999). The lack of MF projection to CA1
parts of CA3 (distal CA3b and CA3a) facing CA1, was confirmed by lesion-induced anterograde ax-
the fibers do, however, bend in the temporal di- onal tracing (rat, Blackstad et al., 1970), antero-
rection to project for additional 1–2 mm along the grade axonal transport of injected tracers, like
longitudinal axis of the hippocampus (Gaarskjaer, horseradish peroxidase (rat, Lynch et al., 1973;
1978, 1981; Amaral and Witter, 1989). MFs orig- West et al., 1982; Claiborne et al., 1986), isotope
inating at septal levels and from the tip of the labelled amino acids (rat, Swanson et al., 1978)
suprapyramidal granule cell layer at other levels and biocytin (man, Lim et al., 1997; rat, Acsády
display the longest temporal descent (Gaarskjaer, et al., 1998). The same conclusion was drawn from
1978b, 1981; Swanson et al., 1978; Tamamaki visualization of the MF terminal projection by
and Nojyo, 1991; Acsády et al., 1998). After 3- Timm staining in rat (Fig. 2A) (Haug, 1974), sev-
dimensional reconstruction of seven biocytin- eral strains of mice (Fig. 3) (Barber et al., 1974;
injected granule cells from dorsal levels of the Fredens, 1981), guinea pig (Fig. 2B) (Blackstad,
fascia dentata in adult Sprague-Dawley rats, 1963; Geneser-Jensen et al., 1974), rabbit (Hjorth-
Acsády et al. (1998) measured the MF axon in Simonsen, 1977), dog (Fig. 2E), domestic pig
CA3 to be in avarage 3.24 mm long, of which (Holm and Geneser., 1991b) and man (Fig. 2F)
0.80 mm was in CA3c, 1.05 mm in CA3b and (Danscher and Zimmer, 1978; Cassell and Brown,
1.39 mm in CA3a. 1984). The continuation of tufts of Timm stained,
In guinea pigs, a rather abrupt turn of MFs thread-like elements with minor terminal-like
followed by a steep, longitudinal descent in tem- swellings from the suprapyramidal MF layer of
poral direction takes place just close to and within CA3 into CA2 in several species both confirms
the so-called ‘‘end bulb’’, which is a characteristic that CA2 acts as a transitional zone and that MFs
bulge on the very distal part of the suprapyramidal do not reach into CA1.
MF layer in this species, as demonstrated very A dogmatic generalization of the absence of an
clearly by Timm staining (Fig. 2B). MF projection to CA1 to all species, was, how-
ever, proven to be wrong by the discovery that
MFs occasionally project into CA1 in some do-
Mossy fibers projections to CA1 (regio superior) mestic cats (Fig. 2C) (Laurberg and Zimmer, 1980)
and is normal in European hedgehogs (Fig. 2D)
Based on Golgi-impregnated brain sections from (Gaarskjaer et al., 1982; West et al., 1984). In the
rabbit, guinea pig, macaque and rodents, Ramón y occasional cat with MFs in CA1, the projection
Cajal (1893) and Lorente de Nó (1934) described was variable, being most extensive or only present
the large, irregular MF terminals as ‘‘boutons en at septodorsal levels (Fig. 2C). The MF projection
passant’’ which contacted the proximal dendrites to CA1 in European hedgehogs was present at
of the large pyramidal cells of regio inferior (CA3), midposterior (Fig. 2D) to temporal levels only,
and characterized them by the presence of ‘‘thorny or along the entire septotemporal extent of the
excrescences’’ or complex spines. In particular, hippocampus.
Lorente de Nó (1934) noted that MFs did not By demonstrating an overlap of MF terminal-
proceed to contact the smaller pyramidal cells in like, Timm stained elements and pyramidal cells
regio superior (CA1). Located between the two with small cell nuclei corresponding to CA1
subfields is, in most species, a narrow transitional pyramidal cells, Gaarskjaer (1986) later provided
area or ‘‘Mischzone’’ (Doinikow, 1908), which evidence that the most temporal levels of the rat
Lorente de Nó (1934) defined as a separate CA1 also normally receive a MF projection, a
92

phenomenon which might therefore appear to be


rather common among species.
Distinct guidance or stop cues for MF innerva-
tion of CA1 have not been identified so far. Cer-
tainly an increased number of granule cells relative
to target CA3 pyramidal cells at septal levels in
rodents does not by itself cause MFs to enter CA1,
even that this is paralleled by a more extensive
infrapyramidal termination in CA3 (see above),
and also may be a causal factor behind the more
extensive MF projections into CA1 in some cats
at septodorsal levels (Fig. 2C) (Laurberg and
Zimmer, 1980). Occasional observations of MF
projections to CA1 in rats with early postnatal
lesions of the CA3 to CA2/CA1 transition area
suggest that cells in this area act as a barrier
(Fig. 5; see also Cook and Crutcher, 1985). Clearly
cells of the CA2 subfield have been shown to differ
from the surrounding CA3 and CA1 subfields
both with regard to neuronal birthdate (Angevine,
1965; Bayer, 1980) and gene expression profile
(Lein et al., 2005).

Mossy fiber connections in the dentate hilus (CA4)

The cytoarchitectural organization of the dentate


hilus displays significant differences among spe-
cies, ranging from a laminar organization in the Fig. 5. Timm stained section from dorsoposterior level of
guinea pigs (Geneser-Jensen et al., 1974) and pigs young adult, rat hippocampus. When newborn, the rat received
a mechanical lesion of the CA3–CA1 transition (les) as well a
(Sus scrofu domesticus) (Holm and Geneser, 1991)
transection of the entorhinal perforant path to the fascia den-
to an almost confluent mixture of cell types in ro- tata. The CA3–CA1 lesion induced an aberrant ingrowth of
dents (Blackstad, 1956; Amaral, 1978). Observa- MFs into CA1, illustrated by presence of small, Timm stained
tions of the hilar distribution of MF collaterals MF terminals (asterisk) along the CA1 pyramidal cell layer.
and terminals in one species may accordingly not The removal of the perforant path projection to the outer mo-
lecular layer (m) induced spread of associational-commissural
apply to another species, as exemplified by laminar
projections from inner part of the layer, and induction of sup-
distribution of large and small Timm stained MF ragranular MF terminals (arrow) (see Zimmer, 1973, 1974).
terminals in the dentate hilus of guinea pig Abbreviations: g, granule cell layer. Scale bar: 500 mm (See Color
(Fig. 2B) compared to the confluent distribution Plate 5.5 in color plate section.)
in the dentate hilus of rat and European hedgehog
(Fig. 2A and D). Another obstacle to comparisons MF axons from the two cells give off several ex-
of dentate hilar cell and neuropil layers between tensively branching collaterals, starting at the level
species is the use of different nomenclatures. of the layer of polymorphic cells and further along
As one of the first illustrations of the exten- their trajectory toward the hippocampal pyrami-
sive MF projections within the dentate hilus, dal cell layer. Primarily based on observations in
(Koelliker, 1896, figure 786; also reproduced by Golgi-stained sections of one-week- to one-month-
Gaarskjaer, 1986) depicted two Golgi-impregnated old rabbits, guinea pigs and mice, Ramón y Cajal
granule cells from a three-day-old cat. The main (1893, 1911) described and illustrated the emission
93

of 4–8 collaterals from main MF axons just below all species examined, presumably contacting the
the granule cell layer in what he called the plexi- pyramidal-shaped basket cells in the limiting sub-
form layer (or outer part of the polymorphic cell zone (cf. Ribak and Peterson, 1991). Among the
layer), as well as deeper into the hilus close to the studied collateral networks only one collateral was
hippocampal pyramidal layer from where they re- reported to enter the granule cell layer and none to
turned back into the hilus. Referring primarily to reach the dentate molecular layer. This contrasts
Ramón y Cajal’s well-known schematic diagram to the observations of several such MF collateral
of hippocampal connections (Figure 479 in terminals in Timm staining in normal rats
Ramón y Cajal, 1911), the MF collateral projec- (Fig. 6A) and guinea pigs (Laurberg and Zimmer,
tions in the (rodent) hilus were, however, long 1981; Sloviter et al., 2006), as dealt with below.
taken to be of limited extension, until Soriano
et al. (1983), also using Golgi-impregnation,
described extensive MF collateral projections en- Timm staining and species differences, including
gaging major parts of the rat dentate hilus. This dentate hilus
spurred Claiborne et al. (1986) to trace MFs and
collaterals from single and small group of granule ‘‘Bulk’’ visualization of the zinc-rich MF terminals
cells by intracellular and small extracellular injec- by the Timm method, which intensely stains both
tion of horseradish peroxidase in acute hippocam- the characteristic large terminals of the main fibers
pal slices from young adult Sprague-Dawley rats. in CA3 and the corresponding large terminals as
Later Acsády et al. (1998) combined in vivo intra- well as the smaller MF collateral terminals in the
cellular biocytin injections of single granule cells in dentate hilus, have been performed for several
the dorsal fascia dentata of adult Sprague-Dawley species, including rat (Figs. 2A and 4A, B) (Haug,
rats with EM immunocytochemistry and 3D re- 1974; Gaarskjaer, 1978b; Laurberg and Zimmer
consruction to provide the most comprehensive 1981; Dimitrieva et al., 1993), several strains of
and detailed single study of the trajectory and mice (Fig. 3) (Barber et al., 1974; Fredens, 1981,
synaptic connections of single MF main axons and including C57/BL/6J, DBA/2J, NMRI, BALB/c/
collaterals so far, including description of post- A, C3H/Tif, A/J, AKR/A mouse strains), guinea
synaptic target cells. pig (Fig. 2B) (Blackstad, 1963; Geneser-Jensen et
According to the combined observations of al., 1974), rabbit (Hjorth-Simonsen, 1977), Euro-
Claiborne et al. (1986) and Acsády et al. (1998) pean hedgehog (Fig. 2D) (Gaarskjaer et al., 1982;
in the rat, each MF main axon emits between 5 West et al., 1984), cat (Fig. 2C) (Laurberg and
and 12 collaterals, less than 0.2 mm thick, at var- Zimmer, 1980), dog (Fig. 2E), domestic pig (Holm
ious distances along its course through the hilus. and Geneser, 1991b), and man (Fig. 2F) (Danscher
Including secondary and tertiary branches, the to- and Zimmer, 1978; Cassell and Brown, 1984). At
tal length of the collateral plexus formed by one low magnification, the Timm staining pattern re-
MF amounts to approx. 2.3 mm (Claiborne et al., veals both the general similarities and the differ-
1986). In the transverse plane, collaterals from a ences between the species in the both the
single MF cover up to one third of the total hilar hippocampal and the hilar MF projection.
extent nearest the location of the parent granule Within the dentate hilus one difference among
cell, with a total transverse extent of 0.4–0.7 mm, species is a more distinct layering of hilar neurons
depending on the location of the parent granule (not illustrated) and Timm stained MF terminals
cell. Along the longitudinal axis of the dentate hi- in guinea pig (Fig. 2B), pig (Holm and Geneser,
lus, the spread of collaterals never seems to exceed 1991b) and in part man (Fig. 2F), as compared to
0.6 mm, with 90% of the collateral network being other species like rat (Fig. 2A). The mentioned
within a 0.4 mm high lamina (Acsády et al., 1998). species are thus characterized by having a virtually
Regarding relations to the granule cell layer, in- MF terminal free zone (corresponding to Ramón
dividual collaterals running close to the deep bor- y Cajal’s (1911) intermediate or plexiforme sub-
der of the granule cell layer have been described in layer), which is located between a narrow irregular
94

Fig. 6. A: Timm stained MF (collateral) terminals, found in normal adult rats to accompany dendrites (arrows), arising from neurons
in the limiting subzone just below the dentate granule cell layer or in the dentate hilus, and extending into and sometimes through the
dentate granule cell layer (g) into the, commissural-associational terminal zone (c/a). B: Timm stained, dense supragranular MF
terminal projection (sgr) found normally at temporal levels of the cat fascia dentata as well as in other species. h, dentate hilus.
Scale bar: 50 mm (A); 100 mm (B). (See Color Plate 5.6 in color plate section.)

band of MF collateral terminal staining just below infrapyramidal blade (Haug, 1974; Laurberg and
the granule cell layer (Ramón y Cajal’s limiting Zimmer, 1981; Sloviter et al., 2006), but become
subzone) and a more deeply located, densely more widespread and numerous at temporal levels
stained layer containing large and small MF ter- (Gaarskjaer, 1978a). Intra- and supragranular MF
minals (Ramón y Cajal’s deep polymorphic cell terminals can be very prominent at temporal
layer). levels, as shown in Fig. 6B for the cat.
Aberrant growth of supragranular MF termi-
nals into the commissural-associational terminal
Intra- and supragranular mossy fiber collaterals — zone, which normally cover the inner approxi-
a normal feature and subject to plasticity mately one third of the dentate molecular layer,
can be induced: (a) by sufficiently strong denerva-
Focusing on the limiting subzone, it is possible in tion of the dentate molecular layer in developing
all species at high magnification to observe small, and adult animals (Fig. 5) (Zimmer, 1973, 1974;
presumably MF collateral terminals extending Laurberg and Zimmer, 1981; Frotscher and
along dendrites from the hilus into the granule Zimmer, 1983), including excitotoxic loss of hilar
cell layer and sometimes as far as the molecular mossy cells (see below, Sloviter et al., 2006); (b) in
layer. In the rat, these intra- and supragranular intracerebral rat dentate allografts and mouse
Timm-positive terminals are most frequent at the dentate xenografts (Sunde and Zimmer, 1981;
dentate crest (Fig. 6A) and toward the tip of the Jensen et al., 1984; Frotscher and Zimmer, 1986;
95

Fig. 7. (A) Timm stained section from midposterior level of 5-day-old rat, demonstrating a distinct Timm stained MF layer in CA3
and the hilus at this age (arrows). (B, C) Parallel sections of hippocampal slice culture, derived from 7-day-old rat and grown for
3 weeks. Timm stain reveals the distribution of aberrant supragranular and normal CA3 MFs (arrow and mf, respectively, in B) and
thionine cell staining to visualize general cellular organization (C). Abbreviations: g, granule cell layer. Scale bars: 500 mm (See Color
Plate 5.7 in color plate section.)

Zimmer et al., 1988a, b), depending on the density (1998) extended the historical perspective by show-
of afferent host brain fiber innervation; (c) in ing that each MF makes approx. 120–150 excita-
developing hippocampal tissue slices grown as tory synaptic connections with predominantly
organotypic explant cultures (Fig. 7B) (Zimmer inhibitory interneurons via axon collaterals in the
and Gahwiler, 1984); and (d) in animals with dentate hilus, 7–12 excitatory, large terminal
pilocarpine- or kainic acid-induced seizures synaptic contacts with hilar mossy cells and
(Sutula et al., 1988; Mello et al., 1993; Frotscher 11–18 corresponding contacts with proximal py-
et al., 2006; Sloviter et al., 2006) and in epileptic ramidal cell dendrites during the transverse MF
patients (Sutula et al., 1989; Blümcke et al., 2000). trajectory in CA3, plus an additional number of
This appears to be a response to loss of dentate terminal contacts made by the more distal and
hilar mossy cells normally projecting to the dentate longitudinally running parts of the MFs. These
molecular layer. Transection of the main MF pro- two studies and a number of related ones (see
jection to CA3 will not by itself induce excessive Frotscher et al., 2006 and references therein) have
supragranular MF collateral projection, excluding changed the classical view of dentate-CA3 MF
an axonal pruning or compensatory sprouting connectivity from primarily being a direct synaptic
effect (Laurberg and Zimmer, 1980, 1981). interaction between the excitatory, large MF ter-
minals and the CA3 pyramidal cells and hilar
mossy cells, into a pathway with a prominent
Structure of synaptic mossy fiber connections in the indirect inhibitory effect not only in the dentate
dentate hilus and CA3 hilus, but also in CA3, apparently acting to refine
the activation of CA3 pyramidal cells.
As outlined by Ramón y Cajal (1911), refined by The findings were not entirely surprising, how-
Lorente de Nó (1934) and demonstrated by EM ever, because short, thick diverging processes and
(Blackstad and Kjaerheim, 1961), typical MF the thin, up to 30 mm-long filamentous or filopo-
synapses are made by the large boutons en pass- dial processes had been noted by Ramón y Cajal
ant in contact with the complex spines (thorny ex- (1911) as extending from the angles of the large,
crescences) found on the proximal dendrites of the irregularly shaped MF terminals in the dentate
CA3 pyramidal cells (Fig. 1AB) and hilar mossy hilus and CA3. These extensions appear to con-
cells, as well as by MF collateral terminals con- stitute a separate, major presynaptic MF element,
tacting different types of neurons in the dentate contacting defined types of inhibitory GABAergic
hilus. Claiborne et al. (1986) and Acsády et al. and peptidergic interneurons in the dentate hilus
96

and the CA3 (Amaral, 1979; Claiborne et al., 1986; section about Timm staining, above). Interest-
Frotscher et al., 1994; Acsády et al., 1998; Henze ingly, dentate granule cells and MFs also express
et al., 2000; Vida and Frotscher, 2000, and refer- both glutamic acid decarboxylase 67 and GABA,
ences therein). Based on the data provided by particularly after seizures (Sandler and Smith,
Acsády et al., 1998, the synaptic contacts made 1991; Sloviter et al., 1996).
by a single MF can accordingly be grouped into Each large terminal synaptic complex includes
contacts with hilar mossy cells (made by as few as up to 35 synaptic membrane specializations or ac-
7–12 large terminals), CA3 pyramidal cells (11–18 tive zones (Fig. 1) as reported by Chicurel and
large terminals), hilar interneurons (as many as Harris (1992), which together with the high density
120–150 small terminal contacts) and interneurons and composition of presynaptic sodium and cal-
in CA3 (40–50 filopodial contacts). cium ion channels provides each terminal with a
Regarding the ultrastructural appearance of the large excitatory drive on the postsynaptic CA3
MF synapses, the well known, classical type is pyramidal cells (see review by Bischofberger et al.,
made by the 4–10 mm large boutons en passant, or 2006).
in several cases attached to the main MF axon by a While the large MF terminals normally appear
short thin extension (Fig. 1) (Ramón y Cajal, 1911; to exclusively make synaptic contact with CA3
Blackstad and Kjaerheim, 1961; Amaral and Dent, pyramidal cells and hilar mossy cells (Acsády
1981). In the hilus these terminals contact the hilar et al., 1998), other presynaptic MF terminals con-
mossy cells. In CA3 they occur at intervals of tact interneurons in hilus and CA3. These presy-
approx. 135 mm along the main MF, which has naptic elements are smaller, do not have multiple
been estimated to be every 6th or 7th of the CA3 active zones, and are located en passant or at the
pyramidal cell that it meets along its trajectory end of short or long thin filopodia extending from
(Claiborne et al., 1986). the giant MF boutons (Acsády et al., 1998), often
The contact between the giant MF terminals up to nine per giant bouton (Amaral, 1979;
and the ‘‘thorny excresences’’ on CA3 pyramidal Amaral and Dent, 1981; Frotscher, 1985). In de-
cells and hilar mossy cells are structurally intimate veloping rats, some filopodia may extend as far
and complex, with spinules projecting from as 30 mm from the giant boutons in the hilus,
the excresences into the terminals (sp, Fig. 1) but shorten to approximately 12 mm in adult rats
(Blackstad and Kjaerheim, 1961; Amaral and (Amaral, 1979).
Dent, 1981; Ramón y Cajal, 1911). The terminals
themselves are anchored to the dendritic shaft of
CA3 pyramidal cells and to each other by puncta New emerging structural plasticity of individual
adhaerentia. mossy fiber terminals
Within the giant MF terminals (Fig. 1), three
different types of vesicles have been described. The Based on the structural data on individual MF
majority are small clear vesicles (40 nm) contain- axonal and terminal organization provided by the
ing glutamate (Storm-Mathisen, 1981), but the studies cited above, and advanced imaging tech-
boutons also contain large dense-core vesicles niques employed on in vivo hippocampal tissue and
containing neuropeptides such as dynorphin organotypic hippocampal slice cultures derived
(McGinty et al., 1983), enkephalin (Commons from transgenic mice expressing membrane tar-
and Milner, 1996), cholecystokinin (Chandy et al., geted green fluorescent protein (GFP) in few neu-
1995), neuropeptide Y (Gall et al., 1990) and rons (Thy1-mGFPs ), Caroni and coworkers have
neurokinin-B (Schwarzer et al., 1995) and large engaged in detailed studies on activity dependent
clear vesicles (200 nm; Amaral and Dent, 1981; plasticity of individual MF termination in CA3
Chicurel and Harris, 1992; Henze et al., 2000). (De Paola et al., 2006; Galimberti et al., 2006).
Besides glutamate and neuropeptides, MF bou- Results obtained so far both in vivo and in long
tons also contain the neuromodulator ATP/ term slice cultures have shown that one large MF
adenosine (Terrian et al., 1989) and zinc (see terminal per MF (and only one) typically is larger
97

than the others, and that this one, besides of hav- more of the granule cells form after birth, and
ing elongated along its postsynaptic CA3 pyram- predominantly during the first two postnatal
idal cell dendrite in complex structural contact weeks, with ongoing, although declining forma-
with complex spines, also can have side branches tion of new granule cells into adulthood and old
with other large, satellite MF terminals, contacting age (Bayer, 1982; Cameron et al., 1993; Kuhn
the same and neighboring pyramidal cells, thereby et al., 1996). Neurogenesis is regulated by genetic
forming an local and interconnected, large MF (Kempermann et al., 1997, 2006) and epigenetic
terminal cluster or complex. The formation of in- factors, including normal activity dependent ones
dividual larger MF terminals and new satellite and seizures (Gould et al., 1998; Kempermann
large MF terminals was found to be activity- et al., 1998; Derrick et al., 2000; Gage, 2000;
dependent, regulated by age and housing in enriched Blümcke et al., 2001; Song et al., 2002; Lehmann
environments. With increasing age single MF ter- et al., 2005; Thom et al., 2005; McCloskey et al.,
minals typically attained larger size by elongating 2006, and references therein). Before dealing with
along the postsynaptic CA3 pyramidal cell dend- the ontogenic development of the MF projection
rite, at the same time expressing more presynaptic to the dentate hilus and CA3, it should be kept in
release sites and engaging in more complex struc- mind that late-forming dentate granule cells also
tural relations with the increasingly complex have been shown to extend MFs into the hilus and
postsynaptic spines developing during the same CA3 (Stanfield and Trice, 1988; Hastings and
period. Placing both young adult and older adult Gould, 1999; van Praag et al., 2002) as well as
mice in enriched environments for a period of time becoming functionally integrated, as exemplified
on the other hand made the individual large MF by the studies of Jessberger and Kempermann
terminals respond by formation of large MF ter- (2003), Santarelli et al. (2003), Bruel-Jungerman
minal satellites. The mechanisms behind the MF et al. (2005) and Raineteau et al. (2006).
plasticity elicited by select large MF terminals in a
focal manner, as demonstrated by Caroni and co-
workers, and its role in dentate-hippocampal func- Development of MF axonal and synaptic
tional interaction are not known at present. It is, connections
however, tempting to believe that the focal
expression and strengthening of synaptic contacts The developmental growth of MFs into the dent-
include a postsynaptic feedback from those CA3 ate hilus and CA3 with accompanying terminal
pyramidal cells, which at the time of induction are and synaptic maturation have been studied in
activated by a net excitatory effect of the direct Wistar and Long Evans rats, mice and rabbits by
stimulatory and indirect inhibitory MF input. Timm staining or modifications thereof (Zimmer
and Haug, 1978; Gaarskjaer, 1985; Sanchez-
Andres et al., 1997; Slomianka and Geneser,
Development of the dentate mossy fiber system 1997), supplemented by retrograde axonal tracing
(Gaarskjaer, 1985), Golgi impregnation and EM
Dentate neurogenesis in brief (Amaral and Dent, 1981) or just by EM (Stirling
and Bliss, 1978). In rats, the first sign of MF de-
The dentate MF system has a unique developmen- velopment by Timm staining appears at the day of
tal profile relative to other central nervous projec- birth, where they appear in the hilus subjacent to
tions, due to the late start and thereafter extended the early developed parts of the suprapyramidal
neurogenesis of the parent dentate granule popu- granule cell layer, and just above the proximal,
lation, lasting the entire lifetime (Altman and Das, CA3c pyramidal cell layer (at temporal and mid
1965; Angevine, 1965; Bayer, 1980; Eckenhoff and septo-temporal levels). Development at septal
Rakic, 1988; Kuhn et al., 1996; Eriksson et al., levels seems to be relatively delayed. Within the
1998; Gould et al., 1998; Dawirs et al., 2000; Gage, next 3–5 days, a full-length Timm-stained supra-
2002; Guidi et al., 2005). In rodents, two third or pyramidal MF layer develops, corresponding to
98

the formation of more granule cells and a widening the average number of symmetrical membrane
and infrapyramidal (medial) extension of cell layer junctions (punctae adhaerentia) to seven per ter-
(Fig. 7A). At this time, Golgi staining revealed minal profile at day 21. Asymmetrical synapses
dentate granule cells with a relatively mature ap- at the same time reached the adult level of an
pearance, especially in the suprapyramidal blade, average number of nine per terminal profile,
with MF axons extending into the hilus and send- counted in two dimensions only. It should also
ing collaterals to form an infragranular plexus be recognized that, after the average number of
(Amaral and Dent, 1981). By postnatal day 3, invaginating spinules per terminal profile increased
bundles of MFs were present above the CA3 py- from two at day 21 to five in more than one-year-
ramidal cell layer. By retrograde axonal tracing old rats.
following microinjection of the fluorescent dye As illustrated above, the rat MF system is sub-
True Blue in CA3, covering different proximo- ject to considerable development during the first
distal distances from the hilus, Gaarskjaer (1985) three postnatal weeks and even thereafter, and this
demonstrated a rapid elongation of MF along the should of course be taken into consideration in
transverse axis of CA3 during the very first post- experimental and other studies of this system.
natal week, with some retrograde labeling of gran- Specific points to note may be that MF synapse
ule cells in the suprapyramidal layer from the most formation precedes the development of thorny
distal CA1 close part of CA3 (as early as postnatal excrecenses on CA3 pyramidal cells. Moreover,
day 1). The rate of MF outgrowth during the first asymmetrical synapses seem to develop at a steady
days was estimated to be at least 0.2 mm at day 1, pace, unrelated to the dramatic structural consol-
and nearly 1 mm at day 4. Retrograde tracing idation expressed by a prominent increase in the
showed a continued growth along the septotem- non-synaptic puncta adhaerentia from day 14 to
poral axis with a septotemporal descent relative to day 21.
the parent cell location of about 0.5 mm at day 1,
reaching nearly 1 mm at day 4 and adult length
of 1.6 mm during the next 3–4 weeks. After 10–
12 days postnatal growth appeared to slow. Experimental studies on structural and connective
Ultrastructurally, MF boutons identified at day 1 plasticity of dentate mossy fibers, including
contained both clear and dense core vesicles and dentate transplants
formed both symmetrical and asymmetrical synap-
tic contacts with the apical dendrites of CA3 The parent dentate granule cells, the presynaptic
pyramidal cells, despite the fact that spines were MFs and their terminals and the postsynaptic tar-
not yet present (Amaral and Dent, 1981). During get cells in the dentate hilus and the hippocampal
the following week, the terminals developed a CA3 subfield are each and interdependently
more mature and complex morphology, size and affected by genetic and epigenetic factors and con-
synaptic vesicle content and density. Spines on the ditions, the effects of which are expressed as direct
proximal CA3 pyramidal cell dendrites started to structural and connective abnormalities, devia-
develop about one week postnatal, but a mature, tions from normal or natural variations among
classic, MF terminal and its intimate relations with species, strains and individuals. New observations
‘‘thorny excrescences’’ developed at three weeks on activity dependent plasticity of individual MF
postnatal. Thereafter, both the size of the MF ter- terminals and their postsynaptic CA3 pyramidal
minals, and the density of Timm staining, contin- cells have already been dealt with above, just as
ued to increase. It should be noted that the MF species (and mouse strain) differences have been
system is not mature in two-week-old rats, as presented. This section is therefore devoted to
demonstrated by Amaral and Dent (1981), who other experimental manipulations, ranging from
found the mean area and perimeter of the MF breeding studies and effects of hormonal admin-
terminal profiles to increase substantially during istration to studies of lesion-induced effects and
this period, together with a three-fold increase in neural transplantation.
99

Genetics and breeding experiments MF projection and a significant improvement in


radial-maze learning (Schwegler et al., 1991).
Several studies have investigated the MF system in Stress and glucocorticoid stress hormones also
mice and rats, selectively bred on the basis of be- effect the MF system (McEwen, 1999), and most
havioral criteria such as low and high cognitive interestingly such induced changes seems to be re-
performance levels in defined tests. These studies versible by behavioral training (Sandi et al., 2003).
have demonstrated significant differences in the
organization and size of the MF projection to CA3
(see Bernasconi-Guastalla et al., 1994; Lipp and Lesion-induced MF sprouting and rerouting
Wolfer, 1998; and references therein). Using such a
heritable difference between two mouse strains Several studies referred to above have included
(CH3 and CPB-K) as a starting point, Nek et al. interference with the development or integrity of
(1993) demonstrated that MF terminals in the dentate granule cells, hilar cells and CA3 pyram-
strain with fewer MFs to CA3 were larger and idal cells resulting in changes in the size and dis-
established more spine synapses per terminal than tribution of MFs, like for example induction of
in the strain with more MFs. Accordingly, the supragranular MFs following seizures and dener-
fiber density seemed to have a regulatory effect on vation of the dentate molecular layer (Fig. 5),
the morphology and synaptic contacts of the changes in the size of infrapyramidal projections
terminals. to CA3 or ingrowths into CA1 (Fig. 5). Only a few
Numerous mutants and gene-modified mice studies have addressed the reaction of MFs to
with neurological functional and structural defects transection during development or in the mature
exist, and have been and are currently screened for brain. In two studies on rat MFs by Laurberg and
effects on cognitive hippocampus-related function Zimmer (1980, 1981), supplemented by a subse-
and developmental changes in cellular and con- quent EM study (Frotscher and Zimmer, 1983)
nective organization. Only the mutant Reeler and observations in organotypic slice cultures
mouse will be mentioned here, because the muta- (Zimmer and Gahwiler, 1984, 1987), it was dem-
tion clearly affects the organization of the dentate onstrated that: (1) the main MF projection to
and hippocampal cell layers and as such has CA3, with an decrease in effect over the first 3
involved experimental studies of the hippocampus postnatal weeks, can be rerouted along the septo-
(Förster et al., 2006). As seen in Fig. 3C, the MF temporal axis in response to lesion-induced re-
system is present, but appears disorganized in moval or blockade of access to CA3 pyramidal
Reeler mice. This is, however, not due to change in cells at normal levels of trajectory; (2) the rerouted
cellular specificity in termination, but a conse- MFs terminate in expanded suprapyramidal and
quence of the disordered location of cells. infrapyramidal terminal zones in accessible adja-
cent levels; (3) loss of CA3 target cells not by itself
elicits an axonal pruning effect with compensatory
Hormonal effects on MF development and sprouting of MFs or collaterals into the dentate
cognitive function molecular layer; and (4) lesion-induced sprouting
of MF terminals into the dentate molecular layer
Lauder and Mugnaini (1980) and Lipp et al. (1984) could be induced in both postnatal developing and
early on demonstrated the sensitivity of the rodent young adult rats in a graded manner depending on
MF system to early postnatal hyperthyreoidism, the density of denervation of the dentate molecular
inducing an increase in particular infrapyramidal layer. As possible regrowth of severed or newly
MFs. Interestingly treatment of mice from a strain formed MFs across lesions in CA3 were difficult to
(DBA/2) virtually devoid of infrapyramidal MFs prove in vivo, given the possibility for fibers to
(correlating with poor radial-maze learning) with circumvent the lesion and then invade the normal
daily injections of 2 mg L-thyroxine early postna- target area, transections of the MF layer in hip-
tally, both induced a prominent infrapyramidal pocampal slice cultures and coculturing of dentate
100

and CA3 slice cultures derived from 7-day-old rats versa at all recipient brain ages, glutamatergic
were employed. Under both conditions regrowth projections, including those of the hippocampal
of MFs into distal CA3 required an intimate con- region, are clearly dependent of recipient age with
tact between the proximal transected MF layer regard to exchange of fibers between the always
and the distal, normal target zone, contrasting a developing graft tissue and the host brain. In ac-
vivid growth of MFs onto the substratum of glial cordance with these general observations, MFs
cells surrounding the cultures. were only seen to grow from dentate grafts into the
recipient brain and from the host brain fascia den-
tata into grafts under certain conditions. More
Neural grafting, testing MF growth and plasticity over, except for a minor projection of MFs grow-
ing from dentate grafts into the surrounding CA1
Intracerebral grafting of tissue blocks of develop- area (Sunde and Zimmer, 1981), ingrowth and
ing fascia dentata from newborn (or late fetal) rats termination of MFs only occurred into the normal
and mice have been used to identify factors reg- CA3 target area.
ulating the growth and specificity in the formation Figure 8 illustrates one experimental setup,
of nerve connections in the brain in general and to where a small block of neonatal rat fascia dentata
and from the dentate gyrus specifically (Sunde and had been grafted into the dentate area (Fig. 8B) or
Zimmer, 1981, 1983; Frotscher and Zimmer, 1986; next to CA3 (Fig. 8C) in other newborn rats with
Zimmer et al., 1988a, b; Tønder et al., 1990). In the purpose of examining whether transplanted
contrast to extensive growth of monoaminergic dentata granule cells were able to restore the MF
and cholinergic graft fibers into host brain and vice projection of the recipient rats, which before the

Fig. 8. (A) Timm stained sections of hippocampus from adult rat, subjected to hippocampal x-irradiation as newborn, stop odentate
granule cell formation and lead to few MF terminals. (B) Timm stained section from the contralateral hemisphere of the rat shown in
A, depicting a well-integrated dentate transplant (tpl), derived from a small block of fascia dentata, grafted just after the x-irradiation.
From the graft, an apparently normal, laminar-specific MF projection (mf) developed, connecting dentate granule cells of the graft (g)
with the host CA3. The molecular layer of the graft (m) received a comparably laminar-specific projection of host entorhinal perforant
path projections, normalizing the Timm stained laminar appearance of the layer. (C) Slightly displaced dentate transplant (tpl), grafted
to x-irradiated newborn hippocampus just after irradiation as part of same experiment (Sunde et al., 1984). Encroaching on the
recipient CA3, MFs from the granule cells of this graft has entered adjacent parts of CA3 and project ‘‘downstream’’, but appear to
stop at the border with CA1. Surprisingly, no MFs from the graft projected in the ‘‘upstream’’ direction, i.e., toward the host fascia
dentata. Scale bar: 500 mm (See Color Plate 5.8 in color plate section.)
101

transplantation had selectively x-irradiated over (Fig. 8C), graft MFs entered and innervated the
the hippocampal region, resulting in a severely re- host CA3 in a specific and laminar fashion from
duced number of granule cells and MFs for the the site of entry and ‘‘downstream’’ to CA1. For
rest of the animal’s life (Fig. 8A). In the situation unknown reasons no graft MFs grew in the op-
with both graft and recipient being neonatal and posite of normal direction toward the residual host
the graft positioned corresponding to the recipient fascia dentata.
fascia dentata (Fig. 8B) there was extensive and Using the presence of the peptide cholecystoki-
laminar specific outgrowth of host MFs into the nin (CCK) in mouse, but not rat, MF terminals,
recipient CA3, but no outgrowth ‘‘backward’’ into Zimmer et al. (1988a, b) moreover demonstrated
areas posterior to the graft. Note also in this ho- that mouse MFs, originating from a newborn
motopically placed graft the extensive, laminar mouse dentate xenograft placed next to the recip-
specific ingrowth of host entorhinal perforant fib- ient rat fascia dentata, were able to participate in
ers into the outer parts of the graft molecular the normal development of the recipient MF sys-
layer, illustrated by the presence of normal Timm tem, as illustrated by the presence of mouse xeno-
stained lamination of this layer (m, Fig. 8B.). With graft-derived, CCK immunoreactive MF terminals
placement of the dentate graft in the lateral ven- in the adult host rat CA3 (Fig. 9). With proper
tricle, encroaching on the distal parts of CA3 location of dentate grafts providing direct access

Fig. 9. Innervation of rat MF layer by mouse MF terminals from mouse dentate xenograft, topographically integrated with the
recipient rat fascia dentata. (A) Adult rat MF layer stained immunocytochemically for the peptide cholecystokinin (CCK), disclosing
some small-size, terminal-like structures with a preferential distribution among the CA3 pyramidal cells (p), and likely to arise from the
CCK-immunoreactive neurons (arrow heads). (B) Corresponding level of MF layer in contralateral hippocampus, which early post-
natally had received a xenograft of neonatal mouse fascia dentata, encroaching on the rat recipient fascia dentata for some distance
along the septotemporal axis. Corresponding to the levels of xenograft integration with the rat host fascia dentata and for some
additional distance temporally, mouse MF terminals, distinguished by their normal immunoreactivity for CCK, were present in the
host rat MF layer, as documented by their large MF terminal-like size (arrows), and their preferential suprapyramidal position, unlike
the smaller elements in the pyramidal cell layer (p) from the CCK-reaction intrinsic neurons (arrowheads). Scale bar: 50 mm.
102

Fig. 10. Adjacent Timm (A) and toluidine blue (B) stained sections from the dorsal hippocampus of adult rat, showing a focal lesion of
CA3 containing a small graft of fetal rat CA3 neurons, receiving a Timm stained, MF projection (asterisk, A) from the host fascia
dentata. In a two-step procedure, the rat received a localized injection of ibotenic acid, killing all CA3 neurons at the injection site. One
week later a cell suspension of late fetal rat CA3 neurons were grafted into the lesion site, where the afferent fibers, here dentate MFs,
still remained. Several weeks postgrafting, the CA3 graft (tpl) was found to be well integrated, yet not filling the lesion completely. Part
of lesioned CA3 pyramidal cell layer is thus still visible, but densely gliotic (p3). As demonstrated by Timm staining (asterisk, A), host
MFs have innervated the fetal CA3 graft. Scale bar: 100 mm.

for graft MFs to enter the recipient MF layer, Acknowledgment


graft-derived MFs may accordingly innervate the
CA3 area of the host hippocampus. The authors are grateful for the support received
Ingrowth of adult recipient brain, glutamatergic by former and current colleagues during the writ-
fibers into developing graft tissue are normally ing of this review, primarily in terms of help with
very restricted, apparently because the gluta- provision of relevant references for reading and
matergic fibers, unlike monoaminergic and choli- discussions. Prof. Michael Frotscher, Freiburg,
nergic host fibers, are competing with the intrinsic Germany is particularly acknowledged for provid-
graft fibers, which are in their normal phase of ing the EM pictures used as Fig. 1. Prof. Pico
development. One exception to this is when grafts Caroni, Basel, Switzerland is thanked for personal
are placed in so called ‘‘axon-sparing’’ lesions, communications on the newly emerging MF ter-
where recipient brain neurons have been killed by minal plasticity.
ischemia or excitotoxins like ibotenic acid one or
more days prior to grafting. Using this paradigm,
Tønder et al. (1990) demonstrated that adult re-
References
cipient brain MFs, primed by preceding ibotenic
acid lesion of their normal target CA3 pyramidal Acsády, L., Kamondi, A., Sik, A., Freund, T. and Buzsaki, G.
cells, indeed were able to innervate grafts of (1998) GABAergic cells are the major postsynaptic targets of
late fetal CA3 pyramidal neurons, placed in mossy fibers in the rat hippocampus. J. Neurosci., 18(9):
the lesioned CA3 area (Fig. 10). Adult MFs 3386–3403.
with already established connection area can Altman, J. and Das, G.D. (1965) Post-natal origin of micro-
neurones in the rat brain. Nature, 207(5000): 953–956.
accordingly after preceding axon-sparing lesions Amaral, D.G. (1978) A Golgi study of cell types in the hilar
of CA3 enter and terminate in fetal grafts of CA3 region of the hippocampus in the rat. J. Comp. Neurol.,
neurons. 182(4 Pt 2): 851–914.
103

Amaral, D.G. (1979) Synaptic extensions from the mossy fibers Blümcke, I., Suter, B., Behle, K., Kuhn, R., Schramm, J.,
of the fascia dentata. Anat. Embryol. (Berl.), 155(3): Elger, C.E. and Wiestler, O.D. (2000) Loss of hilar mossy
241–251. cells in Ammon’s horn sclerosis. Epilepsia, 41(Suppl 6):
Amaral, D.G. and Dent, J.A. (1981) Development of the mossy S174–S180.
fibers of the dentate gyrus. I. A light and electron microscopic Bruel-Jungerman, E., Laroche, S. and Rampon, C. (2005) New
study of the mossy fibers and their expansions. J. Comp. neurons in the dentate gyrus are involved in the expression of
Neurol., 195(1): 51–86. enhanced long-term memory following environmental en-
Amaral, D.G. and Witter, M.P. (1989) The three-dimensional richment. Eur. J. Neurosci., 21(2): 513–521.
organization of the hippocampal formation: a review of an- Cameron, H.A., Woolley, C.S., McEwen, B.S. and Gould, E.
atomical data. Neuroscience, 31(3): 571–591. (1993) Differentiation of newly born neurons and glia in the
Andersen, P., Bliss, T.V., Lomo, T., Olsen, L.I. and Skrede, dentate gyrus of the adult rat. Neuroscience, 56(2): 337–344.
K.K. (1969) Lamellar organization of hippocampal excita- Cassell, M.D. and Brown, M.W. (1984) The distribution of
tory pathways. Acta Physiol. Scand., 76(1): 4A–5A. Timm’s stain in the nonsulphide-perfused human hippocam-
Angevine, J.B. Jr., (1965) Time of neuron origin in the hippo- pal formation. J. Comp. Neurol., 222(3): 461–471.
campal region. An autoradiographic study in the mouse. Chandy, J., Pierce, J.P. and Milner, T.A. (1995) Rat hippo-
Exp. Neurol., Suppl 2: 1–70. campal mossy fibers contain cholecystokinin-like immunore-
Barber, R.P., Vaughn, J.E., Wimer, R.E. and Wimer, C.C. activity. Anat. Rec., 243(4): 519–523.
(1974) Genetically-associated variations in the distribution of Chicurel, M.E. and Harris, K.M. (1992) Three-dimensional
dentate granule cell synapses upon the pyramidal cell dend- analysis of the structure and composition of CA3 branched
rites in mouse hippocampus. J. Comp. Neurol., 156(4): dendritic spines and their synaptic relationships with mossy
417–434. fiber boutons in the rat hippocampus. J. Comp. Neurol.,
Barthesaghi, R. and Ravasi, L. (1999) Pyramidal types 325(2): 169–182.
in the field CA2 of the guinea pig. Brain Res. Bull., 50: Claiborne, B.J., Amaral, D.G. and Cowan, W.M. (1986) A light
263–273. and electron microscopic analysis of the mossy fibers of the
Bayer, S.A. (1980) Development of the hippocampal region in rat dentate gyrus. J. Comp. Neurol., 246(4): 435–458.
the rat. I. Neurogenesis examined with 3H-thymidine auto- Commons, K.G. and Milner, T.A. (1996) Ultrastructural rela-
radiography. J. Comp. Neurol., 190(1): 87–114. tionships between leu-enkephalin- and GABA-containing
Bayer, S.A. (1982) Changes in the total number of dentate neurons differ within the hippocampal formation. Brain Res.,
granule cells in juvenile and adult rats: a correlated volumet- 724(1): 1–15.
ric and 3H-thymidine autoradiographic study. Exp. Brain Cook, T.M. and Crutcher, K.A. (1985) Extensive target cell loss
Res., 46(3): 315–323. during development results in mossy fibers in the regio su-
Bernasconi-Guastalla, S., Wolfer, D.P. and Lipp, H.P. (1994) perior (CA1) of the rat hippocampal formation. Brain Res.,
Hippocampal mossy fibers and swimming navigation in mice: 353(1): 19–30.
correlations with size and left-right asymmetries. Hippocam- Danscher, G. (1981) Histochemical demonstration of heavy
pus, 4(1): 53–63. metals. A revised version of the sulphide silver method suit-
Bischofberger, J., Engel, D., Frotscher, M. and Jonas, P. (2006) able for both light and electronmicroscopy. Histochemistry,
Timing and efficacy of transmitter release at mossy fiber 71(1): 1–16.
synapses in the hippocampal network. Pflugers Arch., 453(3): Danscher, G., Howell, G., Perez-Clausell, J. and Hertel, N.
361–372. (1985) The dithizone, Timm’s sulphide silver and the sele-
Blackstad, T.W. (1956) Commissural connections of the hip- nium methods demonstrate a chelatable pool of zinc in CNS.
pocampal region in the rat, with special reference to their A proton activation (PIXE) analysis of carbon tetrachloride
mode of termination. J. Comp. Neurol., 105(3): 417–537. extracts from rat brains and spinal cords intravitally treated
Blackstad, T.W. (1963) Ultrastructural studies on the hippo- with dithizone. Histochemistry, 83(5): 419–422.
campal region. Prog. Brain Res., 3: 122–148. Danscher, G. and Zimmer, J. (1978) An improved Timm sul-
Blackstad, T.W., Brink, K., Hem, J. and Jeune, B. (1970) Dis- phide silver method for light and electron microscopic local-
tribution of hippocampal mossy fibers in the rat. An exper- ization of heavy metals in biological tissues. Histochemistry,
imental study with silver impregnation methods. J. Comp. 55(1): 27–40.
Neurol., 138(4): 433–449. Dawirs, R.R., Teuchert-Noodt, G., Hildebrandt, K. and Fei, F.
Blackstad, T.W. and Kjaerheim, A. (1961) Special axo- (2000) Granule cell proliferation and axon terminal degra-
dendritic synapses in the hippocampal cortex: electron and dation in the dentate gyrus of gerbils (Meriones unguiculatus)
light microscopic studies on the layer of mossy fibers. during maturation, adulthood and aging. J. Neural Transm.,
J. Comp. Neurol., 117: 133–159. 107(6): 639–647.
Blümcke, I., Schewe, J.C., Normann, S., Brustle, O., Schramm, De Paola, V., Arber, S. and Caroni, P. (2006) AMPA receptors
J., Elger, C.E. and Wiestler, O.D. (2001) Increase of nestin- regulat edynamic equilibrium of presynaptic terminals in
immunoreactive neural precursor cells in the dentate gyrus of mature hippocampal networks. Nat. Neurosci., 6: 491–500.
pediatric patients with early-onset temporal lobe epilepsy. Derrick, B.E., York, A.D. and Martinez Jr., J.L. (2000) In-
Hippocampus, 11(3): 311–321. creased granule cell neurogenesis in the adult dentate gyrus
104

following mossy fiber stimulation sufficient to induce long- Experimental analysis of fiber distribution with silver im-
term potentiation. Brain Res., 857(1–2): 300–307. pregnation methods. J. Comp. Neurol., 178(1): 73–88.
Dimitrieva, N.I., Gozzo, S., Dmitriev, I.u.S., Lopatina, N.G. Gaarskjaer, F.B. (1978b) Organization of the mossy fiber sys-
and Kassil’, V.G. (1993) Neuroanatomical characteristics of tem of the rat studied in extended hippocampi. I. Terminal
rat strains differing in the ability to form active avoidance area related to number of granule and pyramidal cells. J.
conditioned reflexes. Dokl. Akad. Nauk SSSR, 272(5): Comp. Neurol., 178(1): 49–72.
1235–1238. Gaarskjaer, F.B. (1981) The hippocampal mossy fiber system of
Doinikow, B. (1908) Beitrag zur histologie des Ammonshorns. the rat studied with retrograde tracing techniques. Correla-
J. Psychol. Neurol. (Leipzig), 13: 166–202. tion between topographic organization and neurogenetic
Duffy, C.J. and Rakic, P. (1983) Differentiation of granule cell gradients. J. Comp. Neurol., 203(4): 717–735.
dendrites in the dentate gyrus of the rhesus monkey: a quan- Gaarskjaer, F.B. (1985) The development of the dentate area
titative Golgi study. J. Comp. Neurol., 214(2): 224–237. and the hippocampal mossy fiber projection of the rat. J.
Eckenhoff, M.F. and Rakic, P. (1988) Nature and fate of pro- Comp. Neurol., 241(2): 154–170.
liferative cells in the hippocampal dentate gyrus during the Gaarskjaer, F.B. (1986) The organization and development of
life span of the rhesus monkey. J. Neurosci., 8(8): 2729–2747. the hippocampal mossy fiber system. Brain Res., 396(4):
Eriksson, P.S., Perfilieva, E., Bjork-Eriksson, T., Alborn, A.M., 335–357.
Nordborg, C., Peterson, D.A. and Gage, F.H. (1998) Ne- Gaarskjaer, F.B., Danscher, G. and West, M.J. (1982) Hippo-
urogenesis in the adult human hippocampus. Nat. Med., campal mossy fibers in the regio superior of the European
4(11): 1313–1317. hedgehog. Brain Res., 237(1): 79–90.
Fitzpatrick, D. and Johnson, R.P. (1981) Enkephalin-like Gage, F.H. (2000) Mammalian neural stem cells. Science,
immunoreactivity in the mossy fiber pathway of the hippo- 287(5457): 1433–1438.
campal formation of the tree shrew (Tupaia glis). Neuro- Gage, F.H. (2002) Neurogenesis in the adult brain. J. Neurosci.,
science, 6(12): 2485–2494. 22(3): 612–613.
Förster, E., Jossin, Y., Zhao, X., Frotscher, M. and Goffinet, Galimberti, I., Gogolia, N., Alberi, S., Santos, A.F., Muller, D.
A.M. (2006) Recent progress in understanding the role and Caroni, P. (2006) Long-term rearrangements of hippo-
of Reelin in the radial neuronal migration, with specific campal mossy fiber terminal connectivity in the adult regu-
emphasis on the dentate gyrus. Eur. J. Neurosci., 23: lated by experience. Neuron, 50: 749–763.
901–909. Gall, C. (1984) The distribution of cholecystokinin-like
Fredens, K. (1981) Genetic variation in the histoarchitecture of immunoreactivity in the hippocampal formation of the
the hippocampal region of mice. Anat. Embryol. (Berl.), guinea pig: localization in the mossy fibers. Brain Res.,
161(3): 265–281. 306(1–2): 73–83.
Frotscher, M. (1985) Mossy fibres form synapses with identified Gall, C. (1988) Seizures induce dramatic and distinctly different
pyramidal basket cells in the CA3 region of the guinea-pig changes in enkephalin, dynorphin, and CCK immunoreac-
hippocampus: a combined Golgi-electron microscope study. tivities in mouse hippocampal mossy fibers. J. Neurosci., 6:
J. Neurocytol., 14(2): 245–259. 1852–1862.
Frotscher, M. (1989) Mossy fiber synapses on glutamate dec- Gall, C., Lauterborn, J., Isackson, P. and White, J. (1990) Sei-
arboxylase-immunoreactive neurons: evidence for feed-for- zures, neuropeptide regulation, and mRNA expression in the
ward inhibition in the CA3 region of the hippocampus. Exp. hippocampus. Prog. Brain Res., 83: 371–390.
Brain Res., 75(2): 441–445. Ganser, S. (1882) Vergleichend-anatomische studien über das
Frotscher, M., Jonas, P. and Sloviter, R.S. (2006) Synapses gehirn des maulwurfs. Morphol. Jahrb., 7: 591–725.
formed by normal and abnormal hippocampal mossy fibers. Geneser-Jensen, F.A., Haug, F.M. and Danscher, G. (1974)
Cell Tissue Res., 326(2): 361–367. Distribution of heavy metals in the hippocampal region of
Frotscher, M. and Seress, L. (1991) Basket cells in the monkey the guinea pig. A light microscope study with Timm’s sulfide
fascia dentata: a Golgi/electron microscopic study. J. Ne- silver method. Z. Zellforsch. Mikrosk. Anat., 147(4):
urocytol., 20(11): 915–928. 441–478.
Frotscher, M., Soriano, E. and Misgeld, U. (1994) Divergence Golgi, C. (1886) Sulla Fina Anatomia Degli Organi Centrali
of hippocampal mossy fibers. Synapse, 16(2): 148–160. Del Sistema Nervoso. U. Hoepli, Milano, p. 215.
Frotscher, M. and Zimmer, J. (1983) Lesion-induced mossy Gould, E., Tanapat, P., McEwen, B.S., Flugge, G. and Fuchs,
fibers to the molecular layer of the rat fascia dentata: iden- E. (1998) Proliferation of granule cell precursors in the dent-
tification of postsynaptic granule cells by the Golgi-EM ate gyrus of adult monkeys is diminished by stress. Proc.
technique. J. Comp. Neurol., 215(3): 299–311. Natl. Acad. Sci. U.S.A., 95(6): 3168–3171.
Frotscher, M. and Zimmer, J. (1986) Intracerebral transplants Guidi, S., Ciani, E., Severi, S., Contestabile, A. and Bartesaghi,
of the rat fascia dentata: a Golgi/electron microscope R. (2005) Postnatal neurogenesis in the dentate gyrus of the
study of dentate granule cells. J. Comp. Neurol., 246(2): guinea pig. Hippocampus, 15(3): 285–301.
181–190. Hastings, N.B. and Gould, E. (1999) Rapid extension of axons
Gaarskjaer, F.B. (1978a) Organization of the mossy fiber sys- into the CA3 region by adult-generated granule cells. J.
tem of the rat studied in extended hippocampi. II. Comp. Neurol., 413(1): 146–154.
105

Haug, F.M. (1967) Electron microscopical localization of the Lauder, J.M. and Mugnaini, E. (1980) Infrapyramidal mossy
zinc in hippocampal mossy fibre synapses by a modified fibers in the hippocampus of the hyperthyroid rat. A light
sulfide silver procedure. Histochemie, 8(4): 355–368. and electron microscopic study. Dev. Neurosci., 3(4–6):
Haug, F.M. (1974) Light microscopical mapping of the hippo- 248–265.
campal region, the pyriform cortex and the corticomedial Laurberg, S. and Zimmer, J. (1980) Aberrant hippocampal
amygdaloid nuclei of the rat with Timm’s sulphide silver mossy fibers in cats. Brain Res., 188(2): 555–559.
method. I. Area dentata, hippocampus and subiculum. Laurberg, S. and Zimmer, J. (1981) Lesion-induced sprouting
Z. Anat. Entwicklungsgesch., 145(1): 1–27. of hippocampal mossy fiber collaterals to the fascia dentata
Henze, D.A., Urban, N.N. and Barrionuevo, G. (2000) The in developing and adult rats. J. Comp. Neurol., 200(3):
multifarious hippocampal mossy fiber pathway: a review. 433–459.
Neuroscience, 98(3): 407–427. Lehmann, K., Butz, M. and Teuchert-Noodt, G. (2005) Offer
Hjorth-Simonsen, A. (1977) Distribution of commissural affer- and demand: proliferation and survival of neurons in the
ents to the hippocampus of the rabbit. J. Comp. Neurol., dentate gyrus. Eur. J. Neurosci., 21(12): 3205–3216.
176(4): 495–513. Lein, E.S., Callaway, E.M., Albright, T.D. and Gage, F.H.
Holm, I.E. and Geneser, F.A. (1991) Histochemical (2005) Redefining the boundaries of the hippocampal
demonstration of zinc in the hippocampal region of the do- CA2 subfield in the mouse using gene expression and
mestic pig. III. The dentate area. J. Comp. Neurol., 308(3): 3-dimensional reconstruction. J. Comp. Neurol., 485(1):
409–417. 1–10.
Holm, I.E., Geneser, F.A. and Zimmer, J. (1992) Somatostatin- Lim, C., Blume, H.W., Madsen, J.R. and Saper, C.B. (1997)
and neuropeptide Y-like immunoreactivity in the dentate Connections of the hippocampal formation in humans: I. The
area, hippocampus, and subiculum of the domestic pig. J. mossy fiber pathway. J. Comp. Neurol., 385(3): 325–351.
Comp. Neurol., 322(3): 390–408. Lipp, H.P., Schwegler, H. and Driscoll, P. (1984) Postnatal
Holm, I.E., Geneser, F.A. and Zimmer, J. (1993) Cholecysto- modification of hippocampal circuitry alters avoidance learn-
kinin-, enkephalin-, and substance P-like immunoreactivity in ing in adult rats. Science, 225(4657): 80–82.
the dentate area, hippocampus, and subiculum of the do- Lipp, H.P. and Wolfer, D.P. (1998) Genetically modified mice
mestic pig. J. Comp. Neurol., 331(3): 310–325. and cognition. Curr. Opin. Neurobiol., 8(2): 272–280.
Holm, I.E. and West, M.J. (1994) Hippocampus of the domes- Lorente de Nó, R. (1934) Studies on the structure of the cer-
tic pig: a stereological study of subdivisional volumes and ebral cortex II. Continuation of the study of the ammonic
neuron numbers. Hippocampus, 4(1): 115–125. system. J. Psychol. Neurol. (Leipzig), 46: 113–177.
Jensen, S., Sorensen, T., Møller, A.G. and Zimmer, J. (1984) Lynch, G., Smith, R.L., Mensah, P. and Cotman, C. (1973)
Intraocular grafts of fresh and freeze-stored rat hippocampal Tracing the dentate gyrus mossy fiber system with horserad-
tissue: a comparison of survivability and histological and ish peroxidase histochemistry. Exp. Neurol., 40(2): 516–524.
connective organization. J. Comp. Neurol., 227(4): 559–568. Madeira, M.D., Paula-Barbosa, M., Cadete-Leite, A. and Tav-
Jessberger, S. and Kempermann, G. (2003) Adult-born hippo- ares, M.A. (1988) Unbiased estimate of hippocampal granule
campal neurons mature into activity-dependent responsive- cell numbers in hypothyroid and in sex-age-matched control
ness. Eur. J. Neurosci., 18(10): 2707–2712. rats. J. Hirnforsch., 29(6): 643–650.
Kempermann, G., Chesler, E.J., Lu, L., Williams, R.W. and McCloskey, D.P., Hintz, T.M., Pierce, J.P. and Scharfman,
Gage, F.H. (2006) Natural variation and genetic covariance H.E. (2006) Stereological methods reveal the robust size and
in adult hippocampal neurogenesis. Proc. Natl. Acad. Sci. stability of ectopic hilar granule cells after pilocarpine-in-
U.S.A., 103(3): 780–785. duced status epilepticus in the adult rat. Eur. J. Neurosci.,
Kempermann, G., Kuhn, H.G. and Gage, F.H. (1997) Genetic 24(8): 2203–2210.
influence on neurogenesis in the dentate gyrus of adult mice. McEwen, B.S. (1999) Stress and hippocampal plasticity. Ann.
Proc. Natl. Acad. Sci. U.S.A., 94(19): 10409–10414. Rev. Neurosci., 22: 105–122.
Kempermann, G., Kuhn, H.G. and Gage, F.H. (1998) Expe- McGinty, J.F., Henriksen, S.J., Goldstein, A., Terenius, L. and
rience-induced neurogenesis in the senescent dentate gyrus. J. Bloom, F.E. (1983) Dynorphin is contained within hippo-
Neurosci., 18(9): 3206–3212. campal mossy fibers: immunochemical alterations after
Keuker, J.I., Rochford, C.D., Witter, M.P. and Fuchs, E. kainic acid administration and colchicine-induced neurotox-
(2003) A cytoarchitectonic study of the hippocampal forma- icity. Proc. Natl. Acad. Sci. U.S.A., 80(2): 589–593.
tion of the tree shrew (Tupaia belangeri). J. Chem. Ne- Mello, L.E., Cavalheiro, E.A., Tan, A.M., Kupfer, W.R., Pre-
uroanat., 26(1): 1–15. torius, J.K., Babb, T.L. and Finch, D.M. (1993) Circuit
Koelliker, A. (1896) Handbuch der Gewebelehre des Menchen mechanisms of seizures in the pilocarpine model of chronic
2. Nervensystem des Menchen und der Thiere. Wilhelm En- epilepsy: cell loss and mossy fiber sprouting. Epilepsia, 34(6):
gelmann, Leipzig, p. 874. 985–995.
Kuhn, H.G., Dickinson-Anson, H. and Gage, F.H. (1996) Ne- Nek, N., Schwegler, H., Crusio, WE. and Frotscher, M. (1993)
urogenesis in the dentate gyrus of the adult rat: age-related Are the fine-structural characteristics of mouse hippocampal
decrease of neuronal progenitor proliferation. J. Neurosci., mossy fiber synapses determined by the density of mossy fiber
16(6): 2027–2033. axons? Neurosci. Lett., 158(1): 75–78.
106

van Praag, H., Schinder, A.F., Christie, B.R., Toni, N., Palmer, of dentate gyrus inhibitory interneurons: possible anatomical
T.D. and Gage, F.H. (2002) Functional neurogenesis in the substrate of granule cell hyper-inhibition in chronically
adult hippocampus. Nature, 415(6875): 1030–1034. epileptic rats. J. Comp. Neurol., 494(6): 944–960.
Raineteau, O., Hugel, S., Ozen, I., Rietschin, L., Sigrist, M., Song, H., Stevens, C.F. and Gage, F.H. (2002) Neural
Arber, S. and Gahwiler, B.H. (2006) Conditional labeling of stem cells from adult hippocampus develop essential prop-
newborn granule cells to visualize their integration into es- erties of functional CNS neurons. Nat. Neurosci., 5(5):
tablished circuits in hippocampal slice cultures. Mol. Cell. 438–445.
Neurosci., 32(4): 344–355. Soriano, E., Cobas, A., Lopez, C. and Berbel, P. (1983) Mor-
Ramón y Cajal, S.R. (1893) Estructura del asta de Ammon. phology of mossy fiber collaterals in the hilus of the rat.
Anal. Soc. Esp. Hist. Nat. (Madrid), 22: 53–114. Neurosci. Lett. Suppl., 14: S352.
Ramón y Cajal, S.R. (1911) Histologie du Système Nerveux de Stanfield, B.B. and Trice, J.E. (1988) Evidence that granule cells
l’Homme et des Vertébrés, Vol. II. Maloine, Paris. generated in the dentate gyrus of adult rats extend axonal
Ribak, C.E. and Peterson, G.M. (1991) Intragranular mossy projections. Exp. Brain Res., 72(2): 399–406.
fibers in rats and gerbils form synapses with the somata and Stengaard-Pedersen, K., Fredens, K. and Larsson, L.I. (1983)
proximal dendrites of basket cells in the dentate gyrus. Hip- Comparative localization of enkephalin and cholecystokinin
pocampus, 1(4): 355–364. immunoreactivities and heavy metals in the hippocampus.
Sala, L. (1891) Zur feineren anatomie des grossen seepherde- Brain Res., 273(1): 81–96.
fusses. Z. Wiss. Zoll., 52: 18–45. Stirling, R.V. and Bliss, T.V. (1978) Hippocampal mossy fiber
Sanchez-Andres, J.V., Palop, J.J., Ramirez, C., Nacher, J., development at the ultrastructural level. Prog. Brain Res., 48:
Molowny, A. and Lopez-Gracia, C. (1997) Zinc-positive pre- 191–198.
synaptic boutons of the rabbit hippocampus during early Storm-Mathisen, J. (1981) Glutamate in hippocampal path-
postnatal development. Dev. Brain Res., 103(2): 171–183. ways. Adv. Biochem. Psychopharmacol., 27: 43–55.
Sandi, C., Davies, H.A., Cordero, M.I., Rodriquez, J.J., Popov, Sunde, N., Laurberg, S. and Zimmer, J. (1984) Brain grafts can
V.I. and Stewart, M.G. (2003) Rapid reversal of stress in- restore irradiation-damaged neuronal connections in new-
duced loss of synapses in CA3 of rat hippocampal following born rats. Nature, 310(5972): 51–53.
water maze training. Eur. J. Neurosci., 17: 2447–2456. Sunde, N. and Zimmer, J. (1981) Transplantation of central
Sandler, R. and Smith, A. (1991) Coexistence of GABA and nervous tissue. An introduction with results and implications.
glutamate in mossy fiber terminals of the primate hippocam- Acta Neurol. Scand., 63(5): 323–335.
pus: an ultrastructural study. J. Comp. Neurol., 303(2): Sunde, N.A. and Zimmer, J. (1983) Cellular, histochemical and
177–192. connective organization of the hippocampus and fascia den-
Santarelli, L., Saxe, M., Gross, C., Surget, A., Battaglia, F., tata transplanted to different regions of immature and adult
Dulawa, S., Weisstaub, N., Lee, J., Duman, R., Arancio, O., rat brains. Brain Res., 284(2–3): 165–191.
Belzung, C. and Hen, R. (2003) Requirement of hippocampal Sutula, T., Cascino, G., Cavazos, J., Parada, I. and Ramirez, L.
neurogenesis for the behavioral effects of antidepressants. (1989) Mossy fiber synaptic reorganization in the epileptic
Science, 301(5634): 805–809. human temporal lobe. Ann. Neurol., 26(3): 321–330.
Schaffer, K. (1882) Beitrag zur histologie der Ammonshorn- Sutula, T., He, X.X., Cavazos, J. and Scott, G. (1988) Synaptic
formation. Arch. Mikr. Anat., 39: 611–632. reorganization in the hippocampus induced by abnormal
Schwarzer, C., Williamson, J.M., Lothman, E.W., Vezzani, A. functional activity. Science, 239(4844): 1147–1150.
and Sperk, G. (1995) Somatostatin, neuropeptide Y, ne- Swanson, L.W., Wyss, J.M. and Cowan, W.M. (1978) An au-
urokinin B and cholecystokinin immunoreactivity in two toradiographic study of the organization of intrahippocam-
chronic models of temporal lobe epilepsy. Neuroscience, pal association pathways in the rat. J. Comp. Neurol., 181(4):
69(3): 831–845. 681–715.
Schwegler, H., Crusio, W.E., Lipp, H.P., Brust, I. and Mueller, Tamamaki, N., Abe, K. and Nojyo, Y. (1988) Three-dimen-
G.G. (1991) Early postnatal hyperthyroidism alters hippo- sional analysis of the whole axonal arbors originating from
campal circuitry and improves radial-maze learning in adult single CA2 pyramidal neurons in the rat hippocampus with
mice. J. Neurosci., 11(7): 2102–2106. the aid of a computer graphic technique. Brain Res.,
Slomianka, L. and Geneser, F.A. (1997) Postnatal development 452(1–2): 255–272.
of zinc-containing cells and neuropil in the hippocampal re- Tamamaki, N. and Nojyo, Y. (1991) Crossing fiber arrays in
gion of the mouse. Hippocampus, 7(3): 321–340. the rat hippocampus as demonstrated by three-dimensional
Sloviter, R.S., Dichter, M.A., Rachinsky, T.L., Dean, E., reconstruction. J. Comp. Neurol., 303(3): 435–442.
Goodman, J.H., Sollas, A.L. and Martin, D.L. (1996) Ba- Terrian, D.M., Hernandez, P.G., Rea, M.A. and Peters, R.I.
salin expression and duction of glutamate decarboxylase and (1989) ATP release, adenosine formation, and modulation of
GABA in excitatory granule cells of the rat and monkey dynorphin and glutamic acid release by adenosine analogues
hippocampal dentate gyrus. J. Comp. Neurol., 373(4): in rat hippocampal mossy fiber synaptosomes. J. Neuroc-
593–618. hem., 53(5): 1390–1399.
Sloviter, R.S., Zappone, C.A., Harvey, B.D. and Frotscher, M. Thom, M., Martinian, L., Williams, G., Stoeber, K. and
(2006) Kainic acid-induced recurrent mossy fiber innervation Sisodiya, S.M. (2005) Cell proliferation and granule cell
107

dispersion in human hippocampal sclerosis. J. Neuropathol. neurons in the subdivisions of the rat hippocampus using the
Exp. Neurol., 64(3): 194–201. optical fractionator. Anat. Rec., 231(4): 482–497.
Timm, F. (1958) Histochemistry of the region of Ammon’s Zimmer, J. (1973) Changes in the Timm sulfide silver
horn. Z. Zellforsch. Mikrosk. Anat., 48(5): 548–555. staining pattern of the rat hippocampus and fascia dentata
Tønder, N., Sørensen, T. and Zimmer, J. (1990) Grafting of following early postnatal deafferentation. Brain Res., 64:
fetal CA3 neurons to excitotoxic, axon-sparing lesions of the 313–326.
hippocampal CA3 area in adult rats. Prog. Brain Res., 83: Zimmer, J. (1974) Long term synaptic reorganization in rat
391–409. fascia dentata deafferented at adolescent and adult stages:
Vaughn, J.E., Matthews, D.A., Barber, R.P., Wimer, C.C. and observations with the Timm method. Brain Res., 76(2):
Wimer, R.E. (1977) Genetically-associated variations in the 336–342.
development of hippocampal pyramidal neurons may pro- Zimmer, J., Finsen, B., Sorensen, T. and Poulsen, P.H. (1988a)
duce differences in mossy fiber connectivity. J. Comp. Ne- Xenografts of mouse hippocampal tissue. Exchange of lami-
urol., 173(1): 41–52. nar and neuropeptide specific nerve connections with the host
Vida, I. and Frotscher, M. (2000) A hippocampal interneuron rat brain. Brain Res. Bull., 3: 369–379.
associated with the mossy fiber system. Proc. Natl. Acad. Sci. Zimmer, J., Finsen, B., Sorensen, T. and Poulsen, P.H. (1988b)
U.S.A., 97(3): 1275–1280. Xenografts of mouse hippocampal tissue. Formation of nerve
West, J.R., Van Hoesen, G.W. and Kosel, K.C. (1982) A dem- connections between the graft fascia dentata and the host rat
onstration of hippocampal mossy fiber axon morphology brain. Prog. Brain Res., 78: 271–280.
using the anterograde transport of horseradish peroxidase. Zimmer, J. and Gahwiler, B.H. (1984) Cellular and connective
Exp. Brain Res., 48(2): 209–216. organization of slice cultures of the rat hippocampus and
West, M.J., Gaarskjaer, F.B. and Danscher, G. (1984) The fascia dentata. J. Comp. Neurol., 228(3): 432–446.
Timm-stained hippocampus of the European hedgehog: a Zimmer, J. and Gahwiler, B.H. (1987) Growth of hippocampal
basal mammalian form. J. Comp. Neurol., 226(4): 477–488. mossy fibers: a lesion and coculture study of organotypic slice
West, M.J. and Gundersen, H.J. (1990) Unbiased stereological cultures. J. Comp. Neurol., 264(1): 1–13.
estimation of the number of neurons in the human hippo- Zimmer, J. and Haug, F.M. (1978) Laminar differentiation of
campus. J. Comp. Neurol., 296(1): 1–22. the hippocampus, fascia dentata and subiculum in developing
West, M.J., Slomianka, L. and Gundersen, H.J. (1991) rats, observed with the Timm sulphide silver method. J.
Unbiased stereological estimation of the total number of Comp. Neurol., 179(3): 581–617.
Plate 5.2. Timm stained hippocampal sections from rat (A), guinea pig (B), cat (C), European hedgehog (D), dog (E) and man (F),
illustrating the distribution of intensely stained, black MF terminals in the different species, as well as other terminal fields and general
organization. Note species specific characteristics, like the ‘‘end bulb’’ at the terminal part of the MF layer in guinea pig (black asterisk,
B) and the layering of the dentate hilus with MF terminal-free, intermediate or plexiforme sublayer in the same species (white asterisk,
B), and MF projections into CA1 in some cats (arrows, C) and European hedgehog (D). Except for the black MF terminal staining, the
neuropil of the human hippocampus is virtually unstained, due to the use of special Timm staining procedures for non-perfused tissue
(Danscher and Zimmer, 1978). Abbreviations: CA1, hippocampal regio superior or subfield CA1; CA3, hippocampal regio inferior or
subfield CA3; FD, fascia dentata (or denate gyrus); g, dentate granule cell layer; h, dentate hilus or CA4; m, dentate molecular layer;
mf, MF layer; Sub, subiculum. Scale bars: 500 mm (For B/W version, see page 88 in the volume.)
Plate 5.3. Timm stained mouse hippocampal sections, illustrating strain differences in MFs terminal projections between BALB7C (A)
and C57B mice (B) and Reeler mutants (C). BALB/c mice have no infrapyramidal MFs, but a short intrapyramidal bundle (black
asterisk, A), compared to long infrapyramidal bundles in C57B mice (black asterisk, B). In Reeler mutant mice, granule and pyramidal
cell layers are disorganized, resulting in a corresponding distribution of MF terminals in the confluent granule cell layer/hilar area and
CA3. However, MFs still avoid CA1, similar to normal mice. Abbreviations: g, dentate granule cell layer; h, dentate hilus (CA4); m,
dentate molecular layer. Scale bar: 500 mm (For B/W version, see page 89 in the volume.)

Plate 5.5. Timm stained section from dorsoposterior level of young adult, rat hippocampus. When newborn, the rat received a
mechanical lesion of the CA3–CA1 transition (les) as well a transection of the entorhinal perforant path to the fascia dentata. The
CA3–CA1 lesion induced an aberrant ingrowth of MFs into CA1, illustrated by presence of small, Timm stained MF terminals
(asterisk) along the CA1 pyramidal cell layer. The removal of the perforant path projection to the outer molecular layer (m) induced
spread of associational-commissural projections from inner part of the layer, and induction of supragranular MF terminals (arrow)
(see Zimmer, 1973, 1974). Abbreviations: g, granule cell layer. Scale bar: 500 mm (For B/W version, see page 92 in the volume.)
Plate 5.6. A: Timm stained MF (collateral) terminals, found in normal adult rats to accompany dendrites (arrows), arising from
neurons in the limiting subzone just below the dentate granule cell layer or in the dentate hilus, and extending into and sometimes
through the dentate granule cell layer. (g) into the, commissural-associational terminal zone (c/a). B: Timm stained, dense supra-
granular MF terminal projection (sgr) found normally at temporal levels of the cat fascia dentata as well as in other species. h, dentate
hilus. Scale bar: 50 mm (A); 100 mm (B). (For B/W version, see page 94 in the volume.)

Plate 5.7. (A) Timm stained section from midposterior level of 5-day-old rat, demonstrating a distinct Timm stained MF layer in CA3
and the hilus at this age (arrows). (B, C) Parallel sections of hippocampal slice culture, derived from 7-day-old rat and grown for 3
weeks. Timm stain reveals the distribution of aberrant supragranular and normal CA3 MFs (arrow and mf, respectively, in B) and
thionine cell staining to visualize general cellular organization (C). Abbreviations: g, granule cell layer. Scale bars: 500 mm (For B/W
version, see page 95 in the volume.)
Plate 5.8. (A) Timm stained sections of hippocampus from adult rat, subjected to hippocampal x-irradiation as newborn, stop
odentate granule cell formation and lead to few MF terminals. (B) Timm stained section from the contralateral hemisphere of the rat
shown in A, depicting a well-integrated dentate transplant (tpl), derived from a small block of fascia dentata, grafted just after the x-
irradiation. From the graft, an apparently normal, laminar-specific MF projection (mf) developed, connecting dentate granule cells of
the graft (g) with the host CA3. The molecular layer of the graft (m) received a comparably laminar-specific projection of host
entorhinal perforant path projections, normalizing the Timm stained laminar appearance of the layer. (C) Slightly displaced dentate
transplant (tpl), grafted to x-irradiated newborn hippocampus just after irradiation as part of same experiment (Sunde et al., 1984).
Encroaching on the recipient CA3, MFs from the granule cells of this graft has entered adjacent parts of CA3 and project ‘‘down-
stream’’, but appear to stop at the border with CA1. Surprisingly, no MFs from the graft projected in the ‘‘upstream’’ direction, i.e.,
toward the host fascia dentata. Scale bar: 500 mm (For B/W version, see page 100 in the volume.)

Plate 6.1. Interneurons are the primary target of DG granule cells. Diagram representing the multiple types of excitatory contacts
made by granule cells. Within the dentate/hilar region collaterals of the mossy fibers (MFs) innervate both mossy cells and inhibitory
basket cells. In area CA3, the MFs contact pyramidal neurons via large expansions, but also excite interneurons through either
filopodial extensions from the large boutons or smaller en passant expansions along MF axons. (For B/W version, see page 110 in the
volume.)
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 6

Mossy fiber synaptic transmission: communication


$
from the dentate gyrus to area CA3

David B. Jaffe1, and Rafael Gutiérrez2

1
Department of Biology, University of Texas at San Antonio, One UTSA Circle, San Antonio, TX 78249, USA
2
Department of Physiology, Biophysics and Neurosciences, Center for Research and Advanced Studies, Mexico City,
Apartado Postal 14-740, Mexico D.F. 07000, Mexico

Abstract: Communication between the dentate gyrus (DG) and area CA3 of the hippocampus proper is
transmitted via axons of granule cells — the mossy fiber (MF) pathway. In this review we discuss and
compare the properties of transmitter release from the MFs onto pyramidal neurons and interneurons. An
examination of the anatomical connectivity from DG to CA3 reveals a surprising interplay between ex-
citation and inhibition for this circuit. In this respect it is particularly relevant that the major targets of the
MFs are interneurons and that the consequence of MF input into CA3 may be inhibitory or excitatory,
conditionally dependent on the frequency of input and modulatory regulation. This is further complicated
by the properties of transmitter release from the MFs where a large number of co-localized transmitters,
including GABAergic inhibitory transmitter release, and the effects of presynaptic modulation finely tune
transmitter release. A picture emerges that extends beyond the hypothesis that the MFs are simply ‘‘det-
onators’’ of CA3 pyramidal neurons; the properties of synaptic information flow from the DG have more
subtle and complex influences on the CA3 network.

Keywords: mossy fibers; synaptic transmission; CA3; co-transmitters; plasticity

Neural output from the dentate gyrus (DG) is originally) and Lorente de No both suggested that
transmitted via axons of granule cells called the the large varicosities made synaptic contacts onto
mossy fiber (MF) pathway. The MFs not only CA3 pyramidal neurons (Ramon y Cajal, 1911;
project into area CA3 of the hippocampus, but also Lorente de No, 1934), later to be confirmed with
synapse proximally onto DG basket cells (provid- electron microscopy (Blackstad and Kjaerheim,
ing local recurrent inhibition in the dentate) and 1961), thereby making this pathway an integral
pyramidal-like neurons in the hilus (Johnston and component within the so-called, and overly sim-
Amaral, 2004). They are so named because of the plistic, ‘‘tri-synaptic’’ circuit.
large (up to 5 mm diameter) varicosities along the In this review we will discuss and compare the
axon. Ramon y Cajal (who named the pathway properties of transmitter release at the most well-
characterized MF synapses, those in area CA3
onto pyramidal neurons and interneurons. Our
$
Both authors contributed equally to this work. discussion will focus on the potential roles of
Corresponding author. Tel.: +1 210 458 5843; co-localized transmitters and modulators as well
Fax: +1 210 458 5658; E-mail: david.jaffe@utsa.edu as how the properties of synaptic transmission

DOI: 10.1016/S0079-6123(07)63006-4 109


110

from the DG influence the primary target of onto inhibitory interneurons, rather than pyram-
DG output — the CA3 region of the hippo- idal neurons (Fig. 1).
campus. The large terminals of the MFs are unique
structures in a variety of ways. In addition to the
filopodial-like synaptic contacts onto interneu-
MF anatomy: does form follow function? rons, mentioned above, the large boutons are
filled with a high density of vesicles (Blackstad and
The MFs form three types of synaptic contacts Kjaerheim, 1961), express more than one active
onto its target neurons. First, and most notably, zone, and encase a complex branched dendritic
are the large expansions that synapse onto CA3 spine emanating from CA3 pyramidal neurons —
pyramidal neurons. The large boutons appear at generally referred to as a thorny excrescence
approximately 150 mm intervals (Blackstad et al., (Amaral and Dent, 1981; Chicurel and Harris,
1970) and a single granule cell contacts approxi- 1992). Putative synaptic sites on these boutons
mately 15 pyramidal neurons (each terminal were originally identified as both asymmetric and
synapses onto a single pyramidal neuron). One symmetric, suggesting differences in their respec-
CA3 pyramidal neuron may receive up to a total tive properties of transmission (Hamlyn, 1961).
of approximately 50 MF inputs only (Claiborne Another important anatomical aspect of the MF
et al., 1986; Amaral et al., 1990). Second, from the pathway is that they are generally restricted to
large expansions there may extend 2–3 filopodia a narrow band running within stratum lucidum
that make synaptic contacts onto interneurons where the large terminals make contacts onto the
(Amaral, 1979; Acsady et al., 1998). These most proximal portions of pyramidal neuron
so-called filopodia are motile and regulated by apical dendrites (Brown and Johnston, 1983). Ad-
glutamatergic neuromodulation (Tashiro et al., ditionally, there is a less extensive infrapyramidal
2003). Third, small terminals, resembling boutons MF projection onto the proximal basal dendrites
of other hippocampal neurons, also contact CA3 of pyramidal neurons in CA3b and CA3c.
interneurons. As a result, it appears that by sheer The first recordings of MF synaptic transmis-
numbers the major target of the MFs in CA3 is sion were extracellular field potential recordings

Fig. 1. Interneurons are the primary target of DG granule cells. Diagram representing the multiple types of excitatory contacts made
by granule cells. Within the dentate/hilar region collaterals of the mossy fibers (MFs) innervate both mossy cells and inhibitory basket
cells. In area CA3, the MFs contact pyramidal neurons via large expansions, but also excite interneurons through either filopodial
extensions from the large boutons or smaller en passant expansions along MF axons. (See Color Plate 6.1 in color plate section.)
111

made in vivo. Stimulation of the DG triggered a contacts directly or via its filopodia up to three
current sink restricted to s. lucidum (Gloor et al., interneurons, and (iii) 15 interneurons are inner-
1963), confirming the anatomical evidence that vated by the smaller boutons one-granule cell will
MF input was restricted to a discrete layer, contact approximately 60 interneurons compared
and supporting the view that the synapse was with only 15 pyramidal neurons (a 4:1 ratio).
excitatory. Therefore, it is conceivable that under certain con-
Two conflicting ideas emerge out of the ana- ditions the firing of granule cells would have a net
tomical connectivity from the DG to area CA3, inhibitory effect on pyramidal neurons in CA3
discussed above. On the one hand, the proximal (Bragin et al., 1995a, b; Penttonen et al., 1997;
location of the MF large boutons, and the fact that Acsady et al., 1998).
their large terminals are copiously supplied with
vesicles, suggests that a single action potential
from a granule cell might have a strong excitatory Excitatory–inhibitory conductance sequence
influence on CA3 pyramidal neurons. That is, a
single spike invading a single terminal might be Yamamoto (1972) was the first to utilize brain slice
capable of triggering an action potential in a CA3 methods and intracellular recording to study MF
pyramidal neuron. This is the so-called ‘‘detonator synaptic transmission onto CA3 pyramidal neu-
synapse’’ hypothesis. The circuitry of area CA3 rons. Stimulation of the granule cell layer triggered
resembles that of an auto-associative network a biphasic response composed of a small com-
(Marr, 1971) reflected in the large number of pound excitatory postsynaptic potential (EPSP)
recurrent excitatory connections between CA3 followed by a larger, overlapping compound in-
pyramidal neurons (Johnston and Amaral, 2004). hibitory postsynaptic potential (IPSP), presumably
The concept of the MFs as detonators led to the mediated either by feed-forward/feed-back inhibi-
hypothesis that these sparse inputs might serve as a tion or the direct stimulation of inhibitory inter-
‘‘teacher’’ signal underlying forms of associative neurons (Yamamoto, 1972). The IPSP itself was
learning (McNaughton and Morris, 1987; Rolls, biphasic, comprised of an initial GABAA receptor
1989). The presence of a large supply of vesicles and later GABAB receptor-mediated phase (Ogata
suggests a large readily releasable pool (Hallermann and Ueno, 1976; Knowles et al., 1984).
et al., 2003), and that the degree of transmitter One of Yamamoto’s most important early
release could be maintained over time by the large observations was that paired stimulation of the
reserve pool of vesicles. The proximal location fibers at short intervals (20–100 ms) markedly
of these synapses relative to the spike generating potentiated the second EPSP. While low-frequency
zone also would limit any loss of depolarization stimulation (LFS) rarely triggered spikes,
due to cable filtering (Johnston and Brown, 1983). frequency facilitation of the compound EPSP
That individual MF synapses would have large was generally capable of reaching spike threshold,
unitary synaptic strength would also be a require- in spite of the overlapping IPSPs. Frequency
ment if the pathway as a whole were to be effective facilitation occurred over a very wide range of
at driving the postsynaptic cell, because — as frequencies of stimulation (0.05–1 Hz), and could
discussed above — each CA3 pyramidal neuron potentiate responses as much as threefold (Salin
receives only about 50 mossy inputs (Amaral et al., et al., 1996).
1990). Johnston and Brown (1983) applied voltage-
On the other hand, the MFs also make direct clamp methods to the study of MF synaptic
synaptic contact onto inhibitory interneurons in transmission, recognizing that the advantageous
area CA3 (Fig. 1). Moreover, the ratio of synaptic proximal location of MFs relative to the soma
contacts onto interneurons is approximately four- permitted reasonable voltage-control and, in turn,
fold higher than onto pyramidal neurons. Assum- space clamp issues were minimized (Johnston and
ing that (i) a single granule cell contacts 15 Brown, 1983; Spruston and Johnston, 1992; Sprus-
pyramidal neurons, (ii) each large expansion ton et al., 1993). Measuring synaptic currents
112

under voltage-clamp also allowed better resolution To directly study presynaptic action potentials
of the mixed excitatory–inhibitory conductance and their interaction with voltage-gated ion chan-
sequence, even though it was recognized that there nels, patch-clamp methods have recently been
was still considerable overlap between the excita- applied to the large boutons of the MFs (Bischof-
tion and inhibition (Brown and Johnston, 1983; berger et al., 2006). One of the first notable
Barrionuevo et al., 1986). Two important obser- observations was that the action potential, while
vations were made in these early studies. First, for of short duration during LFS, widens during
compound postsynaptic currents, the GABAA- high-frequency stimulation (HFS) due to the rapid
mediated inhibitory conductance was almost inactivation of a voltage-gated K+ conductance
fivefold larger than the excitatory conductance. (Geiger and Jonas, 2000). Frequency-dependent
Second, the onset of the inhibitory conductance spike broadening, and the associated increase in
appeared to be very close to that of the excitatory Ca2+ influx, provides for, in part, a mechanism
response. A disynaptic GABAergic response, from underlying frequency-facilitation (Salin et al.,
a feed-forward or feed-back circuit, should retard 1996). The density of Na+ channels in the boutons
the onset of the inhibitory response relative to the appears relatively large, and may be important
excitatory conductance. It is therefore possible for the efficient propagation of action potentials
that the paradigms used to stimulate the MF (Engel and Jonas, 2005). The load resulting from
pathway might directly trigger the release of the greater surface area of the large boutons could
GABA, most likely from inhibitory neurons, onto potentially extinguish or slow down spike propa-
CA3 pyramidal neurons. gation in these unmyelinated fibers. Computer
simulations suggested that the high density and
properties of these Na+ channels ensures the
Properties of transmitter release from MF synapses reliability of synaptic transmission.
Ca2+ in the presynaptic terminal is also affected
Granule cells discharge action potentials down by other mechanisms. Intracellular Ca2+ stores
the MFs at basal rates less than 0.5 Hz (Jung and regulate presynaptic Ca2+ concentration (Liang
McNaughton, 1993), though firing rates may reach et al., 2002). Ca2+-induced Ca2+ release may
up to 50 Hz during certain types of behaviors work in concert with spike-mediated Ca2+ influx
(Skaggs et al., 1996; Wiebe and Staubli, 1999; to raise intraterminal Ca2+ concentration. Inter-
Henze et al., 2002b) and conduction velocity is estingly, although Ca2+-induced Ca2+ release ac-
approximately 7 m/s, consistent with the MFs counts for a large proportion of the presynaptic
being an unmyelinated pathway (Langdon et al., signal (Scott and Rusakov, 2006), depletion of
1993). Upon reaching synaptic terminals, presy- Ca2+ stores has no effect on fast trans-
naptic action potentials trigger synaptic transmis- mitter release. Release of Ca2+ may, however,
sion by eliciting Ca2+ influx through multiple play a role in frequency facilitation (Lauri et al.,
types of voltage-gated Ca2+ channels. Blockade 2003).
of N-, P-, and R-type Ca2+ channels decreases fast In a recent study, it was reported that EPSPs in
transmitter release, while dihydropyridine-sensi- dentate granule cells, surprisingly, propagated
tive L-type Ca2+ channels do not (Kamiya et al., down hundreds of microns along the MFs and
1988; Castillo et al., 1994; Nicoll et al., 1994; were capable of modulating synaptic transmission
Yamamoto et al., 1994; Wu et al., 1998; Miyazaki (Alle and Geiger, 2006). Furthermore, spike-
et al., 2005). That said, L-type channels are present evoked EPSCs in CA3 pyramidal neurons were
in MF presynaptic terminals and allow Ca2+ entry enhanced when paired with a presynaptic wave-
when MFs are stimulated (Tokunaga et al., 2004). form that mimicked an EPSP in the bouton. The
Although L-channels do not appear to participate mechanism underling this type of modulation may
in fast transmitter release, they may play a role in be related to a similar phenomenon observed
transmission during repetitive stimulation (Reuter, at the Calyx of Held (Awatramani et al., 2005),
1995). but other studies suggest it may also be due to
113

voltage and not solely due to resting Ca2+ levels and Madison, 1999). Similarly large unitary EPSPs
(Ruiz et al., 2003). have been demonstrated for MF-mossy cell
synapses (Scharfman et al., 1990).

Quantal nature of transmission at the MF-CA3


pyramidal neuron synapse Evidence for multiple neurotransmitters/modulators
in the granule cells
The strong frequency-dependent facilitation of
MF synaptic transmission onto CA3 pyramidal Glutamate is believed to be the primary excitatory
neurons implies that the initial probability of neurotransmitter released from the MFs (Crawford
transmitter release is small. If the probability of and Connor, 1973; Terrian et al., 1988), and MF
release were high, either a ceiling effect would limit EPSPs are blocked by glutamate receptor antago-
an increase in release (assuming that there is a nists (Sawada et al., 1983). The neuronal gluta-
large reserve pool of primed vesicles) or, alterna- mate transporter (EAAC1) is the most abundant
tively, one could observe synaptic depression due uptake mechanism and is selectively enriched in
to the refractory period produced by a lack of hippocampal principal neurons, including DG
primed vesicles. Indeed, quantal analysis of uni- granule cells (Rothstein et al., 1994). The presy-
tary synaptic responses onto CA3 pyramidal neu- naptic and postsynaptic actions of glutamate re-
rons is consistent with the hypothesis that the leased from the MFs are discussed below.
initial release probability is in fact low (von Granule cells also contain and release several
Kitzing et al., 1994; Lawrence et al., 2004). This neuromodulators of different chemical composi-
is in spite of the fact that the number of release tion (Fig. 2). The most prominent are those of
sites at a single MF synapse is very high (Chicurel peptidic nature, which are contained and released
and Harris, 1992), and that release at these from large dense-core vesicles. Dynorphin and
synapses is most likely multiquantal (Henze enkephalin, as well as their receptors, are present
et al., 1997). Indeed, in the study by Jonas et al. in the MFs in rodents and humans, but there is
(1993) quantal content (the number of quanta variation between mammalian species (Gall et al.,
released per action potential) ranged from 2–16, 1981; McGinty et al., 1983; Chavkin et al., 1985;
with a mean of 8, values similar to that first McLean et al., 1987; Terrian et al., 1988; Houser
reported by Yamamoto (Yamamoto et al., 1991). et al., 1990; Chavkin, 2000). The actions of opioid
Others have found that the probability of release peptides in the hippocampus are inhibitory, medi-
was much lower, o0.3 (Lawrence et al., 2004). ated by inhibition of Ca++ currents, in the case
There is also evidence, however, that large synap- of dynorphin, or activation of K+ currents, in
tic currents arise from the release of single quanta the case of enkephalins, and are a consequence
(Henze et al., 2002a). This is in contrast to com- of activation of G-protein-coupled receptors
missural/associational (C/A) synapses, where the (Zieglgänsberger et al., 1979). In the rat MF sys-
probability of release is higher and, concomitantly, tem, activation of m receptors facilitates MF LTP
frequency facilitation is weaker (Salin et al., 1996). in an indirect fashion. Activation of m receptors
Mean quantal size at MF-CA3 pyramidal depresses GABA release from interneurons, which
neuron synapses is approximately 150 pS, corre- in turn leads to a failure of GABA to inhibit
sponding to about 17 AMPA receptors. These glutamate release from MFs (Jin and Chavkin,
numbers are consistent with unitary responses 1999). On the other hand, the m opioid receptor
(unitary conductance ¼ quantal size  quantal agonist DAMGO inhibits low-frequency stimu-
content), which generally have magnitudes of lated MF responses in Sprague-Dawley rats, as it
approximately 1 nS. Corresponding unitary EPSP does in the guinea pig (Salin et al., 1995). Direct
amplitudes can be up to 10 mV (Jonas et al., 1993), actions of dynorphin on MF transmission have
in contrast to C/A unitary responses that are been described in the guinea pig, but not in the rat.
typically o1 mV (Debanne et al., 1998; Pavlidis Indeed the MFs possess k receptors, which, upon
114

Fig. 2. Summary of the pre- and post-synaptic constituents of the different MF terminals and of their plasticity. (A) Schematic
representation of the giant MF boutons, which contact CA3 pyramidal cells. These terminals are characterized by low probability of
basal release and multiple release sites. (B) Schematic representation of filopodial extensions and en passant contacts, which synapse on
to interneurons. These have a high probability of basal release and single release sites. Both types of MF terminals contain several
neuromodulators and the neurotransmitters glutamate and GABA. Note that both the presynaptic and postsynaptic sites contain
several ionotropic GluRs and metabotropic GluRs, which confer to the MF a high degree of plasticity and the capacity for synaptic
integration. (C) Relative expression of the different releasable contents and of the receptors and transporters during development, in
the adult (D), and after epileptic activity (E). Arrows and font size indicate relative differences between the three states. (See Color
Plate 6.2 in color plate section.)

activation, depress neurotransmitter release. Thus, However, in the rat, neither exogenous nor en-
for example, in the guinea pig high frequency dogenous dynorphin affect MF neurotransmis-
stimulation of the MF causes a transient heterosy- sion, which is consistent with the finding that
naptic inhibition of neighboring MF or perforant k-receptor binding in this projection is dense in the
path synapses in the dentate, which is mediated by guinea pig and sparse in the Sprague-Dawley rat
the synaptic release of dynorphin that activates (see Salin et al., 1995). Dynorphin significantly
presynaptic k receptors (Weisskopf et al., 1993). inhibits MF responses in the hamster, mouse, and
115

in Long-Evans rats (Salin et al., 1995). Interest- block GABAA receptors composed by a and b but
ingly, hippocampal opioid peptides are upregulat- not g subunits (Draguhn et al., 1990). It has been
ed after seizure activity (McGinty et al., 1983; shown that synaptically released Zn++ from the
Gall, 1988; Gall et al., 1990), whereas a decrease in MF pathway strongly modulates NMDA (Vogt
k receptors has been observed in epileptic human et al., 2000) and GABAA receptors in CA3 (Ruiz
cells (Jeub et al., 1999). Thus, opioid peptides are et al., 2004). Because MF release, besides gluta-
localized in the MF, presumably for effective con- mate, GABA and Zn++ (Gutierrez, 2005), this
trol of neurotransmitter release (Weisskopf et al., synapse is most suitable to study the effects of the
1993; Castillo et al., 1996; Drake et al., 1996; synaptically released Zn++ on GABA responses
Simmons and Chavkin, 1996). elicited by the same terminals. Indeed, it was
Other peptides present in the MF are neuropep- found that Zn++ tonically depresses the inhibitory
tide Y (NPY), neurokinin B and cholecystokinin actions of GABA, as it reaches a high concentra-
(Gall, 1984; Gall et al., 1990; Holm et al., 1993; tion in the vicinity of the GABAA receptor on
Tonder et al., 1994; Chandy et al., 1995; Schwarzer CA3 pyramidal cells. Consequently, chelation of
and Sperk, 1995; Makiura et al., 1999). Like opioid Zn++ relieves this inhibition (Ruiz et al., 2004).
peptides, these peptides have been shown to exert This is particularly important for developmental
inhibitory actions. Neuropeptide Y is synthesized, and pathological processes, in particular, epilepsy.
stored in, and released from the MFs, and inhibits Indeed, a higher input of Zn++ from sprouted
MF transmission by a presynaptic mechanism MFs after epileptic activity contributes to inhibit
(McQuiston and Colmers, 1996; McCarthy et al., reverberating excitatory activity in the DG
1998) and their receptors are normally present in the (Nadler, 2003; Tu et al., 2005).
MFs themselves (Widdowson, 1993; Jacques Other interesting signaling molecules present in
et al., 1997). Seizures increase the release of NPY the MF are brain derived neurotrophic factor
which in turn, tonically inhibits MF synaptic trans- (BDNF) and nerve growth factor (NGF) (Gall and
mission (Tu et al., 2005). Besides, seizures upregu- Isackson, 1989; Gall and Lauterborn, 1992;
late NPY Y2 receptors, which appear to mediate the Lowenstein and Arsenault, 1996; Conner et al.,
effects on MF (Vezzani and Sperk, 2004), 1997; Smith et al., 1997). Granule cells contain
although Y5 receptors could play a role (Marsh relatively high concentrations of BDNF and NGF,
et al., 1999; Ho et al., 2000). This upregulation re- which can be released in vivo by the MFs and
flects a neuroprotective function of NPY that has dendrites to affect other MFs or other cells, re-
been shown in animal models of epilepsy and in the spectively (Blöchl and Thoenen, 1995; Lowenstein
epileptic human (Schwarzer et al., 1998; Patrylo and Arsenault, 1996; Conner et al., 1997). Den-
et al., 1999; Furtinger et al., 2001; Vezzani and dritic targeting of BDNF mRNA and its transla-
Sperk, 2004). By contrast, substance P, which is tion to protein can account for the release of this
contained in and released from the MFs activates neurotrophin to neighboring granule cell dendrites
its receptors in the same MF, increases glutamate (Tongiorgi et al., 2004). This suggests that the lig-
release (Borhegyi and Leranth, 1997; Liu et al., and–receptor interaction occurs by means of an
1999). autocrine/paracrine mechanism. On the other
The MFs also contain high levels of Zn++, hand, anterograde transport of BDNF (Conner
which can be released together with glutamate et al., 1997) would effectively provide a mechanism
in an activity-dependent manner (Stengaard- to release BDNF from MF terminals, where it
Pedersen et al., 1981; Wenzel et al., 1997; Molnar can increase neurotransmitter release (Altar and
and Nadler, 2001; Qian and Noebels, 2005). DiStefano, 1998; Elmer et al., 1998). As with most
Exogenously applied Zn++ blocks glutamate signaling molecules in the granule cells, BDNF
and GABA receptors (Westbrook and Mayer, expression is increased after limbic seizures (Isack-
1987; Draguhn et al., 1990; Smart et al., 1994). son et al., 1991). In the MF, a consequence of this
Indeed, Zn++ occupies a high-affinity binding site would be the activation of TrkB receptors located
on NMDA receptors (Vogt et al., 2000) and can in the same fibers (He et al., 2002; Scharfman,
116

2005). TrkB activation is thought to promote Finally, the inhibitory amino acid GABA is
Ca++ influx into terminals (Berninger et al., 1993) probably the most important and controversial
and this is the mechanism that leads to enhanced addition to the list of modulators and transmitters
neurotransmitter release. Eventually, this produces that are co-localized in the granule cells. Sandler
an enhancement of excitability that can lead to the and Smith (1991) provided the first data that sug-
generation of epileptiform activity (Thoenen, gested the presence GABA in the ‘‘glutamatergic’’
1995; Huang and Reichardt, 2001; Scharfman MF in monkey and human hippocampi (Sandler
et al., 2002). and Smith, 1991). They found GABA immunore-
Importantly, it has been shown that BDNF can activity in MF terminals that made asymmetric
exert a rapid depolarization in hippocampal neu- synaptic contacts with spines arising from large
rons by promoting Na+ influx through TTX- dendrites of CA3 pyramidal cells. They also
insensitive channels (Kafitz et al., 1999; Rose et al., showed the colocalization of GABA and gluta-
2003). Another important issue is that BDNF may mate within the same terminals with electron
be packaged preferentially in the large MF microscopy. From this anatomical evidence, and
boutons, although it does reside in the small ter- the idea that GAD was not present in the MF,
minals also, so it can potentially modulate both they concluded that GABA had to be either
excitation and feed-forward inhibition in CA3 incorporated from the extracellular space or orig-
(Danzer and McNamara, 2004). inated from an alternative route of synthesis such
Besides a possible involvement of hippocampal as g-hydroxybutyrate. Second, they suggested that
BDNF in excitability and synaptic plasticity at least a component of the inhibitory synaptic
(Thoenen, 1995, 2000), neurotrophins have a potentials evoked in pyramidal cells by DG stim-
differentiating effect on interneurons in the hip- ulation had to be of MF origin. In this way,
pocampus (Marty et al., 1996a, b) and cortex GABA released by the MF could modulate
(Marty et al., 1997; Patz et al., 2003), by regulating the normal MF glutamatergic responses. On the
the expression of GABAergic markers. Neuronal other hand, studies carried out on MF synapto-
activity is the main activator of GAD expression somes provided the first neurochemical evidence
by neurotrophins, differentially modulating tran- showing that MF terminals contained and re-
scription and translation in a context-dependent leased GABA (Terrian et al., 1988; Taupin et al.,
manner (Patz et al., 2003). Interestingly, this is the 1994a, b). These authors, in agreement with
mechanism shown to underlie the maturation of Sandler and Smith (1991), suggested that the
the GABAergic phenotype in hippocampal inter- amino acid could be synthesized in the granule
neurons (Marty et al., 1996a, b) and, as recently cells from a route different from the GAD-
shown, the expression of all GABAergic markers dependent pathway.
in granule cells (Gomez-Lira et al., 2005). Evidence clarifying the presence and origin of
Granule cells also contain high levels of adeno- GABA in the MF was provided by Sloviter et al.
sine (Braas et al., 1986) and its A1 receptor (1996). These authors conclusively demonstrated
(Rivkees et al., 1995; Swanson et al., 1995). It that immunoreactivity for the amino acid and its
has been suggested that endogenous tonic activa- synthesizing enzyme, GAD, was normally present
tion of A1 receptors underlies the low basal prob- in the MFs of rats, monkeys, and humans (Sloviter
ability of neurotransmitter release from the MFs et al., 1996). Therefore, if the granule cells had the
(Moore et al., 2003; also see Kukley et al., 2005). necessary enzyme for the synthesis of GABA and
Their activation produces a Gi/o protein-depend- GABA itself, the granule cells indeed synthesized
ent sustained inhibition of neurotransmitter GABA that could probably be used as a neuro-
release (Moore et al., 2003). Other set of recep- transmitter. Complementing and extending the
tors present in the MF terminals, and which initial findings of Sandler and Smith (1991) and
activation controls neurotransmitter release, are Sloviter et al. (1996), Bergersen et al. (2003)
the ATP receptors P2X (Armstrong et al., 2002) recently confirmed with immunogold the coexist-
and a adrenergic receptors (Scanziani et al., 1993). ence of glutamate and GABA in MF synapses,
117

which also contained the respective receptors in Contrary to the data showing the upregulation
apposition to the presynaptic terminal (Bergersen of both isoforms of GAD by seizures (Sloviter
et al., 2003). This study demonstrated that both et al., 1996), several reports have shown that
amino acids coexist in all MF terminals examined GAD67 (and its mRNA) and not GAD65 is reg-
and that they have a close spatial relation to ulated in granule cells by increased activity
synaptic vesicles. Indeed, both GABA and gluta- (Schwarzer and Sperk, 1995; Szabo et al., 2000;
mate were shown to be located at a distance that Ramirez and Gutierrez, 2001; Maqueda et al.,
suggests their presence inside vesicles and in the 2003), Ca++ entry and BDNF activation (Gomez-
release zones. The estimated concentration of Lira et al., 2005). It has been proposed that the
GABA, however, was much lower than that cellular distribution of both enzymes differed and
of glutamate within the MF terminals and even this could be reflected in distinctive functions
lower than that of inhibitory types of terminals within neurons (Erlander and Tobin, 1991), i.e.,
(Bergersen et al., 2003). GAD65 may be in synaptic terminals, while
Interestingly, it was demonstrated that provok- GAD67 is present in terminals, somata and dend-
ing seizures by stimulation of the perforant path- rites (Kaufman et al., 1991). This suggests that one
way for 24 upregulated the content of both isoform synthesizes a metabolic pool of GABA
isoforms of GAD and GABA in the MFs, whereas (in the soma) and the other the releasable pool
no changes were observed in area CA1 (Sloviter (in the terminals). Despite the possible differences
et al., 1996). Other authors (Schwarzer and Sperk, in GAD65- and GAD67-originated GABA, there is
1995; Lehmann et al., 1996) showed that GAD67 a strong correspondence of the expression of
and its mRNA (but not GAD65) were transiently GAD67 to that of GABA present within the gran-
upregulated after seizures provoked by kainic acid ule cells and their MFs (Gomez-Lira et al., 2002).
(KA) or by the kindling method in the rat. On the All these studies suggest that GABA synthesis in
other hand, the upregulation of GAD67 and its the MFs is a means to counteract the enhanced
mRNA in granule cells was further confirmed with excitability caused by epileptic activity.
other seizure- and epilepsy-induction methods that
use chemical convulsants or electrical stimulation
(Ding et al., 1998; Makiura et al., 1999; Szabo Presynaptic modulation
et al., 2000; Ramirez and Gutierrez, 2001). It was
also established that GAD67 could be upregulated As mentioned above, a number of neuromodula-
in an activity-dependent manner, in the absence of tors control transmitter release from the MFs.
epileptiform activity (Ramirez and Gutierrez, Glutamate and GABA are also important presy-
2001; Gutierrez, 2002). Recently, a direct link be- naptic modulators of MF synaptic transmission
tween the presence of seizures and the concomitant transmission. In particular, one of the unique
upregulation of GAD and endogenous GABA was properties of MF synaptic transmission, in con-
found in MF terminals of epileptic rats (Gomez- trast to transmitter release from recurrent synapses
Lira et al., 2002). An analysis of the expression of onto CA3 pyramidal neurons, is its sensitivity to
GAD67 at different ages has revealed that GAD67 metabotropic glutamate receptor (mGluR) ago-
(but not GAD65 or GAD65 RNA) is also devel- nists; mGluR agonists depress MF synaptic trans-
opmentally regulated in the MFs (Maqueda et al., mission (Manzoni et al., 1995; Yoshino et al.,
2003). Indeed, it was shown that GAD67 is 1996). Because of this response, sensitivity to
expressed in the MFs early in life and then down- mGluR agonists is a widely used assay to deter-
regulated by days 23–24, after completion of mine if a synaptic response is of MF origin. There
development (Gutierrez, 2003; Maqueda et al., are species-specific differences with respect to pre-
2003). Likewise, GABA-immunoreactive cells with synaptic mGluRs. MF glutamatergic neurotrans-
characteristics of granule cells are found in the mission in the rat is sensitive to the group II
stratum granulare of the DG of developing rats mGluR agonist, DCG-IV, but not to the group III
but not of adults. mGluR agonist, L-AP4 (Lanthorn et al., 1984;
118

Kamiya et al., 1996; Maccaferri et al., 1998). The modulate plasticity at this synapse in a complex
opposite is true for the guinea pig where L-AP4 manner (Schmitz et al., 2000).
strongly depresses MF transmission (Lanthorn Presynaptic GABAB receptors effectively inhibit
et al., 1984; Manzoni et al., 1995; Tong et al., 1996; glutamate release from the MFs (Thompson and
Min et al., 1998). Interestingly, the granule cells Gahwiler, 1992), and it has recently shown that
and MF of the rat DG express groups II/III ionotropic GABAA receptors also inhibit gluta-
mGluR mRNA (Ohishi et al., 1993; Ohishi et al., mate (Ruiz et al., 2003) and GABA release
1995) and mGluR2,4,7 (Bradley et al., 1996; (Treviño and Gutierrez, 2005) and control the
Shigemoto et al., 1997; Lie et al., 2000). Electron excitability of this pathway (Kullmann et al., 2005;
microscopy has revealed immunolabeling for the Treviño and Gutierrez, 2005). Other subcortical
group III mGluR predominantly in presynaptic modulators also control the release of transmitter
active zones of asymmetrical and symmetrical from the MFs. Muscarinic acetylcholine receptors
synapses, whereas mGluR-II immunolabeling inhibit transmitter release (Williams and Johnston,
was found in preterminal rather than terminal 1988, 1990) while nicotinic receptor activation
portions of axons (Shigemoto et al., 1996, 1997). raises presynaptic Ca2+ levels and thereby may
Although it has been well established that acti- enhance transmitter release (Gray et al., 1996; also
vation of group II mGluR almost completely de- see Vogt and Regehr, 2001).
presses MF glutamatergic transmission in the rat,
several reports show that MF GABAergic trans-
mission is strongly inhibited through activation of Co-localization of plasma membrane transporters
group III mGluR (with L-AP4) both in the guinea of glutamate and GABA
pig and the rat (Gutierrez, 2000; Walker et al.,
2001; Gutierrez, 2002, 2003; Romo-Parra et al., Glutamate transport is the major mechanism con-
2003; Kasyanov et al., 2004; Safiulina et al., 2006). trolling extracellular glutamate levels, preventing
This pharmacological profile is not only consistent excitotoxicity, and averting neural damage associ-
with neurotransmission of MF origin but it also ated with hyperexcitability. As mentioned above,
demonstrates that the modulation of MF the neuronal glutamate transporter (EAAC1) is
GABAergic transmission differs from MF glut- expressed in granule cells and, surprisingly, in a
amatergic transmission. This physiological evi- number of GABAergic neurons (Rothstein et al.,
dence gives a functional significance to the 1994; He et al., 2002; Sepkuty et al., 2002). There-
anatomical findings showing that both types of fore, it has been suggested that besides controlling
receptors are present in rat granule cells with a extracellular glutamate levels, its function is linked
distinct localization and function (Ohishi et al., to GABA metabolism, because capture of gluta-
1995; Shigemoto et al., 1996, 1997). These data, mate is essential for GABA synthesis (Sepkuty
together with the finding that activation of mGluR et al., 2002). It is not surprising that seizures and
produces a downregulation of the exocytotic epilepsy have direct consequences on EAAC1 ex-
machinery (Kamiya and Ozawa, 1999), suggest pression and function, as has been demonstrated
that group III mGluR are associated with mech- (Gorter et al., 2002; Zhang et al., 2004). On the
anisms of GABA release, and also that there other hand, it has been proposed that the presence
is presynaptic segregation of mGluR receptors of the membrane transporter of GABA, GAT-1, is
according to the class of neurotransmitter to be restricted to neurons that synthesize and release
released. GABA, and glial cells (Iversen and Kelly, 1975;
Kainic acid has long been used as an epilepto- Radian et al., 1990; Ribak et al., 1996). Some
genic agent and was recognized that it binds to the GABAergic cells, immunocytochemically charac-
MF with high affinity (Monaghan and Cotman, terized by the presence of GAD67 or GABA, do
1982; Represa et al., 1987). It is now clear that the not contain or contain traces of GAT-1, but not
several KA receptors that are present in the MF vice versa (Rattray and Priestley, 1993). However,
(Wisden and Seeburg, 1993; Darstein et al., 2003) it has been found that GAT-1 is also localized in
119

the glutamatergic granule cells (Frahm et al., 2000; (Lanthorn et al., 1984; Neuman et al., 1988; Ito
Sperk et al., 2003), and it controls GABA uptake and Sugiyama, 1991; Jonas et al., 1993). Voltage-
to the MF terminals (Gomez-Lira et al., 2002) but clamp analysis of unitary and compound excitatory
not GABA release at this synapse (Vivar and postsynaptic currents (EPSC) onto CA3 pyramidal
Gutierrez, 2005; Safiulina et al., 2006). neurons finds the equilibrium potential close to
0 mV, consistent with ionotropic glutamate recep-
tors (Brown and Johnston, 1983). Based on the fast
Co-localization of the vesicular transporters for
rise times and decay kinetics of the current, the re-
glutamate (VGlut-1) and GABA (VGAT)
sponses are consistent with the properties of AMPA
receptors and therefore reflects an electrotonically
Because glutamate is a general metabolic substrate
close synapse (Williams and Johnston, 1991; Jonas
and serves as the precursor of inhibitory transmit-
et al., 1993).
ter GABA, glutamate immunoreactivity is not
Early studies using radio-labeled NMDA found
specific to glutamatergic neurons. Therefore, the
that the MF terminal field contains a low density
detection of glutamate vesicular transporter(s) has
of NMDA receptors relative to other regions of
been used to establish the glutamatergic phenotype
the hippocampal formation (Monaghan and
of neurons. As expected for glutamatergic neu-
Cotman, 1985). Although lower than at other
rons, the MF terminals of the granule cells contain
synaptic sites, glutamate release from the MFs
the glutamate vesicular transporter VGlut-1
activates a small, but measurable, NMDA com-
(Bellocchio et al., 1998; Kaneko et al., 2002). In
ponent that, as expected, exhibits voltage-depend-
accordance with immunohistological data showing
ence and slower kinetics, which are characteristics
that the granule cells express GAD and GABA, it
of NMDA receptor-mediated responses (Jonas
was found that they express VGAT mRNA in an
et al., 1993; Weisskopf and Nicoll, 1995).
activity-dependent manner (Lamas et al., 2001;
In contrast to NMDA receptors, the MF ter-
Gutierrez, 2003; Gomez-Lira et al., 2005). How-
minal field contains a sizable density of high-
ever, these cells do not contain the transporter
affinity sites for KA binding (Foster et al., 1981;
protein (Chaudhry et al., 1998; Sperk et al., 2003;
Monaghan and Cotman, 1982) and KA channels
also see Safiulina et al., 2006). Thus, the lack of
are expressed in CA3 pyramidal neurons (Egebjerg
detection of VGAT indicates that its expression is
et al., 1991; Werner et al., 1991). Focal application
too low to be revealed with immunohistochemis-
of kainate into the MF terminal field induces a
try, or the existence of a yet unidentified trans-
strong depolarization in CA3 pyramidal neurons
porter. Recent experiments show that glutamate
(Sawada et al., 1988). Postsynaptic activation of
and GABA are in close relation to vesicles in MFs
kainate receptors requires repetitive stimulation
terminals (Bergersen et al., 2003) strongly suggest-
due to their slow kinetics (Castillo et al., 1997;
ing that VGAT or another related transporter
Bortolotto et al., 2003). This, combined with the
protein must be present in these terminals
strong frequency-dependent facilitation of release,
(Chaudhry et al., 1998; Gomez-Lira et al., 2005).
suggests that KA receptors contribute to the
activity-dependent enhancement of transmission
Postsynaptic responses of CA3 pyramidal neurons from granule cells to CA3 pyramidal neurons
to MF glutamatergic input discussed above.
Because of the arguments above that MFs re-
As mentioned above, glutamate is believed to be the lease GABA and it acts as a classical neurotrans-
primary excitatory neurotransmitter released from mitter, one would expect that GABA receptors
the MFs (Crawford and Connor, 1973; Terrian would be located in apposition to MF terminals.
et al., 1988). The primary ionotropic glutamate re- In support of this prediction, there is evidence that
ceptors mediating the fast synaptic response at the GABAA receptors cluster with glutamatergic
MF synapse are a-amino-3-hydroxy-5-methyl-4- receptors in pyramidal cells in apposition to
isoxazolepropionic acid (AMPA)-type receptors both glutamatergic and GABAergic terminals
120

(Rao et al., 2000). It was later found, with im- 1990), a difference that may contribute to hilar cell
munogold detection, that AMPA and GABAA vulnerability to excitotoxicity.
receptors also colocalize at MF synapses in the The MF-to-CA3 synapse can therefore be con-
hilar region (Bergersen et al., 2003). sidered as ‘‘conditional detonators’’ in two do-
Finally, glutamate released from MFs triggers mains. The first, as described above, is the
the release of Ca++ from intracellular stores of temporal domain where presynaptic facilitation,
CA3 pyramidal cells via mGluR receptor activa- the slow membrane time constant of CA3 pyram-
tion (Miller et al., 1996; Jaffe and Brown, 1997). idal neurons (Spruston and Johnston, 1992), and
Tetanic stimulation of the MFs triggers waves of possibly postsynaptic amplification by KA recep-
spike-independent increases in Ca++ concentra- tors combine to boost EPSP amplitude and fire the
tion in the proximal apical dendrites of CA3 cell. The second domain reflects the spatial sum-
pyramidal neurons mediated by the activation of mation of inputs concomitantly with the MFs.
type I mGluR receptors (Kapur et al., 2001). For example, if a MF EPSP occurred coincidently
Single presynaptic spikes appear to be insuffi- with a perforant path EPSP, the summation of
cient to elicit Ca++ release (Reid et al., 2001). these two would then be suprathreshold (Urban
Frequency-dependent mGluR release of Ca++ from and Barrionuevo, 1998). Although all synapses
intracellular stores may be important for long- could be considered as conditional detonators
lasting changes in MF synaptic efficacy, discussed given this definition, the mean strength of unitary
below (Yeckel et al., 1999; Wang et al., 2004). MF EPSPs endows them with a greater ability to
move the membrane potential closer to threshold.

Are MF synapses onto CA3 pyramidal neurons


detonators? The granule cells simultaneously release glutamate
and GABA: electrophysiological evidence
It is not surprising that LFS of the MFs fails to
routinely trigger spikes, given many of the points Indirect but compelling evidence has accumulated
made above. For example, the major target of the over the last years of the co-release of glutamate
MFs in area CA3 involves the activation of feed- and GABA from the MFs. The first electrophys-
forward inhibitory circuits. Glutamate is co- iological evidence of GABAergic transmission
released with GABA, and a wide-array of other from the MFs to CA3 agreed with the
neuromodulators, which potentially inhibit MF immunohistochemical observations showing that
transmission. CA3 pyramidal cells also have a seizures transiently upregulated the expression of
threshold that is 10–15 mV from resting potential GAD65 and GAD67 (Schwarzer and Sperk, 1995;
(Podlogar and Dietrich, 2006). There is a large Sloviter et al., 1996). Indeed, it was shown that
amplitude unitary EPSP (Jonas et al., 1993), so the stimulation of the MFs produced monosynaptic
MFs may not relay synapses with a high safety GABA-mediated transmission in pyramidal cells
factor. An elegant demonstration of the ‘‘condi- of CA3 in kindled epileptic but not in control
tional’’ nature of MF excitation of CA3 pyramidal healthy rats (Gutierrez, 2000; Gutierrez and
neurons was shown by Henze et al. (2002b). With Heinemann, 2001).
low frequency stimulation (o40 Hz), the likeli- As mentioned above, feed-forward inhibition
hood that a CA3 pyramidal neuron would dis- results from activation of local inhibitory inter-
charge following a granule cell was low. Only at neurons by MF glutamatergic, excitatory synapses
higher frequencies did CA3 pyramidal neurons (Crawford and Connor, 1973; Acsady et al.),
follow granule cell firing. Interestingly, this may which in turn inhibit pyramidal cells (Yamamoto,
not be the case for all MF targets. Scharfman and 1972; Brown and Johnston, 1983; Buzsaki, 1984).
colleagues showed faithful following of granule Thus, activation of the DG leads to mono-
cell firing when the targets of MFs were hilar synaptic excitation and disynaptic inhibition on
mossy cells or hilar interneurons (Scharfman et al., CA3 pyramidal neurons that, with intracellular
121

recordings, are observed as an EPSP/IPSP discarding the possibility that they are packaged
sequence. Both these conductances are typically in single vesicles. Moreover, the evidence of MF
blocked in the presence of blockers of glut- GABA transmission onto both pyramidal cells
amatergic transmission. In kindled epileptic ani- (Gutierrez, 2000; Gutierrez and Heinemann, 2001;
mals, however, a bicuculline-sensitive IPSP is still Walker et al., 2001), and interneurons of CA3
elicited in the presence of the glutamate receptor (Romo-Parra et al., 2003; Safiulina et al., 2006)
blockers. This IPSP had the same latency as the and onto the mossy cells in the hilar region
control EPSP and could be inhibited by activation (Bergersen et al., 2003) excludes the possible
of metabotropic glutamate receptors (mGluR). segregation of neurotransmitters according to the
Expression of the monosynaptic IPSP was tran- target cell.
sient because it could be observed 24–48 h after the It has been shown that the putative release of
last kindled seizure but was not present if the GABA from the MF also occurs during develop-
experiment was carried out a month after the last ment, when the granule cells express all the mark-
seizure. The kindled epileptic state is not necessary ers of the GABAergic phenotype, and provoke
for this monosynaptic IPSP, because a single sei- GABA receptor-mediated responses in pyramidal
zure (Gutierrez, 2000) or repeated LTP-like stim- cells and interneurons of CA3. This occurs during
ulation could elicit the response (Gutierrez, 2002). the first weeks of age, after which the GABAergic
Moreover, the monosynaptic IPSP produced by phenotype shuts off in a clear-cut manner. Sup-
the activation of the DG persisted when perfusing porting this notion, recent work by Gutiérrez et al.
a medium with a low [Ca++] or the GABAB ago- (2003), Kasyanov et al. (2004) and Safiulina et al.
nist, baclofen. These manipulations depressed the (2006) have shown that GABA is the main neuro-
DG-evoked IPSP amplitude without altering its transmitter released from MF terminals during the
onset latency or slope. This unequivocally estab- first postnatal days (Kasyanov et al., 2004;
lished that the IPSP was a monosynaptic response Safiulina et al., 2006). Thus, MFs contain two
because, under these conditions, the ability to different sets of low- and high-threshold fibers that
recruit local interneurons quickly enough to release GABA and GABA plus glutamate, respec-
trigger such a short-latency IPSP is unlikely. The tively. The first would disappear with maturation,
possibility of recruiting inhibitory interneurons whereas the second would persist longer or would
by activating electrical synapses was also dis- reappear in pathological conditions, such as in ep-
carded (Gutierrez, 2000). Independently, Walker ilepsy (Gutierrez, 2003; Safiulina et al., 2006). It is
et al. (2001) demonstrated that monosynaptic possible that signaling by both, synaptic and non-
GABAergic responses could be normally evoked synaptic release of GABA, may play a crucial role
in CA3 pyramidal cells by MF activation in slices in tuning hippocampal network during postnatal
of young guinea pigs (Walker et al., 2001). They development (Gutierrez, 2003, 2005; Safiulina
showed that these MF GABAergic responses had et al., 2006). While the molecular mechanisms
the same pharmacological and plastic properties as that shut off the GABAergic phenotype at matu-
those reported for glutamatergic MF transmission, rity have not been disclosed, upregulation of
i.e., there was strong frequency-dependent the GABAergic phenotype in the adult depends
potentiation, and they were sensitive to an on protein synthesis, Ca++ entry, and acti-
antagonist of the metabotropic glutamate recep- vation of TrkB receptors by BDNF (Gutierrez,
tor that is expressed preferentially by MFs, L-AP4. 2002; Romo-Parra et al., 2003; Gomez-Lira et al.,
They were able to show that minimal stimulation 2005).
of the MF evoked glutamate only, GABA only,
and compound glutamate-GABA currents in py-
ramidal cells. This indicated that both responses, Long-term plasticity at the MF CA3 synapse
glutamatergic and GABAergic, had a common
origin. In addition, both neurotransmitters could Like other excitatory synapses in the hippo-
be released synchronously or asynchronously, campal formation, the MF synapse expresses
122

LTP in response to a brief episode of HFS B-ephrins reverse signaling (Contractor et al.,
(Yamamoto et al., 1980). As described above, the 2002; Armstrong et al., 2006).
MF terminal field contains a lower density of ‘‘What goes up must go down’’ is an apt phrase
NMDA receptors compared with other areas of the to describe MF synaptic plasticity. LFS triggers
hippocampus. This observation motivated Harris long-term depression (LTD) of MF synaptic in-
and Cotman (1986) to test whether LTP at the MF puts onto CA3 pyramidal neurons (Kobayashi
synapse was dependent on NMDA receptors. They et al., 1996; Yokoi et al., 1996) where presynaptic
found that in the presence of NMDA receptor activation of mGluR2 receptors depresses cAMP
antagonists, HFS of the MFs was still capable of levels, countering the expression of LTP.
triggering LTP (Harris and Cotman, 1986).
Although there is agreement that Ca++ is nec-
essary for the induction process, the mechanism CA3-interneuron synapses
underlying the induction of NMDA-independent
LTP has been a source of controversy (Nicoll and As discussed above, the majority of MF synaptic
Schmitz, 2005). Briefly, many experiments point to contacts are onto GABAergic interneurons
an induction mechanism that is solely presynaptic. (Acsady et al., 1998) of which there are a wide
Here the working hypothesis is that HFS triggers variety of subtypes, typically characterized based
an increase in presynaptic Ca++ that, via calmod- on a combination of parameters including cell
ulin, activates adenylyl cyclase (Zalutsky and body location, axonal projection, morphology, co-
Nicoll, 1992; Weisskopf et al., 1994; Villacres localized peptides, and calcium binding protein
et al., 1998). An alternative presynaptic mecha- content (Parra et al., 1998). Of particular interest
nism is that presynaptic KA receptor activation are a class of bipolar interneurons whose dendrites
leads to Ca++ influx that triggers release of Ca++ primarily reside along and within s. lucidum
from intracellular stores (Contractor et al., 2001; (Spruston et al., 1997; Vida and Frotscher, 2000),
Lauri et al., 2001; Bortolotto et al., 2003; Lauri in contrast to other interneurons where the den-
et al., 2003), although this remains controversial dritic tree projects orthogonally against the MFs
(Breustedt and Schmitz, 2004). that receive most of their innervation from other
Alternatively, there is evidence for a postsynaptic pathways. The studies discussed below were made
component to the induction mechanism (Williams primarily from this specific subtype of interneuron.
and Johnston, 1989; Jaffe and Johnston, 1990). Like the pyramidal cell synapse, MF transmis-
Here, an interaction between Ca++ entry through sion onto interneurons is reduced by the activation
voltage-gated Ca++ channels and the activation of presynaptic group II mGluR receptors
of mGluRs takes place (Kapur et al., 1998; Yeckel (Maccaferri et al., 1998; Toth et al., 2000); Pelkey
et al., 1999; Kapur et al., 2001; Wang et al., 2004). It et al., 2005), although other types of mGluRs may
may be that depending on stimulus conditions, in- play a role in synaptic plasticity at this synapse
duction may have either a presynaptic or post- (discussed below). In contrast to the pyramidal
synaptic locus (Urban and Barrionuevo, 1996). neuron synapse, at MF-interneuron synapses
In contrast to the induction process, there is lit- AMPA receptor expression is heterogeneous.
tle controversy regarding the locus for the expres- Glutamate released from the MFs may activate
sion of MF LTP. Numerous studies indicate that receptors with or without GluR2 subunits, thereby
synaptic potentiation is mediated by a presynaptic forming receptors with either Ca++ permeable or
change in transmitter release (Zalutsky and Nicoll, impermeable subunits (Toth and McBain, 1998;
1990; Yamamoto et al., 1992; Zalutsky and Nicoll, Toth et al., 2000).
1992; Xiang et al., 1994; Reid et al., 2004). If Also different from the pyramidal neuron MF
induction of MF LTP has a postsynaptic locus, synapse is the frequency response of the interneu-
then expression must require a retrograde com- ron, which is highly variable. The response can
munication. One possible mechanism involves range from a weak frequency-dependent facilita-
postsynaptic ephrinB receptors and presynaptic tion (much less pronounced than at the pyramidal
123

neuron synapse) to synaptic depression (Toth Communication from DG to CA3: exciting


et al., 2000). yet inhibiting
Quantal transmission at interneuron synapses is
also different than at the pyramidal neuron A picture is emerging that relates the contribution
synapse. As expected from anatomy, these of MF transmission to excitation and inhibition
synapses express a low number of release sites in area CA3, both in terms of the repertoire of
(Acsady et al., 1998). The probability of release at neurotransmitters/modulators released from the
these sites, however, appears higher than at the fibers but also with respect to circuitry that is
pyramidal neuron synapse (Lawrence et al., 2004). required to take into account inhibitory neurons.
This difference in probability of release is also At low frequencies (o0.5 Hz), frequency facilitation
consistent with differences frequency facilitation of the MF-CA3 pyramidal neuron synapse is weak
observed between the two synapses (i.e., a low in- and the probability of eliciting a spike is low (Henze
itial probability of release predicts large facilita- et al., 2002b). Furthermore, in this frequency do-
tion). Interestingly, quantal size at interneuron main granule cells are likely to be firing inhibitory
synapses is within the same ranges, if not higher, interneurons preferentially relative to pyramidal
than for pyramidal neuron synapse. As a result, neurons CA3 (Bragin et al., 1995a, b; Penttonen
and combined with a more depolarized resting et al., 1997). As such, concomitant extrinsic input
potenial, unitary responses of interneuron from entorhinal cortex (via the perforant path) or
synapses are as capable of triggering action po- via recurrent excitation via the C/A pathway should
tentials as MF inputs onto pyramidal neurons fail to excite sets of CA3 pyramidal neurons. But as
(Lawrence et al., 2004), and this also appears to be granule cell firing rates increase, such as when an
the case for MF transmission to hilar interneurons animal traverses a place field, frequency facilitation
(Scharfman et al., 1990). can then depolarize CA3 pyramidal neurons
Long-lasting plasticity at MF-interneuron to spike threshold. Interestingly, after seizures,
synapses also differ from their pyramidal neuron MFs tonically inhibit b/g field oscillations and
counterparts. Most notably, HFS does not elicit spontaneous subthreshold membrane oscillations
LTP at the MF-interneuron synapse (Maccaferri of CA3 interneurons in the CA3 area through
et al., 1998). Rather, HFS triggers LTD GABA-medated signaling. Coincident stimulation
at synapses expressing Ca++-permeable AMPA of the MFs at y and b/g frequencies produces a
receptors, while NMDA receptors trigger the frequency-dependent excitation of interneurons and
induction of LTD at synapses containing Ca++- the inhibition of pyramidal cells. Indeed, these
impermeant AMPA receptors (Lei and McBain, effects are maximal at the b frequency, suggesting
2002). Activation of mGluR7 receptors by L-AP4 a resonance phenomenon (Treviño et al., 2007).
produces a chemical version of LTD at this Because of the properties of synaptic plasticity
synapse that is distinct from pyramidal neuron within the auto-associative circuit of CA3, pattern
contacts (Pelkey et al., 2005) and differentially separation, pattern completion, and the encoding
modulates voltage-gated Ca++ channels in the of new information will then be dependent on the
presynaptic filopodia (Pelkey et al., 2006). If HFS precise timing of granule cell and pyramidal neuron
triggers LTD at the MF-interneuron synapse, firing (Kohonen, 1977; Chattarji et al., 1989;
feed-forward inhibition should be depressed onto Zalutsky and Nicoll, 1990; August and Levy,
CA3 pyramidal neurons following HFS. However, 1999; Nakazawa et al., 2002; Rolls and Kesner,
fast synaptic inhibition onto CA3 pyramidal neu- 2006).
rons is not affected by LTP of the MFs (Griffith
et al., 1986) suggesting that a decrease in feed-for- References
ward inhibition through interneurons of s. lucidum
Acsady, L., Kamondi, A., Sik, A., Freund, T. and Buzsaki, G.
is compensated by enhanced recurrent inhibition (1998) GABAergic cells are the major postsynaptic targets
or other forms of synaptic plasticity within the of mossy fibers in the rat hippocampus. J. Neurosci., 18:
CA3 network. 3386–3403.
124

Alle, H. and Geiger, J.R. (2006) Combined analog and action light microscopic studies on the layer of mossy fibers.
potential coding in hippocampal mossy fibers. Science, 311: J. Comp. Neurol., 117: 133–159.
1290–1293. Blöchl, A. and Thoenen, H. (1995) Characterization of nerve
Altar, C.A. and DiStefano, P.S. (1998) Neurotrophin traffick- growth factor (NGF) release from hippocampal neurons:
ing by anterograde transport. Trends Neurosci., 21: 433–437. evidence for a constitutive and an unconventional sodium-
Amaral, D.G. (1979) Synaptic extensions from the mossy fibers dependent regulated pathway. Eur. J. Neurosci., 7:
of the fascia dentata. Anat. Embryol. (Berl.), 155: 241–251. 1220–1228.
Amaral, D.G. and Dent, J.A. (1981) Development of the mossy Borhegyi, Z. and Leranth, C. (1997) Substance P innervation of
fibers of the DG: I. A light and electron microscopic study of the rat hippocampal formation. J. Comp. Neurol., 384:
the mossy fibers and their expansions. J. Comp. Neurol., 195: 41–58.
51–86. Bortolotto, Z.A., Lauri, S., Isaac, J.T. and Collingridge, G.L.
Amaral, D.G., Ishizuka, N. and Claiborne, B. (1990) Neurons, (2003) Kainate receptors and the induction of mossy fibre
numbers and the hippocampal network. Prog. Brain Res., 83: long-term potentiation. Philos. Trans. R. Soc. Lond. B Biol.
1–11. Sci., 358: 657–666.
Armstrong, J.N., Brust, T.B., Lewis, R.G. and MacVicar, B.A. Braas, K.M., Newby, A.C., Wilson, V.S. and Snyder, S.H.
(2002) Activation of presynaptic P2  7-like receptors de- (1986) Adenosine-containing neurons in the brain localized
presses mossy fiber-CA3 synaptic transmission through p38 by immunocytochemistry. J. Neurosci., 6: 1952–1961.
mitogen-activated protein kinase. J. Neurosci., 22: Bradley, S.R., Levey, A.I., Hersch, S.M. and Conn, P.J. (1996)
5938–5945. Immunocytochemical localization of group III metabotropic
Armstrong, J.N., Saganich, M.J., Xu, N.J., Henkemeyer, M., glutamate receptors in the hippocampus with subtype-specific
Heinemann, S.F. and Contractor, A. (2006) B-ephrin reverse antibodies. J. Neurosci., 16: 2044–2056.
signaling is required for NMDA-independent long-term po- Bragin, A., Jando, G., Nadasdy, Z., Hetke, J., Wise, K. and
tentiation of mossy fibers in the hippocampus. J. Neurosci., Buzsaki, G. (1995a) Gamma (40–100 Hz) oscillation in the
26: 3474–3481. hippocampus of the behaving rat. J. Neurosci., 15: 47–60.
August, D.A. and Levy, W.B. (1999) Temporal sequence com- Bragin, A., Jando, G., Nadasdy, Z., van Landeghem, M.
pression by an integrate-and-fire model of hippocampal area and Buzsaki, G. (1995b) Dentate EEG spikes and
CA3. J. Comput. Neurosci., 6: 71–90. associated interneuronal population bursts in the hippo-
Awatramani, G.B., Price, G.D. and Trussell, L.O. (2005) Mod- campal hilar region of the rat. J. Neurophysiol., 73:
ulation of transmitter release by presynaptic resting potential 1691–1705.
and background calcium levels. Neuron, 48: 109–121. Breustedt, J. and Schmitz, D. (2004) Assessing the role of
Barrionuevo, G., Kelso, S.R., Johnston, D. and Brown, T.H. GLUK5 and GLUK6 at hippocampal mossy fiber synapses.
(1986) Conductance mechanism responsible for long-term J. Neurosci., 24: 10093–10098.
potentiation in monosynaptic and isolated excitatory synap- Brown, T.H. and Johnston, D. (1983) Voltage-clamp analysis
tic inputs to hippocampus. J. Neurophysiol., 55: 540–550. of mossy fiber synaptic input to hippocampal neurons.
Bellocchio, E.E., Hu, H., Pohorille, A., Chan, J., Pickel, V.M. J. Neurophysiol., 50: 487–507.
and Edwards, R.H. (1998) The localization of the brain- Buzsaki, G. (1984) Feed-forward inhibition in the hippocampal
specific inorganic phosphate transporter suggests a specific formation. Prog. Neurobiol., 22: 131–153.
presynaptic role in glutamatergic transmission. J. Neurosci., Castillo, P.E., Malenka, R.C. and Nicoll, R.A. (1997) Kainate
18: 8648–8659. receptors mediate a slow postsynaptic current in hippocam-
Bergersen, L., Ruiz, A., Bjaalie, J.G., Kullmann, D.M. and pal CA3 neurons. Nature, 388: 182–186.
Gundersen, V. (2003) GABA and GABAA receptors at Castillo, P.E., Salin, P.A., Weisskopf, M.G. and Nicoll, R.A.
hippocampal mossy fibre synapses. Eur. J. Neurosci., 18: (1996) Characterizing the site and mode of action of dynor-
931–941. phin at hippocampal mossy fiber synapses in the guinea pig.
Berninger, B., Garcia, D.E., Inagaki, N., Hahnel, C. and Lind- J. Neurosci., 16: 5942–5950.
holm, D. (1993) BDNF and NT-3 induce intracellular Ca++ Castillo, P.E., Weisskopf, M.G. and Nicoll, R.A. (1994) The
elevation in hippocampal neurones. Neuroreport, 4: role of Ca2+ channels in hippocampal mossy fiber synaptic
1303–1306. transmission and long-term potentiation. Neuron, 12:
Bischofberger, J., Engel, D., Frotscher, M. and Jonas, P. (2006) 261–269.
Timing and efficacy of transmitter release at mossy fiber Chandy, J., Pierce, J.P. and Milner, T.A. (1995) Rat hippo-
synapses in the hippocampal network. Pflugers Arch., 453: campal mossy fibers contain cholecystokinin-like immunore-
361–372. activity. Anat. Rec., 243: 519–523.
Blackstad, T.W., Brink, K., Hem, J. and Jeune, B. (1970) Dis- Chattarji, S., Stanton, P.K. and Sejnowski, T.J. (1989) Com-
tribution of hippocampal mossy fibers in the rat. An exper- missural synapses, but not mossy fiber synapses, in hippo-
imental study with silver impregnation methods. J. Comp. campal field CA3 exhibit associative long-term potentiation
Neurol., 138: 433–449. and depression. Brain Res., 495: 145–150.
Blackstad, T.W. and Kjaerheim, A. (1961) Special axo- Chaudhry, F.A., Reimer, R.J., Bellocchio, E.E., Danbolt, N.C.,
dendritic synapses in the hippocampal cortex: electron and Osen, K.K., Edwards, R.H. and Storm-Mathisen, J. (1998)
125

The vesicular GABA transporter, VGAT, localizes to synap- Drake, C.T., Patterson, T.A., Simmons, M.L., Chavkin, C. and
tic vesicles in sets of glycinergic as well as GABAergic neu- Milner, T.A. (1996) Kappa opioid receptor-like immunore-
rons. J. Neurosci., 18: 9733–9750. activity in guinea pig brain: ultrastructural localization in
Chavkin, C. (2000) Dynorphins are endogenous opioid peptides presynaptic terminals in hippocampal formation. J. Comp.
released from granule cells to act neurohumorly and inhibit Neurol., 370: 377–395.
excitatory neurotransmission in the hippocampus. Prog. Egebjerg, J., Bettler, B., Hermans-Borgmeyer, I. and Heine-
Brain Res., 125: 363–367. mann, S. (1991) Cloning of a cDNA for a glutamate receptor
Chavkin, C., Shoemaker, W.J., McGinty, J.F., Bayon, A. and subunit activated by kainate but not AMPA. Nature, 351:
Bloom, F.E. (1985) Characterization of the prodynorphin 745–748.
and proenkephalin neuropeptide systems in rat hippo- Elmer, E., Kokaia, Z., Kokaia, M., Carnahan, J., Nawa, H.
campus. J. Neurosci., 5: 808–816. and Lindvall, O. (1998) Dynamic changes of brain-derived
Chicurel, M.E. and Harris, K.M. (1992) Three-dimensional neurotrophic factor protein levels in the rat forebrain after
analysis of the structure and composition of CA3 branched single and recurring kindling-induced seizures. Neuroscience,
dendritic spines and their synaptic relationships with mossy 83: 351–362.
fiber boutons in the rat hippocampus. J. Comp. Neurol., 325: Engel, D. and Jonas, P. (2005) Presynaptic action potential
169–182. amplification by voltage-gated Na+ channels in hippocampal
Claiborne, B.J., Amaral, D.G. and Cowan, W.M. (1986) A light mossy fiber boutons. Neuron, 45: 405–417.
and electron microscopic analysis of the mossy fibers of the Erlander, M.G. and Tobin, A.J. (1991) The structural and
rat DG. J. Comp. Neurol., 246: 435–458. functional heterogeneity of glutamic acid decarboxylase: a
Conner, J.M., Lauterborn, J.C., Yan, Q., Gall, C.M. and review. Neurochem. Res., 16: 215–226.
Varon, S. (1997) Distribution of brain-derived neurotrophic Foster, A.C., Mena, E.E., Monaghan, D.T. and Cotman, C.W.
factor (BDNF) protein and mRNA in the normal adult rat (1981) Synaptic localization of kainic acid binding sites.
CNS: evidence for anterograde axonal transport. J. Neuro- Nature, 289: 73–75.
sci., 17: 2295–2313. Frahm, C., Engel, D., Piechotta, A., Heinemann, U. and Drag-
Contractor, A., Rogers, C., Maron, C., Henkemeyer, M., uhn, A. (2000) Presence of gamma-aminobutyric acid trans-
Swanson, G.T. and Heinemann, S.F. (2002) Trans-synaptic porter mRNA in interneurons and principal cells of rat
Eph receptor-ephrin signaling in hippocampal mossy fiber hippocampus. Neurosci. Lett., 288: 175–178.
LTP. Science, 296: 1864–1869. Furtinger, S., Pirker, S., Czech, T., Baumgartner, C., Ransm-
Contractor, A., Swanson, G. and Heinemann, S.F. (2001) ayr, G. and Sperk, G. (2001) Plasticity of Y1 and Y2 recep-
Kainate receptors are involved in short- and long-term plas- tors and neuropeptide Y fibers in patients with temporal lobe
ticity at mossy fiber synapses in the hippocampus. Neuron, epilepsy. J. Neurosci., 21: 5804–5812.
29: 209–216. Gall, C. (1984) The distribution of cholecystokinin-like
Crawford, I.L. and Connor, J.D. (1973) Localization and re- immunoreactivity in the hippocampal formation of the
lease of glutamic acid in relation to the hippocampal mossy guinea pig: localization in the mossy fibers. Brain Res., 306:
fibre pathway. Nature, 244: 442–443. 73–83.
Danzer, S.C. and McNamara, J.O. (2004) Localization of Gall, C. (1988) Seizures induce dramatic and distinctly different
brain-derived neurotrophic factor to distinct terminals of changes in enkephalin, dynorphin, and CCK immunoreac-
mossy fiber axons implies regulation of both excitation and tivities in mouse hippocampal mossy fibers. J. Neurosci., 8:
feedforward inhibition of CA3 pyramidal cells. J. Neurosci., 1852–1862.
24: 11346–11355. Gall, C., Brecha, N., Karten, H.J. and Chang, K.J. (1981) Lo-
Darstein, M., Petralia, R.S., Swanson, G.T., Wenthold, R.J. calization of enkephalin-like immunoreactivity to identified
and Heinemann, S.F. (2003) Distribution of kainate receptor axonal and neuronal populations of the rat hippocampus. J.
subunits at hippocampal mossy fiber synapses. J. Neurosci., Comp. Neurol., 198: 335–350.
23: 8013–8019. Gall, C. and Lauterborn, J. (1992) The DG: a model system for
Debanne, D., Gahwiler, B.H. and Thompson, S.M. (1998) studies of neurotrophin regulation. Epilepsy Res. Suppl., 7:
Long-term synaptic plasticity between pairs of individual 171–185.
CA3 pyramidal cells in rat hippocampal slice cultures. J. Gall, C., Lauterborn, J., Isackson, P. and White, J. (1990) Sei-
Physiol., 507(Pt 1): 237–247. zures, neuropeptide regulation, and mRNA expression in the
Ding, R., Asada, H. and Obata, K. (1998) Changes in extra- hippocampus. Prog. Brain Res., 83: 371–390.
cellular glutamate and GABA levels in the hippocampal CA3 Gall, C.M. and Isackson, P.J. (1989) Limbic seizures increase
and CA1 areas and the induction of glutamic acid dec- neuronal production of messenger RNA for nerve growth
arboxylase-67 in dentate granule cells of rats treated with factor. Science, 245: 758–761.
kainic acid. Brain Res., 800: 105–113. Geiger, J.R. and Jonas, P. (2000) Dynamic control of presy-
Draguhn, A., Verdorn, T.A., Ewert, M., Seeburg, P.H. and naptic Ca(2+) inflow by fast-inactivating K(+) channels in
Sakmann, B. (1990) Functional and molecular distinction hippocampal mossy fiber boutons. Neuron, 28: 927–939.
between recombinant rat GABAA receptor subtypes by Gloor, P., Vera, C.L. and Sperti, L. (1963) Electrophysiological
Zn2+. Neuron, 5: 781–788. Studies of Hippocampal Neurons. I. Configuration and
126

laminar analysis of the ‘‘resting’’ potential gradient, of the currents in hippocampal CA3 pyramidal neurons are of
main-transient response to perforant path, fimbrial and mossy fiber origin. J. Neurophysiol., 77: 1075–1086.
mossy fiber volleys and of ‘‘spontaneous’’ activity. Elect- Henze, D.A., McMahon, D.B., Harris, K.M. and Barrionuevo,
roencephalogr. Clin. Neurophysiol., 15: 353–378. G. (2002a) Giant miniature EPSCs at the hippocampal mossy
Gomez-Lira, G., Lamas, M., Romo-Parra, H. and Gutierrez, fiber to CA3 pyramidal cell synapse are monoquantal. J.
R. (2005) Programmed and induced phenotype of the hip- Neurophysiol., 87: 15–29.
pocampal granule cells. J. Neurosci., 25: 6939–6946. Henze, D.A., Wittner, L. and Buzsaki, G. (2002b) Single gran-
Gomez-Lira, G., Trillo, E., Ramirez, M., Asai, M., Sitges, M. ule cells reliably discharge targets in the hippocampal CA3
and Gutierrez, R. (2002) The expression of GABA in mossy network in vivo. Nat. Neurosci., 5: 790–795.
fiber synaptosomes coincides with the seizure-induced ex- Ho, M.W., Beck-Sickinger, A.G. and Colmers, W.F. (2000)
pression of GABAergic transmission in the mossy fiber Neuropeptide Y(5) receptors reduce synaptic excitation in
synapse. Exp. Neurol., 177: 276–283. proximal subiculum, but not epileptiform activity in rat hip-
Gorter, J.A., Van Vliet, E.A., Proper, E.A., De Graan, P.N., pocampal slices. J. Neurophysiol., 83: 723–734.
Ghijsen, W.E., Lopes Da Silva, F.H. and Aronica, E. (2002) Holm, I.E., Geneser, F.A. and Zimmer, J. (1993) Cholecysto-
Glutamate transporters alterations in the reorganizing DG kinin-, enkephalin-, and substance P-like immunoreactivity in
are associated with progressive seizure activity in chronic the dentate area, hippocampus, and subiculum of the do-
epileptic rats. J. Comp. Neurol., 442: 365–377. mestic pig. J. Comp. Neurol., 331: 310–325.
Gray, R., Rajan, A.S., Radcliffe, K.A., Yakehiro, M. and Dani, Houser, C.R., Miyashiro, J.E., Swartz, B.E., Walsh, G.O.,
J.A. (1996) Hippocampal synaptic transmission enhanced by Rich, J.R. and Delgado-Escueta, A.V. (1990) Altered pat-
low concentrations of nicotine. Nature, 383: 713–716. terns of dynorphin immunoreactivity suggest mossy fiber re-
Griffith, W.H., Brown, T.H. and Johnston, D. (1986) Voltage- organization in human hippocampal epilepsy. J. Neurosci.,
clamp analysis of synaptic inhibition during long-term po- 10: 267–282.
tentiation in hippocampus. J. Neurophysiol., 55: 767–775. Huang, E.J. and Reichardt, L.F. (2001) Neurotrophins: roles in
Gutierrez, R. (2000) Seizures induce simultaneous GABAergic neuronal development and function. Annu. Rev. Neurosci.,
and glutamatergic transmission in the DG-CA3 system. J. 24: 677–736.
Neurophysiol., 84: 3088–3090. Isackson, P.J., Huntsman, M.M., Murray, K.D. and
Gutierrez, R. (2002) Activity-dependent expression of simulta- Gall, C.M. (1991) BDNF mRNA expression is increased
neous glutamatergic and GABAergic neurotransmission in adult rat forebrain after limbic seizures: temporal
from the mossy fibers in vitro. J. Neurophysiol., 87: patterns of induction distinct from NGF. Neuron, 6:
2562–2570. 937–948.
Gutierrez, R. (2003) The GABAergic phenotype of the ‘‘glut- Ito, I. and Sugiyama, H. (1991) Roles of glutamate receptors in
amatergic’’ granule cells of the DG. Prog. Neurobiol., 71: long-term potentiation at hippocampal mossy fiber synapses.
337–358. Neuroreport, 2: 333–336.
Gutierrez, R. (2005) The dual glutamatergic-GABAergic phe- Iversen, L.L. and Kelly, J.S. (1975) Uptake and metabolism of
notype of hippocampal granule cells. Trends Neurosci., 28: gamma-aminobutyric acid by neurones and glial cells. Bioc-
297–303. hem. Pharmacol., 24: 933–938.
Gutierrez, R. and Heinemann, U. (2001) Kindling induces Jacques, D., Dumont, Y., Fournier, A. and Quirion, R. (1997)
transient fast inhibition in the DG-CA3 projection. Eur. J. Characterization of neuropeptide Y receptor subtypes in the
Neurosci., 13: 1371–1379. normal human brain, including the hypothalamus. Neuro-
Gutierrez, R., Romo-Parra, H., Maqueda, J., Vivar, C., Ra- science, 79: 129–148.
mirez, M., Morales, M.A. and Lamas, M. (2003) Plasticity of Jaffe, D. and Johnston, D. (1990) Induction of long-term po-
the GABAergic phenotype of the ‘‘glutamatergic’’ granule tentiation at hippocampal mossy-fiber synapses follows a
cells of the rat DG. J. Neurosci., 23: 5594–5598. Hebbian rule. J. Neurophysiol., 64: 948–960.
Hallermann, S., Pawlu, C., Jonas, P. and Heckmann, M. (2003) Jaffe, D.B. and Brown, T.H. (1997) Calcium dynamics in
A large pool of releasable vesicles in a cortical glutamatergic thorny excrescences of CA3 pyramidal neurons. J. Ne-
synapse. Proc. Natl. Acad. Sci. U.S.A., 100: 8975–8980. urophysiol., 78: 10–18.
Hamlyn, L.H. (1961) Electron microscopy of mossy fibre end- Jeub, M., Lie, A., Blumcke, I., Elger, C.E. and Beck, H. (1999)
ings in Ammon’s horn. Nature, 190: 645–646. Loss of dynorphin-mediated inhibition of voltage-dependent
Harris, E.W. and Cotman, C.W. (1986) Long-term potentiation Ca2+ currents in hippocampal granule cells isolated from
of guinea pig mossy fiber responses is not blocked by N- epilepsy patients is associated with mossy fiber sprouting.
methyl D-aspartate antagonists. Neurosci. Lett., 70: 132–137. Neuroscience, 94: 465–471.
He, X.P., Minichiello, L., Klein, R. and McNamara, J.O. Jin, W. and Chavkin, C. (1999) Mu opioids enhance mossy fiber
(2002) Immunohistochemical evidence of seizure-induced ac- synaptic transmission indirectly by reducing GABAB recep-
tivation of trkB receptors in the mossy fiber pathway of adult tor activation. Brain Res., 821: 286–293.
mouse hippocampus. J. Neurosci., 22: 7502–7508. Johnston, D. and Amaral, D.G. (2004) Hippocampus. In:
Henze, D.A., Card, J.P., Barrionuevo, G. and Ben-Ari, Y. Shepherd G.M. (Ed.), The Synaptic Organization of the
(1997) Large amplitude miniature excitatory postsynaptic Brain. Oxford University Press, New York.
127

Johnston, D. and Brown, T.H. (1983) Interpretation of voltage- Kukley, M., Schwan, M., Fredholm, B.B. and Dietrich, D.
clamp measurements in hippocampal neurons. J. Ne- (2005) The role of extracellular adenosine in regulating mossy
urophysiol., 50: 464–486. fiber synaptic plasticity. J. Neurosci., 25: 2832–2837.
Jonas, P., Major, G. and Sakmann, B. (1993) Quantal compo- Kullmann, D.M., Ruiz, A., Rusakov, D.M., Scott, R., Semya-
nents of unitary EPSCs at the mossy fibre synapse on CA3 nov, A. and Walker, M.C. (2005) Presynaptic, extrasynaptic
pyramidal cells of rat hippocampus. J. Physiol., 472: 615–663. and axonal GABAA receptors in the CNS: where and why?
Jung, M.W. and McNaughton, B.L. (1993) Spatial selectivity Prog. Biophys. Mol. Biol., 87: 33–46.
of unit activity in the hippocampal granular layer. Lamas, M., Gomez-Lira, G. and Gutierrez, R. (2001) Vesicular
Hippocampus, 3: 165–182. GABA transporter mRNA expression in the DG and in
Kafitz, K.W., Rose, C.R., Thoenen, H. and Konnerth, A. mossy fiber synaptosomes. Brain Res. Mol. Brain Res., 93:
(1999) Neurotrophin-evoked rapid excitation through TrkB 209–214.
receptors. Nature, 401: 918–921. Langdon, R.B., Johnson, J.W. and Barrionuevo, G. (1993) As-
Kamiya, H. and Ozawa, S. (1999) Dual mechanism for presy- ynchrony of mossy fibre inputs and excitatory postsynaptic
naptic modulation by axonal metabotropic glutamate recep- currents in rat hippocampus. J. Physiol., 472: 157–176.
tor at the mouse mossy fibre-CA3 synapse. J. Physiol., Lanthorn, T.H., Ganong, A.H. and Cotman, C.W. (1984)
518(Pt 2): 497–506. 2-Amino-4-phosphonobutyrate selectively blocks mossy
Kamiya, H., Sawada, S. and Yamamoto, C. (1988) Synthetic fiber-CA3 responses in guinea pig but not rat hippocampus.
omega-conotoxin blocks synaptic transmission in the hippo- Brain Res., 290: 174–178.
campus in vitro. Neurosci. Lett., 91: 84–88. Lauri, S.E., Bortolotto, Z.A., Bleakman, D., Ornstein, P.L.,
Kamiya, H., Shinozaki, H. and Yamamoto, C. (1996) Activa- Lodge, D., Isaac, J.T. and Collingridge, G.L. (2001) A crit-
tion of metabotropic glutamate receptor type 2/3 suppresses ical role of a facilitatory presynaptic kainate receptor in
transmission at rat hippocampal mossy fibre synapses. J. mossy fiber LTP. Neuron, 32: 697–709.
Physiol., 493(Pt 2): 447–455. Lauri, S.E., Bortolotto, Z.A., Nistico, R., Bleakman, D.,
Kaneko, T., Fujiyama, F. and Hioki, H. (2002) Immunohisto- Ornstein, P.L., Lodge, D., Isaac, J.T. and Collingridge,
chemical localization of candidates for vesicular glutamate G.L. (2003) A role for Ca2+ stores in kainate receptor-
transporters in the rat brain. J. Comp. Neurol., 444: 39–62. dependent synaptic facilitation and LTP at mossy fiber
Kapur, A., Yeckel, M. and Johnston, D. (2001) Hippocampal synapses in the hippocampus. Neuron, 39: 327–341.
mossy fiber activity evokes Ca2+ release in CA3 pyramidal Lawrence, J.J., Grinspan, Z.M. and McBain, C.J. (2004)
neurons via a metabotropic glutamate receptor pathway. Quantal transmission at mossy fibre targets in the CA3
Neuroscience, 107: 59–69. region of the rat hippocampus. J. Physiol., 554: 175–193.
Kapur, A., Yeckel, M.F., Gray, R. and Johnston, D. (1998) Lehmann, H., Ebert, U. and Loscher, W. (1996) Immunocyto-
L-type calcium channels are required for one form of chemical localization of GABA immunoreactivity in dentate
hippocampal mossy fiber LTP. J. Neurophysiol., 79: granule cells of normal and kindled rats. Neurosci. Lett., 212:
2181–2190. 41–44.
Kasyanov, A.M., Safiulina, V.F., Voronin, L.L. and Cherubini, Lei, S. and McBain, C.J. (2002) Distinct NMDA receptors
E. (2004) GABA-mediated giant depolarizing potentials as provide differential modes of transmission at mossy fiber-
coincidence detectors for enhancing synaptic efficacy in the interneuron synapses. Neuron, 33: 921–933.
developing hippocampus. Proc. Natl. Acad. Sci. U.S.A., 101: Liang, Y., Yuan, L.L., Johnston, D. and Gray, R. (2002) Cal-
3967–3972. cium signaling at single mossy fiber presynaptic terminals in
Kaufman, D.L., Houser, C.R. and Tobin, A.J. (1991) Two the rat hippocampus. J. Neurophysiol., 87: 1132–1137.
forms of the gamma-aminobutyric acid synthetic enzyme Lie, A.A., Becker, A., Behle, K., Beck, H., Malitschek, B.,
glutamate decarboxylase have distinct intraneuronal distri- Conn, P.J., Kuhn, R., Nitsch, R., Plaschke, M., Schramm, J.,
butions and cofactor interactions. J. Neurochem., 56: Elger, C.E., Wiestler, O.D. and Blumcke, I. (2000) Up-reg-
720–723. ulation of the metabotropic glutamate receptor mGluR4 in
von Kitzing, E., Jonas, P. and Sakmann, B. (1994) Quantal hippocampal neurons with reduced seizure vulnerability.
analysis of excitatory postsynaptic currents at the hippocam- Ann. Neurol., 47: 26–35.
pal mossy fiber-CA3 pyramidal cell synapse. Adv. Second Liu, H., Mazarati, A.M., Katsumori, H., Sankar, R. and
Messenger Phosphoprotein Res., 29: 235–260. Wasterlain, C.G. (1999) Substance P is expressed in hippo-
Knowles, W.D., Schneiderman, J.H., Wheal, H.V., Stafstrom, campal principal neurons during status epilepticus and plays
C.E. and Schwartzkroin, P.A. (1984) Hyperpolarizing poten- a critical role in the maintenance of status epilepticus. Proc.
tials in guinea pig hippocampal CA3 neurons. Cell Mol. Natl. Acad. Sci. U.S.A., 96: 5286–5291.
Neurobiol., 4: 207–230. Lorente de No, E. (1934) Studies on the structure of the cer-
Kobayashi, K., Manabe, T. and Takahashi, T. (1996) Presy- ebral cortex II. Continuation of the study of the ammonic
naptic long-term depression at the hippocampal mossy fiber- system. J. Psychol. Neurol., 46: 113–177.
CA3 synapse. Science, 273: 648–650. Lowenstein, D.H. and Arsenault, L. (1996) The effects of
Kohonen, T. (1977) Associative Memory: A System Theoretic growth factors on the survival and differentiation of cultured
Approach. Springer-Verlag, New York. DG neurons. J. Neurosci., 16: 1759–1769.
128

Maccaferri, G., Toth, K. and McBain, C.J. (1998) Target- afferent stimulation of CA3 pyramidal neurons in hippo-
specific expression of presynaptic mossy fiber plasticity. campal slices. J. Neurophysiol., 76: 554–562.
Science, 279: 1368–1370. Min, M.Y., Rusakov, D.A. and Kullmann, D.M. (1998)
Makiura, Y., Suzuki, F., Chevalier, E. and Onteniente, B. Activation of AMPA, kainate, and metabotropic receptors
(1999) Excitatory granule cells of the DG exhibit a double at hippocampal mossy fiber synapses: role of glutamate
inhibitory neurochemical content after intrahippocampal ad- diffusion. Neuron, 21: 561–570.
ministration of kainate in adult mice. Exp. Neurol., 159: Miyazaki, K., Ishizuka, T. and Yawo, H. (2005) Synapse-to-
73–83. synapse variation of calcium channel subtype contributions
Manzoni, O.J., Castillo, P.E. and Nicoll, R.A. (1995) Pharma- in large mossy fiber terminals of mouse hippocampus.
cology of metabotropic glutamate receptors at the mossy Neuroscience, 136: 1003–1014.
fiber synapses of the guinea pig hippocampus. Neurophar- Molnar, P. and Nadler, J.V. (2001) Synaptically-released zinc
macology, 34: 965–971. inhibits N-methyl-D-aspartate receptor activation at recurrent
Maqueda, J., Ramirez, M., Lamas, M. and Gutierrez, R. (2003) mossy fiber synapses. Brain Res., 910: 205–207.
Glutamic acid decarboxylase (GAD)67, but not GAD65, is Monaghan, D.T. and Cotman, C.W. (1982) The distribution of
constitutively expressed during development and transiently [3H]kainic acid binding sites in rat CNS as determined by
overexpressed by activity in the granule cells of the rat. Ne- autoradiography. Brain Res., 252: 91–100.
urosci. Lett., 353: 69–71. Monaghan, D.T. and Cotman, C.W. (1985) Distribution of
Marr, D. (1971) Simple memory: a theory of archicortex. Phi- N-methyl-D-aspartate-sensitive L-[3H]glutamate-binding
los. Trans. R. Soc. Lond., 262: 23–81. sites in rat brain. J. Neurosci., 5: 2909–2919.
Marsh, D.J., Baraban, S.C., Hollopeter, G. and Palmiter, R.D. Moore, K.A., Nicoll, R.A. and Schmitz, D. (2003) Adenosine
(1999) Role of the Y5 neuropeptide Y receptor in limbic sei- gates synaptic plasticity at hippocampal mossy fiber
zures. Proc. Natl. Acad. Sci. U.S.A., 96: 13518–13523. synapses. Proc. Natl. Acad. Sci. U.S.A., 100: 14397–14402.
Marty, S., Berninger, B., Carroll, P. and Thoenen, H. (1996a) Nadler, J.V. (2003) The recurrent mossy fiber pathway of the
GABAergic stimulation regulates the phenotype of hippo- epileptic brain. Neurochem. Res., 28: 1649–1658.
campal interneurons through the regulation of brain-derived Nakazawa, K., Quirk, M.C., Chitwood, R.A., Watanabe, M.,
neurotrophic factor. Neuron, 16: 565–570. Yeckel, M.F., Sun, L.D., Kato, A., Carr, C.A., Johnston, D.,
Marty, S., Berzaghi Mda, P. and Berninger, B. (1997) Ne- Wilson, M.A. and Tonegawa, S. (2002) Requirement for
urotrophins and activity-dependent plasticity of cortical in- hippocampal CA3 NMDA receptors in associative memory
terneurons. Trends Neurosci., 20: 198–202. recall. Science, 297: 211–218.
Marty, S., Carroll, P., Cellerino, A., Castren, E., Staiger, V., Neuman, R.S., Ben-Ari, Y., Gho, M. and Cherubini, E. (1988)
Thoenen, H. and Lindholm, D. (1996b) Brain-derived ne- Blockade of excitatory synaptic transmission by 6-cyano-7-
urotrophic factor promotes the differentiation of various nitroquinoxaline-2,3-dione (CNQX) in the hippocampus in
hippocampal nonpyramidal neurons, including Cajal-Retzius vitro. Neurosci. Lett., 92: 64–68.
cells, in organotypic slice cultures. J. Neurosci., 16: 675–687. Nicoll, R.A., Castillo, P.E. and Weisskopf, M.G. (1994) The
McCarthy, J.B., Walker, M., Pierce, J., Camp, P. and White, role of Ca2+ in transmitter release and long-term potentia-
J.D. (1998) Biosynthesis and metabolism of native and ox- tion at hippocampal mossy fiber synapses. Adv. Second
idized neuropeptide Y in the hippocampal mossy fiber sys- Messenger Phosphoprotein Res., 29: 497–505.
tem. J. Neurochem., 70: 1950–1963. Nicoll, R.A. and Schmitz, D. (2005) Synaptic plasticity at hip-
McGinty, J.F., Henriksen, S.J., Goldstein, A., Terenius, L. and pocampal mossy fibre synapses. Nat. Rev. Neurosci., 6:
Bloom, F.E. (1983) Dynorphin is contained within hippo- 863–876.
campal mossy fibers: immunochemical alterations after Ogata, N. and Ueno, S. (1976) Mode of activation of pyramidal
kainic acid administration and colchicine-induced neurotox- neurons by mossy fiber stimulation in thin hippocampal slices
icity. Proc. Natl. Acad. Sci. U.S.A., 80: 589–593. in vitro. Exp. Neurol., 53: 567–584.
McLean, S., Rothman, R.B., Jacobson, A.E., Rice, K.C. and Ohishi, H., Akazawa, C., Shigemoto, R., Nakanishi, S. and
Herkenham, M. (1987) Distribution of opiate receptor sub- Mizuno, N. (1995) Distributions of the mRNAs for L-2-
types and enkephalin and dynorphin immunoreactivity in the amino-4-phosphonobutyrate-sensitive metabotropic gluta-
hippocampus of squirrel, guinea pig, rat, and hamster. J. mate receptors, mGluR4 and mGluR7, in the rat brain. J.
Comp. Neurol., 255: 497–510. Comp. Neurol., 360: 555–570.
McNaughton, B.L. and Morris, R.G.M. (1987) Hippocampal Ohishi, H., Shigemoto, R., Nakanishi, S. and Mizuno, N.
synaptic enhancement and information storage within a dis- (1993) Distribution of the mRNA for a metabotropic gluta-
tributed memory system. Trends Neurosci., 10: 408–415. mate receptor (mGluR3) in the rat brain: an in situ hybrid-
McQuiston, A.R. and Colmers, W.F. (1996) Neuropeptide Y2 ization study. J. Comp. Neurol., 335: 252–266.
receptors inhibit the frequency of spontaneous but not min- Parra, P., Gulyas, A.I. and Miles, R. (1998) How many subtypes
iature EPSCs in CA3 pyramidal cells of rat hippocampus. J. of inhibitory cells in the hippocampus? Neuron, 20: 983–993.
Neurophysiol., 76: 3159–3168. Patrylo, P.R., van den Pol, A.N., Spencer, D.D. and
Miller, L.D., Petrozzino, J.J., Golarai, G. and Connor, J.A. Williamson, A. (1999) NPY inhibits glutamatergic excitation
(1996) Ca2+ release from intracellular stores induced by in the epileptic human DG. J. Neurophysiol., 82: 478–483.
129

Patz, S., Wirth, M.J., Gorba, T., Klostermann, O. and Ribak, C.E., Tong, W.M. and Brecha, N.C. (1996) GABA
Wahle, P. (2003) Neuronal activity and neurotrophic factors plasma membrane transporters, GAT-1 and GAT-3, display
regulate GAD-65/67 mRNA and protein expression in different distributions in the rat hippocampus. J. Comp. Ne-
organotypic cultures of rat visual cortex. Eur. J. Neurosci., urol., 367: 595–606.
18: 1–12. Rivkees, S.A., Price, S.L. and Zhou, F.C. (1995) Immunohisto-
Pavlidis, P. and Madison, D.V. (1999) Synaptic transmission in chemical detection of A1 adenosine receptors in rat brain
pair recordings from CA3 pyramidal cells in organotypic with emphasis on localization in the hippocampal formation,
culture. J. Neurophysiol., 81: 2787–2797. cerebral cortex, cerebellum, and basal ganglia. Brain Res.,
Pelkey, K.A., Lavezzari, G., Racca, C., Roche, K.W. and 677: 193–203.
McBain, C.J. (2005) mGluR7 is a metaplastic switch con- Rolls, E.T. (1989) Functions of neuronal networks in the hip-
trolling bidirectional plasticity of feedforward inhibition. pocampus and neocortex in memory. In: Byrne J.H. and
Neuron, 46: 89–102. Berry W.O. (Eds.), Neural Models of Plasticity. Academic
Pelkey, K.A., Topolnik, L., Lacaille, J.C. and McBain, C.J. Press, San Diego, CA.
(2006) Compartmentalized Ca2+ channel regulation at Rolls, E.T. and Kesner, R.P. (2006) A computational theory of
divergent mossy-fiber release sites underlies target cell- hippocampal function, and empirical tests of the theory.
dependent plasticity. Neuron, 52: 497–510. Prog. Neurobiol., 79: 1–48.
Penttonen, M., Kamondi, A., Sik, A., Acsady, L. and Buzsaki, Romo-Parra, H., Vivar, C., Maqueda, J., Morales, M.A. and
G. (1997) Feed-forward and feed-back activation of the DG Gutierrez, R. (2003) Activity-dependent induction of multi-
in vivo during dentate spikes and sharp wave bursts. Hip- transmitter signaling onto pyramidal cells and interneurons
pocampus, 7: 437–450. of hippocampal area CA3. J. Neurophysiol., 89: 3155–3167.
Podlogar, M. and Dietrich, D. (2006) Firing pattern of rat Rose, C.R., Blum, R., Pichler, B., Lepier, A., Kafitz, K.W. and
hippocampal neurons: a perforated patch clamp study. Brain Konnerth, A. (2003) Truncated TrkB-T1 mediates neurotro-
Res., 1085: 95–101. phin-evoked calcium signalling in glia cells. Nature, 426:
Qian, J. and Noebels, J.L. (2005) Visualization of transmitter 74–78.
release with zinc fluorescence detection at the mouse hippo- Rothstein, J.D., Martin, L., Levey, A.I., Dykes-Hoberg, M.,
campal mossy fibre synapse. J. Physiol., 566: 747–758. Jin, L., Wu, D., Nash, N. and Kuncl, R.W. (1994) Local-
Radian, R., Ottersen, O.P., Storm-Mathisen, J., Castel, M. and ization of neuronal and glial glutamate transporters. Neuron,
Kanner, B.I. (1990) Immunocytochemical localization of the 13: 713–725.
GABA transporter in rat brain. J. Neurosci., 10: 1319–1330. Ruiz, A., Fabian-Fine, R., Scott, R., Walker, M.C., Rusakov,
Ramirez, M. and Gutierrez, R. (2001) Activity-dependent ex- D.A. and Kullmann, D.M. (2003) GABAA receptors at hip-
pression of GAD67 in the granule cells of the rat hippocam- pocampal mossy fibers. Neuron, 39: 961–973.
pus. Brain Res., 917: 139–146. Ruiz, A., Walker, M.C., Fabian-Fine, R. and Kullmann, D.M.
Ramon y Cajal, S. (1911) Histology of the nervous system of (2004) Endogenous zinc inhibits GABA(A) receptors in a
man and vertebrates. Translated from Spanish to French by hippocampal pathway. J. Neurophysiol., 91: 1091–1096.
L. Azoulay, from the French to English by N. Swanson and Safiulina, V.F., Fattorini, G., Conti, F. and Cherubini, E.
L.W. Swanson (1995) Oxford University Press, New York. (2006) GABAergic signaling at mossy fiber synapses in neo-
Rao, A., Cha, E.M. and Craig, A.M. (2000) Mismatched natal rat hippocampus. J. Neurosci., 26: 597–608.
appositions of presynaptic and postsynaptic components in Salin, P.A., Scanziani, M., Malenka, R.C. and Nicoll, R.A.
isolated hippocampal neurons. J. Neurosci., 20: 8344–8353. (1996) Distinct short-term plasticity at two excitatory
Rattray, M. and Priestley, J.V. (1993) Differential expression of synapses in the hippocampus. Proc. Natl. Acad. Sci.
GABA transporter-1 messenger RNA in subpopulations of U.S.A., 93: 13304–13309.
GABA neurones. Neurosci. Lett., 156: 163–166. Salin, P.A., Weisskopf, M.G. and Nicoll, R.A. (1995) A com-
Reid, C.A., Dixon, D.B., Takahashi, M., Bliss, T.V. and Fine, parison of the role of dynorphin in the hippocampal mossy
A. (2004) Optical quantal analysis indicates that long-term fiber pathway in guinea pig and rat. J. Neurosci., 15:
potentiation at single hippocampal mossy fiber synapses is 6939–6945.
expressed through increased release probability, recruitment Sandler, R. and Smith, A.D. (1991) Coexistence of GABA and
of new release sites, and activation of silent synapses. J. glutamate in mossy fiber terminals of the primate hippocam-
Neurosci., 24: 3618–3626. pus: an ultrastructural study. J. Comp. Neurol., 303:
Reid, C.A., Fabian-Fine, R. and Fine, A. (2001) Postsynaptic 177–192.
calcium transients evoked by activation of individual hippo- Sawada, S., Higashima, M. and Yamamoto, C. (1988) Kainic
campal mossy fiber synapses. J. Neurosci., 21: 2206–2214. acid induces long-lasting depolarizations in hippocampal
Represa, A., Tremblay, E. and Ben-Ari, Y. (1987) Kainate neurons only when applied to stratum lucidum. Exp. Brain
binding sites in the hippocampal mossy fibers: localization Res., 72: 135–140.
and plasticity. Neuroscience, 20: 739–748. Sawada, S., Takada, S. and Yamamoto, C. (1983) Selective
Reuter, H. (1995) Measurements of exocytosis from single pre- activation of synapses near the tip of drug-ejecting micro-
synaptic nerve terminals reveal heterogeneous inhibition by electrode, and effects of antagonists of excitatory amino acids
Ca(2+)-channel blockers. Neuron, 14: 773–779. in the hippocampus. Brain Res., 267: 156–160.
130

Scanziani, M., Gahwiler, B.H. and Thompson, S.M. (1993) Smith, M.A., Zhang, L.X., Lyons, W.E. and Mamounas, L.A.
Presynaptic inhibition of excitatory synaptic transmission (1997) Anterograde transport of endogenous brain-derived
mediated by alpha adrenergic receptors in area CA3 of the neurotrophic factor in hippocampal mossy fibers. Neurore-
rat hippocampus in vitro. J. Neurosci., 13: 5393–5401. port, 8: 1829–1834.
Scharfman, H.E., Goodman, J.H., Sollas, A.L. and Croll, S.D. Sperk, G., Schwarzer, C., Heilman, J., Furtinger, S., Reimer,
(2002) Spontaneous limbic seizures after intrahippocampal R.J., Edwards, R.H. and Nelson, N. (2003) Expression of
infusion of brain-derived neurotrophic factor. Exp. Neurol., plasma membrane GABA transporters but not of the vesic-
174: 201–214. ular GABA transporter in dentate granule cells after kainic
Scharfman, H.E. (2005) Bain-derived neurotrophic factor and acid seizures. Hippocampus, 13: 806–815.
epilepsy-A missing link? Epilepsy Curr., 5:83–88. Spruston, N., Jaffe, D.B., Williams, S.H. and Johnston, D.
Scharfman, H.E., Kunkel, D.D. and Schwartzkroin, P.A. (1993) Voltage- and space-clamp errors associated with the
(1990) Synaptic connections of dentate granule cells and measurement of electrotonically remote synaptic events. J.
hilar neurons: results of paired intracellular recordings and Neurophysiol., 70: 781–802.
intracellular horseradish peroxidase injections. Neuroscience, Spruston, N. and Johnston, D. (1992) Perforated patch-clamp
37: 693–707. analysis of the passive membrane properties of three classes
Schmitz, D., Frerking, M. and Nicoll, R.A. (2000) Synaptic of hippocampal neurons. J. Neurophysiol., 67: 508–529.
activation of presynaptic kainate receptors on hippocampal Spruston, N., Lubke, J. and Frotscher, M. (1997) Interneurons
mossy fiber synapses. Neuron, 27: 327–338. in the stratum lucidum of the rat hippocampus: an anatom-
Schwarzer, C., Kofler, N. and Sperk, G. (1998) Up-regulation ical and electrophysiological characterization. J. Comp.
of neuropeptide Y-Y2 receptors in an animal model of tem- Neurol., 385: 427–440.
poral lobe epilepsy. Mol. Pharmacol., 53: 6–13. Stengaard-Pedersen, K., Fredens, K. and Larsson, L.I. (1981)
Schwarzer, C. and Sperk, G. (1995) Hippocampal granule cells Enkephalin and zinc in the hippocampal mossy fiber system.
express glutamic acid decarboxylase-67 after limbic seizures Brain Res., 212: 230–233.
in the rat. Neuroscience, 69: 705–709. Swanson, T.H., Drazba, J.A. and Rivkees, S.A. (1995) Adeno-
Scott, R. and Rusakov, D.A. (2006) Main determinants of sine A1 receptors are located predominantly on axons in the
presynaptic Ca2+ dynamics at individual mossy fiber-CA3 rat hippocampal formation. J. Comp. Neurol., 363: 517–531.
pyramidal cell synapses. J. Neurosci., 26: 7071–7081. Szabo, G., Kartarova, Z., Hoertnagl, B., Somogyi, R. and
Sepkuty, J.P., Cohen, A.S., Eccles, C., Rafiq, A., Behar, K., Sperk, G. (2000) Differential regulation of adult and embry-
Ganel, R., Coulter, D.A. and Rothstein, J.D. (2002) A onic glutamate decarboxylases in rat dentate granule cells
neuronal glutamate transporter contributes to neurotrans- after kainate-induced limbic seizures. Neuroscience, 100:
mitter GABA synthesis and epilepsy. J. Neurosci., 22: 287–295.
6372–6379. Tashiro, A., Dunaevsky, A., Blazeski, R., Mason, C.A. and
Shigemoto, R., Kinoshita, A., Wada, E., Nomura, S., Ohishi, Yuste, R. (2003) Bidirectional regulation of hippocampal
H., Takada, M., Flor, P.J., Neki, A., Abe, T., Nakanishi, S. mossy fiber filopodial motility by kainate receptors: a two-
and Mizuno, N. (1997) Differential presynaptic localization step model of synaptogenesis. Neuron, 38: 773–784.
of metabotropic glutamate receptor subtypes in the rat hip- Taupin, P., Ben-Ari, Y. and Roisin, M.P. (1994a) Subcellular
pocampus. J. Neurosci., 17: 7503–7522. fractionation on Percoll gradient of mossy fiber synapto-
Shigemoto, R., Kulik, A., Roberts, J.D., Ohishi, H., Nusser, Z., somes: evoked release of glutamate, GABA, aspartate and
Kaneko, T. and Somogyi, P. (1996) Target-cell-specific con- glutamate decarboxylase activity in control and degranulated
centration of a metabotropic glutamate receptor in the pre- rat hippocampus. Brain Res., 644: 313–321.
synaptic active zone. Nature, 381: 523–525. Taupin, P., Zini, S., Cesselin, F., Ben-Ari, Y. and Roisin, M.P.
Simmons, M.L. and Chavkin, C. (1996) k-Opioid receptor ac- (1994b) Subcellular fractionation on Percoll gradient of
tivation of a dendrotoxin-sensitive potassium channel medi- mossy fiber synaptosomes: morphological and biochemical
ates presynaptic inhibition of mossy fiber neurotransmitter characterization in control and degranulated rat hippocam-
release. Mol. Pharmacol., 50: 80–85. pus. J. Neurochem., 62: 1586–1595.
Skaggs, W.E., McNaughton, B.L., Wilson, M.A. and Barnes, Terrian, D.M., Johnston, D., Claiborne, B.J., Ansah-Yiadom,
C.A. (1996) Theta phase precession in hippocampal neuronal R., Strittmatter, W.J. and Rea, M.A. (1988) Glutamate and
populations and the compression of temporal sequences. dynorphin release from a subcellular fraction enriched in
Hippocampus, 6: 149–172. hippocampal mossy fiber synaptosomes. Brain Res. Bull., 21:
Sloviter, R.S., Dichter, M.A., Rachinsky, T.L., Dean, E., 343–351.
Goodman, J.H., Sollas, A.L. and Martin, D.L. (1996) Basal Thoenen, H. (1995) Neurotrophins and neuronal plasticity.
expression and induction of glutamate decarboxylase and Science, 270: 593–598.
GABA in excitatory granule cells of the rat and monkey Thoenen, H. (2000) Neurotrophins and activity-dependent
hippocampal DG. J. Comp. Neurol., 373: 593–618. plasticity. Prog. Brain Res., 128: 183–191.
Smart, T.G., Xie, X. and Krishek, B.J. (1994) Modulation of Thompson, S.M. and Gahwiler, B.H. (1992) Comparison of the
inhibitory and excitatory amino acid receptor ion channels by actions of baclofen at pre- and postsynaptic receptors in the
zinc. Prog. Neurobiol., 42: 393–441. rat hippocampus in vitro. J. Physiol., 451: 329–345.
131

Tokunaga, T., Miyazaki, K., Koseki, M., Mobarakeh, J.I., Vivar, C. and Gutierrez, R. (2005) Blockade of the membranal
Ishizuka, T. and Yawo, H. (2004) Pharmacological dissection GABA transporter potentiates GABAergic responses evoked
of calcium channel subtype-related components of strontium in pyramidal cells by mossy fiber activation after seizures.
inflow in large mossy fiber boutons of mouse hippocampus. Hippocampus, 15: 281–284.
Hippocampus, 14: 570–585. Vogt, K., Mellor, J., Tong, G. and Nicoll, R. (2000) The actions
Tonder, N., Kragh, J., Finsen, B.R., Bolwig, T.G. and Zimmer, of synaptically released zinc at hippocampal mossy fiber
J. (1994) Kindling induces transient changes in neuronal synapses. Neuron, 26: 187–196.
expression of somatostatin, neuropeptide Y, and calbindin in Vogt, K.E. and Regehr, W.G. (2001) Cholinergic modulation of
adult rat hippocampus and fascia dentata. Epilepsia, 35: excitatory synaptic transmission in the CA3 area of the hip-
1299–1308. pocampus. J. Neurosci., 21: 75–83.
Tong, G., Malenka, R.C. and Nicoll, R.A. (1996) Long-term Walker, M.C., Ruiz, A. and Kullmann, D.M. (2001) Monosy-
potentiation in cultures of single hippocampal granule cells: a naptic GABAergic signaling from dentate to CA3 with a
presynaptic form of plasticity. Neuron, 16: 1147–1157. pharmacological and physiological profile typical of mossy
Tongiorgi, E., Armellin, M., Giulianini, P.G., Bregola, G., fiber synapses. Neuron, 29: 703–715.
Zucchini, S., Paradiso, B., Steward, O., Cattaneo, A. and Wang, J., Yeckel, M.F., Johnston, D. and Zucker, R.S. (2004)
Simonato, M. (2004) Brain-derived neurotrophic factor Photolysis of postsynaptic caged Ca2+ can potentiate and
mRNA and protein are targeted to discrete dendritic lami- depress mossy fiber synaptic responses in rat hippocampal
nas by events that trigger epileptogenesis. J. Neurosci., 24: CA3 pyramidal neurons. J. Neurophysiol., 91: 1596–1607.
6842–6852. Weisskopf, M.G., Castillo, P.E., Zalutsky, R.A. and
Toth, K. and McBain, C.J. (1998) Afferent-specific innervation Nicoll, R.A. (1994) Mediation of hippocampal mossy fiber
of two distinct AMPA receptor subtypes on single hippo- long-term potentiation by cyclic AMP. Science, 265:
campal interneurons. Nat. Neurosci., 1: 572–578. 1878–1882.
Toth, K., Suares, G., Lawrence, J.J., Philips-Tansey, E. and Weisskopf, M.G. and Nicoll, R.A. (1995) Presynaptic changes
McBain, C.J. (2000) Differential mechanisms of transmission during mossy fibre LTP revealed by NMDA receptor-medi-
at three types of mossy fiber synapse. J. Neurosci., 20: ated synaptic responses. Nature, 376: 256–259.
8279–8289. Weisskopf, M.G., Zalutsky, R.A. and Nicoll, R.A. (1993) The
Treviño, M. and Gutierrez, R. (2005) The GABAergic projec- opioid peptide dynorphin mediates heterosynaptic depression
tion of the DG to hippocampal area CA3 of the rat: pre- of hippocampal mossy fibre synapses and modulates long-
and postsynaptic actions after seizures. J. Physiol., 567: term potentiation. Nature, 365: 188.
939–949. Wenzel, H.J., Cole, T.B., Born, D.E., Schwartzkroin, P.A. and
Treviño, M., Vivar, C. and Gutierrez, R. (2007) b/g Oscillatory Palmiter, R.D. (1997) Ultrastructural localization of zinc
activity in the CA3 hippocampal area is depressed by aber- transporter-3 (ZnT-3) to synaptic vesicle membranes within
rant GABAergic transmission from the DG after seizures. J. mossy fiber boutons in the hippocampus of mouse and mon-
Neurosci., 27: 251–259. key. Proc. Natl. Acad. Sci. U.S.A., 94: 12676–12681.
Tu, B., Timofeeva, O., Jiao, Y. and Nadler, J.V. (2005) Spon- Werner, P., Voigt, M., Keinanen, K., Wisden, W. and Seeburg,
taneous release of neuropeptide Y tonically inhibits recurrent P.H. (1991) Cloning of a putative high-affinity kainate re-
mossy fiber synaptic transmission in epileptic brain. J. Ne- ceptor expressed predominantly in hippocampal CA3 cells.
urosci., 25: 1718–1729. Nature, 351: 742–744.
Urban, N.N. and Barrionuevo, G. (1996) Induction of hebbian Westbrook, G.L. and Mayer, M.L. (1987) Micromolar con-
and non-hebbian mossy fiber long-term potentiation by dis- centrations of Zn2+ antagonize NMDA and GABA re-
tinct patterns of high-frequency stimulation. J. Neurosci., 16: sponses of hippocampal neurons. Nature, 328: 640–643.
4293–4299. Widdowson, P.S. (1993) Quantitative receptor autoradiography
Urban, N.N. and Barrionuevo, G. (1998) Active summation of demonstrates a differential distribution of neuropeptide-Y
excitatory postsynaptic potentials in hippocampal CA3 py- Y1 and Y2 receptor subtypes in human and rat brain. Brain
ramidal neurons. Proc. Natl. Acad. Sci. U.S.A., 95: Res., 631: 27–38.
11450–11455. Wiebe, S.P. and Staubli, U.V. (1999) Dynamic filtering of rec-
Vezzani, A. and Sperk, G. (2004) Overexpression of NPY and ognition memory codes in the hippocampus. J. Neurosci., 19:
Y2 receptors in epileptic brain tissue: an endogenous neuro- 10562–10574.
protective mechanism in temporal lobe epilepsy? Neuropep- Williams, S. and Johnston, D. (1988) Muscarinic depression of
tides, 38: 245–252. long-term potentiation in CA3 hippocampal neurons. Sci-
Vida, I. and Frotscher, M. (2000) A hippocampal interneuron ence, 242: 84–87.
associated with the mossy fiber system. Proc. Natl. Acad. Sci. Williams, S. and Johnston, D. (1989) Long-term potentiation of
U.S.A., 97: 1275–1280. hippocampal mossy fiber synapses is blocked by postsynaptic
Villacres, E.C., Wong, S.T., Chavkin, C. and Storm, D.R. injection of calcium chelators. Neuron, 3: 583–588.
(1998) Type I adenylyl cyclase mutant mice have impaired Williams, S. and Johnston, D. (1990) Muscarinic depression of
mossy fiber long-term potentiation. J. Neurosci., 18: synaptic transmission at the hippocampal mossy fiber
3186–3194. synapse. J. Neurophysiol., 64: 1089–1097.
132

Williams, S.H. and Johnston, D. (1991) Kinetic properties Yamamoto, C., Sawada, S. and Ohno-Shosaku, T. (1994) Sup-
of two anatomically distinct excitatory synapses in hippo- pression of hippocampal synaptic transmission by the spider
campal CA3 pyramidal neurons. J. Neurophysiol., 66: toxin omega-agatoxin-IV-A. Brain Res., 634: 349–352.
1010–1020. Yeckel, M.F., Kapur, A. and Johnston, D. (1999) Multiple
Wisden, W. and Seeburg, P.H. (1993) A complex mosaic of forms of LTP in hippocampal CA3 neurons use a common
high-affinity kainate receptors in rat brain. J. Neurosci., 13: postsynaptic mechanism. Nat. Neurosci., 2: 625–633.
3582–3598. Yokoi, M., Kobayashi, K., Manabe, T., Takahashi, T., Sa-
Wu, L.G., Borst, J.G. and Sakmann, B. (1998) R-type Ca2+ kaguchi, I., Katsuura, G., Shigemoto, R., Ohishi, H., No-
currents evoke transmitter release at a rat central synapse. mura, S., Nakamura, K., Nakao, K., Katsuki, M. and
Proc. Natl. Acad. Sci. U.S.A., 95: 4720–4725. Nakanishi, S. (1996) Impairment of hippocampal mossy fiber
Xiang, Z., Greenwood, A.C., Kairiss, E.W. and Brown, T.H. LTD in mice lacking mGluR2. Science, 273: 645–647.
(1994) Quantal mechanism of long-term potentiation in hip- Yoshino, M., Sawada, S., Yamamoto, C. and Kamiya, H.
pocampal mossy-fiber synapses. J. Neurophysiol., 71: (1996) A metabotropic glutamate receptor agonist DCG-IV
2552–2556. suppresses synaptic transmission at mossy fiber pathway of
Yamamoto, C. (1972) Activation of hippocampal neurons by the guinea pig hippocampus. Neurosci. Lett., 207: 70–72.
mossy fiber stimulation in thin brain sections in vitro. Exp. Zalutsky, R.A. and Nicoll, R.A. (1990) Comparison of two
Brain Res., 14: 423–435. forms of long-term potentiation in single hippocampal neu-
Yamamoto, C., Higashima, M., Sawada, S. and Kamiya, H. rons. Science, 248: 1619–1624.
(1991) Quantal components of the synaptic potential induced Zalutsky, R.A. and Nicoll, R.A. (1992) Mossy fiber long-term
in hippocampal neurons by activation of granule cells, and potentiation shows specificity but no apparent cooperativity.
the effect of 2-amino-4-phosphonobutyric acid. Hippocam- Neurosci. Lett., 138: 193–197.
pus, 1: 93–106. Zhang, G., Raol, Y.S., Hsu, F.C. and Brooks-Kayal, A.R.
Yamamoto, C., Matsumoto, K. and Takagi, M. (1980) Potent- (2004) Long-term alterations in glutamate receptor and
iation of excitatory postsynaptic potentials during and after transporter expression following early-life seizures are asso-
repetitive stimulation in thin hippocampal sections. Exp. ciated with increased seizure susceptibility. J. Neurochem.,
Brain Res., 38: 469–477. 88: 91–101.
Yamamoto, C., Sawada, S. and Kamiya, H. (1992) Enhance- Zieglgänsberger, W., French, E.D., Siggins, G.R. and Bloom,
ment of postsynaptic responsiveness during long-term F.E. (1979) Opioid peptides may excite hippocampal pyram-
potentiation of mossy fiber synapses in guinea pig hippo- idal neurons by inhibiting adjacent inhibitory interneurons.
campus. Neurosci. Lett., 138: 111–114. Science, 205: 415–417.
Plate 5.8. (A) Timm stained sections of hippocampus from adult rat, subjected to hippocampal x-irradiation as newborn, stop
odentate granule cell formation and lead to few MF terminals. (B) Timm stained section from the contralateral hemisphere of the rat
shown in A, depicting a well-integrated dentate transplant (tpl), derived from a small block of fascia dentata, grafted just after the x-
irradiation. From the graft, an apparently normal, laminar-specific MF projection (mf) developed, connecting dentate granule cells of
the graft (g) with the host CA3. The molecular layer of the graft (m) received a comparably laminar-specific projection of host
entorhinal perforant path projections, normalizing the Timm stained laminar appearance of the layer. (C) Slightly displaced dentate
transplant (tpl), grafted to x-irradiated newborn hippocampus just after irradiation as part of same experiment (Sunde et al., 1984).
Encroaching on the recipient CA3, MFs from the granule cells of this graft has entered adjacent parts of CA3 and project ‘‘down-
stream’’, but appear to stop at the border with CA1. Surprisingly, no MFs from the graft projected in the ‘‘upstream’’ direction, i.e.,
toward the host fascia dentata. Scale bar: 500 mm (For B/W version, see page 100 in the volume.)

Plate 6.1. Interneurons are the primary target of DG granule cells. Diagram representing the multiple types of excitatory contacts
made by granule cells. Within the dentate/hilar region collaterals of the mossy fibers (MFs) innervate both mossy cells and inhibitory
basket cells. In area CA3, the MFs contact pyramidal neurons via large expansions, but also excite interneurons through either
filopodial extensions from the large boutons or smaller en passant expansions along MF axons. (For B/W version, see page 110 in the
volume.)
Plate 6.2. Summary of the pre- and post-synaptic constituents of the different MF terminals and of their plasticity. (A) Schematic
representation of the giant MF boutons, which contact CA3 pyramidal cells. These terminals are characterized by low probability of
basal release and multiple release sites. (B) Schematic representation of filopodial extensions and en passant contacts, which synapse on
to interneurons. These have a high probability of basal release and single release sites. Both types of MF terminals contain several
neuromodulators and the neurotransmitters glutamate and GABA. Note that both the presynaptic and postsynaptic sites contain
several ionotropic GluRs and metabotropic GluRs, which confer to the MF a high degree of plasticity and the capacity for synaptic
integration. (C) Relative expression of the different releasable contents and of the receptors and transporters during development, in
the adult (D), and after epileptic activity (E). Arrows and font size indicate relative differences between the three states. (For B/W
version, see page 114 in the volume.)
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 7

Development of cell and fiber layers in


the dentate gyrus

Michael Frotscher, Shanting Zhao and Eckart Förster

Institute of Anatomy and Cell Biology, University of Freiburg, Albertstr. 17, D-79104 Freiburg, Germany

Abstract: This chapter deals with the laminated organization of the dentate gyrus, particularly with the
molecular signals controlling its development. First, sites of granule cell generation, their modes and routes
of migration are described. This is followed by an analysis of the molecular determinants governing the
formation of a tightly packed granule cell layer that is normal in rodents and primates. Reelin, a protein
of the extracellular matrix, plays an important role for the proper migration and lamination of the granule
cells during development and for the maintenance of a laminated dentate gyrus in adulthood. Granule cell
positioning is crucial for the laminated termination of commissural/associational fibers to the dentate
gyrus, suggesting that the granule cells carry positional signals for these fibers. In contrast, not signals
of the target cells but molecules of the extracellular matrix, such as hyaluronan, underlie the layer-specific
termination of fibers from the entorhinal cortex. The molecular determinants controlling axonal pathfind-
ing and target recognition of the profusely terminating cholinergic and GABAergic subcortical afferents
still need to be elucidated.

Keywords: granule cell; reelin; entorhinal fibers; commissural/associational fibers; cholinergic fibers

Introduction et al., 2006; Christie and Cameron, 2006; Tashiro


et al., 2006; Zhao et al., 2006). It has been a focus of
When compared to other brain regions, which these recent studies how postnatal granule cell for-
largely develop prenatally, the dentate gyrus is char- mation can be modified by external stimuli and how
acterized by an ongoing postnatal development. the newly generated granule cells become integrated
This enduring neurogenesis of dentate granule cells into the mature network of the dentate gyrus.
was first described in the early studies by Altman In this chapter, we will not directly deal with the
and coworkers (Altman, 1966; Altman and Das, very interesting phenomenon of postnatal neuro-
1966; Altman and Bayer, 1975; Bayer, 1980; Bayer genesis in the dentate gyrus and the factors that
and Altman, 1987) and has recently been further can modify this process, and the reader is referred
elaborated (e.g., Kempermann et al., 1997, 1998, to the articles cited above. Here, we will address
2003, 2004, review; Eriksson et al., 1998; Markakis fundamental issues of dentate gyrus development,
and Gage, 1999; van Praag et al., 2002, 2005; such as where granule cells are generated, how
Schmidt-Hieber et al., 2004; Lie et al., 2005; Aimone they migrate to reach their destination, how the
formation of a compact granule cell layer is con-
Corresponding author. Tel.: +(49) 761 203-5056; Fax: +(49) trolled, and how major afferents to the dentate
761 203-5054; E-mail: Michael.Frotscher@anat.uni-freiburg.de gyrus are instructed to form distinct, laminated

DOI: 10.1016/S0079-6123(07)63007-6 133


134

termination zones for the contact with defined cell movement. It is not clear which of the two
segments of the granule cell’s dendritic arbor. It is forms predominates in the dentate gyrus and
likely that many of the signals involved in these whether the temporal sequence assumed for neo-
processes apply to both early-generated granule cortical neurons (Nadarajah and Parnavelas, 2002)
cells and granule cells generated in the adult holds true for dentate granule cells. It is known,
organism. As far as the development of afferent however, that a radial glial scaffold forms in the
projections to the granule cells is concerned, dentate gyrus that persists long into the postnatal
we will deal with the projection from the ent- period, suggesting that radial glia-guided migra-
orhinal cortex to the dentate gyrus, commissural/ tion takes place in the dentate gyrus despite
associational (C/A) projections, and subcortical the relatively short distance that granule cells
projections, i.e., the cholinergic and GABAergic have to migrate from the secondary proliferation
projections from the medial septum to the dentate zone to their destination in the granule cell layer.
gyrus. Studies in mutants with defects in the reelin signa-
ling cascade provide further evidence for a role of
the radial glial scaffold in the proper migration of
Sites of granule cell generation dentate granule cells. In these mutants, the radial
glial scaffold is altered to a varying extent, and this
Like the pyramidal cells of the hippocampus is accompanied by migration defects of the granule
proper, the first granule cells are generated in the cells (Förster et al., 2002, 2006a, b; Frotscher et al.,
ventricular zone. From there they migrate to form 2003; Weiss et al., 2003).
the suprapyramidal blade of the dentate gyrus,
which develops before the infrapyramidal blade.
Later on in development, the secondary prolifer- Formation of a compact granule cell layer
ation zone emerges within the future hilus. Obvi-
ously, this region contains stem cells from the The lamination of neuronal cell bodies and fibers
ventricular zone that have retained their prolifer- is a characteristic feature of cortical organization.
ative capacity. Cells derived from this secondary In the dentate gyrus, where the granule cells form
proliferation zone form the late developing infra- a densely packed cell layer, this lamination is par-
pyramidal blade of the dentate gyrus and are ticularly striking. Moreover, the granule cells show
the source of all ongoing postnatal granule cell a clear, bipolar orientation with their dendrites
neurogenesis. extending into the molecular layer and their axons,
How do early- and late-generated granule cells the mossy fibers, invading the hilar region. This
reach their final destinations in the suprapyrami- characteristic organization of the dentate gyrus
dal and infrapyramidal blades of the dentate and its predominant neurons, the granule cells, has
gyrus, respectively? No definitive answer to this been recognized since the first Golgi studies of the
question is possible, since real-time microscopy hippocampus (Fig. 1; Golgi, 1886; Koelliker, 1896;
studies of newly generated, migrating granule cells Ramón y Cajal, 1911). What are the molecular
are not yet available. From studies in the neocor- signals that control this well-ordered arrangement
tex, we know that there are two major modes of the granule cells?
of neuronal migration, nuclear translocation and Much of our knowledge about the formation of
glia-guided neuronal migration (Nadarajah and granule cell lamination derives from the study
Parnavelas, 2002). Movement of the nucleus of various mouse mutants. In reeler mice lacking
within early-generated processes is generally as- the extracellular matrix protein reelin, the granule
sumed to be an early form of neuronal migration, cells do not form a compact cell layer but are
allowing for a migration over short distances in the scattered all over the dentate gyrus. Moreover,
yet small, immature embryonic brain. Later their dendrites have lost the uniform orientation
on, when large distances need to be bridged, glia- toward the hippocampal fissure and extend to var-
guided migration becomes the dominant form of ious directions (Stanfield and Cowan, 1979;
135

Fig. 1. Original drawing of the dentate gyrus and hippocampus proper by Camillo Golgi (Golgi, 1886). While Golgi was aware of the
laminated, bipolar arrangement of the granule cells, he did not correctly draw the course of the mossy fibers. As is known from
numerous more recent tracer studies, granule cell axons mainly run in stratum lucidum of CA3 and impinge on proximal dendritic
segments of pyramidal neurons. (See Color Plate 7.1 in color plate section.)

Drakew et al., 2002). Granule cell axons do is not ubiquitously present in the tissue but
not form the normal compact projection to stra- at specific sites. It came as a surprise that lipo-
tum lucidum of CA3 and appear defasciculated protein receptors, the very low-density lipopro-
(Drakew et al., 2002). However, like in wildtype tein receptor (VLDLR) and the apolipoprotein
mice, mossy fibers in the reeler mutant respect the E receptor 2 (ApoER2) are receptors for reelin.
border with CA1. As mentioned above, the radial Double-knockouts lacking these two lipopro-
glial scaffold is also altered in mutants with defects tein receptors show a reeler-like phenotype
in the reelin signaling pathway. (Trommsdorff et al., 1999). Binding of reelin to
How can one explain these multiple effects its lipoprotein receptors results in the phosphor-
of reelin on the radial glial scaffold, granule cell ylation of disabled 1 (dab1), an adapter protein
migration, and granule cell orientation? An expla- interacting with the intracellular domains of
nation is easier after a brief description of the ApoER2 and VLDLR, respectively. The reelin
reelin signaling cascade. Reelin is synthesized and signaling cascade eventually involves cytoskeletal
secreted by early-generated Cajal-Retzius cells proteins such as tau. For details and a survey of
populating the marginal zone of the cortex, which recent literature, the reader is referred to relevant
in the dentate gyrus corresponds to the outer review articles (Tissir and Goffinet, 2003; Förster
molecular layer (Del Rio et al., 1997). Thus, reelin et al., 2006a, b).
136

Recent studies have provided evidence for a dual tyrosine 220 and 232 leads to a detachment of the
role of reelin in the development of the dentate gyrus migrating neuron from the radial fiber, thus termi-
(Zhao et al., 2004; Förster et al., 2006a, b). It has nating the migration process. By coculturing slices
been suggested that reelin is a positional signal for of reeler hippocampus with wildtype slices, such that
the directed growth of radial glial fibers in the dent- a reelin-containing marginal zone was proximal to
ate gyrus (Förster et al., 2002; Zhao et al., 2004). In the reeler dentate gyrus, Zhao et al. (2004) were able
the absence of reelin, its lipoprotein receptors or to show that a compact granule cell layer could
dab1, no regular radial glial scaffold is formed be rescued in the reeler dentate gyrus, leaving the
(Weiss et al., 2003). Moreover, glial fibrillary acid molecular layer free of neurons. No rescue was
protein (GFAP)-positive glial cells of the hippo- achieved when a reeler slice was cocultured with
campus were found to express molecules of the stratum oriens of CA1, which did not contain reelin-
reelin signaling cascade and showed a preference for synthesizing Cajal-Retzius cells (Fig. 2). These au-
reelin in a stripe choice assay (Förster et al., 2002; thors also showed that a regularly oriented radial
Frotscher et al., 2003). It was concluded that the glial scaffold could be rescued under these in vitro
migration defects in mutants of the reelin pathway conditions. It remains to be determined whether
were due, at least in part, to a malformation of the reelin’s effect on radial glial cells or direct effect
radial glial scaffold (Frotscher et al., 2003). on neurons, or both, is crucial for the formation of
A second effect of reelin appears to be on neurons a compact granule cell layer.
directly. In the reeler mutant and in ApoER2 / /
VLDLR / mutants, the marginal zone is densely
Formation of afferent fiber lamination in the dentate
populated by neurons, whereas it is almost cell-free
gyrus
in wildtype animals. Thus, reelin was regarded as a
stop signal, preventing migrating neurons from in-
With the exception of some subcortical afferents,
vading the marginal zone. Sanada et al. (2004) have
major projections to the dentate gyrus show
recently shown that phosphorylation of dab1 at
a preference for distinct layers. Fibers originating
from neurons in layer 2 of the entorhinal cortex
(entorhinal fibers) are known to terminate in the
outer two-thirds of the molecular layer (Blackstad,
1958; Steward and Scoville, 1976; Amaral and
Witter, 1995). C/A fibers, derived from hilar mossy
cells projecting ipsilaterally and contralaterally,
terminate in the inner molecular layer (Blackstad
1956; Deller et al., 1995), and there is no overlap of
these two major projections to the dentate gyrus.
In the following paragraphs, we will summarize
what is known about the development of these
Fig. 2. Rescue of granule cell lamination in a slice culture of
projections and the factors determining their
reeler dentate gyrus is achieved by coculturing with a wildtype
culture providing a reelin-containing marginal zone. Two reeler laminar specificities.
cultures (rl / 1 and rl / 2) are cocultured with a rat hippo-
campal slice. A compact cell layer (arrow) has only formed in
rl / 1, which was cocultured next to the outer molecular layer Entorhinal afferents
of the rat dentate gyrus containing reelin-synthesizing Cajal-
Retzius cells. In rl / 2, which was cultured next to the stratum Fibers from the entorhinal cortex are among the
oriens of CA1, the reeler-specific loose distribution of neurons first afferents to reach the dentate gyrus in the
in the dentate gyrus is retained (arrowhead). Dashed lines rep-
rodent (at E18/E19; Super and Soriano, 1994;
resent borders between cultures. CA1, CA3, hippocampal re-
gions CA1 and CA3; DG, dentate gyrus; g, granule cell layer. Ceranik et al., 1999). This early arrival implies that
Scale bar: 200 mm (from Zhao et al., 2004, with permission). their regional specificity and layer-specific termi-
(See Color Plate 7.2 in color plate section.) nation in the outer two-thirds of molecular layer
137

cannot be controlled by their definitive target cells, arrive in their termination zone, the inner molec-
the granule cells, which are mainly generated after ular layer, after the fibers from the entorhinal cor-
birth (see above). How do the entorhinal axons tex, i.e., at P2. This implies that many granule cells
find their long way from the entorhinal cortex have already developed proximal dendrites for the
across the hippocampal fissure and the subiculum? contact with C/A fibers, and no transient target
Recent studies have indicated that early-generated neurons, as in the case of the entorhinal projec-
Cajal-Retzius cells located in the marginal zone tion, are required. In fact, the laminated projection
give rise to an early projection to the entorhinal of C/A fibers mirrors the laminated arrangement
cortex, thereby providing a template for the later of the granule cells in wildtype animals. Accord-
outgrowth of entorhinal fibers. In fact, intracellu- ingly, in mutants with defects in the reelin signa-
lar labeling studies as well as retrograde and ling pathway showing migration defects of the
anterograde tracing substantiated this early dent- granule cells and no compact granular layer, the
ate-entorhinal projection of Cajal-Retzius cells projection of C/A fibers is diffuse and not lami-
(Ceranik et al., 1999). Moreover, Cajal-Retzius nated (Deller et al., 1999; Gebhardt et al., 2002;
cells seem to serve as transient targets of early Zhao et al., 2003). A diffuse termination of C/A
arriving entorhinal fibers in the absence of the fibers is most pronounced in the reeler mutant and
granule cells. Electron microscopic studies showed in double-knockouts lacking ApoER2 and
that entorhinal axons established synapses with VLDLR, but is less prominent in single receptor
reelin-immunoreactive Cajal-Retzius cells in the mutants which only show minor migration defects
outer molecular layer, and selective removal of of the granule cells. Zhao et al. (2003) used slice
Cajal-Retzius cells prevented the ingrowth of ent- cocultures to study the development of the com-
orhinal afferents (Del Rio et al., 1997; Frotscher missural projection to the dentate gyrus. By
et al., 2001). Collectively, these findings provide coculturing slices of reeler hippocampus and wild-
evidence for a role of Cajal-Retzius cells as a tem- type hippocampus and tracing the ‘‘commissural’’
plate for outgrowing entorhinal axons. However, projection developing under these in vitro condi-
reelin, secreted by Cajal-Retzius cells and forming tions, they showed that the loose distribution of C/
a component of the extracellular matrix in the A fibers is not a cell-autonomous effect of the
outer molecular layer, does not seem to be reeler mutation on the C/A projection. C/A fibers
required for the layer-specific ingrowth of ent- originating from a reeler hippocampal culture
orhinal axons. In the reeler mutant lacking reelin, terminated with correct laminar specificity in the
entorhinal axons find their correct laminar posi- inner molecular layer in a cocultured wildtype
tion (Del Rio et al., 1997; Zhao et al., 2003). slice, whereas C/A fibers from a wildtype culture
Recent studies have provided evidence that other projecting to a reeler culture terminated profusely,
molecules of the extracellular matrix such as thus reflecting the scattered distribution of the
hyaluronan and molecules bound to it play granule cells in the reeler culture (Fig. 3). Finally,
an important role in the laminar specificity of ent- rescue of granule cell lamination in a reeler culture
orhinal fibers. Treating slice cultures of hippo- by coculturing to a wildtype slice (see above) also
campus with hyaluronidase, which degrades rescued the laminated termination of C/A fibers
hyaluronan, abolishes the layer-specific termina- (Zhao et al., 2004; Förster et al., 2006a, b). To-
tion of entorhinal fibers but not of C/A axons gether these data indicate that the granule cells
(Zhao et al., 2003). carry positional signals for the laminated termina-
tion of C/A fibers. This contrasts to the guidance
cues relevant for the proper termination of ent-
Commissural/associational fibers orhinal fibers. As described, they are guided
to their termination zone by transient pioneer
C/A fibers represent ipsilateral and contralateral neurons (Cajal-Retzius cells) and molecules of the
projections of hilar mossy cells (Ribak et al., 1985; extracellular matrix. Thus, as one would expect, in
Frotscher et al., 1991; Frotscher, 1992). C/A fibers the reeler mutant with a scattered distribution of
138

the granule cells, the C/A fibers terminate diffu- types of molecular signal control laminar specifi-
sely, but the entorhino-dentate fibers terminate city of commissural and entorhinal fibers to the
with laminar specificity in the outer molecular dentate gyrus, molecules associated with the cell
layer forming a sharp border (Zhao et al., 2003). membranes of the granule cells and extracellular
Collectively, these findings show that different matrix molecules, respectively. In light of these
results, the temporal sequence of fiber ingrowth
during development does not seem to determine
the layered termination of entorhinal and com-
missural afferents as was hypothesized by Bayer
and Altman (1987). This attractive temporal
hypothesis was challenged by Frotscher and
Heimrich (1993) when they used sequential cocul-
tures of hippocampus with its afferent regions.
Reversing the sequence of arrival of entorhinal
and commissural fibers did not change their layer-
specific termination in the dentate gyrus.

Cholinergic and GABAergic afferents from the


medial septum

While both the entorhinal fibers and C/A fibers are


characterized by their segregated, layer-specific
termination in the dentate gyrus, no such strictly
laminar termination is observed with subcortical

Fig. 3. Cocultures of two hippocampal slices to allow for the


formation of ‘‘commissural’’ projections, traced by anterograd-
ely transported biocytin injected into the hilar region of one of
the two hippocampal slices (sites of biocytin injections labeled
by asterisks). (A) Coculturing of two wildtype slices results in
the formation of a compact ‘‘commissural’’ projection in the
inner molecular layer (black arrow). Open arrow labels ipsilat-
eral mossy fiber projection in stratum lucidum of CA3. CA1,
CA3, hippocampal regions CA1 and CA3; g, granule cell layer.
(B) Coculture of wildtype (wt) hippocampus and reeler hippo-
campus. Biocytin-labeled ‘‘commissural’’ fibers from the wild-
type culture terminate profusely in the reeler culture, thus
reflecting the scattered distribution of their target neurons, the
granule cells. Open arrow labels the mossy fiber projection in
the wildtype culture. DG, dentate gyrus of the reeler culture
with scattered granule cells; P, pyramidal layer in the wildtype
culture; P1, P2, double pyramidal layer in the reeler culture.
(C) ‘‘Commissural’’ fibers from a reeler culture form a compact
projection in the inner molecular layer of the wildtype culture
(black arrow), indicating that reeler commissural fibers
project to the inner molecular layer as is normal, provided that
their target cells, the granule cells, form a tightly packed
layer. g, granule cell layer. Scale bar: 100 mm (applies to A–C)
(modified from Zhao et al., 2003, with permission; copyright by
the Society for Neuroscience). (See Color Plate 7.3 in color
plate section.)
139

afferents, such as the cholinergic and GABAergic projections. The hippocampo-septal projection de-
fibers from the medial septum and cat- velops early, clearly before the septohippocampal
echolaminergic brain stem fibers (the latter not projection (Linke et al., 1995) with hippocampal
dealt with in this chapter). The large cholinergic fibers reaching the septum around E16. Septal fib-
neurons of the medial septum/diagonal band (MS/ ers arrive in the hippocampus 2–3 days later.
DB) complex project via the fimbria-fornix to the Moreover, it has been shown that the septal fibers
hippocampus where they terminate profusely in all grow along hippocampo-septal axons (Linke and
hippocampal layers (Frotscher and Leranth, Frotscher, 1993). The determinants of the different
1985). A concentration of cholinergic (choline target cell specificities of cholinergic and GAB-
acetyltransferase-immunoreactive) fibers is regu- Aergic septohippocampal fibers remain to be
larly observed in the cell body layers, in the dent- elucidated.
ate gyrus it is mainly in the subgranular zone. The
fine, varicose cholinergic fibers establish synaptic
contacts with a variety of postsynaptic structures Functional considerations
including cell bodies, dendritic shafts, and spines
(Frotscher and Leranth, 1986). The functional significance of the laminated organ-
The GABAergic septohippocampal fibers, not ization of the dentate gyrus is not known. In tem-
showing a laminated termination pattern like the poral lobe epilepsy associated with Ammon’s horn
cholinergic fibers, specifically form synapses with sclerosis, the lamination of the dentate gyrus is dis-
hippocampal GABAergic interneurons, thus serv- rupted, and the granule cells do not form a tightly
ing a disinhibitory function (Freund and Antal, packed cell layer (granule cell dispersion, see
1988). The cholinergic afferents that terminate Houser, 1990). These structural changes may indi-
on both the hippocampal principal cells and cate that a laminated dentate gyrus with a clear
GABAergic interneurons do not have such a cell- segregation of the input site (granule cell dendrites
specific termination. in the molecular layer) from the output site (granule
Little is known about the molecular signals con- cell axons in the hilus) is required for the proper
trolling the development of the septohippocampal function of this brain region. This hypothesis im-
projection and the different target cell specificities plies that structural alterations, for instance migra-
of cholinergic and GABAergic septohippocampal tion defects during development, underlie the
fibers. Sema3C and its receptor neuropilin 2 have development of the epileptic disorder.
been suggested to play a role in the guidance of Recent studies of animal models of epilepsy
septal fibers to the hippocampus (Chedotal et al., have suggested a different scenario. Unilateral
1998; Steup et al., 2000; Skutella and Nitsch, injections of the glutamate agonist kainate into the
2001). Cholinergic fibers express the p75 neurotro- hippocampus of mice induced status epilepticus
phin receptor (p75NTR), and nerve growth factor and, with some delay, a characteristic granule cell
(NGF), one of the ligands for this receptor, is ex- dispersion (Bouilleret et al., 1999; Heinrich et al.,
pressed by GABAergic interneurons in the hippo- 2006), thus suggesting that the dispersion of gran-
campus. Since GABAergic hippocampal neurons ule cells is a result of seizure activity rather than
project to the septum, they may thus provide both being its source. How to explain, then, the loss of
a guiding scaffold and a neurotrophic source for granule cell lamination following seizure activity?
the ingrowing cholinergic fibers. The diffuse ter- Haas et al. (2002) showed that the number
mination of septohippocampal fibers would then of reelin-synthesizing neurons in tissue samples of
result from the scattered distribution of GAB- hippocampus removed from epileptic patients for
Aergic interneurons in the hippocampus and dent- therapeutical reasons was dramatically decreased
ate gyrus. That hippocampal neurons projecting when compared to control samples. A similar loss
to the septum are involved in the guidance of of reelin-synthesizing cells was observed following
septohippocampal fibers is supported by tracer unilateral kainate injection into the hippocampus
studies on the development of these two of adult mice (Heinrich et al., 2006). Granule cell
140

dispersion thus seems to be accompanied by a loss Altman, J. (1966) Autoradiographic and histological studies of
of reelin-expressing cells both in human temporal postnatal neurogenesis. II. A longitudinal investigation of the
lobe epilepsy and in an animal model of the dis- kinetics, migration and transformation of cells incorporating
tritiated thymidine in infant rats, with special reference
ease. During development, reelin is required for to postnatal neurogenesis in some brain regions. J. Comp.
the formation of a tightly packed granule cell Neurol., 128: 431–474.
layer. The findings from epileptic tissue indicate Altman, J. and Bayer, S.A. (1975) Postnatal development of
that reelin is also required for the maintenance of the hippocampal dentate gyrus under normal and experi-
a compact granule cell layer during adulthood. mental conditions. In: Isaacson R.L. and Pribram K.H.
(Eds.), The Hippocampus, Vol. 2. Plenum Press, New
Cajal-Retzius cells synthesizing reelin are very sen- York, pp. 95–122.
sitive to glutamate agonists (Del Rio et al., 1997). Altman, J. and Das, G.D. (1966) Autoradiographic and histo-
Hence, epileptic seizures associated with excessive logical studies of postnatal neurogenesis. I. A longitudinal
glutamate release may affect Cajal-Retzius cells, investigation of the kinetics, migration and transformation of
resulting in decreased reelin expression. In turn, cells incorporating tritiated thymidine in neonate rats, with
special reference to postnatal neurogenesis in some brain
loss of reelin, a stop signal for neurons, may allow
regions. J. Comp. Neurol., 126: 337–389.
the granule cells to leave their layer, resulting in Amaral, D.G. and Witter, M.P. (1995) The hippocampal
granule cell dispersion. A role for reelin in the formation. In: Paxinos G. (Ed.), The Rat Nervous System
maintenance of a compact granule cell layer in (2nd Ed). Academic Press, New York, pp. 443–494.
adult animals was confirmed by chronic unilateral Bayer, S.A. (1980) Development of the hippocampal region in
the rat. I. Neurogenesis examined with 3H-thymidine auto-
infusion of CR-50, a reelin-blocking antibody, into
radiography. J. Comp. Neurol., 190: 87–114.
the hippocampus of adult, naı̈ve animals. On the Bayer, S.A. and Altman, J. (1987) Directions in neurogenetic
injection side, but not on the control side, granule gradients and patterns of anatomical connections in the
cell dispersion developed. There was no granule telencephalon. Prog. Neurobiol., 29: 57–106.
cell dispersion when the reelin-blocking CR-50 Blackstad, T.W. (1956) Commissural connections of the hip-
pocampal region in the rat, with special reference to their
antibody was replaced by nonspecific IgG
mode of termination. J. Comp. Neurol., 105: 417–537.
(Heinrich et al., 2006). These data, as well as other Blackstad, T.W. (1958) On the termination of some afferents to
recent studies showing that sprouted mossy fibers the hippocampus and fascia dentata: an experimental study
in epileptic animals may exert an inhibitory rather in the rat. Acta Anat. (Basel), 35: 202–214.
than the commonly assumed excitatory effect on Bouilleret, V., Ridoux, V., Depaulis, A., Marescaux, C., Nehlig,
hippocampal principal cells (Sloviter et al., 2006), A. and Le Gal La Salle, G. (1999) Recurrent seizures and
hippocampal sclerosis following intrahippocampal kainate
indicate that structural changes in the hippocam- injection in adult mice: electroencephalography, histopathol-
pus in epilepsy are likely to be the consequence ogy and synaptic reorganization similar to mesial temporal
of the disease rather than its cause. However, the lobe epilepsy. Neuroscience, 89: 717–729.
functional significance of a strictly laminated dent- Ceranik, K., Deng, J., Heimrich, B., Lübke, J., Zhao, S.,
ate gyrus, first noticed by Golgi and his contem- Förster, E. and Frotscher, M. (1999) Hippocampal Cajal-
Retzius cells project to the entorhinal cortex: retrograde
poraries, still requires explanation. tracing and intracellular labelling studies. Eur. J. Neurosci.,
11: 4278–4290.
Chedotal, A., del Rio, J.A., Ruiz, M., He, Z., Borrell, V., de
Acknowledgments Castro, F., Ezan, F., Goodman, C.S., Tessier-Lavigne, M.,
Sotelo, C. and Soriano, E. (1998) Semaphorins III and IV
Supported by the German Research Foundation repel hippocampal axons via two distinct receptors. Devel-
(SFB 505, TR-3) and Jung Foundation for Science opment, 125: 4313–4323.
Christie, B.R. and Cameron, H.A. (2006) Neurogenesis in the
and Research. adult hippocampus. Hippocampus, 16: 199–207.
Del Rio, J.A., Heimrich, B., Borrell, V., Förster, E., Drakew,
A., Alcantara, S., Nakajima, K., Miyata, T., Ogawa, M.,
References Mikoshiba, K., Derer, P., Frotscher, M. and Soriano, E.
(1997) A role for Cajal-Retzius cells and reelin in the devel-
Aimone, J.B., Wiles, J. and Gage, F.H. (2006) Potential role for opment of hippocampal connections. Nature, 385: 70–74.
adult neurogenesis in the encoding of time in new memories. Deller, T., Drakew, A. and Frotscher, M. (1999) Different
Nat. Neurosci., 9: 723–727. primary target cells are important for fiber lamination
141

in the fascia dentata: a lesson from reeler mutant mice. Exp. study of their fine structure and synaptic connections in ro-
Neurol., 156: 239–253. dents and primates. J. Comp. Neurol., 312: 145–163.
Deller, T., Nitsch, R. and Frotscher, M. (1995) Phaseolus vul- Gebhardt, C., Del Turco, D., Drakew, A., Tielsch, A., Herz, J.,
garis-Leucoagglutinin tracing of commissural fibers to the rat Frotscher, M. and Deller, T. (2002) Abnormal positioning of
dentate gyrus: evidence for a previously unknown commis- granule cells alters afferent fiber distribution in the mouse
sural projection to the outer molecular layer. J. Comp. fascia dentata: morphologic evidence from reeler, apolipo-
Neurol., 352: 55–68. protein E receptor 2-, and very low density lipoprotein re-
Drakew, A., Deller, T., Heimrich, B., Gebhardt, C., Del Turco, ceptor knockout mice. J. Comp. Neurol., 445: 278–292.
D., Tielsch, A., Förster, E., Herz, J. and Frotscher, M. (2002) Golgi, C. (1886) Sulla fina anatomica degli organi centrali del
Dentate granule cells in reeler mutants and VLDLR and sistema nervoso. Hoepli, Milan.
ApoER2 knockout mice. Exp. Neurol., 176: 12–24. Haas, C.A., Dudeck, O., Kirsch, M., Huszka, C., Kann, G.,
Eriksson, P.S., Perfilieva, E., Bjork-Eriksson, T., Alborn, A.M., Pollak, S., Zentner, J. and Frotscher, M. (2002) Role for
Nordborg, C., Peterson, D.A. and Gage, F.H. (1998) Ne- reelin in the development of granule cell dispersion in tem-
urogenesis in the adult human hippocampus. Nat. Med., 4: poral lobe epilepsy. J. Neurosci., 22: 5797–5802.
1313–1317. Heinrich, C., Nitta, N., Flubacher, A., Müller, M., Fahrner, A.,
Förster, E., Jossin, Y., Zhao, S., Chai, X., Frotscher, M. and Kirsch, M., Freiman, T., Suzuki, F., Depaulis, A., Frotscher,
Goffinet, A.M. (2006a) Recent progress in understand- M. and Haas, C.A. (2006) Reelin deficiency and displacement
ing the role of Reelin in radial neuronal migration, with of mature neurons, but not neurogenesis, underlie the for-
specific emphasis on the dentate gyrus. Eur. J. Neurosci., mation of granule cell dispersion in the epileptic hippocam-
23: 901–909. pus. J. Neurosci., 26: 4701–4713.
Förster, E., Tielsch, A., Saum, B., Weiss, K.H., Johanssen, C., Houser, C.R. (1990) Granule cell dispersion in the dentate
Graus-Porta, D., Müller, U. and Frotscher, M. (2002) Reel- gyrus of humans with temporal lobe epilepsy. Brain Res.,
in, Disabled 1, and beta1 integrins are required for the for- 535: 195–204.
mation of the radial glial scaffold in the hippocampus. Proc. Kempermann, G., Gast, D., Kronenberg, G., Yamaguchi, M.
Natl. Acad. Sci. U.S.A., 99: 13178–13183. and Gage, F.H. (2003) Early determination and long-term
Förster, E., Zhao, S. and Frotscher, M. (2006b) Laminating the persistence of adult-generated new neurons in the hippocam-
hippocampus. Nat. Rev. Neurosci., 7: 259–267. pus of mice. Development, 130: 391–399.
Freund, T.F. and Antal, M. (1988) GABA-containing neurons Kempermann, G., Jessberger, S., Steiner, B. and Kronenberg,
in the septum control inhibitory interneurons in the hippo- G. (2004) Milestones of neuronal development in the adult
campus. Nature, 336: 170–173. hippocampus. Trends Neurosci., 27: 447–452.
Frotscher, M. (1992) Application of the Golgi/electron Kempermann, G., Kuhn, H.G. and Gage, F.H. (1997) More
microscopy technique for cell identification in immunocyto- hippocampal neurons in adult mice living in an enriched en-
chemical, retrograde labeling, and developmental studies vironment. Nature, 386: 493–495.
of hippocampal neurons. Microsc. Res. Technol., 23: Kempermann, G., Kuhn, H.G. and Gage, F.H. (1998) Expe-
306–323. rience-induced neurogenesis in the senescent dentate gyrus. J.
Frotscher, M., Haas, C.A. and Förster, E. (2003) Reelin con- Neurosci., 18: 3206–3212.
trols granule cell migration in the dentate gyrus by acting on Koelliker, A.v. (1896) Handbuch der Gewebelehre des Mensc-
the radial glial scaffold. Cereb. Cortex, 13: 634–640. hen. Nervensystem des Menschen und der Thiere (6th Ed).
Frotscher, M. and Heimrich, B. (1993) Formation of layer- Engelmann, Leipzig, Germany.
specific fiber projections to the hippocampus in vitro. Proc. Lie, D.C., Colamarino, S.A., Song, H.J., Desire, L., Mira, H.,
Natl. Acad. Sci. U.S.A., 90: 10400–10403. Consiglio, A., Lein, E.S., Jessberger, S., Lansford, H., Dear-
Frotscher, M. and Leranth, C. (1985) Cholinergic innervation ie, A.R. and Gage, F.H. (2005) Wnt signalling regulates adult
of the rat hippocampus as revealed by choline acetyltransf- hippocampal neurogenesis. Nature, 437: 1370–1375.
erase immunocytochemistry: a combined light and electron Linke, R. and Frotscher, M. (1993) Development of the rat
microscopic study. J. Comp. Neurol., 239: 237–246. septohippocampal projection: tracing with DiI and electron
Frotscher, M. and Leranth, C. (1986) The cholinergic innerva- microscopy of identified growth cones. J. Comp. Neurol.,
tion of the rat fascia dentata: identification of target struc- 332: 69–88.
tures on granule cells by combining choline acetyltransferase Linke, R., Pabst, T. and Frotscher, M. (1995) Development
immunocytochemistry and Golgi impregnation. J. Comp. of the hippocamposeptal projection in the rat. J. Comp.
Neurol., 243: 58–70. Neurol., 351: 602–616.
Frotscher, M., Seress, L., Abraham, H. and Heimrich, B. (2001) Markakis, E.A. and Gage, F.H. (1999) Adult-generated neu-
Early generated Cajal-Retzius cells have different functions rons in the dentate gyrus send axonal projections to field CA3
in cortical development. In: Miyan J.A., Thorndyke M., and are surrounded by synaptic vesicles. J. Comp. Neurol.,
Beesley P.W. and Bannister C.M. (Eds.), Brain Stem Cells. 406: 449–460.
BIOS Scientific Publishers Ltd., Oxford, pp. 43–49. Nadarajah, B. and Parnavelas, J.G. (2002) Modes of neuronal
Frotscher, M., Seress, L., Schwerdtfeger, W.K. and Buhl, E. migration in the developing cerebral cortex. Nat. Rev.
(1991) The mossy cells of the fascia dentata: a comparative Neurosci., 3: 423–432.
142

van Praag, H., Schinder, A.F., Christie, B.R., Toni, N., Palmer, Steward, O. and Scoville, S.A. (1976) Cells of origin of
T.D. and Gage, F.H. (2002) Functional neurogenesis in the entorhinal cortical afferents to the hippocampus and fascia
adult hippocampus. Nature, 415: 1030–1034. dentata of the rat. J. Comp. Neurol., 169: 347–370.
van Praag, H., Shubert, T., Zhao, C. and Gage, F.H. (2005) Super, H. and Soriano, E. (1994) The organization of the
Exercise enhances learning and hippocampal neurogenesis in embryonic and early postnatal murine hippocampus. II.
aged mice. J. Neurosci., 25: 8680–8685. Development of entorhinal, commissural and septal connec-
Ramón y Cajal, S. (1911) Histologie du Système Nerveux de tions studied with the lipophilic tracer DiI. J. Comp. Neurol.,
l’Homme et des Vertébrés, Vol. 2. Maloine, Paris. 344: 101–120.
Ribak, C.E., Seress, L. and Amaral, D.G. (1985) The develop- Tashiro, A., Sandler, V.M., Toni, N., Zhao, C. and Gage, F.H.
ment, ultrastructure and synaptic connections of the mossy (2006) NMDA-receptor-mediated, cell-specific integration of
cells of the dentate gyrus. J. Neurocytol., 14: 835–857. new neurons in adult dentate gyrus. Nature, 442: 929–933.
Sanada, K., Gupta, A. and Tsai, L.H. (2004) Disabled-1- Tissir, F. and Goffinet, A.M. (2003) Reelin and brain develop-
regulated adhesion of migrating neurons to radial glial fiber ment. Nat. Rev. Neurosci., 4: 496–505.
contributes to neuronal positioning during early corticogen- Trommsdorff, M., Gotthardt, M., Hiesberger, T., Shelton, J.,
esis. Neuron, 42: 197–211. Stockinger, W., Nimpf, J., Hammer, R.E., Richardson, J.E.
Schmidt-Hieber, C., Jonas, P. and Bischofberger, J. (2004) and Herz, J. (1999) Reeler/disabled-like disruption of neu-
Enhanced synaptic plasticity in newly generated granule cells ronal migration in knockout mice lacking the VLDL receptor
of the adult hippocampus. Nature, 429: 184–187. and ApoE receptor 2. Cell, 97: 689–701.
Skutella, T. and Nitsch, R. (2001) New molecules for hippo- Weiss, K.H., Johanssen, C., Tielsch, A., Herz, J., Deller, T.,
campal development. Trends Neurosci., 24: 107–113. Frotscher, M. and Förster, E. (2003) Malformation of the
Sloviter, R.S., Zappone, C.A., Harvey, B.D. and Frotscher, radial glial scaffold in the dentate gyrus of reeler mice,
M. (2006) Kainic acid-induced recurrent mossy fiber inner- scrambler mice, and ApoER2/VLDLR-deficient mice. J.
vation of dentate gyrus inhibitory interneurons: possi- Comp. Neurol., 460: 56–65.
ble anatomical substrate of granule cell hyperinhibition Zhao, C., Teng, E.M., Summers Jr., R.G.,, Ming, G.L. and
in chronically epileptic rats. J. Comp. Neurol., 494: Gage, F.H. (2006) Distinct morphological stages of dentate
944–960. granule neuron maturation in the adult mouse hippocampus.
Stanfield, B.B. and Cowan, W.M. (1979) The morphology of J. Neurosci., 26: 3–11.
the hippocampus and dentate gyrus in normal and reeler Zhao, S., Chai, X., Förster, E. and Frotscher, M. (2004) Reelin
mice. J. Comp. Neurol., 185: 393–422. is a positional signal for the lamination of dentate granule
Steup, A., Lohrum, M., Hamscho, N., Savaskan, N.E., Ninne- cells. Development, 131: 5117–5125.
mann, O., Nitsch, R., Fujisawa, H., Püschel, A.W. and Zhao, S., Förster, E., Chai, X. and Frotscher, M. (2003)
Skutella, T. (2000) Sema3C and netrin-1 differentially affect Different signals control laminar specificity of commissural
axon growth in the hippocampal formation. Mol. Cell and entorhinal fibers to the dentate gyrus. J. Neurosci.,
Neurosci., 15: 141–155. 23: 7351–7357.
Plate 7.1. Original drawing of the dentate gyrus and hippocampus proper by Camillo Golgi (Golgi, 1886). While Golgi was aware of
the laminated, bipolar arrangement of the granule cells, he did not correctly draw the course of the mossy fibers. As is known from
numerous more recent tracer studies, granule cell axons mainly run in stratum lucidum of CA3 and impinge on proximal dendritic
segments of pyramidal neurons. (For B/W version, see page 135 in the volume.)

Plate 7.2. Rescue of granule cell lamination in a slice culture of reeler dentate gyrus is achieved by coculturing with a wildtype culture
providing a reelin-containing marginal zone. Two reeler cultures (rl/1 and rl/2) are cocultured with a rat hippocampal slice. A
compact cell layer (arrow) has only formed in rl/1, which was cocultured next to the outer molecular layer of the rat dentate gyrus
containing reelin-synthesizing Cajal-Retzius cells. In rl/2, which was cultured next to the stratum oriens of CA1, the reeler-specific
loose distribution of neurons in the dentate gyrus is retained (arrowhead). Dashed lines represent borders between cultures. CA1, CA3,
hippocampal regions CA1 and CA3; DG, dentate gyrus; g, granule cell layer. Scale bar: 200 mm (from Zhao et al., 2004, with
permission). (For B/W version, see page 136 in the volume.)
Plate 7.3. Cocultures of two hippocampal slices to allow for the formation of ‘‘commissural’’ projections, traced by anterogradely
transported biocytin injected into the hilar region of one of the two hippocampal slices (sites of biocytin injections labeled by asterisks).
(A) Coculturing of two wildtype slices results in the formation of a compact ‘‘commissural’’ projection in the inner molecular layer
(black arrow). Open arrow labels ipsilateral mossy fiber projection in stratum lucidum of CA3. CA1, CA3, hippocampal regions CA1
and CA3; g, granule cell layer. (B) Coculture of wildtype (wt) hippocampus and reeler hippocampus. Biocytin-labeled ‘‘commissural’’
fibers from the wildtype culture terminate profusely in the reeler culture, thus reflecting the scattered distribution of their target
neurons, the granule cells. Open arrow labels the mossy fiber projection in the wildtype culture. DG, dentate gyrus of the reeler culture
with scattered granule cells; P, pyramidal layer in the wildtype culture; P1, P2, double pyramidal layer in the reeler culture.
(C) ‘‘Commissural’’ fibers from a reeler culture form a compact projection in the inner molecular layer of the wildtype culture (black
arrow), indicating that reeler commissural fibers project to the inner molecular layer as is normal, provided that their target cells, the
granule cells, form a tightly packed layer. g, granule cell layer. Scale bar: 100 mm (applies to A–C) (modified from Zhao et al., 2003,
with permission; copyright by the Society for Neuro-science). (For B/W version, see page 138 in the volume.)
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 8

Genetic regulation of dentate gyrus morphogenesis

Guangnan Li and Samuel J. Pleasure

Department of Neurology, Programs in Neuroscience, Developmental Biology, University of California at San Francisco,
1550 4th Street, Rock Hall Room 448c, San Francisco, CA 94158, USA

Abstract: The dentate gyrus is one of the small number of forebrain areas that have continued adult
neurogenesis. During development the dentate gyrus acquires the capacity for neurogenesis by generating a
new neurogenic stem cell niche at the border between the hilus and dentate granule cell layer. This is in
distinction to the other prominent zone of continued neurogenesis in the subventricular zone where neurons
are born in a structure directly descended from the mid-gestation subventricular zone. The ability to
generate this newly formed dentate neurogenic niche is controlled by the action of a number of genes during
prenatal and early postnatal development that regulate the fate, survival, migration, expansion, and
differentiation of the cellular components of the dentate neurogenic niche. In this review, we provide an
updated framework discussing the molecular steps and genes involved in these early stages of dentate gyrus
formation. We previously described a molecular framework for dentate gyrus morphogenesis that can be
associated with specific gene defects (Li, G., Pleasure, S.J. (2005). Dev. Neurosci., 27, 93–99), and here we
add additional recently described molecular players and discuss this framework.

Neurogenesis in the adult dentate gyrus requires a the most significant mouse mutants with defects in
specialized stem cell niche in the subgranular zone dentate development.
at the boundary between the hilus and dentate
granule cell layer. It is now clear that this niche
includes quiescent multipotent stem cells, transit- The neuroanatomy of the developing dentate gyrus
amplifying precursor cells, and immature neurons,
and that the transitions between these stages of Studies using classic neuroanatomic labeling meth-
neuronal maturation are regulated by signaling ods revealed specialized aspects of dentate gyrus
molecules from families of proteins also power- morphogenesis related to the development of the
fully involved in dentate development (Pozniak postnatal neurogenic niche (Nowakowski and
and Pleasure, 2006). We believe that the develop- Rakic, 1979, 1981; Eckenhoff and Rakic, 1984;
ment of this neurogenic stem cell niche is first es- Altman and Bayer, 1990a, b). The first granule
tablished in prenatal and early postnatal life. In neurons are born and the dentate precursor pool
this review we will first provide an overview of the expands in a specialized region of subventricular
neuroanatomy of dentate development and then zone that first becomes apparent adjacent to the
turn to consideration of the phenotypes of some of dentate notch at mid-gestation. Quickly after that,
the earliest dentate granule cells are born, migrate
to the nascent dentate gyrus, and provide some of
Corresponding author. Tel.: þ 1 (415)502-5683; the scaffolding for the developing dentate granule
Fax: þ 1 (415)502-4335; E-mail: sam.pleasure@ucsf.edu cell layers. Contemporaneously, dividing precursor

DOI: 10.1016/S0079-6123(07)63008-8 143


144

cells destined to settle in the hilus and subgranular released from localized signaling centers (Grove
zone migrate from the dentate notch in a subpial et al., 1998; Fukuchi-Shimogori and Grove, 2001;
route toward the dentate anlage (in developmental Garel et al., 2003; Shimogori et al., 2004; Storm
terms the anlage is the site of an incipient structure). et al., 2006). The cortical hem is the most caudo-
En route and in the dentate anlage these cells con- medial of these signaling centers and occupies a
tinue to divide and produce additional granule small zone of neuroepithelium adjacent to the re-
neurons locally. The proliferative matrix in the hilus gion of neuroepithelium where the dentate gyrus
beneath the condensing dentate granule cell layers arises (Fig. 1). The cortical hem releases Wnt and
was termed the tertiary matrix by (Altman and BMP family members and a variety of axon guid-
Bayer, 1990a, b) and is the site of a massive pro- ance molecules that are candidates for regulating
duction of granule neurons in the first week of the development of the dentate gyrus and hippo-
postnatal life. During the first postnatal week, the campus at early stages (Furuta et al., 1997; Grove
hilus develops a radial glial matrix with radially et al., 1998; Bagri et al., 2002). In addition, mice
oriented fibers traversing the dentate granule cell with early deletion of the cortical hem had dra-
layer with their cell bodies in the subgranular zone matic loss of cortical volume and no identifiable
and their endfeet attached to the pial surface. By hippocampus, implying that factors from the hem
P10, the tertiary matrix has reorganized so that are necessary for cortical expansion (Monuki
most precursor cells are dividing in the subgranular et al., 2001; Yoshida et al., 2006).
zone at the border between the condensing granule Interestingly, the cortical hem was also recently
cell layer and the now identifiable hilus. Granule shown to be the source of most Cajal-Retzius neu-
cells continue to be born in this specialized niche, rons that are generated in a wave of neurogenesis
although in decreasing numbers, throughout adult- very early in cortical development and then migrate
hood. Our current growing understanding of the tangentially to cover the cerebral cortex in the most
molecular steps of dentate development (Li and superficial cortical layer, called the marginal zone
Pleasure, 2005) is founded on this strong neuroan- (Takiguchi-Hayashi et al., 2004; Bielle et al., 2005;
atomic underpinning (see Fig. 1 for an overview of Yoshida et al., 2006; Zhao et al., 2006). Thus, many
this process and some of the molecules involved). Cajal-Retzius cells, critical regulators of cortical
migration via the release of reelin, are actually the
earliest born neurons from the medial pallium and
The cortical hem in initial development of the are produced from the neuroepithelium immedi-
dentate gyrus and hippocampus ately adjacent to the region of neuroepithelium that
will generate the dentate granule neurons. Recent
The initial formation of the hippocampus and studies have now shown that the tangential migra-
dentate gyrus shares features with the development tion of these neurons is controlled by SDF1 pro-
of the cortex as a whole since the hippocampal duced by the meninges. and this allows these cells to
formation occupies the caudomedial — most area maintain their superficial cortical position (Borrell
of the cortex and there is no evidence for a strict and Marin, 2006; Paredes et al., 2006). Interestingly
compartmental boundary between the hippocam- the same secreted chemokine (SDF-1) later controls
pus and neocortex. Recent studies have shown that the migration of cells to the dentate gyrus (Bagri
gradients of transcription factor expression are et al., 2002; Lu et al., 2002).
critically important in specifying the cortex and
driving development of four pallial compartments
(Wilson and Rubenstein, 2000; Sur and Ruben- Transcription factors pattern the cortex and
stein, 2005). The hippocampus is derived from the hippocampus
most medial pallial domain, adjacent to the region
forming the neocortex. Numerous recent studies Mice with mutations in transcription factors nor-
demonstrate that the gradients of these transcrip- mally expressed at their highest levels in the most
tion factors are controlled by diffusible cues caudomedial cortex have been found to have
A B C D

Lhx2/FoxG1/Emx2/Lef1
Hip

DG
Wnts

p73
Hem
Bmps

Lhx5
CP

granule cells Cajal-Retzius cells


Secreted Transcription precursors
hippocampus fissure
Factors Factors
meninges

dentate notch fimbria

E11.5 E13.5 E17.5 P4


Fig. 1. Schematic diagram showing the anatomic events and some of the genes involved in dentate morphogenesis. (A) At E11.5 the cortical hem (Hem) is adjacent to the
dentate neuroepithelium (DG) and the development of both are regulated by Wnts, while the hem and choroid plexus (CP) are also regulated by BMPs. The expression
domains for Wnts and BMPs are shown with arrows to the left of the neuroepithelium. The dentate and hippocampal neuroepithelia (Hip) are regulated by the function of
a number of transcription factors (Foxg1, Lhx2, Emx2, and Lef1) discussed in the text, while hem development is controlled by p73 and Lhx5. The expression domains for
transcription factors are shown to the right of the neuroepithelium and their extent indicated with arrows. (B) At E13.5 the initial migration of precursors is underway
from the dentate notch to the forming dentate gyrus, this migration may be targeted in part by SDF1 expressed in the meninges. By this time, Cajal-Retzius cells, derived
from the hem, are already present in the marginal zone and their position is maintained by SDF1 signaling. (C) By E17.5 precursor cells and granule cells have begun to
mix, forming the tertiary matrix in the nascent hilus and there is continued dentate precursor migration along the subpial migratory course. (D) By P4 the radial
reorganization of the dentate is largely accomplished with condensing of the granule cell layers and positioning of precursors in the subgranular zone.

145
146

major defects in hippocampal development. Emx2 either of these leads to cortical hem expansion
is a homeobox gene expressed in a high caudome- (Porter et al., 1997; Bulchand et al., 2001; Monuki et
dial to low anterolateral gradient, and mice with al., 2001; Hanashima et al., 2004; Muzio and Malla-
Emx2 mutations have reduced size of caudal cor- maci, 2005; Zhao et al., 2006). Amazingly, in the
tical structures like the hippocampus and visual case of Foxg1, mutant mice have a massive expan-
cortex with compensatory increases in anterior sion of all medial cortical structures so that essen-
cortical structures (Bishop et al., 2003; Muzio and tially all residual cortical structures have molecular
Mallamaci, 2003), while mice with overexpression markers consistent with a hippocampal origin and
of Emx2 have the reciprocal phenotype, with ex- there is no discernible neocortex at all (Muzio and
pansion of caudal cortical structures at the expense Mallamaci, 2005).
of anterior structures (Hamasaki et al., 2004).
Consistent with their loss of posterior structures,
Emx2 null mice also lack a clearly distinguishable The role of hem-derived signals in hippocampal and
dentate gyrus (the dentate gyrus forms from al- dentate development
most the most embryologically caudal portion of
the cortical neuroepithelium) and have reductions The studies discussed above indicate the central
in the production of Cajal-Retzius cells from the importance of the cortical hem in controlling hip-
cortical hem (the most caudal region of cortical pocampal and dentate development at the earliest
neuroepithelium) (Pellegrini et al., 1996; Yoshida stages, but what are the factors that the hem
et al., 1997; Mallamaci et al., 2000; Shinozaki makes? In addition to producing Cajal-Retzius
et al., 2002). Another interesting patterning mol- cells, the cortical hem is a source of multiple se-
ecule is the homeobox gene Lhx5, which in the creted Wnt proteins. The Wnts are a large family
cortex is expressed solely in the cortical hem and (19 ligands in mice) of secreted glycoprotein mol-
Cajal-Retzius cells derived from the hem. Mouse ecules that act potently during neural development
mutants for Lhx5 lack hem development entirely (Ciani and Salinas, 2005). The Wnt signaling path-
and have dramatic secondary hippocampal defects way is very complex and significant discussion of it
due to the loss of hem-derived patterning signals is beyond the scope of this review, but most of the
(Zhao et al., 1999), as well as defects in hippo- hippocampal phenotypes of Wnt signaling mu-
campal dependent learning (Paylor et al., 2001). In tants or Wnt ligands thus far have been most
addition, the transcription factor p73, which is se- clearly associated with defects in the so-called ‘‘ca-
lectively expressed in the cortical hem and Cajal- nonical’’ Wnt signaling pathway. Both of the most
Retzius cells derived from it also appear to have studied Wnt signaling pathways involve Wnt lig-
major hippocampal and cortical proliferation de- and binding to one of the cognate receptor types,
fects due at least in large part due to disruption of the Frizzled genes, but after this the two pathways
normal cortical hem development, although this diverge substantially. The canonical pathway re-
phenotype is less severe than the Lhx5 phenotype quires an additional receptor from the LRP family
with the most severe defects being dramatic (LRP6) to stabilize beta-catenin, which is then
hypoplasia of the lower blade of the dentate gyrus transported to the nucleus where it acts with the
(Yang et al., 2000; Meyer et al., 2004). family of Tcf/Lef transcription factors to drive al-
Loss of function mutations of transcription fac- terations in gene expression. The non-canonical
tors expressed in the hem lead to hem defects (and pathway has some transcriptional output but is
secondary hippocampal defects, e.g. the Lhx5 mu- also involved in local cytoskeletal rearrangements
tant mice), but in contrast, mutation of transcription and axon guidance decisions and also utilizes
factors normally excluded from the cortical hem newly recognized receptors from the Ryk family in
may result in phenotypes that seem to be almost the conjunction with Frizzled receptors.
opposite. Foxg1 is a forkhead transcription factor During early cortical development the earliest
and Lhx2 a homeobox gene, both of which are nor- Wnt ligand expressed in the cortical hem is Wnt3a
mally excluded from the hem, and mutation of (Grove et al., 1998) and Wnt3a mutants entirely
147

lack discernible hippocampal morphogenesis, most demonstrating that Twisted Gastrulation mutants
prominently due to a failure to expand neural pre- (combined with BMP4) have no dorsomedial telen-
cursors in the entire medial pallium (Lee et al., cephalic structures (Zakin and De Robertis, 2004).
2000). Mutants in one of the downstream tran- Since Twisted Gastrulation is an important extra-
scription factors, Lef1, which is expressed exclu- cellular mediator that enhances BMP function (thus
sively in the medial pallium in the cortex, almost loss of Twisted Gastrulation would mimic loss of
completely lack dentate granule neurons (Galceran function for BMP signaling), this establishes a crit-
et al., 2000; Zhou et al., 2004). Interestingly, mice ical early role for BMPs in the developing medial
with mutations in the canonical signaling pathway pallium and hippocampus, whether other later roles
co-receptor LRP6, which are hypomorphic for for the BMPs can be distinguished remains to be
multiple Wnt signaling pathways during develop- seen.
ment, have a very similar phenotype to that of the
Lef1 mutants (Zhou et al., 2004). The detailed
analysis of both the Lef1 and LRP6 mutant phe- Reforming the neurogenic niche in the dentate gyrus
notypes is most consistent with a relatively specific
deficit in expansion of early dentate precursors Once the dentate neuroepithelium is specified by the
leading to a failure to populate the dentate with actions of the signals discussed above, the next step
stem cells, and also a related defect in the prenatal is the process of moving a cohort of neurogenic
radial glial scaffolding (Zhou et al., 2004). stem cells with continued capacity for expansion to
In addition to the production of Wnts, the cor- the developing dentate gyrus. Previous studies have
tical hem also produces members of the extended shown that dentate precursors, along with early-
BMP family, although they are produced in higher born granule neurons, migrate in a subpial stream
quantities by the choroid plexus epithelium adjacent at the boundary between the fimbria and the me-
to the hem (Furuta et al., 1997; Grove et al., 1998; ninges during mid-gestation (Nowakowski and
Shimogori et al., 2004). Previous studies using BMP Rakic, 1979; Altman and Bayer, 1990b; Bagri
ligand mutants demonstrated that there are defects et al., 2002). Genes that control either migration
in the development of the medial cortical wall, but of these cells directly or the organization of the ra-
early studies did not fully evaluate cortical develop- dial glial network in this region disrupt morpho-
ment in these mutants (Furuta et al., 1997). These genesis of the dentate. One recent example of a
studies were quite difficult though, because of the relatively specific medial cortical radial glial defect
large number of BMPs expressed at the cortical comes from conditional FGFR1 mutants. These
midline. However, a recent analysis of conditional mice have diminished precursor proliferation in the
mutants, lacking all BMP receptor signaling during dentate and defects in the radial glial scaffolding
cortical development after E10, showed that BMP projecting to the dentate (Ohkubo et al., 2004).
signaling was particularly important in development Even more interesting are the results of a recent in
of the choroid plexus (Hebert et al., 2002). However, depth analysis of these mice showing a defect in
previous studies showing that BMP signaling is im- radial glial precursor somal translocation in the
portant earlier in regulating Foxg1 expression entire medial cortex (Smith et al., 2006). Since one
(Furuta et al., 1997; Monuki et al., 2001) do dem- of the main distinctions between dentate develop-
onstrate that earlier in cortical development as the ment and that of the rest of the cortex is that stem
cortex is initially patterned, there is likely to be an cell/precursors must relocate to the developing
important role for BMPs. This may not have been dentate from their original location in the neuro-
apparent in the analysis of the BMP receptor con- epithelium, this implies that these mice may actually
ditional mice because the Cre-driver line used was have a failure to properly seed the dentate gyrus
Foxg1 and this does not start expressing until the with appropriate numbers of dentate precursors.
telencephalon is already partially specified (Hebert This is somewhat similar to the canonical Wnt
et al., 2002). The important early role of BMPs in signaling dentate defects seen from loss of Lef1 and
cortical patterning was confirmed in a recent study LRP6, which have proliferative failure of dentate
148

precursors and defects in the radial glial network as has shown that this complex process depends on
well, but in these Wnt mutants it is not possible to intersecting networks of interacting transcription
exclude earlier patterning defects (Galceran et al., factors regulated by combinations of extracellular
2000; Zhou et al., 2004). We also recently described signaling cues. It has been difficult thus far to de-
a different Wnt signaling mechanism leading to loss termine the specific roles of these signaling path-
of some dentate precursors in the Frizzled9 mutant ways in granule cell differentiation because of their
mice, which have increased apoptotic cell death of pleiotrophic effects during earlier development
putative dentate precursors as they migrate prena- and at different stages of dentate morphogenesis.
tally, leading to an adult defect in granule cell Mice with mutations in NeuroD, a basic helix-
numbers (Zhao et al., 2005). This is a quite different loop-helix transcription factor, have quite selective
phenotype than in the Lef1 and LRP6 mutants, and defects in the differentiation of dentate granule
we are continuing to address whether this pheno- neurons without affecting other hippocampal neu-
type is due to a defect in non-canonical signaling or rons (Miyata et al., 1999; Liu et al., 2000). This
compensation by other Frizzleds that uncovers a selectivity is likely to be due to the expression of
later developmental role for canonical Wnt signa- Math2 and NeuroD2 (two closely related family
ling in other aspects of dentate development. members) in essentially all other cortical regions
The subpial migratory route used by the dentate (Pleasure et al., 2000). In NeuroD mutants there are
precursors and immature granule cells to reach the still small numbers of neurons produced that even-
forming dentate gyrus is also a region with very tually express Prox1 at low levels, and these cells
strong expression of the chemokine ligand SDF1, still project axons toward CA3 (the appropriate
which may regulate migration of these cells (Bagri target) and express Calbindin (another marker of
et al., 2002; Lu et al., 2002). Indeed, we and others mature granule neurons) (Liu et al., 2000). This in-
demonstrated that SDF1 serves as a direct chemo- dicates that there are likely to be other regulatory
attractant for cells migrating in the subpial migra- genes involved in specifying granule neuron fate
tory route and that the dentate gyrus is and differentiation in addition to NeuroD.
depopulated of precursors without SDF-1, either
because of a migratory deficit or because of a pro-
liferative failure of the precursors (Bagri et al., Regulation of dentate granule cell layer density
2002; Lu et al., 2002). Interestingly, mice with de-
fects in integrin receptors or signaling systems The normal granule cell layer of mammals is a dis-
downstream of integrins in the migrating cells (e.g. crete, compact layer with few granule cells located
focal adhesion kinase — Fak) also have defects in outside its narrow boundaries. Reelin mutants
the organization of the subpial migratory pathway (and mice with mutations in Reelin receptors —
frequently resulting in major defects of the lower VLDLR and ApoER3 — or downstream signaling
blade of the dentate (Forster et al., 2002; Beggs molecules — Disabled) have a dispersed granule cell
et al., 2003; Niewmierzycka et al., 2005). layer, and a major defect in proliferative capacity of
dentate progenitors, but do not appear to have a
primary migratory defect because the earliest gran-
Mutants with defects in dentate granule cell ule cells find the dentate gyrus just as they would
differentiation under normal conditions (Forster et al., 2002;
Frotscher et al., 2003; Zhao et al., 2004). The ab-
During dentate gyrus development, from the ear- normalities in the Reelin mutant may be due to di-
liest embryonic stages to adulthood, the final step rect effects of Reelin on granule neurons and their
is cell type-specific differentiation into neurons of organization, but it is also likely that the mecha-
the appropriate phenotype with the appropriate nisms are related to a failure to properly organize
specialized markers [in the case of granule cells the the radially organized radial glial network in the
most specific marker is the divergent homeobox early postnatal dentate gyrus. This is because there
gene Prox1 — (Pleasure et al., 2000)]. Recent work are also direct effects of Reelin on the radial glial
Fig. 2. NeuroD mutants have radial glial scaffolding defects in the postnatal dentate. We examined the organization of the GFAP+ radially oriented dentate precursors
usually residing in the subgranular zone in NeuroD mutant mice. In control mice, Prox1 clearly marks the dentate granule cell layer with the adjacent subgranular zone
that serves as the focus of radially oriented GFAP+ fibers. In contrast, in NeuroD mutant mice there is only a small cap of Prox1+ granule cells in the dentate of P21
mutants and there are essentially no GFAP+ processes present in the mutant dentate. (See Color Plate 8.2 in color plate section.)

149
150

cells themselves (Frotscher et al., 2003). Interest- in adult neurogenesis (Pozniak and Pleasure, 2006).
ingly, recent studies of intractable epilepsy patients Sonic Hedgehog (Shh) is a crucial ligand in the es-
(Haas et al., 2002) and animal models of epilepsy tablishment of the dentate neurogenic niche during
(Heinrich et al., 2006) showed that dentate granule development (Machold et al., 2003) and clearly acts
cell layer dispersion is associated with loss of Reelin- directly on migratory precursors in the subpial mi-
expressing cells in the dentate marginal zone, im- gration pathway (Ahn and Joyner, 2005). It has now
plying a crucial role for Reelin in maintaining dent- become apparent that Shh regulates the prolifera-
ate architecture in adulthood. However, since there tion of rapidly dividing adult dentate precursors and
was no correlation with severity or duration of ep- quiescent stem cells as well (Lai et al., 2003; Ahn
ilepsy comparing this group to the other patients and Joyner, 2005). In addition, Wnts have now been
with pharmacologically tractable or other forms of shown to regulate the proliferation of committed
temporal lobe epilepsy one can not rule out the role dentate precursors and Wnt signaling is required for
of preexisting developmental anomalies. Thus, in adult neurogenesis (Lie et al., 2005). It is likely that
some subsets of epilepsy patients, granule cell dis- over the next few years roles will become established
persion in temporal lobe epilepsy could be due to a for most, if not all, developmental modulators of
neurodevelopmental defect that led to a loss of dentate development in adult neurogenesis as well.
Reelin, or due to abnormalities in Cajal-Retzius cell
genesis or distribution (Haas et al., 2002).
We recently also noted that the residual granule Acknowledgments
cells in NeuroD mutant mice are diffusely distrib-
uted as compared to the residual granule cells found The authors would like to acknowledge the other
in Lef1 mutants. This implies that the few granule members of the Pleasure Lab for helpful discus-
cells made when NeuroD is missing either have a sions. This work was funded by K02 MH074958,
cell autonomous defect in compaction or that their R01 MH066084 and P50 NS35902 to S.J.P.
might be a non-autonomous defect similar to the
radial glial phenotype in Reelin mice. In other
words, newborn granule cells lacking NeuroD may References
fail to properly organize the forming dentate radial
glial scaffolding and secondarily cause a granule cell Ahn, S. and Joyner, A.L. (2005) In vivo analysis of quiescent
dispersion phenotype. This would presumably be adult neural stem cells responding to Sonic hedgehog. Na-
ture, 437: 894–897.
non-autonomous because radial glial cells do not Altman, J. and Bayer, S.A. (1990a) Migration and distribution
express NeuroD (whereas they do express Reelin of two populations of hippocampal granule cell precursors
receptors) and might point to a role for newborn during the perinatal and postnatal periods. J. Comp. Neurol.,
granule cells in organizing their own lamination. If 301: 365–381.
Altman, J. and Bayer, S.A. (1990b) Mosaic organization of the
this is true, we would predict that NeuroD mice
hippocampal neuroepithelium and the multiple germinal
would display radial glial defects even though these sources of dentate granule cells. J. Comp. Neurol., 301:
cells never express NeuroD. While it is difficult to 325–342.
resolve this issue without more specifically ex- Bagri, A., Gurney, T., He, X., Zou, Y.R., Littman, D.R.,
pressed markers and conditional alleles, we have Tessier-Lavigne, M. and Pleasure, S.J. (2002) The chemokine
found that NeuroD mice do have radial glial SDF1 regulates migration of dentate granule cells. Develop-
ment, 129: 4249–4260.
scaffolding defects (Fig. 2). Beggs, H.E., Schahin-Reed, D., Zang, K., Goebbels, S., Nave,
K.A., Gorski, J., Jones, K.R., Sretavan, D. and Reichardt,
L.F. (2003) FAK deficiency in cells contributing to the basal
Continued use of developmental signaling systems in lamina results in cortical abnormalities resembling congenital
muscular dystrophies. Neuron, 40: 501–514.
the adult dentate gyrus
Bielle, F., Griveau, A., Narboux-Neme, N., Vigneau, S., Sigrist,
M., Arber, S., Wassef, M. and Pierani, A. (2005) Multiple
Recent studies have begun to establish that cues origins of Cajal-Retzius cells at the borders of the developing
regulating dentate development continue to function pallium. Nat. Neurosci., 8: 1002–1012.
151

Bishop, K.M., Garel, S., Nakagawa, Y., Rubenstein, J.L. and Heinrich, C., Nitta, N., Flubacher, A., Muller, M., Fahrner, A.,
O’leary, D.D. (2003) Emx1 and Emx2 cooperate to regulate Kirsch, M., Freiman, T., Suzuki, F., Depaulis, A., Frotscher,
cortical size, lamination, neuronal differentiation, develop- M. and Haas, C.A. (2006) Reelin deficiency and displacement
ment of cortical efferents, and thalamocortical pathfinding. of mature neurons, but not neurogenesis, underlie the for-
J. Comp. Neurol., 457: 345–360. mation of granule cell dispersion in the epileptic hippocam-
Borrell, V. and Marin, O. (2006) Meninges control tangential pus. J. Neurosci., 26: 4701–4713.
migration of hem-derived Cajal-Retzius cells via CXCL12/ Lai, K., Kaspar, B.K., Gage, F.H. and Schaffer, D.V. (2003)
CXCR4 signaling. Nat. Neurosci., 9: 1284–1293. Sonic hedgehog regulates adult neural progenitor prolifera-
Bulchand, S., Grove, E.A., Porter, F.D. and Tole, S. (2001) tion in vitro and in vivo. Nat. Neurosci., 6: 21–27.
LIM-homeodomain gene Lhx2 regulates the formation of the Lee, S.M., Tole, S., Grove, E. and Mcmahon, A.P. (2000) A
cortical hem. Mech. Dev., 100: 165–175. local Wnt-3a signal is required for development of the mam-
Ciani, L. and Salinas, P.C. (2005) WNTs in the vertebrate malian hippocampus. Development, 127: 457–467.
nervous system: from patterning to neuronal connectivity. Li, G. and Pleasure, S.J. (2005) Morphogenesis of the dentate
Nat. Rev. Neurosci., 6: 351–362. gyrus: what we are learning from mouse mutants. Dev. Ne-
Eckenhoff, M.F. and Rakic, P. (1984) Radial organization of urosci., 27: 93–99.
the hippocampal dentate gyrus: a Golgi, ultrastructural, and Lie, D.C., Colamarino, S.A., Song, H.J., Desire, L., Mira, H.,
immunocytochemical analysis in the developing rhesus mon- Consiglio, A., Lein, E.S., Jessberger, S., Lansford, H., Dear-
key. J. Comp. Neurol., 223: 1–21. ie, A.R. and Gage, F.H. (2005) Wnt signalling regulates adult
Forster, E., Tielsch, A., Saum, B., Weiss, K.H., Johanssen, C., hippocampal neurogenesis. Nature, 437: 1370–1375.
Graus-Porta, D., Muller, U. and Frotscher, M. (2002) Reelin, Liu, M., Pleasure, S.J., Collins, A.E., Noebels, J.L., Naya, F.J.,
Disabled 1, and beta 1 integrins are required for the formation Tsai, M.J. and Lowenstein, D.H. (2000) Loss of BETA2/
of the radial glial scaffold in the hippocampus. Proc. Natl. NeuroD leads to malformation of the dentate gyrus and
Acad. Sci. U.S.A., 99: 13178–13183. epilepsy. Proc. Natl. Acad. Sci. U.S.A., 97: 865–870.
Frotscher, M., Haas, C.A. and Forster, E. (2003) Reelin con- Lu, M., Grove, E.A. and Miller, R.J. (2002) Abnormal devel-
trols granule cell migration in the dentate gyrus by acting on opment of the hippocampal dentate gyrus in mice lacking the
the radial glial scaffold. Cereb. Cortex, 13: 634–640. CXCR4 chemokine receptor. Proc. Natl. Acad. Sci. U.S.A.,
Fukuchi-Shimogori, T. and Grove, E.A. (2001) Neocortex pat- 99: 7090–7095.
terning by the secreted signaling molecule FGF8. Science, Machold, R., Hayashi, S., Rutlin, M., Muzumdar, M.D., Nery,
294: 1071–1074. S., Corbin, J.G., Gritli-Linde, A., Dellovade, T., Porter, J.A.,
Furuta, Y., Piston, D.W. and Hogan, B.L. (1997) Bone Rubin, L.L., Dudek, H., Mcmahon, A.P. and Fishell, G.
morphogenetic proteins (BMPs) as regulators of dorsal fore- (2003) Sonic hedgehog is required for progenitor cell mainte-
brain development. Development, 124: 2203–2212. nance in telencephalic stem cell niches. Neuron, 39: 937–950.
Galceran, J., Miyashita-Lin, E.M., Devaney, E., Rubenstein, Mallamaci, A., Mercurio, S., Muzio, L., Cecchi, C., Pardini,
J.L. and Grosschedl, R. (2000) Hippocampus development C.L., Gruss, P. and Boncinelli, E. (2000) The lack of Emx2
and generation of dentate gyrus granule cells is regulated by causes impairment of Reelin signaling and defects of neu-
LEF1. Development, 127: 469–482. ronal migration in the developing cerebral cortex. J. Neuro-
Garel, S., Huffman, K.J. and Rubenstein, J.L. (2003) Molecular sci., 20: 1109–1118.
regionalization of the neocortex is disrupted in Fgf8 hypo- Meyer, G., Cabrera Socorro, A., Perez Garcia, C.G., Martinez
morphic mutants. Development, 130: 1903–1914. Millan, L., Walker, N. and Caput, D. (2004) Developmental
Grove, E.A., Tole, S., Limon, J., Yip, L. and Ragsdale, C.W. roles of p73 in Cajal-Retzius cells and cortical patterning.
(1998) The hem of the embryonic cerebral cortex is defined by J. Neurosci., 24: 9878–9887.
the expression of multiple Wnt genes and is compromised in Miyata, T., Maeda, T. and Lee, J.E. (1999) NeuroD is required
Gli3-deficient mice. Development, 125: 2315–2325. for differentiation of the granule cells in the cerebellum and
Haas, C.A., Dudeck, O., Kirsch, M., Huszka, C., Kann, G., hippocampus. Genes Dev., 13: 1647–1652.
Pollak, S., Zentner, J. and Frotscher, M. (2002) Role for Monuki, E.S., Porter, F.D. and Walsh, C.A. (2001) Patterning
reelin in the development of granule cell dispersion in tem- of the dorsal telencephalon and cerebral cortex by a roof
poral lobe epilepsy. J. Neurosci., 22: 5797–5802. plate-Lhx2 pathway. Neuron, 32: 591–604.
Hamasaki, T., Leingartner, A., Ringstedt, T. and O’leary, D.D. Muzio, L. and Mallamaci, A. (2003) Emx1, emx2 and pax6 in
(2004) EMX2 regulates sizes and positioning of the primary specification, regionalization and arealization of the cerebral
sensory and motor areas in neocortex by direct specification cortex. Cereb. Cortex, 13: 641–647.
of cortical progenitors. Neuron, 43: 359–372. Muzio, L. and Mallamaci, A. (2005) Foxg1 confines Cajal-Re-
Hanashima, C., Li, S.C., Shen, L., Lai, E. and Fishell, G. (2004) tzius neuronogenesis and hippocampal morphogenesis to the
Foxg1 suppresses early cortical cell fate. Science, 303: dorsomedial pallium. J. Neurosci., 25: 4435–4441.
56–59. Niewmierzycka, A., Mills, J., St-Arnaud, R., Dedhar, S. and
Hebert, J.M., Mishina, Y. and Mcconnell, S.K. (2002) BMP Reichardt, L.F. (2005) Integrin-linked kinase deletion from
signaling is required locally to pattern the dorsal telence- mouse cortex results in cortical lamination defects resembling
phalic midline. Neuron, 35: 1029–1041. cobblestone lissencephaly. J. Neurosci., 25: 7022–7031.
152

Nowakowski, R.S. and Rakic, P. (1979) The mode of migration Storm, E.E., Garel, S., Borello, U., Hebert, J.M., Martinez, S.,
of neurons to the hippocampus: a Golgi and electron micro- Mcconnell, S.K., Martin, G.R. and Rubenstein, J.L. (2006)
scopic analysis in foetal rhesus monkey. J. Neurocytol., 8: Dose-dependent functions of Fgf8 in regulating telencephalic
697–718. patterning centers. Development, 133: 1831–1844.
Nowakowski, R.S. and Rakic, P. (1981) The site of origin and Sur, M. and Rubenstein, J.L. (2005) Patterning and plasticity of
route and rate of migration of neurons to the hippocampal the cerebral cortex. Science, 310: 805–810.
region of the rhesus monkey. J. Comp. Neurol., 196: Takiguchi-Hayashi, K., Sekiguchi, M., Ashigaki, S., Taka-
129–154. matsu, M., Hasegawa, H., Suzuki-Migishima, R., Yoko-
Ohkubo, Y., Uchida, A.O., Shin, D., Partanen, J. and Vac- yama, M., Nakanishi, S. and Tanabe, Y. (2004) Generation
carino, F.M. (2004) Fibroblast growth factor receptor 1 is of reelin-positive marginal zone cells from the caudomedial
required for the proliferation of hippocampal progenitor cells wall of telencephalic vesicles. J. Neurosci., 24: 2286–2295.
and for hippocampal growth in mouse. J. Neurosci., 24: Wilson, S.W. and Rubenstein, J.L. (2000) Induction and
6057–6069. dorsoventral patterning of the telencephalon. Neuron, 28:
Paredes, M.F., Li, G., Berger, O., Baraban, S.C. and Pleasure, 641–651.
S.J. (2006) Stromal-derived factor-1 (CXCL12) regulates la- Yang, A., Walker, N., Bronson, R., Kaghad, M., Oosterwegel,
minar position of Cajal-Retzius cells in normal and dysplastic M., Bonnin, J., Vagner, C., Bonnet, H., Dikkes, P., Sharpe,
brains. J. Neurosci., 26: 9404–9412. A., Mckeon, F. and Caput, D. (2000) p73-deficient mice have
Paylor, R., Zhao, Y., Libbey, M., Westphal, H. and Crawley, neurological, pheromonal and inflammatory defects but lack
J.N. (2001) Learning impairments and motor dysfunctions in spontaneous tumours. Nature, 404: 99–103.
adult Lhx5-deficient mice displaying hippocampal disorgan- Yoshida, M., Assimacopoulos, S., Jones, K.R. and Grove, E.A.
ization. Physiol. Behav., 73: 781–792. (2006) Massive loss of Cajal-Retzius cells does not disrupt
Pellegrini, M., Mansouri, A., Simeone, A., Boncinelli, E. and neocortical layer order. Development, 133: 537–545.
Gruss, P. (1996) Dentate gyrus formation requires Emx2. Yoshida, M., Suda, Y., Matsuo, I., Miyamoto, N., Takeda, N.,
Development, 122: 3893–3898. Kuratani, S. and Aizawa, S. (1997) Emx1 and Emx2 func-
Pleasure, S.J., Collins, A.E. and Lowenstein, D.H. (2000) tions in development of dorsal telencephalon. Development,
Unique expression patterns of cell fate molecules delineate 124: 101–111.
sequential stages of dentate gyrus development. J. Neurosci., Zakin, L. and De Robertis, E.M. (2004) Inactivation of mouse
20: 6095–6105. Twisted gastrulation reveals its role in promoting Bmp4 ac-
Porter, F.D., Drago, J., Xu, Y., Cheema, S.S., Wassif, C., tivity during forebrain development. Development, 131:
Huang, S.P., Lee, E., Grinberg, A., Massalas, J.S., Bodine, 413–424.
D., Alt, F. and Westphal, H. (1997) Lhx2, a LIM homeobox Zhao, C., Aviles, C., Abel, R.A., Almli, C.R., Mcquillen, P. and
gene, is required for eye, forebrain, and definitive erythrocyte Pleasure, S.J. (2005) Hippocampal and visuospatial learning
development. Development, 124: 2935–2944. defects in mice with a deletion of frizzled 9, a gene in the
Pozniak, C.D. and Pleasure, S.J. (2006) A tale of two signals: Williams syndrome deletion interval. Development, 132:
Wnt and Hedgehog in dentate neurogenesis. Sci. STKE, 319: 2917–2927.
pe5. Zhao, C., Guan, W. and Pleasure, S.J. (2006) A transgenic
Shimogori, T., Banuchi, V., Ng, H.Y., Strauss, J.B. and Grove, marker mouse line labels Cajal-Retzius cells from the cortical
E.A. (2004) Embryonic signaling centers expressing BMP, hem and thalamocortical axons. Brain Res., 1077: 48–53.
WNT and FGF proteins interact to pattern the cerebral cor- Zhao, S., Chai, X., Forster, E. and Frotscher, M. (2004) Reelin
tex. Development, 131: 5639–5647. is a positional signal for the lamination of dentate granule
Shinozaki, K., Miyagi, T., Yoshida, M., Miyata, T., Ogawa, cells. Development, 131: 5117–5125.
M., Aizawa, S. and Suda, Y. (2002) Absence of Cajal-Retzius Zhao, Y., Sheng, H.Z., Amini, R., Grinberg, A., Lee, E.,
cells and subplate neurons associated with defects of tangen- Huang, S., Taira, M. and Westphal, H. (1999) Control of
tial cell migration from ganglionic eminence in Emx1/2 dou- hippocampal morphogenesis and neuronal differentiation by
ble mutant cerebral cortex. Development, 129: 3479–3492. the LIM homeobox gene Lhx5. Science, 284: 1155–1158.
Smith, K.M., Ohkubo, Y., Maragnoli, M.E., Rasin, M.R., Zhou, C.J., Zhao, C. and Pleasure, S.J. (2004) Wnt signaling
Schwartz, M.L., Sestan, N. and Vaccarino, F.M. (2006) mutants have decreased dentate granule cell production and
Midline radial glia translocation and corpus callosum for- radial glial scaffolding abnormalities. J. Neurosci., 24:
mation require FGF signaling. Nat. Neurosci., 9: 787–797. 121–126.
Plate 8.2. NeuroD mutants have radial glial scaffolding defects in the postnatal dentate. We examined the organization of the
GFAP+ radially oriented dentate precursors usually residing in the subgranular zone in NeuroD mutant mice. In control mice, Prox1
clearly marks the dentate granule cell layer with the adjacent subgranular zone that serves as the focus of radially oriented GFAP+
fibers. In contrast, in NeuroD mutant mice there is only a small cap of Prox1+ granule cells in the dentate of P21 mutants and there
are essentially no GFAP+ processes present in the mutant dentate. (For B/W version, see page 149 in the volume.)
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 9

Ultrastructure and synaptic connectivity of cell types


in the adult rat dentate gyrus

Charles E. Ribak and Lee A. Shapiro

Department of Anatomy & Neurobiology, University of California at Irvine, Irvine, CA 92697-1275, USA

Abstract: The rat hippocampal dentate gyrus is an extensively studied structural component of the limbic
system. It is the first station in the classical tri-synaptic circuit of the hippocampus in that its major input
arises from the entorhinal cortex via the perforant pathway. The second part of this circuit arises from the
projection cells of the dentate gyrus, the granule cells, which send their axons to the pyramidal cells of CA3.
Within the dentate gyrus, there also is an extensive inhibitory network of cells that are involved in syn-
chronizing the rhythmic firing of the granule cells. This chapter provides a review of the ultrastructural
features and synaptic connectivity of both projection cells and local circuit neurons in the dentate gyrus.

Keywords: granule cells; basket cells; mossy cells; chandelier cells; commissural cells; fusiform cells;
GABAergic interneurons

Granule cells: the principal cells of the dentate gyrus 1963; Seress and Ribak, 1985; Lubbers and Frotsc-
her, 1987). The nucleus is usually round or slightly
The principal cells of the dentate gyrus are the oval and has heterochromatin clumping against its
granule cells, and they are mainly found in the nuclear envelope. It is uncommon to observe any
granule cell layer, which is up to 7–8 cells thick infoldings of the nucleus of granule cells.
(Ramon y Cajal, 1911; Lorente de Nó, 1934; Processes arise from the somata of granule cells
Laatsch and Cowan, 1966). Granule cells are and can be easily distinguished by the poles from
reported to exist infrequently in the two other parts which they arise. The axons of the granule cells,
of the dentate gyrus, the molecular layer and hilus the well-described mossy fibers, arise from the hi-
(Amaral, 1978; Gaarskjaer and Laurberg, 1983; lar pole of the granule cell body (Ramon y Cajal,
Marti-Subirana et al., 1986; Dashtipour et al., 1911), although one report indicated that the axon
2001). It should be noted that the vast majority of can arise from the apical dendrite (Yan et al.,
the cell bodies in the granule cell layer are granule 2001). At their origin, they display the typical fea-
cells, but not all of the cell bodies are of this type tures of an axon initial segment with few organ-
(Fig. 1). In the adult rat, granule cells (Fig. 1) dis- elles, bundles of microtubules, a subaxolemmal
play a round nucleus 10–12 mm in diameter, a thin density and infrequent axon initial segment
shell of perikaryal cytoplasm, and both symmetric synapses (Steward and Ribak, 1986). It is known
and asymmetric axosomatic synapses (Blackstad, that mossy fibers arborize in the hilus and stratum
lucidum of CA3. In addition, mossy fibers can be
Corresponding author. Tel.: +1 949-824-5494 & +1 949-824- found either intragranular or supragranular. The
4558; Fax: +1 949-824-8549; E-mail: ribak@uci.edu intragranular fibers are often observed orthogonal

DOI: 10.1016/S0079-6123(07)63009-X 155


156

NN

GC
D

GC GL
A

Fig. 1. An electron micrograph obtained from the part of the granule cell layer (GL) adjacent to the hilus (H) illustrates the
ultrastructural features of granule cells (GC), radial glial cells (A) and a doublecortin-immunolabeled newborn neuron (NN). The
mature granule cells (GC) located within the GL display a round or oval nucleus and a thin shell of perikaryal cytoplasm. The cell body
of the doublecortin-immunolabeled cell (NN) is much smaller than that of the granule cells (GC). Note that a doublecortin-labeled
apical process (white arrow) is adjacent to the radial process of an astrocyte (A), identified as a radial glial cell based on other
characteristics. Another astrocyte (A) with similar ultrastructural features is shown at the lower part of this image within the GL where
a proximal apical dendrite (D) from another mature granule cell apposes it. Many myelinated axons are present in the neuropil of the
hilus (H) and the lower part of the GL. Scale bar ¼ 2 mm.

to the granule cell layer where they form asym- hilar basal dendrites observed following seizures
metric synapses with the apical dendrites and so- are also postsynaptic to mossy fibers (Ribak et al.,
mata of basket cells (Ribak and Peterson, 1991). 2000). Thus, seizures can alter the morphology and
The supragranular fibers are located in the inner connectivity of granule cells.
molecular layer and are found to sprout there after The apical dendrites of granule cells extend ra-
seizures, but are infrequent in the normal adult rat dially through the granule cell layer (Fig. 1) and
(Nadler et al., 1980; Buckmaster et al., 2002). then arborize in the molecular layer and their tips
These supragranular mossy fibers target the den- eventually grow to reach the hippocampal fissure.
drites of granule cells to contribute to recurrent These dendrites arise as a thick process with or-
excitatory circuits after seizures. In addition, the ganelles that are typical of the perikaryal
157

cytoplasm (Fig. 1), including cisternae of the gran- Radial glial cells in the adult dentate gyrus: the
ular endoplasmic reticulum and Golgi complex mothers of adult born granule cells
(Blackstad, 1963; Laatsch and Cowan, 1966). The
apical dendrites also display spines that are more Of the many types of progenitor cells in the adult
common in the molecular layer, where perforant rat dentate gyrus, the radial glial cells appear to
pathway axons target their outer two-thirds por- have an intimate relationship with the newborn
tion. In addition, there are several types of inter- granule cells. These radial glial cells are immuno-
neurons that synapse on the apical dendrites of reactive for GFAP, are found in the hilus, sub-
granule cells in the molecular layer and the granule granular zone, in the granule cell layer, at the
cell layer. These interneurons will be described in border between the granule cell layer and the mo-
detail later in this chapter. lecular layer and in the molecular layer (Kosaka
In light of the relatively recent discovery of and Hama, 1986). In this relationship, the radial
functional neurogenesis of granule cells in the glial cell and its non-radial watery cytoplasmic
adult rat dentate gyrus, it is necessary to describe processes have been shown to encompass most of
the ultrastructural features of this population of the newborn granule cell at the border between the
newborn neurons. The newborn granule cells have subgranular zone and the granule cell layer (Fig.
been described using doublecortin (DCX)- 1). These glial processes contain only sparse or-
immunolabeled preparations and electron micros- ganelles and bundles of intermediate glial filaments
copy (Shapiro et al., 2005). These newborn granule (Shapiro et al., 2005). The region of the newborn
cells are initially found in the subgranular zone neuron that is not surrounded by the radial glial
with no dendritic processes and are cradled by an cell is shown to give rise to an apical process that
astrocyte, considered to be a radial glial cell grows toward the granule cell layer. In this rela-
(Shapiro et al., 2005). The diameter of these tionship, the radial process of the radial glial cell
DCX-labeled cells is only 6 mm, which is smaller provides a scaffold for the apical dendrite from the
than that of the mature granule cells in the layer newborn neuron to grow through the granule cell
(Fig. 1). Also, these newborn cells have only a thin layer and into the molecular layer (Shapiro et al.,
shell of perikaryal cytoplasm with few organelles. 2005).
As the DCX-labeled newborn neuron migrates to-
ward the granule cell layer, it extends an apical
dendritic process along the radial process of the Basket cells: a plexus of inhibitory axons in the
radial glial cells (Fig. 1). At this stage, the DCX- granule cell layer
labeled cell body has a larger diameter and a
thicker shell of perikaryal cytoplasm, relative to There are five distinct types of basket cells that
the DCX-labeled cells that lack processes in the have been described in the rat dentate gyrus
subgranular zone. It should be noted that the (Ribak and Seress, 1983). These basket cells (Figs.
dendrites in the molecular layer are much thinner 2 and 3) are located in the granule cell layer or the
than the apical dendrites of the mature granule hilus, within 50 mm of the base of the granule cell
cells (Shapiro et al., 2007). Although basal den- layer and may express glutamate decarboxylase,
drites are not normally present on mature granule GABA, or the calcium-binding protein, parvalbu-
cells in the adult rat dentate gyrus, they are a min (Ribak et al., 1978, 1990; Seress and Ribak,
transient feature of newborn granule cells in the 1983; Lubbers and Frotscher, 1987; Nitsch et al.,
young and adult rat (Seress and Pokorny, 1981; 1990; Halasy and Somogyi, 1993). The shapes of
Ribak and Seress, 1990; Ribak et al., 2004), and these cells can be pyramidal (Fig. 2), horizontal,
persist following seizures (Ribak et al., 2000). It is fusiform or multipolar. The ultrastructural fea-
pertinent to note that in human and non-human tures of all five types of basket cells are very similar
primates, basal dendrites are a persistent feature of (Ribak and Anderson, 1980; Ribak and Seress,
some mature granule cells in the adult dentate 1983). They have a somal size of 20–30 mm with
gyrus (Seress and Mrzljak, 1987). large nuclei containing euchromatin and
158

B
G
G

Fig. 2. (A) shows the cell body and two proximal basal dendrites (arrows) of a pyramidal basket cell that was processed by the Golgi/
electron microscopy (EM) method and was previously identified at the light microscopic level. Note the gold particles subjacent to its
plasma membrane. Only a small part of its nucleus (N) is shown but its perikaryal cytoplasm is packed with many organelles. Note the
relatively larger cell body of the basket cell as compared to that of the neighboring granule cells (G). Cell bodies of two oligodendroglia
(O) flank this basket cell at the border between the granule cell layer and the hilus (H). (B) is another basket cell processed with the
Golgi/EM method to show the characteristic features of the nuclei of these cells. These features include nuclear infoldings (black
arrows), a prominent nucleolus and intranuclear filaments (arrowhead). Like the basket cell in (A), this cell also has gold particles
(white arrow) beneath its cell membrane while the surrounding granule cells (G) are not labeled. Scale bars ¼ 2 mm in A and 3 mm in B.
(Reprinted with permission from Ribak and Seress, 1983.)
159

intranuclear rods and sheets, and displaying ex- Ribak, 1984; Zipp et al., 1989). In addition to
tensive nuclear infoldings (Figs. 2 and 3). Their forming synapses on granule cell apical dendrites,
thick shell of perikaryal cytoplasm has abundant these latter afferents also form synapses onto
cisternae of granular endoplasmic reticulum, basket cell apical dendrites. These connections
which aggregate to form Nissl bodies, a well-de- were suggested to function as a feed-forward in-
veloped Golgi complex, numerous mitochondria, hibitory circuit for granule cells (Seress and Ribak,
free ribosomes and lysosomes. Golgi-electron mi- 1984; Zipp et al., 1989; Kneisler and Dingledine,
croscopic preparations revealed details about the 1995a). Other studies have shown that CA3 and
basket cell axons, showing that they have an ex- the septum provide excitatory and inhibitory in-
tensive arborization concentrated mostly in the put, respectively, to basket cells (Gulyas et al.,
granule cell layer, and to a lesser extent at the 1990; Scharfman, 1994; Kneisler and Dingledine,
border with the molecular layer. The axon termi- 1995b). Therefore, the GABAergic inhibitory bas-
nals of basket cells contain numerous mi- ket cells mediate both feedback and feed-forward
tochondria and a sparse number of pleomorphic inhibition in the dentate gyrus (Kneisler and
synaptic vesicles. These terminals were shown to Dingledine, 1995a).
form symmetric synapses, most commonly with
the somata and proximal apical dendrites of gran-
ule cells (Ribak and Seress, 1983). In rare instances Mossy cells: the cells with great convergence of
the basket cell axon terminal was observed to form mossy fibers
a synapse with a spine (Ribak and Seress, 1983). It
is important to emphasize that through these con- The mossy cells represent a distinct population of
nections, these basket cells provide feedback inhi- neurons that dominate the landscape of the dentate
bition to the granule cells of the dentate gyrus. In gyrus hilar region (Ribak et al., 1985; Frotscher
contrast, basket cells have not been observed to et al., 1991). The mossy cells are characterized by
synapse on axon initial segments in adult rats, large triangular or multipolar somata (25–30 mm di-
however, examples of this were documented in ameter). These somata contain a round nucleus
young (postnatal 10–16 days) rats (Seress and without infoldings, a thick shell of perikaryal cyto-
Ribak, 1990). plasm and abundant organelles (Fig. 4). Arising
The dendrites of basket cells are aspinous or from these somata are three or four thick primary
sparsely spinous and are found in all layers of the dendrites (Fig. 4), which bifurcate to form an elab-
dentate gyrus (Ribak and Seress, 1983). The asp- orate dendritic plexus that is mostly restricted to the
inous dendrites have a mixture of asymmetric and hilus. These cells have complex spines known as
symmetric synapses on their surfaces. In the hilus, thorny excrescences on their somata and proximal
the basal dendrites of pyramidal and fusiform dendrites (Fig. 4), while their distal dendrites have
basket cells have the greatest concentration of typical spines. The majority of axon terminals form-
axon terminals apposed to their surfaces, and most ing synapses onto these dendrites arise from mossy
of these axon terminals contain round synaptic fibers of granule cells (Fig. 4). Therefore, this mas-
vesicles and form asymmetric synapses. Some of sive granule cell input provides a large convergence
the axon terminals that synapse on the basal to mossy cells, and this input was shown to be ex-
dendrites of pyramidal basket cells were identified citatory to mossy cells (Scharfman et al., 1990).
as mossy fiber terminals (Ribak and Seress, 1983). The axons of these mossy cells bifurcate and the
The basket cells receive afferents from at least two major projection is to the distant portions of the
other fiber systems besides the mossy fibers from inner molecular layer (Soltesz et al., 1993). The
granule cells. They include the commissural axons axon terminals of mossy cells were observed to
from the contralateral dentate gyrus that terminate make asymmetric synapses onto postsynaptic tar-
in the inner molecular layer and beneath the gran- gets in the hilus and molecular layer of the dentate
ule cell layer, and perforant path fibers that ter- gyrus (Ribak et al., 1985; Scharfman et al., 1990)
minate in the outer molecular layer (Seress and and showed immunoreactivity primarily for
160

Fig. 3. Illustrations of a pyramidal basket cell that was immunolabeled for parvalbumin. (A) shows a semi-thin section of this cell in
the lower part of the granule cell layer (GL) next to the hilus (H). Its apical dendrite branches in the inner molecular layer (ML). The
small punctate labeling in the GL probably represents the basket cell axons that outline the cell bodies of unlabeled granule cells. The
electron micrograph in (B) shows the soma of this parvalbumin-immunolabeled pyramidal basket cell with the typical features of this
cell type. These include nuclear infoldings (arrowheads), intranuclear rod (open arrow), axosomatic synapses (closed arrow) and a
thick perikaryal cytoplasm. Note the size difference between this soma and the somata of the adjacent granule cells (G). (C) is an
enlargement of the axosomatic symmetric synapses (arrows) denoted in (B) with an arrow. The axon terminals (T) forming these
synapses with the parvalbumin-immunolabeled basket cell body have pleomorphic vesicles and several mitochondria. (D) is another
example of axosomatic synapses with the same basket cell body but these are located at the transition with its apical dendrite and these
terminals (T) form asymmetric synapses (arrows). Note that this part of the soma displays more immunolabeling. Scale bar in A ¼
10 mm in A, 2.5 mm in B, 0.1 mm in C and D. (Reprinted with permission from Ribak et al., 1990.)
161

Fig. 4. (A) shows the cell body and two thick proximal dendrites (D) of a mossy cell that was processed by the Golgi/EM method and
was previously identified at the light microscopic level. Most of its cell body is occupied by a large round nucleus (N). A large somal
spine (arrow) is found to the right of the cell body. Two dendrites (D) arise from the cell body. (B) shows a portion of a dendrite (D)
from this mossy cell. Note the numerous gold-labeled spines arising from this dendrite. Many of these spines are postsynaptic to large
axon terminals (M) that resemble mossy fiber tufts. (C) is an enlargement of a mossy fiber (M and outlined by dashed line) that forms a
synapse (arrow) with a gold-labeled dendrite of this mossy cell and with one of its gold-labeled spines (S). Scale bars ¼ 4 mm in A,
1.5 mm in B and 0.5 mm in C. (Reprinted with permission from Ribak et al., 1985.)

glutamate, but never for GABA (Soriano and Fusiform cells (spiny and aspiny) in the hilus
Frotscher, 1994). Within the hilus, glutamate-pos-
itive mossy cell axon terminals targeted GABA- The fusiform cells of the rat dentate gyrus are lo-
positive dendritic shafts of hilar interneurons cated in the hilus, within 100 mm of the base of the
and GABA-negative dendritic spines. Mossy cell granule cell layer (Fig. 5). These cells are one of the
axons in the inner molecular layer form asymmet- most prevalent cell types in the subgranular zone
ric synapses with dendritic spines associated of the dentate gyrus (Ribak and Seress, 1988). The
with GABA-negative, granule cell dendrites. Thus, fusiform cells are bipolar, with oval cell bodies,
excitatory (glutamatergic) mossy cell termi- and their dendrites typically run parallel to the
nals contact GABAergic interneurons and non- granule cell layer, and within the subgranular zone
GABAergic neurons in the hilar region and (Fig. 5). The dendrites of these cells are either
GABA-negative dendrites of granule cells in the spiny or sparsely spiny. The spiny fusiform cell
molecular layer. This pattern of connectivity is receives the majority of its input from the axon
consistent with the hypothesis that mossy cells collaterals of the mossy fiber axons from the gran-
provide excitatory feedback to granule cells in a ule cells. Interestingly, the spines are not only ob-
dentate gyrus associational network and also ac- served on the dendrites, but are also observed on
tivate local hilar inhibitory elements (Wenzel et al., the soma (Fig. 5). The spiny fusiform cell has a
1997). soma that is similar in appearance to the mossy
162

Fig. 5. Illustrations of two types of fusiform cells that were processed by the Golgi/EM method and were previously identified at the
light microscopic level. (A) is an electron micrograph of a sparsely spiny fusiform cell body and one of its proximal dendrites (D). The
nucleus (N) displays a prominent nucleolus and numerous infoldings (large arrows). The perikaryal cytoplasm contains many cisternae
of granular endoplasmic reticulum (ER) that are organized into stacks or Nissl bodies. A few small somal spines (small arrows) are also
shown. (B) is a light micrograph of the same sparsely spiny fusiform cell shown in (A) to illustrate its flattened soma (large arrow) and
its dendrites that run parallel to the granule cell layer (small arrows). (C) shows a spiny fusiform cell body with many gold-labeled
somal spines (arrows). The nucleus (N) in this section occupies less area than would the nucleus obtained from a section through the
center of the soma. The cisternae of the granular endoplasmic reticulum (ER) do not form parallel stacks. Electron-dense lysosomes as
well as mitochondria are randomly distributed in the perikaryal cytoplasm. A small capillary (C) is also shown in the upper left corner
of the micrograph. (D) is an enlargement of the two somal spines (S) indicated by the top arrow in (C). Large axon terminals packed
with agranular round vesicles form asymmetric synapses (large arrows) with the gold-labeled spines. In contrast, the somal surface is
contacted by a small terminal that appears to form a symmetric synapse (small arrow). Scale bars ¼ 20 mm in (A), 2 mm in (B), 2 mm in
(C) and 0.5 mm in (D). (Reprinted with permission from Ribak and Seress, 1988.)

cell because it displays Nissl bodies and little or no axodendritic synapses, suggesting a more diverse
nuclear infolding (Fig. 5). In contrast, the fusiform synaptic input than that of the spiny fusiform cell.
cells with sparsely spiny dendrites have somata Several studies have demonstrated the presence
with nuclear infoldings (Fig. 5) like that of basket of somatostatin immunolabeling in hilar cells with
cells and few Nissl bodies. It should also be noted the morphology of fusiform cells (Bakst et al.,
that this latter cell type displays a variety of 1986; Milner and Bacon, 1989; Leranth et al.,
163

1990). These studies provided a valuable insight the hilus. The three chandelier cell types observed
into the axonal projections of the fusiform cell. It in the hilus projected to granule cells at the same
was shown that the somatostatin-labeled axons septo–temporal level where their cell bodies were
formed a dense plexus in the outer molecular layer located. It should also be noted that the chandelier
where they made symmetric synapses onto the cells exhibit immunolabeling for glutamate de-
granule cell dendrites (Bakst et al., 1986; Milner carboxylase, GABA-transporter-1 and parvalbu-
and Bacon, 1989; Leranth et al., 1990). In addi- min (Ribak et al., 1990, 1996; Soriano et al., 1990).
tion, somatostatin-labeled axon terminals in the Thus, the dentate gyrus chandelier cell provides a
hilus were also reported (Milner and Bacon, 1989). spatially selective innervation of granule cells and
It should be noted that more than 90% of the hilar complements that of the basket cell that provides
neurons expressing somatostatin are GABAergic GABAergic inhibition to cell bodies and proximal
(Esclapez and Houser, 1995) and that 50% of dendrites.
them express neuropeptide Y (Deller and Leranth,
1990).
Commissural neurons: communication and
synchronization across hemispheres involving
Chandelier cells: a GABAergic cell-type targeting excitatory and inhibitory neurons
axon initial segments of granule cells
Within the dentate gyrus, there are at least two
Chandelier cells are observed throughout the classes of cells that send projections to the con-
cerebral cortex, but were first described in the tralateral dentate gyrus. Using retrograde labeling
hippocampus by Kosaka (1980) and Somogyi et al. with horseradish peroxidase (HRP), Seroogy et al.
(1983a, b), and later in the dentate gyrus by (1983) showed different electron microscopic fea-
Soriano and Frotscher (1989). All chandelier cells tures for two classes of labeled commissural neu-
share two features that are unique among inhib- ron. The first type consisted of cells with somata
itory interneurons; their axons form rows of that exhibited round or oval nuclei with no intra-
boutons and they make symmetric synapses ex- nuclear inclusions and had exclusively symmetric
clusively on axon initial segments. synapses on their somata. The main dendrites of
The chandelier cells of the dentate gyrus have those neurons were thick and tapering. This type
their somata located within or immediately adja- had features that resembled the morphology of the
cent to the granule cell layer. The cell bodies and mossy cell. A subsequent study using combined
dendrites of these neurons exhibit several of the retrograde transport of HRP with Golgi/electron
characteristic ultrastructural features of non-gran- microscopy (EM) confirmed this suggestion
ule cells, including a large perikaryal cytoplasm, (Frotscher, 1992). The second type of labeled neu-
nuclear infoldings, intranuclear inclusions and a ronal soma had infolded nuclei containing intra-
large number of synapses on the soma and aspiny nuclear rods or sheets, displayed both symmetric
dendrites. Within the dentate gyrus, chandelier cell and asymmetric synapses on its soma and had
axons densely innervate the granule cell layer with dendrites that were less thick and generally asp-
radially oriented terminal rows, and also form an inous. This type of commissurally labeled cell had
extensive plexus in the hilus. The dendrites of the features that were similar to the dentate gyrus
chandelier cells extend radially through the mo- basket cell, a local circuit neuron associated with
lecular layer and can reach as far as the hippo- GABAergic inhibition, as described above. An-
campal fissure (Soriano and Frotscher, 1989; other line of evidence supported the possibility
Soriano et al., 1990; Han et al., 1993; Buhl et al., that GABAergic neurons had commissural pro-
1994). Occasionally, basal dendrites from chande- jections. In a quantitative study of GABAergic
lier cells cross the granule cell layer toward the neurons in the hilus of the dentate gyrus, Seress
hilus (Soriano et al., 1990). There are also three and Ribak (1983) showed that 60% of the hilar
classes of chandelier cells that have their somata in neurons are GAD-positive. Because previous
164

studies indicated that 80% of hilar neurons give layer and its dendritic and axonal domains were
rise to associational and commissural pathways, confined to the perforant path terminal zone. This
many GABAergic neurons in the hilus were sug- cell type was referred to as the molecular layer/
gested to be projection neurons, and a subsequent perforant pathway (MOPP) cell and its axon made
combined tracer and immunofluorescence study symmetric synapses exclusively onto dendritic
showed several double-labeled GABAergic neu- shafts, 60% of which were shown to emit spines
rons in the hilus contralateral to the injection site (Halasy and Somogyi, 1993). The smooth dend-
(Ribak et al., 1986). Another confirmation of this rites of the MOPP cell were also restricted to the
conclusion came with the elegant intracellular outer two-thirds of the molecular layer, where they
labeling study of Han et al. (1993). In that study, received both GABA-negative and GABA-positive
the axons of several classes of GABAergic inter- synaptic inputs. Similar to the HIPP cells, ultra-
neurons were mapped in the rat dentate gyrus, in- structural details about the MOPP cell’s soma
cluding one that had its axon terminals distributed were not included in the description.
in the commissural and associational pathway ter-
mination field. Because they named the cell using
both its cell body location and axon terminal field, Conclusion
and the cell body of these neurons resided in the
hilus, this cell type was referred to as the hilus/ There have been many studies on the cell types in
commissural-associational pathways (HICAP) the dentate gyrus and their synaptic connections.
cell (Han et al., 1993). Therefore, the commissu- Because the dentate gyrus is the first station in
ral projection in the rat includes both excitatory the classical tri-synaptic circuit of the hippocam-
(mossy cells) and inhibitory (HICAP cells) pus, understanding its anatomical organization is
neurons. essential to understand the function of this struc-
ture. Here we have outlined the ultrastructural
features and synaptic connections of the principal
Other GABAergic interneuron types cell type (the granule cells) and the projection and
local circuit neurons that make up the dentate
Two other GABAergic neuron types were observed gyrus.
in the dentate gyrus and were described simulta-
neously with the HICAP cell (Han et al., 1993).
One of these is another type of hilar cell that has an Acknowledgments
axon, which ramifies in the perforant path terminal
field in the outer two-thirds of the molecular layer. The authors are grateful to the following collab-
This hilar cell with axon terminals distributed in orators who contributed significantly to one or
the perforant path termination field was named the several studies in this review: Drs. David Amaral,
hilus/perforant pathway (HIPP) cell. Electron mi- Nich Brecha, Khashayar Dashtipour, James Fal-
croscopy of this cell type revealed sparsely spiny lon, Michael Frotscher, Mathew Korn, J. Victor
dendrites that were covered with many synaptic Nadler, Robert Nitsch, Andre Obenaus, Maxine
boutons on both their shafts and their spines but Okazaki, Gary Peterson, Kihachi Saito, Larry
no details were provided about the HIPP cell’s Schmued, Laszlo Seress, Kim Seroogy, Igor Spi-
somal features (Han et al., 1993). It should be gelman, Oswald Steward, Winnie Tong, James
noted that this cell type has its axon terminal field Vaughn and Xiao-Xin Yan. This review is dedi-
in the same location as the somatostatin-labeled cated to the mentor of C.E.R., Alan Peters, who
cell described above in the fusiform cell section. perfected the Golgi/EM method for the analysis
That cell type was also GABAergic (Esclapez and of short axonal connections. We acknowledge
Houser, 1995). the support from NIH grant R01-NS38331 (to
The other GABAergic cell type described by C.E.R.) and NIH training grant T32-NS45540 (for
Han et al. (1993) had its cell body in the molecular L.A.S.).
165

References dentate basket cells in the rat hippocampus. Hippocampus, 5:


151–164.
Amaral, D.G. (1978) A Golgi study of cell types in the hilar Kneisler, T.B. and Dingledine, R. (1995b) Synaptic input from
region of the hippocampus in the rat. J. Comp. Neurol., 195: CA3 pyramidal cells to dentate basket cells in rat hippocam-
851–914. pus. J. Physiol., 487: 125–146.
Bakst, I., Avendano, C., Morrison, J.H. and Amaral, D.G. Kosaka, T. (1980) The axon initial segment as a synaptic site:
(1986) An experimental analysis of the origins of somatosta- ultrastructure and synaptology of the initial segment of the
tin-like immunoreactivity in the dentate gyrus of the rat. J. pyramidal cell in the rat hippocampus (CA3 region). J. Ne-
Neurosci., 6: 1452–1462. urocytol., 9: 861–882.
Blackstad, T. (1963) Ultrastructural studies on the hippocampal Kosaka, T. and Hama, K. (1986) Three-dimensional structure
region. Prog. Brain Res., 3: 122–148. of astrocytes in the rat dentate gyrus. J. Comp. Neurol., 249:
Buckmaster, P.S., Zhang, G.F. and Yamawaki, R. (2002) Axon 242–260.
sprouting in a model of temporal lobe epilepsy creates a pre- Laatsch, R.H. and Cowan, W.M. (1966) Electron microscopic
dominantly excitatory feedback circuit. J. Neurosci., 22: studies of the dentate gyrus of the rat. I. Normal structure
6650–6658. with special reference to synaptic organization. J. Comp.
Buhl, E.H., Han, Z.S., Lorinczi, Z., Stezhka, V.V., Karnup, Neurol., 128: 359–396.
S.V. and Somogyi, P. (1994) Physiological properties of an- Leranth, C., Malcolm, A.J. and Frotscher, M. (1990) Afferent
atomically identified axo-axonic cells in the rat hippocampus. and efferent synaptic connections of somatostatin-immuno-
J. Neurophysiol., 71: 1289–1307. reactive neurons in the rat fascia dentata. J. Comp. Neurol.,
Dashtipour, K., Tran, P.H., Okazaki, M.M., Nadler, J.V. and 295: 111–122.
Ribak, C.E. (2001) Ultrastructural features and synaptic Lorente de Nó, R. (1934) Studies on the structure of the
connections of hilar ectopic granule cells in the rat dentate cerebral cortex. II. Continuation of the study of the ammonic
gyrus are different from those of granule cells in the granule system. J. Psychol. Neurol. Lpz., 46: 113–177.
cell layer. Brain Res., 890: 261–271. Lubbers, K. and Frotscher, M. (1987) Fine structure and
Deller, T. and Leranth, C. (1990) Synaptic connections of synaptic connections of identified neurons in the rat fascia
neuropeptide Y (NPY) immunoreactive neurons in the hilar dentata. Anat. Embryol. (Berl.), 177: 1–14.
area of the rat hippocampus. J. Comp. Neurol., 300: Marti-Subirana, A., Soriano, E. and Garcia-Verdugo, J.M.
433–447. (1986) Morphological aspects of the ectopic granule-like cel-
Esclapez, M. and Houser, C.R. (1995) Somatostatin neurons lular populations in the albino rat hippocampal formation: a
are a subpopulation of GABA neurons in the rat dentate Golgi study. J. Anat., 144: 31–47.
gyrus: evidence from colocalization of pre-prosomatostatin Milner, T.A. and Bacon, C.E. (1989) Ultrastructural localiza-
and glutamate decarboxylase messenger RNAs. Neurosci- tion of somatostatin-like immunoreactivity in the rat dentate
ence, 64: 339–355. gyrus. J. Comp. Neurol., 290: 544–560.
Frotscher, M. (1992) Application of the Golgi/electron micros- Nadler, J.V., Perry, B.W. and Cotman, C.W. (1980) Selective
copy technique for cell identification in immunocytochemi- reinnervation of hippocampal area CA1 and the fascia den-
cal, retrograde labeling, and developmental studies of tata after destruction of CA3-CA4 afferents with kainic acid.
hippocampal neurons. Microsc. Res. Tech., 23: 306–323. Brain Res., 182: 1–9.
Frotscher, M., Seress, L., Schwerdtfeger, W.K. and Buhl, E. Nitsch, R., Soriano, E. and Frotscher, M. (1990) The parval-
(1991) The mossy cells of the fascia dentata: a comparative bumin-containing nonpyramidal neurons in the rat hippo-
study of their fine structure and synaptic connections in ro- campus. Anat. Embryol. (Berl.), 181: 413–425.
dents and primates. J. Comp. Neurol., 312: 145–163. Ramon y Cajal, S. (1911) Histologie du Systeme Nerveux de
Gaarskjaer, F.B. and Laurberg, S. (1983) Ectopic granule cells l’Homme et des Vertebres, Vol. 2. Maloine, Paris.
of hilus fasciae dentatae projecting to the ipsilateral regio Ribak, C.E. and Anderson, L. (1980) Ultrastructure of the py-
inferior of the rat hippocampus. Brain Res., 274: 11–16. ramidal basket cells in the dentate gyrus of the rat. J. Comp.
Gulyas, A.I., Gorcs, T.J. and Freund, T.F. (1990) Innervation Neurol., 192: 903–916.
of different peptide-containing neurons in the hippocampus Ribak, C.E., Korn, M.J., Shan, Z. and Obenaus, A. (2004)
by GABAergic septal afferents. Neuroscience, 37: 31–44. Dendritic growth cones and recurrent basal dendrites are
Halasy, K. and Somogyi, P. (1993) Subdivisions in the multiple typical features of newly generated dentate granule cells in
GABAergic innervation of granule cells in the dentate gyrus the adult hippocampus. Brain Res., 1000: 195–199.
of the rat hippocampus. Eur. J. Neurosci., 5: 411–429. Ribak, C.E., Nitsch, R. and Seress, L. (1990) Proportion of
Han, Z.S., Buhl, E.H., Lorinczi, Z. and Somogyi, P. (1993) A parvalbumin-positive basket cells in the GABAergic inner-
high degree of spatial selectivity in the axonal and dendritic vation of pyramidal and granule cells of the rat hippocampal
domains of physiologically identified local-circuit neurons in formation. J. Comp. Neurol., 300: 449–461.
the dentate gyrus of the rat hippocampus. Eur. J. Neurosci., Ribak, C.E. and Peterson, G.M. (1991) Intragranular mossy
5: 395–410. fibers in rats and gerbils form synapses with the somata and
Kneisler, T.B. and Dingledine, R. (1995a) Spontaneous and proximal dendrites of basket cells in the dentate gyrus.
synaptic input from granule cells and the perforant path to Hippocampus, 1: 355–364.
166

Ribak, C.E. and Seress, L. (1983) Five types of basket cell in the Seress, L. and Ribak, C.E. (1990) The synaptic connections of
hippocampal dentate gyrus: a combined Golgi and electron basket cell axons in the developing rat hippocampal forma-
microscopic study. J. Neurocytol., 12: 577–597. tion. Exp. Brain Res., 81: 500–508.
Ribak, C.E. and Seress, L. (1988) A Golgi-electron microscopic Seroogy, K.B., Seress, L. and Ribak, C.E. (1983) Ultrastructure
study of fusiform neurons in the hilar region of the dentate of commissural neurons of the hilar region in the hippocam-
gyrus. J. Comp. Neurol., 271: 67–78. pal dentate gyrus. Exp. Neurol., 82: 594–608.
Ribak, C.E. and Seress, L. (1990) Postnatal development of the Shapiro, L.A., Korn, M.J., Shan, Z. and Ribak, C.E. (2005)
light and electron microscopic features of basket cells in the GFAP-expressing radial glia-like cell bodies are involved in a
hippocampal dentate gyrus of the rat. Anat. Embryol. (Berl.), one-to-one relationship with doublecortin-immunolabeled
181: 547–565. newborn neurons in the adult dentate gyrus. Brain Res.,
Ribak, C.E., Seress, L. and Amaral, D.G. (1985) The develop- 1040: 81–91.
ment, ultrastructure and synaptic connections of the mossy Shapiro, L.A., Upadhyaya, P. and Ribak, C.E. (2007) Spatio-
cells of the dentate gyrus. J. Neurocytol., 14: 835–857. temporal profile of dendritic outgrowth from newly born
Ribak, C.E., Seress, L., Peterson, G.M., Seroogy, K.B., Fallon, granule cells in the adult rat dentate gyrus. Brain Res., 1149:
J.H. and Schmued, L.C. (1986) A GABAergic inhibitory 30–37.
component within the hippocampal commissural pathway. J. Soltesz, I., Bourassa, J. and Deschenes, M. (1993) The behavior
Neurosci., 6: 3492–3498. of mossy cells of the rat dentate gyrus during theta oscilla-
Ribak, C.E., Tong, W.M. and Brecha, N.C. (1996) GABA tions in vivo. Neuroscience, 57: 555–564.
plasma membrane transporters, GAT-1 and GAT-3, display Somogyi, P., Nunzi, M.G., Gorio, A. and Smith, A.D. (1983a)
different distributions in the rat hippocampus. J. Comp. A new type of specific interneuron in the monkey hippocam-
Neurol., 367: 595–606. pus forming synapses exclusively with the axon initial seg-
Ribak, C.E., Tran, P.H., Spigelman, I., Okazaki, M.M. and ments of pyramidal cells. Brain Res., 259: 137–142.
Nadler, J.V. (2000) Status epilepticus-induced hilar basal Somogyi, P., Smith, A.D., Nunzi, M.G., Gorio, A., Takagi, H.
dendrites on rodent granule cells contribute to recurrent ex- and Wu, J.Y. (1983b) Glutamate decarboxylase immunore-
citatory circuitry. J. Comp. Neurol., 428: 240–253. activity in the hippocampus of the cat: distribution of
Ribak, C.E., Vaughn, J.E. and Saito, K. (1978) Immunocyto- immunoreactive synaptic terminals with special reference to
chemical localization of glutamic acid decarboxylase in neu- the axon initial segment of pyramidal neurons. J. Neurosci.,
ronal somata following colchicine inhibition of axonal 3: 1450–1468.
transport. Brain Res., 140: 315–332. Soriano, E. and Frotscher, M. (1989) A GABAergic axo-axonic
Scharfman, H.E. (1994) Synchronization of area CA3 hippo- cell in the fascia dentata controls the main excitatory hippo-
campal pyramidal cells and non-granule cells of the dentate campal pathway. Brain Res., 503: 170–174.
gyrus in bicuculline-treated rat hippocampal slices. Neuro- Soriano, E. and Frotscher, M. (1994) Mossy cells of the rat
science, 59: 245–257. fascia dentata are glutamate-immunoreactive. Hippocampus,
Scharfman, H.E., Kunkel, D.D. and Schwartzkroin, P.A. 4: 65–69.
(1990) Synaptic connections of dentate granule cells and hi- Soriano, E., Nitsch, R. and Frotscher, M. (1990) Axo-axonic
lar neurons: results of paired intracellular recordings and in- chandelier cells in the rat fascia dentata: Golgi-electron mi-
tracellular horseradish peroxidase injections. Neuroscience, croscopy and immunocytochemical studies. J. Comp.
37: 693–707. Neurol., 293: 1–25.
Seress, L. and Mrzljak, L. (1987) Basal dendrites of granule Steward, O. and Ribak, C.E. (1986) Polyribosomes associated
cells are normal features of the fetal and adult dentate gyrus with synaptic specializations on axon initial segments: local-
of both monkey and human hippocampal formations. Brain ization of protein-synthetic machinery at inhibitory synapses.
Res., 405: 169–174. J. Neurosci., 6: 3079–3085.
Seress, L. and Pokorny, J. (1981) Structure of the granular layer Wenzel, H.J., Buckmaster, P.S., Anderson, N.L., Wenzel, M.E.
of the rat dentate gyrus. J. Anat., 133: 181–195. and Schwartzkroin, P.A. (1997) Ultrastructural localization
Seress, L. and Ribak, C.E. (1983) GABAergic cells in the dent- of neurotransmitter immunoreactivity in mossy cell axons
ate gyrus appear to be local circuit and projection neurons. and their synaptic targets in the rat dentate gyrus. Hippo-
Exp. Brain Res., 50: 173–182. campus, 7: 559–570.
Seress, L. and Ribak, C.E. (1984) Direct commissural connec- Yan, X.X., Spigelman, I., Tran, P.H. and Ribak, C.E. (2001)
tions to the basket cells of the hippocampal dentate gyrus: Atypical features of rat dentate granule cells: recurrent basal
anatomical evidence for feed-forward inhibition. J. Neuro- dendrites and apical axons. Anat. Embryol. (Berl.), 203:
cytol., 13: 215–225. 203–209.
Seress, L. and Ribak, C.E. (1985) A substantial number of Zipp, F., Nitsch, R., Soriano, E. and Frotscher, M. (1989)
asymmetric axosomatic synapses is a characteristic of the Entorhinal fibers form synaptic contacts on parvalbumin-
granule cells of the hippocampal dentate gyrus. Neurosci. immunoreactive neurons in the rat fascia dentata. Brain Res.,
Lett., 56: 21–26. 495: 161–166.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 10

Morphological development and maturation of


granule neuron dendrites in the rat dentate gyrus

Omid Rahimi1 and Brenda J. Claiborne2,

1
Department of Cellular and Structural Biology, University of Texas Health Science Center at San Antonio, San Antonio,
TX 78229, USA
2
Department of Biology, University of Texas at San Antonio, One UTSA Circle, San Antonio, TX 78249, USA

Abstract: The first granule neurons in the dentate gyrus are born during late embryogenesis in the rodent,
and the primary period of granule cell neurogenesis continues into the second postnatal week. On the day of
birth in the rat, the oldest granule neurons are visible in the suprapyramidal blade and exhibit rudimentary
dendrites extending into the molecular layer. Here we describe the morphological development of the
dendritic trees between birth and day 14, and we then review the process of dendritic remodeling that occurs
after the end of the second week. Data indicate that the first adult-like granule neurons are present on day
7, and, furthermore, physiological recordings demonstrate that some granule neurons are functional at this
time. Taken together, these results suggest that the dentate gyrus may be incorporated into the hippo-
campal circuit as early as the end of the first week. The dendritic trees of the granule neurons, however,
continue to increase in size until day 14. After that time, the dendritic trees of the oldest granule neurons are
sculpted and refined. Some dendrites elongate while others are lost, resulting in a conservation of total
dendritic length. We end this chapter with a review of the quantitative aspects of granule cell dendrites in
the adult rat and a discussion of the relationship between the morphology of a granule neuron and the
location of its cell body within stratum granulosum and along the transverse axis of the dentate gyrus.

Keywords: dendritic trees; spines; filopodia; neonates; hippocampus

Introduction forms synapses on pyramidal neurons in the CA3


region of the hippocampus proper. The apical
Granule neurons are the principal cell type in the dendrites of the granule neurons bifurcate as they
dentate gyrus, and their cell bodies are located in traverse the molecular layer, and the vast majority
stratum granulosum of the suprapyramidal and in- of terminal branches reach the top of the layer in the
frapyramidal blades. Dendrites extend from the adult. The dendritic trees of most granule neurons
apical pole of the granule cell body into the over- are elliptical, and all dendrites of granule neurons in
lying molecular layer, and the axon, or mossy fiber, the adult dentate gyrus are covered with spines.
exits from the basal pole. The axon gives rise to The primary period of granule cell neurogenesis
collateral branches in the hilar region and then occurs over a two- to three-week period in the ro-
dent, beginning in late embryogenesis and contin-
Corresponding author. Tel.: +1 210 458 5487; uing through the second postnatal week. In the rat,
Fax: 1 210 458 5669; E-mail: brenda.claiborne@utsa.edu although a few granule neurons are born as early as

DOI: 10.1016/S0079-6123(07)63010-6 167


168

embryonic day 14, over 80% are born after the Thus the oldest granule neurons are most likely
birth of the animal (which occurs at about embry- to be found at the top of stratum granulosum near
onic day 21) and neurogenesis peaks near the end of the distal tip of the suprapyramidal blade at the
the first week of life (Bayer and Altman, 1974; septal pole, whereas the youngest neurons are lo-
Schlessinger et al., 1975). It is worth noting that the cated predominantly in the infrapyramidal blade
granule neurons are the last cells to be generated in near the temporal pole and in the deeper portions
the hippocampal formation, and it is well known of stratum granulosum along the entire extent of
that granule cell neurogenesis continues into adult- the transverse axis of the dentate gyrus. In the rat,
hood (Altman and Das, 1965; Kaplan and Hinds, the suprapyramidal blade begins to form in late
1977). Here we focus on granule neurons that are embryogenesis as the first granule neurons are
generated in the neonatal rat. From recent evidence born — it is visible as a separate structure on the
in the mouse, it appears that adult-generated gran- day of birth and consists of a cell-body layer and a
ule neurons progress through a similar set of stages relatively thin molecular layer (Cowan et al.,
as they develop and mature (Zhao et al., 2006). 1980). The molecular layer increases greatly in
Any description of the development and matu- width over the first several weeks. It is less than
ration of the granule neurons must take into ac- 100 mm at day 4 and increases to just over 200 mm
count the temporal and spatial gradients of at day 14; it averages approximately 300 mm in
granule cell neurogenesis (Schlessinger et al., width in young adult rats (Loy et al., 1977; Clai-
1975; Cowan et al., 1980, 1981). The earliest born borne et al., 1990; Rihn and Claiborne, 1990). The
granule neurons form stratum granulosum in the infrapyramidal blade is barely visible on the day of
septal portion of the dentate gyrus, and neurons birth and grows more slowly than the suprapy-
that are generated later form the more temporal ramidal blade during the first week, increasing
portions of the dentate gyrus. This gradient is re- from approximately 45 mm in width on day 4 to
ferred to as the septotemporal gradient. A second approximately 110 mm on day 10; it measures be-
gradient exists along the transverse axis of the tween 205 and 240 mm in young adult rats (Loy et
dentate gyrus and is of considerable importance al., 1977; Claiborne et al., 1990).
for developmental and morphological studies. As Here we describe the development and matura-
the first granule neurons are born, they form the tion of the dendritic trees of the granule neurons in
cell-body layer at the tip of the suprapyramidal the rat, and we review the quantitative data on
blade. As additional neurons are generated, they dendritic morphology in the adult. We consider
form the cell-body layer in the middle of the sup- the developmental period to encompass the time
rapyramidal blade and then the portions of stra- from the birth of the animal through day 14. By
tum granulosum closest to the crest region. This day 14, the oldest granule neurons have assumed
gradient continues as more neurons are generated their adult form and size. From day 14 to 60,
such that the youngest neurons make up stratum however, the neurons go through a period of mat-
granulosum in the infrapyramidal blade. A third uration during which the dendritic tree is sculpted
gradient exists within stratum granulosum. The and refined and the density of spines continues to
neurons that are born first move into their final increase. The dendritic trees of granule neurons
position at the top of the cell-body layer near the appear to be mature by day 60: unpublished data
molecular layer, and the younger neurons move from our lab indicate that they do not undergo any
into position beneath them such that they are lo- quantitative changes between 60 and 180 days.
cated in the bottom portion of stratum granulo-
sum near the hilar border. This developmental
pattern is in contrast to the ‘‘inside-out’’ pattern Development of granule neuron dendrites
found in other areas of the mammalian cerebral
cortex in which the later-generated neurons move Because of the prolonged time-course of granule
through the earlier-generated cells to occupy po- cell neurogenesis in neonatal rats, a wide range of
sitions at the top of the cell-body layer. dendritic morphologies are observed on any one
169

day during the first and second postnatal weeks variety of ages between days 0 and 180, and Zafirov
(Fricke, 1975; Seress and Pokorny, 1981; Wenzel et al. (1994) described granule neurons at days 5,
et al., 1981; Lübbers and Frotscher, 1988; Liu 10, 15, and 20. Trommer and colleagues (Liu et al.,
et al., 1996, 2000; Ye et al., 2000; Jones et al., 1996, 2000; Ye et al., 2000) injected granule neurons
2003). It is possible, however, to identify distinct in slices from rats between the ages of 5 and 32 days
stages in the development of granule cell dendritic with biocytin, whereas Jones et al. (2003) used ret-
trees, either by comparing the various neuronal rograde labeling with DiI to analyze the dendritic
structures present on a single day or by examining structures of granule neurons in rats between the
the progression of dendritic morphologies over the ages of 2 and 9 days. The latter group focused on
course of the neonatal period. As an example of the oldest granule neurons — those that were lo-
the first approach, Lübbers and Frotscher (1988) cated near the tip of the suprapyramidal blade and
identified several stages of developing granule neu- in the top portion of stratum granulosum.
rons in a 5-day-old rat. Neurons in the earliest On the day of birth and on postnatal day 1 in the
stages of development had only rudimentary den- rat, granule neurons are visible in the suprapyram-
dritic trees, each consisting of two or more pri- idal blade of Golgi-stained tissue and exhibit rudi-
mary apical dendrites exiting from the cell body, mentary trees. Some have only a few short, stubby
along with one or more basal branches. The pri- branches, whereas others have longer and more
mary apical dendrites gave rise to higher order numerous dendrites (Fricke, 1975; Wenzel et al.,
branches that were relatively short and thin and 1981). The dendrites are smooth and growth cones
that exhibited varicosities and the occasional are not typically observed. Wenzel et al. (1981)
growth cone. Spines were not present on these considered granule neurons on the day of birth to
immature cells. In contrast, the most mature gran- be in the first stage of development and labeled
ule neurons on day 5 had a more elaborate apical them as primitive or early neuroblasts. As noted by
tree with longer branches that exhibited numerous Fricke (1975), it is a bit surprising that all of the
varicosities as well as occasional growth cones and stained granule neurons at this age have such sparse
filopodia. A few spines were present on the apical dendritic trees — although most granule neurons
dendrites. are likely to be only a few days old at this time, a
Based on data from Lübbers and Frotscher few neurons are born as early as embryonic day 14
(1988) and a number of other investigators, here and are already 7 days old on the day of birth.
we describe a sequence of stages that characterize On postnatal days 2 and 3, the oldest granule
the morphological development of granule neuron neurons, in the suprapyramidal blade, have one or
dendrites in the young rat. A variety of techniques more primary apical dendrites and varying num-
have been used to examine developing granule neu- bers of shorter, higher-order branches (Fig. 1A
rons during the first few weeks of life, including and B; Jones et al., 2003). Granule cells located at
Golgi impregnation (Fricke, 1975; Duffy and the very top of stratum granulosum have several
Teyler, 1978a, b; Wenzel et al., 1981; Lübbers and primary dendrites whereas those located a bit
Frotscher, 1988; Zafirov et al., 1994), intracellular deeper in the layer are more likely to have only one
injection (Liu et al., 1996, 2000; Ye et al., 2000), and thick primary dendrite emerging from the cell
retrograde labeling (Jones et al., 2003). Fricke body, as described previously for granule neurons
(1975) characterized the dendritic structures of in adult rodents (Fricke, 1975; Desmond and
Golgi-stained granule neurons in tissue from rats Levy, 1982; Green and Juraska, 1985; Claiborne et
at postnatal days 1, 2, 4, 8, 12, and 20, as well as in al., 1990). In the developing rat, the primary dend-
tissue from adult rats over the age of 60 days. Other rites of both the superficial and the deep granule
labs using Golgi-impregnated material have exam- cells tend to branch at or near the top of stratum
ined a similar range of ages. Duffy and Teyler granulosum, and the diameters of the dendrites
(1978a, b) analyzed granule neurons in sections change abruptly at branch points. The majority of
from rats at 7, 14, 30, 60, and 210 days of age, branches are still relatively smooth at this age,
Wenzel et al. (1981) determined morphologies at a with only occasional varicosities or growth cones.
170

thinner than apical dendrites but exhibit the same


immature features, including growth cones, vari-
cosities, and filopodia. Basal dendrites are con-
sidered to be an immature feature of granule
neurons in the rodent — they are not commonly
found on adult granule neurons (Seress and Po-
korny, 1981; Lübbers and Frotscher, 1988).
As development proceeds, the most mature
granule neurons exhibit more extensive dendritic
trees with longer branches and a full complement
of immature features (Fricke, 1975; Duffy and
Teyler, 1978a, b; Wenzel et al., 1981; Lübbers and
Frotscher, 1988; Jones et al., 2003). On day 4,
dendrites have larger varicosities and an increased
number of filopodia as compared to neurons in
younger animals (Jones et al., 2003). Diameter
changes are abrupt at branch points, and most
dendrites terminate before reaching the top of the
molecular layer. Basal dendrites are present on the
majority of cells. Although Jones et al. (2003) did
not report any spines on granule neuron dendrites
at day 4, Fricke (1975) noted the occasional spine
on Golgi-stained dendrites at this age.
On day 5, the oldest granule neurons are char-
acterized by numerous apical branches with many
filopodia and varicosities and several basal
branches (Fig. 2A; Wenzel et al., 1981; Lübbers
and Frotscher, 1988; Zafirov et al., 1994; Jones et
al., 2003). Abrupt diameter changes and growth
cones are present but are less numerous than on
Fig. 1. Photomontage (A) and drawings (B and C) made from day 4. Spines are scattered throughout the dendri-
serial micrographs of DiI-labeled granule neurons from a 3- tic trees of most of the older neurons on day 5,
day-old (A, B) and a 6-day-old rat (C). Note the presence of although they are found in clusters on a few dend-
immature features on dendrites at this age. Solid-curved arrow, rites on a small number of cells (Seress and Po-
growth cone; solid-straight arrow, varicosity; open-straight ar-
korny, 1981; Wenzel et al., 1981; Lübbers and
row, filopodia; long-filled arrow, abrupt diameter change; open
arrowheads, continuation of axon; a, axon; c, axon collateral. Frotscher, 1988; Zafirov et al., 1994; Jones et al.,
The montage and the drawings are shown at the same magni- 2003). Wenzel et al. (1981) characterized granule
fication. Scale bar ¼ 50 mm. (Adapted with permission from neurons between 5 and 8 days of age as interme-
Jones et al., 2003.) diate neuroblasts and considered this period to be
the second stage of granule neuron development.
The most mature granule neurons observed on
A few dendrites, however, exhibit numerous day 6 differ somewhat from those in the younger
growth cones, varicosities, and filopodia. Most in- animals (Fig. 1C; Jones et al., 2003). Growth cones
vestigators report the absence of dendritic spines and varicosities are smaller and less numerous,
at these early ages, although Wenzel et al. (1981) whereas the number of filopodia is greater (Fig.
noted ‘‘spine-resembling’’ structures at day 0 and a 2B). Abrupt diameter changes and basal dendrites
few spines at day 2. Basal dendrites are common are less frequent. Most dendrites on these neurons
on neurons at days 2 and 3. They tend to be reach the hippocampal fissure at the top of the
171

Fricke (1975) illustrated a granule neuron with nu-


merous spines from an 8-day-old rat. Wenzel et al.
(1981) noted that granule neurons in 10-day-old rats
resembled adult neurons, although they reported
the presence of growth cones and varicosities and
suggested that such features were indicative of con-
tinued dendritic growth. They considered granule
cells at this age to be in the third stage of devel-
opment and labeled them maturing or young neu-
rons. Zafirov et al. (1994) reported that some
granule neurons display extensive dendritic trees
with spines and without immature features on day
10. Similarly, Liu et al. (2000) described granule
neurons with at least some adult-like features on
day 7 and illustrated an adult-like granule cell from
a 12-day-old rat.
Thus it is now clear that a small proportion of
the oldest granule neurons exhibit adult-like mor-
phological features by the end of the first postnatal
week in the rat. These data, in combination with
results from other studies, suggest that the dentate
gyrus may be functional at this early age. For ex-
Fig. 2. Stacked series of individual images taken with a con- ample, our in vivo studies demonstrated that long-
focal laser-scanning microscope. Dendrites are from a 5-day- term potentiation (LTP) and long-term depression
old rat (A) and a 6-day-old rat (B). The dendrite shown in B is (LTD) can be elicited at medial perforant path
from one of the most mature neurons seen on day 6. Although synapses onto the granule neurons at day 7
many spines are present, immature features are still observed on
(O’Boyle et al., 2004), confirming earlier in vitro
day 6. Open arrow, filopodia; filled arrow, spines. Scale
bar ¼ 10 mm. (Adapted with permission from Jones et al., studies (Duffy and Teyler, 1978b; Trommer et al.,
2003.) 1995). It is also worth noting that the granule cell
afferents are in their approximate adult locations
at this time, and that the dendritic trees of hilar
molecular layer, and spines are present in varying interneurons are well developed (Cowan et al.,
densities and on varying numbers of dendritic 1980; Seay-Lowe and Claiborne, 1992). Further-
branches. more, the axons (or mossy fibers) of the oldest
Jones et al. (2003) observed the first adult-like granule neurons reach region CA3 of the hippo-
granule neurons on day 7 (Fig. 3). These neurons campus proper on the day of birth or soon there-
were considered to be adult-like because they ex- after (Minkwitz, 1976; Stirling and Bliss, 1978;
hibited the three primary characteristics of adult Jones et al., 2003), and granule cell stimulation can
granule neurons: they were devoid of immature induce LTD in CA3 pyramidal neurons by day 7
features except for occasional varicosities or filopo- (Battistin and Cherubini, 1994). Taken together,
dia, the vast majority of the dendrites reached the these data suggest that the traditional view of the
top of the molecular layer, and all dendrites were dentate gyrus as a ‘‘late developing’’ structure may
covered with spines. The cell bodies of the adult-like be incorrect; at least some of the granule neurons
granule neurons were located towards the top of the are capable of functioning within the hippocampal
granule cell layer in the suprapyramidal blade and circuit by the end of the first postnatal week, at
were most likely between 7 and 10 days old at this about the same time that adult-like properties are
time. Other investigators had suggested that adult- first observed in the pyramidal neurons of the hip-
like granule neurons were present at about this time. pocampus proper.
172

Fig. 3. Drawing made from serial micrographs of a DiI-labeled granule neuron from a 7-day-old rat. Note the absence of immature
features and the prevalence of spines. A varicosity is visible on the axon (short, filled arrow); open arrowhead, continuation of axon.
Scale bar ¼ 50 mm. (Adapted with permission from Jones et al., 2003.)

Although granule neurons with adult-like fea- approximately 465 mm on day 10, with no signifi-
tures are present by the end of the first week, it is cant increase between days 10 and 15.
important to note that granule neuron dendrites As granule neuron dendrites continue to elon-
continue to elongate during the second postnatal gate after day 7, spine densities and synaptic con-
week (Fricke, 1975; Duffy and Teyler, 1978a, b; tacts also increase. As noted above, spines are
Seress and Pokorny, 1981; Wenzel et al., 1981; observed on granule neurons on days 4 and 5 in
Zafirov et al., 1994). Using one of the first com- the rat, and in some cases, spines have been re-
puter-microscope systems designed for quantita- ported as early as days 2 or 3. Jones et al. (2003)
tive, three-dimensional morphological studies of considered spines to be protrusions that were less
single neurons, Fricke (1975) analyzed the dendri- than 2 mm in length (whereas longer protrusions
tic trees of developing Golgi-impregnated granule were considered to be filopodia) and noted that
neurons over the first postnatal month in the rat. dendritic spines on neonatal granule neurons
He found that the total dendritic length (defined as ranged in shape from short, stubby protrusions,
the sum of the lengths of all individual dendritic either with or without heads, to those with long,
segments) of a granule neuron increased between thin necks ending in definitive spine heads (Des-
day 8 and day 12, from an average of approxi- mond and Levy, 1985; Trommald and Hulleberg,
mately 500 mm on day 8 to an average of approx- 1997). Counts of spines on DiI-labeled dendrites
imately 1000 mm on day 12, with a wide variability located in the middle of the molecular layer re-
at both ages. Duffy and Teyler (1978a, b) also re- vealed an average of 0.40 spines/mm in 5-day-old
ported that granule neurons increased dramati- rats and 0.57 spines/mm in 6-day-old animals
cally in size between day 7 and day 14, and Zafirov (Jones et al., 2003). Densities increased to 0.81
et al. (1994) showed that total dendritic lengths spines/mm by day 7, although they were still far
increased from approximately 300 mm at day 5 to below the values reported for dendrites in the same
173

region in the adult rat (1.66 spines/mm; Desmond 40% are onto shafts. The percentage of shaft
and Levy, 1985). Zafirov et al. (1994) also dem- synapses continues to decline into adulthood, and
onstrated that spine densities increased after day 5. by day 41, only approximately 10% of synapses
These densities, however, were slightly lower than are found on dendritic shafts with the remainder
those reported by Jones et al. (2003), perhaps be- contacting dendritic spines. In addition to
cause they calculated spine densities across the en- synapses in the molecular layer, synaptic contacts
tire dendritic tree. Based on measurements of the are found on granule cell bodies in the neonatal
total dendritic length and the total number of rat. A fair proportion is symmetric and stain for
spines per cell, they found that spine densities in- glutamate-decarboxylase (Lubbers and Frotscher,
creased from day 5 through day 15, with 0.11 1988; Seress et al., 1989).
spines/mm on day 5, 0.47 spines/mm on day 10, and In summary, during the first week of life, granule
0.67 spines/mm on day 15. Duffy and Teyler neuron dendrites undergo a sequence of morpho-
(1978a, b) also reported that the number of spines logical changes. Immature features appear and then
per granule neuron more than doubled between regress, dendrites elongate, and spines and synapses
day 7 and day 14, increasing from approximately develop. By day 7, the oldest granule neurons ex-
100 spines/cell on day 7 to approximately 255 hibit adult-like characteristics immature features
spines/cell on day 14. Data from the above studies have regressed, the vast majority of dendrites reach
are substantiated by the detailed analysis of spine the top of the molecular layer, and all dendrites are
densities on dendrites of various orders over the covered with spines. During the second postnatal
course of granule neuron development and matu- week, dendrites continue to elongate and spine and
ration done by Wenzel et al. (1981). For example, synaptic densities increase dramatically. By day 14,
they reported that densities on 4th order branches a considerable number of granule neurons have at-
increased from 0.05 spines/mm on day 5 to 0.12 tained adult-like characteristics. It also appears that
spines/mm on day 8 and to 0.18 spines/mm on the large en passant boutons of the mossy fibers
day 10. have attained an adult-like shape and complexity by
Electron microscope studies demonstrate that day 14, even as the number of boutons continue to
only a few synapses are present in the molecular increase into young adulthood (Stirling and Bliss,
layer of the suprapyramidal blade on postnatal 1978; Amaral and Dent, 1981).
day 1 (0.4 synapses/100 mm2; Cowan et al., 1980).
There is, however, a dramatic increase in synapses
over the first 10 days (Cowan et al., 1980). On days Maturation of granule neuron dendritic trees
4 and 5, there are approximately 2–3 synapses/
100 mm2 and, by day 10, densities have increased to While granule neurons in 14-day-old rats qualita-
11.0 synapses/100 mm2 (Crain et al., 1973; Cowan tively resemble adult cells, quantitative studies sug-
et al., 1980). In the molecular layer of the infra- gest that their dendrites continue to change as the
pyramidal blade, synapses are present on day 5 animal matures. Early work indicated that the den-
(0.8 synapses/100 mm2) and synaptic density in- dritic tree might increase in size after the second
creases about sixfold by day 10 (6.1 synapses/ week. Fricke (1975) reported that the total dendritic
100 mm2; Cowan et al., 1980). In addition to this lengths of Golgi-stained granule neurons increased
considerable increase in synapse densities, the mo- after day 12, reaching adult values of a little over
lecular layer also increases in volume approxi- 1500 mm on day 20. Duffy and Teyler (1978a, b)
mately fourfold between days 5 and 10, resulting in also reported an increase in granule cell dendritic
approximately a 16-fold increase in absolute length between days 14 and 30. In addition, they
synapse number. It is also worth noting that a found a slight but not statistically significant de-
high percentage of synapses are found on dendritic crease between days 30 and 60 and another slight
shafts during the neonatal period (Cowan et al., increase at day 210 — the average length at 210
1981). On day 5, about 50% of synapses are onto days was approximately the same as that at day 30.
dendritic shafts whereas on day 10, approximately It is not clear from the methods, however, whether
174

the reported lengths reflect measurements of the majority of dendrites reached the top of the mo-
sum of all dendritic branch lengths or the length of lecular layer in animals of all ages, suggesting that
the longest dendrite. Duffy and Teyler (1978a, b) individual dendrites elongated as the animal ma-
state that they measured ‘‘from mid-soma to the tured. The number of dendritic segments, however,
most distal dendritic process’’, and the reported decreased from an average of 36 segments in the
values are quite similar to the width of the molec- 14- to 19-day-old rats to an average of 28 segments
ular layer (Rihn and Claiborne, 1990). Thus, these in the 50- to 60-day-old animals, with the majority
lengths may reflect the distance from the soma to of the decrease occurring by day 29. (Segments
the most distal dendritic tips. Liu et al. (2000), using were defined as a length of dendrite between an
intracellular labeling techniques, also found that origin and a branch point, between two branch
granule cells enlarged after day 14 although the in- points, or between a branch point and a distal
creases in total dendritic length were not statisti- termination point.) As a consequence of this con-
cally significant. current branch elongation and loss, the average
Whereas the studies described above suggest total dendritic length did not change significantly
that dendritic trees increase in size during the third between days 14 and 60 (3086 and 3417 mm, re-
and fourth weeks, Wenzel et al. (1981) reported spectively). A similar process of dendritic loss with
that granule cell dendritic length and the degree of no change in total dendritic length has been doc-
branching reached adult proportions at day 15. To umented for nonpyramidal neurons in the rat vis-
resolve this issue, Rihn and Claiborne (1990) used ual cortex (Parnavelas and Uylings, 1980).
intracellular labeling techniques and three-dimen- Thus, although granule cells have attained their
sional analyses to quantify the dendritic trees of adult dendritic lengths by postnatal day 14, den-
only the oldest granule neurons in the suprapy- dritic remodeling occurs between days 14 and 60
ramidal blade between days 14 and 60. The somata with the elongation of some branches and the loss
of these neurons were located in the top portion of of others. In addition, the number of spines and
the granule cell layer, between the tip and the synapses increase (Duffy and Teyler, 1978a, b;
middle of the suprapyramidal blade. Neurons Cowan et al., 1981; Wenzel et al., 1981; Zafirov et
from rats of five age groups were examined: the al., 1994). Duffy and Teyler (1978a, b) showed a
first group included rats between the ages of 14 slight increase in the number of spines per cell be-
and 19 days and the oldest group included animals tween days 14 and 30. Values remained stable be-
between 50 and 60 days. The dendritic trees from tween days 30 and 60 and then increased again
rats in each age group exhibited similar structures, between days 60 and 210. Seress and Pokorny
having between 1 and 4 primary apical dendrites (1981) noted that spine densities increased between
that were covered with spines (Fig. 4). Several days 10 and 25, and Zafirov et al. (1994) reported
qualitative differences, however, were apparent that spine densities (calculated on the basis of total
between neurons in the youngest rats and those in spines per cell divided by the total dendritic length)
the oldest animals. For example, neurons in the increased from 0.67 spines/mm at day 15 to 1.16
youngest rats had thicker dendrites in the proximal spines/mm at day 20. Wenzel et al. (1981) found
and middle thirds of the molecular layer, and they that spine densities increased dramatically on
exhibited a higher number of short, terminal seg- dendrites in almost all branch orders between
ments in the distal third of the layer. days 15 and 30. For example, densities increased
Analyses of these maturing granule neurons from approximately 0.2 to 0.59 spines/mm on 4th
demonstrated that both dendritic growth and re- order branches and from 0.14 to 0.56 spines/mm on
gression occurred between days 14 and 60, leading 6th order branches. Furthermore, they found that
to a conservation of total dendritic length during densities also continued to increase slightly after
this period (Rihn and Claiborne, 1990). The mo- day 30; when all branch orders were considered
lecular layer of the suprapyramidal blade increased together, densities increased from approximately
by approximately 50% (from an average of 205 to 0.5 to approximately 0.6 spines/mm between days
305 mm) between days 14 and 60, and the vast 30 and 180. Synaptic density in the molecular layer
175

Fig. 4. Camera lucida drawing of a granule neuron from a 53-day-old rat. Note the significant increase in size as compared to the
adult-like neuron from a 7-day-old rat shown in Fig. 3. Scale bar ¼ 50 mm.

also increased with maturation (Cowan et al., maturation. The axons or mossy fibers of the
1980). In the molecular layer of the suprapyram- granule neurons exhibit filopodial-like extensions
idal blade, densities increased from 11 synapses/ that elongate during the first two postnatal weeks,
100 mm2 on day 10 to 36 synapses/100 mm2 on day reach a peak in length on day 14, and then de-
21. Only a slight increase was seen after day 21; crease to adult lengths by day 28 (Amaral, 1979).
densities were 37 synapses/100 mm2 on day 41. The time course of their growth and retraction is
Similarly, the percentage of spine synapses in- remarkably similar to the time course of dendritic
creased considerably from day 10 to 21, but did branch growth and regression in the oldest granule
not exhibit much of a change between days 21 and neurons — whether the two processes are gov-
41. Spine synapses comprise 47% of all synapses erned by the same or similar mechanisms is not yet
on day 10, 84% on day 21, and 88% on day 41. known. In this regard, preliminary data from our
In summary, the total dendritic lengths of the lab indicate that dendritic branch loss may be
oldest granule neurons in the suprapyramidal affected by incoming neuronal activity. Specifi-
blade appear to be established by day 14. Quan- cally, we found that blockade of N-methyl-D-as-
titative changes in the dendritic tree do occur after partate (NMDA) glutamate receptors between
this time, however; some dendrites elongate while days 14 and 24 did not affect dendritic growth,
others are lost, thus leading to a conservation of but did result in a decrease in dendritic branch loss
total length. Interestingly, there is one other report in the oldest granule neurons (Blake and Clai-
of regression in the hippocampal formation during borne, 1995). Because branches continued to grow
176

and fewer branches were lost, the granule neurons hippocampal formation. They analyzed only those
in the treated animals had greater total dendritic neurons that were completely stained, had no cut
lengths than did those in the controls. It is worth dendrites in the proximal portion of the molecular
noting that the glutamate released from synapses layer, and had fewer than two-severed dendrites in
on granule cell dendrites in the middle third of the the distal two-thirds of the layer.
molecular layer binds to NMDA receptors, and Here we review the quantitative parameters of
that the afferents to this region arise from the ent- adult granule neurons reported by the above in-
orhinal cortex. The entorhinal cortex, in turn, re- vestigators and by others employing similar tech-
ceives inputs from the visual, auditory, and niques. We first discuss the available data on the
somatosensory cortices, and a number of signifi- entire population of granule neurons, and we then
cant events occur in the maturation of these sen- review the relationship between the morphology of
sory systems soon after day 14 in the rat. For a granule neuron and the location of its cell body
example, the eyes open approximately at day 15 within stratum granulosum and within the two
and adult sensitivities to sound are reached be- blades of the dentate gyrus.
tween days 16 and 18 (Tilney, 1933; Crowley and The cell body of an adult granule neuron in the
Hepp-Reymond, 1966). Therefore, it is not sur- rat is approximately 10 mm wide and approxi-
prising that granule cell remodeling begins shortly mately 19 mm long (Claiborne et al., 1990). Be-
after day 14 and may be governed by afferent in- tween 1 and 5 primary apical dendrites exit from
puts from the entorhinal cortex. the apical pole of the cell body and bifurcate rel-
atively close to the soma (Fig. 5). Fricke (1975)
noted that most branching occurs within 100 mm of
Granule neuron dendrites in the adult rat the soma and within the vicinity of the boundary
between stratum granulosum and stratum mole-
Quantitative data on the dendritic trees of adult culare. Desmond and Levy (1982) found that the
granule neurons are based on two- and three-di- majority of primary segments branch within the
mensional measurements of both Golgi-impreg- granule cell layer and first ninth of the molecular
nated and intracellularly labeled cells. Fricke layer. Furthermore, they reported that nearly all
(1975) was the first to make use of a computer- first-order branch points occur within 50 mm of the
microscope system to analyze Golgi-stained neu- cell body and within the first 30 mm of the molec-
rons in three dimensions, whereas Desmond and ular layer. Primary dendrites bifurcate into higher
Levy (1982) developed a novel probabilistic order segments, and up to eighth-order branches
method for quantifying dendritic trees of granule have been reported, with an average maximum
neurons from Golgi-stained tissue in two dimen- branch order of 5.7 (Claiborne et al., 1990). Gran-
sions. They corrected for cut dendrites at the edge ule neurons have a total of approximately 30 den-
of sections by estimating their branching and ter- dritic segments, with reported numbers ranging
mination patterns based on the patterns observed from 22 to 40 (Fricke, 1975; Seress and Pokorny,
for intact dendrites. Corrected values were com- 1981; Desmond and Levy, 1982; Green and
pared with data from tracings of six neurons (two Juraska, 1985; Claiborne et al., 1990; Rihn and
from each region) that were followed through se- Claiborne, 1990).
rial sections. Only granule neurons located in the Based on Golgi impregnations, Fricke (1975)
middle 80% of the longitudinal extent of the hip- reported an average total dendritic length of
pocampus and with a minimum of cut dendrites in 1602 mm (range from 773 to 2445 mm) for adult
the proximal third of the molecular layer were in- granule neurons, whereas Seress and Pokorny
cluded. Claiborne et al. (1990) and Rihn and Clai- (1981) reported an average of 2405 mm, and Green
borne (1990) applied the computational techniques and Juraska (1985) found total dendritic lengths
first used by Fricke (1975) to quantify the dendritic ranging from 1161 to 1279 mm. After correcting for
trees of intracellularly labeled granule neurons in cut dendrites, Desmond and Levy (1982) calcu-
relatively thick transverse slices (400 mm) of the lated an average total dendritic length of 3662 mm
177

Fig. 5. Computer-generated plots of three-dimensional reconstructions of the dendritic trees of granule cells located at various
positions along the transverse axis of the dentate gyrus. Each neuron is from a different animal, and the total dendritic length of each
dendritic tree is indicated. Note that the total dendritic lengths of the trees in the suprapyramidal blade are greater than those of the
trees in the infrapyramidal blade. CA3, field CA3 of the hippocampus proper; GL, granule cell layer. Scale bar ¼ 100 mm. (Adapted
with permission from Claiborne et al., 1990.)

for Golgi-stained neurons. Claiborne and col- terminate in the distal two-thirds of the layer than
leagues reported similar averages of 3221 and for commissural and associational afferents that
3417 mm, respectively, for intracellularly labeled make contacts in the inner third of the layer.
granule neurons analyzed in two different studies The majority of synapses onto the granule neu-
(Fig. 5; Claiborne et al., 1990; Rihn and Claiborne, ron dendrites occur on spines in the adult. As dis-
1990). Desmond and Levy (1982) reported that cussed above, a number of groups have shown that
approximately 12% of the total length was found spine and synaptic densities increase throughout
within stratum granulosum, approximately 25% granule neuron development and maturation, and
within the proximal third of the molecular layer, their reported values for densities in maturing
and the remaining portion was restricted to the and young adult rats are reviewed above (Duffy
distal two-thirds. Similarly, data from Claiborne et and Teyler, 1978a, b; Cowan et al., 1981; Seress and
al. (1990) indicated that 30% of the total length Pokorny, 1981; Wenzel et al., 1981; Zafirov et al.,
was found in the granule cell layer and proximal 1994). Other investigators have reported spine den-
third of the molecular layer, 30% in the middle sities only for granule neurons in young adult rats.
third and 40% in the distal third. These results Fricke (1975) noted that spine densities were low on
suggest that more dendritic length is available for dendrites of adult granule neurons within the first
synapses from the entorhinal afferents that 20 mm of the soma and within 50 mm of dendritic
178

terminations. He found a maximum of approxi- and within the depth of the granule cell layer. Data
mately 1.3 spines/mm on dendrites in the remaining demonstrate that the dendritic trees of neurons in
portion of the dendritic tree. Desmond and Levy the suprapyramidal blade are larger than those in
(1985) reported that spines were infrequent on the infrapyramidal blade (Fig. 5). Fricke (1975)
dendrites within stratum granulosum and that there reported that suprapyramidal neurons have
were three major types of spines on adult granule greater total dendritic lengths than do those in
cell dendrites in the molecular layer: stubby, mush- the infrapyramidal blade (1674 mm vs. 1482 mm),
room-shaped, and thin. They calculated spine den- and Desmond and Levy (1982) confirmed this re-
sities for each type in the proximal, middle, and sult for Golgi-stained granule neurons that were
distal thirds of the molecular layer of the two traced through multiple sections (3107 mm for sup-
blades. In both blades and in all three portions of rapyramidal neurons vs. 2078 mm for infrapyram-
the layer, densities were highest for the thin spines idal blade cells). Similarly, Claiborne et al. (1990)
and were lowest for mushroom-shaped protrusions. showed that suprapyramidal granule neurons have
When all spine types were included and corrections greater total dendritic lengths (3478 mm vs.
made for obscured spines, densities averaged 1.6 2793 mm), more dendritic segments (31 vs. 27),
spines/mm on dendrites of neurons in the suprapy- and wider dendritic spreads in the transverse plane
ramidal blade and 1.3 spines/mm on those in the of the hippocampal formation (347 mm vs. 288 mm)
infrapyramidal blade. These values are similar to than do the infrapyramidal blade neurons. These
those reported by Fricke (1975) but are much results demonstrate that suprapyramidal granule
greater than those reported by Wenzel et al. (1981; neurons have more dendritic length available for
see above) perhaps because different criteria were afferent contacts than do neurons in the opposite
used for the inclusion of spines (Desmond and blade. In addition, Desmond and Levy (1985) re-
Levy, 1985). ported that spine density is greater on dendrites of
In summary, the dendritic trees of granule neu- suprapyramidal neurons (1.6 spines/mm vs. 1.3
rons in young adult rats are composed of approx- spines/mm), and this result, combined with the in-
imately 30 spine-covered segments and have total creased length of the suprapyramidal neurons,
dendritic lengths of approximately 3400 mm. It is suggests that neurons in this blade receive many
not surprising that these total lengths are much more synaptic contacts than do neurons in the in-
less than those of the relatively large pyramidal frapyramidal blade.
neurons in the hippocampus proper. Reports of Overall, these data indicate that the dendritic
the total dendritic lengths of pyramidal neurons in trees of granule neurons in the suprapyramidal
region CA1 vary from an average of 10,800 to blade are likely to be larger than those in the in-
17,400 mm (Ishizuka et al., 1995; Mainen et al., frapyramidal blade and to receive more afferent
1996; Pyapali and Turner, 1996; Pyapali et al., contacts. It is of considerable interest that the axon
1998; Megias et al., 2001), whereas the total trajectories of the granule neurons also vary with
lengths for pyramidal neurons in region CA3 granule cell location. Axons of granule neurons
range from a low of 7000 mm for pyramidal neu- located toward the tip of the suprapyramidal blade
rons in region CA3c to a high of 19,800 mm for traverse stratum radiatum of region CA3 before
those in region CA3a (Ishizuka et al., 1995; Turner entering stratum lucidum, avoiding both the hilar
et al., 1995; Gonzales et al., 2001). It is also not region and the proximal part of the pyramidal cell
surprising that the granule neurons are the most layer, whereas axons of neurons in the crest and
electrotonically compact of the three major classes the infrapyramidal blade travel through the hilar
of hippocampal neurons (Carnevale et al., 1997). region and contact pyramidal neurons in the prox-
Given the gradients of granule cell neurogenesis, imal portion of field CA3 (Claiborne et al., 1986).
it is of interest that several parameters of granule It is likely that such morphological distinctions in
neuron dendrites are correlated with the location the granule neurons, in combination with physio-
of the parent cell body in the granule cell layer — logical differences (Scharfman et al., 2002; Chawla
both along the transverse axis of the dentate gyrus et al., 2005), may lead to functional differences
179

between the two blades. This idea is best exempli- Thus, a number of structural differences exist
fied by recent results showing that induction of the between the dendritic trees of neurons located in
activity-regulated, immediate early gene Arc fol- the superficial portion of stratum granulosum and
lowing behavioral stimulation occurs in the sup- those located in the deeper aspects of the layer. At
rapyramidal blade, but not in the infrapyramidal least one of the differences may have functional
blade (Chawla et al., 2005). consequences. Carnevale et al. (1997) demon-
Not only do some morphological parameters strated that the single primary dendrites of deeper
vary according to the location of the cell body neurons attenuated voltage signals spreading from
along the transverse axis of the dentate gyrus, sev- the soma out to the dendrites, thereby reducing the
eral dendritic parameters also differ according to effect of somatic events, including action poten-
the depth of the parent cell body in the granule cell tials, on molecular layer synapses. Given that
layer. Neurons with somata located at the top of adult-generated granule neurons migrate into the
the layer tend to have multiple primary branches lower portion of the granule cell layer, it will be of
whereas those with cell bodies located deeper in interest to determine whether other functional
the layer tend to have only one primary dendrite differences correlate with the depth of granule cell
(Fricke, 1975; Seress and Pokorny, 1981; Des- bodies in stratum granulosum in adult animals.
mond and Levy, 1982; Green and Juraska, 1985;
Claiborne et al., 1990). Granule neurons in the
superficial half of the granule cell layer also exhibit Summary
nearly twice the density of axosomatic synapses as
do cells in the bottom half of the layer (Lee et al., Granule neurons in the dentate gyrus of the rat
1982). In addition, Green and Juraska (1985) undergo periods of development and maturation
showed that Golgi-stained granule neurons with before reaching their final adult form and size. In
somata in the superficial third of the granule cell the rat, the primary period of neurogenesis begins
layer had greater maximum widths in the trans- during late embryogenesis and continues over the
verse plane of the dentate gyrus, more branches in first two weeks of life. The oldest granule neurons
orders 1 through 3, and fewer branches in orders 4 occupy the suprapyramidal blade whereas later
through 6. These neurons also had greater total generated cells form the infrapyramidal blade. On
dendritic lengths (1279 vs. 1161) than did deeper the day of birth, granule neurons in the suprapy-
neurons; 80% of the neurons in their sample were ramidal blade exhibit a few sparse dendrites. Ru-
located in the suprapyramidal blade. In contrast, dimentary trees appear over the next few days and
Claiborne et al. (1990) did not find a statistically by day 4, dendritic branching is quite extensive.
significant difference between the average total Immature features are abundant at this time and
lengths of neurons with somata in the superficial include abrupt diameter changes, varicosities, filo-
half of the granule cell layer and of those with podia, growth cones, and basilar dendrites. On days
somata in the bottom half of the layer in either 5 and 6, elaborate dendritic trees are present, there
blade. When only neurons in the suprapyramidal is a reduction in the frequency of immature fea-
blade were analyzed, however, they found that tures, and some spines are visible on the most ma-
superficial granule neurons had more primary ture cells. On day 7, the first few adult-like trees are
dendrites (2.4 vs. 1.5) and larger transverse spreads present on granule neurons in the suprapyramidal
(378 mm vs. 293 mm) than deeper neurons. In ad- blade. Their dendrites reach the top of the molec-
dition, 42% of the total length of superficial cells ular layer and no longer exhibit immature features
was in the distal third of the layer as compared to except for an occasional filopodium or varicosity.
37% of the length of the deeper cells. For neurons All of the branches are covered with spines. Phys-
in the infrapyramidal blade, the only significant iological recordings demonstrate that some granule
difference was that superficial cells had larger neurons are functional at this time, suggesting that
transverse spreads than did deeper neurons the dentate gyrus may be incorporated into the
(311 mm vs. 244 mm). hippocampal circuit by the end of the first week.
180

After day 7, the dendritic trees increase in size, and, Carnevale, N.T., Tsai, K.Y., Claiborne, B.J. and Brown, T.H.
by day 14, numerous adult-like granule cells are (1997) Comparative electrotonic analysis of three classes
of rat hippocampal neurons. J. Neurophysiol., 78(2):
present. The oldest granule neurons then undergo a
703–720.
process of refinement. Many dendrites continue to Chawla, M.K., Guzowski, J.F., Ramirez-Amaya, V., Lipa, P.,
elongate during this period, and other branches are Hoffman, K.L., Marriott, L.K., Worley, P.F., McNaughton,
lost; the simultaneous growth and regression results B.L. and Barnes, C.A. (2005) Sparse, environmentally selec-
in a conservation of total dendritic length after day tive expression of Arc RNA in the upper blade of the rodent
14. Spine densities, however, continue to increase. fascia dentata by brief spatial experience. Hippocampus,
15(5): 579–586.
Adult granule neurons in the rat have approxi- Claiborne, B.J., Amaral, D.G. and Cowan, W.M. (1986) A light
mately 30 dendritic segments and total dendritic and electron microscopic analysis of the mossy fibers of the
lengths of approximately 3400 mm. A variety of rat dentate gyrus. J. Comp. Neurol., 246: 435–458.
morphological features are correlated with the lo- Claiborne, B.J., Amaral, D.G. and Cowan, W.M. (1990) Quan-
titative, three-dimensional analysis of granule cell dendrites
cation of the granule cell body, both along the
in the rat dentate gyrus. J. Comp. Neurol., 302: 206–219.
transverse axis of the dentate gyrus and within the Cowan, W.M., Stanfield, B.B. and Amaral, D.G. (1981) Fur-
depth of the granule cell layer. Recent evidence ther observations on the development of the dentate gyrus.
suggests that there are functional distinctions be- In: Cowan W.M. (Ed.), Studies in Developmental Neurobi-
tween the two blades of the dentate gyrus; it will be ology: Essays in Honor of Viktor Hamburger. Oxford Uni-
of considerable interest to determine whether differ- versity Press, New York, pp. 395–435.
Cowan, W.M., Stanfield, B.B. and Kishi, K. (1980) The devel-
ences in granule cell morphologies underlie these opment of the dentate gyrus. Curr. Top. Dev. Biol., 15:
distinctions. 103–157.
Crain, B., Cotman, C., Taylor, D. and Lynch, G. (1973) A
quantitative electron microscopic study of synaptogenesis in
the dentate gyrus of the rat. Brain Res., 63: 195–204.
Acknowledgment Crowley, D.E. and Hepp-Reymond, M.C. (1966) Development
of cochlear function in the ear of the infant rat. J. Comp.
This work was supported by NIH grant #GM08194 Physiol., 62: 427–432.
to BJC. Desmond, N.L. and Levy, W.B. (1982) A quantitative anatom-
ical study of the granule cell dendritic fields of the rat dentate
gyrus using a novel probabilistic method. J. Comp. Neurol.,
212: 131–145.
References Desmond, N.L. and Levy, W.B. (1985) Granule cell dendritic
spine density in the rat hippocampus varies with spine shape
Altman, J. and Das, G.D. (1965) Autoradiographic and histo- and location. Neurosci. Lett., 54: 219–224.
logical evidence of postnatal hippocampal neurogenesis in Duffy, C.J. and Teyler, T.J. (1978a) Development of habitua-
rats. J. Comp. Neurol., 124(3): 319–335. tion in the dentate gyrus of rat: physiology and anatomy.
Amaral, D.G. (1979) Synaptic extensions from the mossy fibers Brain Res. Bull., 3: 305–310.
of the fascia dentata. Anat. Embryol. (Berl.), 155: 241–251. Duffy, C.J. and Teyler, T.J. (1978b) Development of potent-
Amaral, D.G. and Dent, J.A. (1981) Development of the mossy iation in the dentate gyrus of rat: physiology and anatomy.
fibers of the dentate gyrus: I. A light and electron microscopic Brain Res. Bull., 3: 425–430.
study of the mossy fibers and their expansions. J. Comp. Fricke, R.A. (1975) Studies on the morphology and develop-
Neurol., 195: 51–86. ment of the hippocampus and dentate gyrus. Ph.D. Disser-
Battistin, T. and Cherubini, E. (1994) Developmental shift from tation, Washington University, St. Louis, MO.
long-term depression to long-term potentiation at the mossy Gonzales, R.B., DeLeon Galvan, C.J., Rangel, Y.M. and Clai-
fibre synapses in the rat hippocampus. Eur. J. Neurosci., borne, B.J. (2001) Distribution of thorny excrescences on
6(11): 1750–1755. CA3 pyramidal neurons in the rat hippocampus. J. Comp.
Bayer, S.A. and Altman, J. (1974) Hippocampal development Neurol., 430(3): 357–368.
in the rat: cytogenesis and morphogenesis examined with Green, E.J. and Juraska, J.M. (1985) The dendritic morphology
autoradiography and low-level X-irradiation. J. Comp. of hippocampal dentate granule cells varies with their posi-
Neurol., 158(1): 55–79. tion in the granule cell layer: a quantitative Golgi study. Exp.
Blake, N.M.J. and Claiborne, B.J. (1995) Possible role of the N- Brain Res., 59: 582–586.
methyl-D-aspartate (NMDA) receptor in the remodeling of Ishizuka, N., Cowan, W.M. and Amaral, D.G. (1995) A quan-
dendritic arbors during development. Soc. Neurosci. Abstr., titative analysis of the dendritic organization of pyramidal
21: 1292. cells in the rat hippocampus. J. Comp. Neurol., 362: 17–45.
181

Jones, S.P., Rahimi, O., O’Boyle, M.P., Diaz, D.L. and Clai- Scharfman, H.E., Sollas, A.L., Smith, K.L., Jackson, M.B. and
borne, B.J. (2003) Maturation of granule cell dendrites after Goodman, J.H. (2002) Structural and functional asymmetry
mossy fiber arrival in hippocampal field CA3. Hippocampus, in the normal and epileptic rat dentate gyrus. J. Comp. Ne-
13: 413–427. urol., 454: 424–439.
Kaplan, M.S. and Hinds, J.W. (1977) Neurogenesis in the adult Schlessinger, A.R., Cowan, W.M. and Gottlieb, D.I. (1975) An
rat: electron microscopic analysis of light radioautographs. autoradiographic study of the time of origin and the pattern
Science, 197: 1092–1094. of granule cell migration in the dentate gyrus of the rat. J.
Lee, K.S., Gerbrandt, L. and Lynch, G. (1982) Axo-somatic Comp. Neurol., 159(2): 149–175.
synapses in the normal and x-irradiated dentate gyrus: fac- Seay-Lowe, S.L. and Claiborne, B.J. (1992) Morphology of in-
tors affecting the density of afferent innervation. Brain Res., tracellularly labeled interneurons in the dentate gyrus of the
249: 51–56. immature rat. J. Comp. Neurol., 324: 23–36.
Liu, X., Tilwalli, S., Ye, G., Lio, P.A., Pasternak, J.F. and Seress, L., Frotscher, M. and Ribak, C.E. (1989) Local circuit
Trommer, B.L. (2000) Morphologic and electrophysiologic neurons in both the dentate gyrus and Ammon’s horn es-
maturation in developing dentate gyrus granule cells. Brain tablish synaptic connections with principal neurons in five
Res., 856: 202–212. day old rats: a morphological basis for inhibition in early
Liu, Y.B., Lio, P.A., Pasternak, J.F. and Trommer, B.L. (1996) development. Exp. Brain Res., 78: 1–9.
Developmental changes in membrane properties and postsy- Seress, L. and Pokorny, J. (1981) Structure of the granular layer
naptic currents of granule cells in rat dentate gyrus. J. of the rat dentate gyrus. A light microscopic and Golgi study.
Neurophysiol., 76(2): 1074–1088. J. Anat., 133(2): 181–195.
Loy, R., Lynch, G. and Cotman, C.W. (1977) Development of Stirling, R.V. and Bliss, T.V. (1978) Hippocampal mossy fiber
afferent lamination in the fascia dentata of the rat. Brain development at the ultrastructural level. Prog. Brain Res., 48:
Res., 121: 229–243. 191–198.
Lübbers, K. and Frotscher, M. (1988) Differentiation of gran- Tilney, F. (1933) Behavior in its relation to the development of
ule cells in relation to GABAergic neurons in the rat fascia the brain. II. Correlation between the development of the
dentata. Combined Golgi/EM and immunocytochemical brain and behavior in the albino rat from embryonic states to
studies. Anat. Embryol. (Berl.), 178: 119–127. maturity. Bull. Neurol. Inst. N.Y., 3: 252–358.
Mainen, Z.F., Carnevale, N.T., Zador, A.M., Claiborne, B.J. Trommald, M. and Hulleberg, G. (1997) Dimensions and den-
and Brown, T.H. (1996) Electrotonic architecture of hippo- sity of dendritic spines from rat dentate granule cells based
campal CA1 pyramidal neurons based on three-dimensional on reconstructions from serial electron micrographs.
reconstructions. J. Neurophysiol., 76(3): 1904–1923. J. Comp. Neurol., 377(1): 15–28.
Megias, M., Emri, Z., Freund, T.F. and Gulyas, A.I. (2001) Trommer, B.L., Kennelly, J.J., Colley, P.A., Overstreet, L.S.,
Total number and distribution of inhibitory and excitatory Slater, N.T. and Pasternak, J.F. (1995) AP5 blocks LTP in
synapses on hippocampal CA1 pyramidal cells. Neurosci- developing rat dentate gyrus and unmasks LTD. Exp.
ence, 102: 527–540. Neurol., 131: 83–92.
Minkwitz, H.G. (1976) Zur Entwicklung der Neuronenstruktur Turner, D.A., Li, X.G., Pyapali, G.K., Ylinen, A. and Buzsaki,
des Hippocampus während der pär- und postnatalen Onto- G. (1995) Morphometric and electrical properties of recon-
genese der Albinoratte. J. Hirnforsch., 17: 213–231. structed hippocampal CA3 neurons recorded in vivo.
O’Boyle, M.P., Do, V., Derrick, B.E. and Claiborne, B.J. (2004) J. Comp. Neurol., 356: 580–594.
In vivo recordings of long-term potentiation and long-term Wenzel, J., Stender, G. and Duwe, G. (1981) Zur Entwicklung
depression in the dentate gyrus of the neonatal rat. der Neuronenstruktur der Fascia dentata bei der Ratte. Ne-
J. Neurophysiol., 91: 613–622. urohistologish-morphometrische, ultrastrukturelle and ex-
Parnavelas, J.G. and Uylings, H.B. (1980) The growth of non- perimentelle Untersuchungen. J. Hirnforsch., 22: 629–683.
pyramidal neurons in the visual cortex of the rat: a morpho- Ye, G.L., Liu, X.S., Pasternak, J.F. and Trommer, B.L. (2000)
metric study. Brain Res., 193(2): 373–382. Maturation of glutamatergic neurotransmission in dentate
Pyapali, G.K., Sik, A., Penttonen, M., Buzsaki, G. and Turner, gyrus granule cells. Brain Res. Dev. Brain Res., 124:
D.A. (1998) Dendritic properties of hippocampal CA1 py- 33–42.
ramidal neurons in the rat: intracellular staining in vivo and Zafirov, S., Heimrich, B. and Frotscher, M. (1994) Dendritic
in vitro. J. Comp. Neurol., 391: 335–352. development of dentate granule cells in the absence of
Pyapali, G.K. and Turner, D.A. (1996) Increased dendritic their specific extrinsic afferents. J. Comp. Neurol., 345:
extent in hippocampal CA1 neurons from aged F344 rats. 472–480.
Neurobiol. Aging, 17(4): 601–611. Zhao, C., Teng, E.M., Summers Jr., R.G., Ming, G.L. and
Rihn, L.L. and Claiborne, B.J. (1990) Dendritic growth and Gage, F.H. (2006) Distinct morphological stages of dentate
regression in rat dentate granule cells during late postnatal granule neuron maturation in the adult mouse hippocampus.
development. Brain Res. Dev. Brain Res., 54: 115–124. J. Neurosci., 26(1): 3–11.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 11

Physiological studies of human dentate granule cells

Anne Williamson1, and Peter R. Patrylo2

1
Department of Neurosurgery, Yale University School of Medicine, New Haven, CT 06518, USA
2
Department of Physiology, Southern Illinois University, Carbondale, IL 62901, USA

Abstract: The availability of human hippocampi obtained through surgery (usually for treatment of tem-
poral lobe epilepsy) has allowed us to investigate the properties of the human dentate in a way that cannot
be done with other brain regions. The dentate has been the primary focus of these studies because of its
relative preservation in all patient specimens. Moreover, there is extensive synaptic reorganization of
numerous neurotransmitter systems in this the fascia dentate (dentate gyrus and the hilus) in humans with
specific forms of TLE. These changes are not evident in tissue from patients with seizure that begin outside
the hippocampus, and, as a result, this tissue provides an invaluable resource for comparisons. Physio-
logical data using both slices and acutely dissociated cells demonstrate that the granule cells have mem-
brane properties similar to those of rodents although there are specific changes that appear to be associated
with seizures. Similarly, in the non-sclerotic hippocampi, the synaptic properties are similar to those
reported in rodents. There are also a number of parallels between the findings in humans and in status
animal models of temporal lobe epilepsy. This review will cover analyses of membrane properties as well as
of glutamatergic, GABAergic, and neuromodulatory systems. Thus, while there are a number of issues that
invariably arise with studies of pathological human tissue, this tissue is ideally suited to verify and refine
animal models of temporal lobe epilepsy. In addition, one can argue that human tissue provides the only
resource to evaluate the ways that granule cells recorded from laboratory animals approximate human
granule cell physiology.

Keywords: temporal lobe epilepsy; hippocampal slice; mossy fiber sprouting; neuropeptides; inhibition

Introduction the work of Schwartzkroin and Prince (1976), and


extending to the present day. The majority of these
We have a greater understanding of the physio- studies have used physiological methods, using
logical properties of the dentate gyrus than any both slices and acutely dissociated cells. More re-
other region of the human brain because of the cent studies have begun to use molecular biological
availability of tissue resected from patients with and imaging approaches to expand our under-
pharmacoresistant temporal lobe epilepsy (TLE). standing of the human dentate gyrus. This review
Numerous groups have studied resected human will focus on the physiology of the human dentate
hippocampi over the past 20 years, starting with gyrus with special attention to the changes ob-
served in patients with the most common pathol-
ogy associated with temporal lobe epilepsy, medial
Corresponding author. Tel.: +1 203 785 5327; temporal lobe sclerosis (MTS) (de Lanerolle et al.,
Fax: +1 203 737 -2159; E-mail: anne.williamson@yale.edu 2003).

DOI: 10.1016/S0079-6123(07)63011-8 183


184

While there are a number of issues associated loss and no evidence for synaptic reorganization.
with studies of human material, we argue that use See Cohen-Gadol et al. (2005) and de Lanerolle
of human tissue is useful because it provides vir- et al. (2003) for detailed descriptions of this pa-
tually the only resource for comparison to rodent thology. Figure 1A shows quantitative cell counts
tissue. In addition, tissue from patients with for autopsy controls, MTS and PTLE material in
intractable TLE provides a critical foundation several hippocampal regions. Note that there is no
for the development of accurate animal models of significant difference between the cell counts for
TLE. the PTLE group and the non-pathological autopsy
The major issues to consider are variability of controls in all four regions sampled.
the patient population and the lack of true con- A second comparison population includes
trols. Other, more minor concerns, include varia- hippocampi from patients with extrahippocampal
bility in the surgical resection of the tissue and lesions. In these cases, the hippocampus is re-
hence tissue viability, and the fact that the area of moved to limit seizure recurrence. This tissue is
the hippocampus that is resected may be limited. referred to as Mass-Associated TLE (MaTLE).
The vast majority of hippocampi that are avail- The epileptogenic lesions include gliomas (usually
able for physiological study are resected for treat- oligodendrogliomas or low grade astrocytomas),
ment of medically intractable seizures and thus are cavernous malformations, and cysts (Beaumont
not from normal brain. However, within the pop- and Whittle, 2000). We have noted no physiolog-
ulation of TLE patients, there are distinct groups ical differences in tissue from these two non-
that can provide useful comparisons. While rap- sclerotic populations and therefore will group
idly obtained autopsy material can serve as a non- them. For clarity, we will refer to these as com-
pathological control for anatomical studies, true parison hippocampi. Similarly, granule cells from
control material for physiological studies is ex- MTS or comparison hippocampi will be referred
tremely rare. The bulk of available tissue has the to as MTS or comparison granule cells.
pathological features of MTS which include pro-
found neuronal loss with the trend being CA1>
hilusCA3>CA2DG. In severely sclerotic hippo- Membrane properties of human dentate granule
campi, up to 80% of neurons are lost; in a rep- cells
resentative series of patients, the mean neuronal
loss (for all hippocampal regions) was 64% nor- There are several published reports, which describe
malized to autopsy controls (Kim et al., 1990; the membrane properties of human dentate gran-
de Lanerolle et al., 2003). There is a concomitant ule cells relative to rodents, conducted in slices.
reactive gliosis in MTS hippocampi; however, the Based on routinely obtained physiological meas-
role of reactive astrocytes in epileptic tissue is only ures, human granule cells can be successfully
now being investigated (Binder and Steinhauser, recorded in slices and are comparable in their rest-
2006). ing membrane potentials and input resistances to
The best available comparison material for typ- those of rodents (Isokawa et al., 1991; Williamson
ical MTS patients is those patients with no evi- et al., 1993, 1999; Dietrich et al., 1999a). Thus,
dence of a lesion and where there is no atrophy any differences between rodent and human tissue
detectable on MRI. There is no consensus on the cannot be attributed solely to slice ‘‘health.’’
terminology for this patient group. For this dis- However, there are differences in the firing
cussion, these patients will be referred to as par- properties of these cells that may be a feature of
adoxical TLE (PTLE). This group of patients human tissue. Williamson et al. (1993) showed
provides the most reasonable comparison material that, using sharp electrode recordings, there was
for the MTS tissue for several reasons. First, the reduced spike frequency accommodation between
confounds of previous seizure experience and ex- guinea pig and human granule cells from non-
posure to anticonvulsants can be nullified with this sclerotic hippocampi. No other changes were
group. Second, there is little region-specific cell noted between the rodent and comparison tissue.
185

Fig. 1. Characteristics of epileptic human dentate. A compares the cell loss in different pathologies between the dentate granule cells,
CA fields and the hilus. Note that the dentate is less affected than the CA1 region in the MTS tissue, and that there is little difference
between the autopsy controls and the comparison material. Data courtesy of Dr. Jung H. Kim. B shows the presence of graded
bursting activity in response to molecular layer stimulation (stimulus strength in mA). The arrows show examples of spontaneous
excitatory activity, which is another common physiological finding in MTS tissue that is rare in comparison material. The resting
membrane potential in the comparison granule cell was 58 mV and 55 in the MTS tissue. Spontaneous events are likely to be
glutamatergic, not GABAergic, because the RMP is depolarized to ECl.
186

These data suggest that there may be some differ- determine if these differences in Na+ channel
ences in intracellular calcium dynamics between properties reflect treatment with anticonvulsants,
the two species, but could also reflect changes due many of which act on Na+ channels. The changes
to anticonvulsant drug use. in the Na+ channel transcripts that have been
A similar study was carried out by Dietrich et al. reported in human granule cells may account for
(1999) using MTS tissue. In this study, two pop- some of these physiological changes (Lombardo
ulations of granule cells were described that could et al., 1996).
be discriminated by the characteristics of the fast
afterhyperpolarization (AHP) and the loss of a
medium duration AHP. There were predictable K+ currents
differences in spike frequency adaptation between
the two groups of cells that correlated with the With few exceptions, there are few differences
extent of the AHPs. The majority of the cells that between the K+ channels in human granule cells
had shorter AHPs also demonstrated synaptic and those in rats. These currents could be sepa-
hyperexcitability (see below); however, 21% had rated into at least four types by their kinetic and
physiological features comparable to rodents pharmacological properties. These include at least
(and primates (St John et al., 1997)) indicating one voltage-dependent current similar to those
that some of the granule cells in sclerotic hippo- observed in mammalian hippocampal neurons,
campi retain normal membrane and synaptic and two Ca2+-dependent K+ currents that
properties. most probably correspond to SK- and BK-type
Other biophysical properties of human hippo- currents. The sole significant difference was that a
campal neurons have been studied using classical A-type current could be detected in some
acutely dissociated cells. This preparation is ad- patients with MTS but not in comparison material
vantageous because it allows high quality voltage (Beck et al., 1996).
clamp recordings; however, the possible effect of
the enzyme treatment on extracellular proteins is Ca2+ currents
a potential limitation. Using this approach pro-
vides the best means to address the properties of Beck et al. (1998, 1999) and Nagerl et al. (2000)
Na+, Ca2+, and K+ channels in human granule demonstrated that granule cells express N-, L-,
cells. and T-type Ca2+ channels based on their phar-
macological sensitivity. Similar to studies in
epileptic rats, the only significant change in these
Na+ currents properties was an increase in the T-type Ca
channel density. This increased density of a low
Several studies have examined the properties of threshold Ca2+ current may lead to greater Ca2+
Na+ currents in acutely isolated granule cells from entry during synaptic activity, and thus modulate a
epileptic patients. It is notable that there were very variety of systems, including GABAA receptor-
large currents in these cells compared to cells from mediated inhibition (Isokawa, 1998).
rat tissue or from non-focal neocortical tissue. The
other noteable feature of the Na+ currents was
that there was a component of the current that had HCN currents
a slow recovery from inactivation (Reckziegel
et al., 1998). In a subsequent study of granule Finally, the ability to generate rhythmic activity in
cells from epileptic patients, it was shown that neurons can be regulated by the hyperpolariza-
there was a limited effect of the anticonvulsant tion-activated cyclic nucleotide gated current
carbemazepine on these channels (Reckziegel (HCN). While a detailed analysis of this current
et al., 1999). Comparable studies have not been has not been performed in human tissue, there is
conducted in non-epileptic cells, so it is difficult to evidence for enhanced HCN mRNA in human
187

dentate granule cells from sclerotic hippocampi. Examples of this graded bursting activity are
This finding parallels that seen in the rat (Bender shown in Figs. 1B and 2A.
et al., 2003). The consequences of excess HCN The evoked bursting activity in slices is depend-
could lead to granule cells with physiological ent on NMDA receptor activation (Isokawa and
properties closer to CA1 pyramidal cells, Levesque, 1991; Franck et al., 1995; Cohen-Gadol
i.e., more depolarized membrane potentials, and et al., 2005). In our studies, we noted a 53.6% and
cells that are more likely to have rebound activity 47.5% decrease in the area of the evoked response
following recovery from a hyperpolarization. in MTS granule cells following application of the
Moreover, in conjunction with the enhanced T NMDA receptor antagonists amino phosphono-
channel density described above, granule cells valeric acid (APV; 30 mM) or Zn2+ (200 mM), re-
from MTS hippocampi will be more prone to spectively. In contrast, there was only a 6.4%
oscillatory activity, analogous to thalamic relay change after APV and 18.5% decrease following
cells. While prominent oscillations have not been Zn2+in the comparison tissue (Williamson and
reported in human granule cells, subtle changes Spencer, 1995). These data are consistent with
in ion channels may allow for enhanced neural voltage clamp recordings in MTS slices (Isokawa
synchronization. and Levesque, 1991) showing an APV-sensitive
component of the evoked EPSC. In addition,
Synaptic properties studies using acutely dissociated granule cells
described prolonged NMDA channel openings in
Excitatory synaptic activity cells from MTS patients. In this study it was
noteable that calcineurin did not modulate the
A consistent observation in granule cells from open channel conductance in these cells, similar
patients with MTS is that bursts, i.e., a depolar- to kindled rat granule cells (Lieberman and
izing envelope with three or more action potentials Mody, 1999). This may be related to the loss of
riding on it, can be evoked by stimulation in the calbindin D28 K in these cells (Magloczky et al.,
outer molecular layer. This stimulus will primarily 1997).
activate the perforant path, but will also activate
any recurrent mossy fibers. It is difficult to selec-
tively activate the medial vs. lateral perforant path Role of mossy fiber sprouting in granule cell
in the human, so we refer below to the site of excitability
stimulation, rather than a specific fiber branch of
the perforant pathway. The percentage of granule Robust recurrent mossy fiber sprouting into the
cells that display the stimulus-induced bursts inner third of the molecular layer in MTS hippo-
varies widely between patients, as well as in pub- campi was described by a number of authors in the
lished reports (Franck et al., 1995; Williamson early 1980s using a variety of techniques, including
et al., 1995a; Cohen-Gadol et al., 2005). In our Timm stain and dynorphin immunoreactivity
sharp electrode recordings from 190 cells (75 (de Lanerolle et al., 1989; Sutula et al., 1989;
patients), we noted bursting in 70/102 (68.6%) Houser et al., 1990). This is not seen in the com-
cells from 41 patients with MTS. In contrast, this parison tissue or in autopsy material. However,
type of activity was occurred in only 15/72 (20.5%) defining the role of the sprouted mossy fibers
cells recorded from comparison hippocampi. in epileptogenesis continues to remain controver-
However, it is important to note that we never sial (e.g., Patrylo and Dudek, 1998; Wuarin
saw greater than a three action potential burst and Dudek, 2001; Harvey and Sloviter, 2005;
in the comparison material while four or more Frotscher et al., 2006; Sloviter et al., 2006).
were routinely observed in the MTS tissue. Thus, We have used a variety of approaches to inves-
in a large sample, bursting is a characteristic of tigate the possible effect of mossy fiber sprouting
hippocampi from patients with MTS, but is not on human granule cell excitability. As in rat mod-
exclusively found in this group of patients. els of TLE with mossy fiber sprouting, robust
188

Fig. 2. Granule cell hyperexcitability is associated with mossy fiber sprouting. A shows representative traces comparing the response to
molecular layer (orthodromic) and hilar (antidromic) stimulation in a patient with MTS. Note that there is a greater response to
maximal hilar stimulation than to stimulation of outer molecular layer. Moreover, note that hilar stimulation produced a synaptic
response prior to an antidromic spike. These data support the hypothesis that mossy fibers are primarily excitatory in resected tissue. B
shows Timm staining from the region of the dentate where the cell shown in A was recorded. Note the robust supergranular labeling
indicating the presence of sprouted fibers. (C1) Stimulation protocols have the problem of activating multiple fiber systems. Therefore,
in the presence of 30 mM bicuculline, we applied a bolus of glutamate to the granule cell layer at a distance from the recorded cell and
were able to evoke a long lasting barrage of EPSPs. The glutamate was applied at a distance from the recorded cell and thus we were
not observing direct glutamate effects. C2 shows that a similar response was not observed when the glutamate was applied to the hilus.
(See Color Plate 11.2 in color plate section.)

synaptic events can be evoked from granule which then form functional excitatory contacts
cells with hilar (antidromic) stimulation in MTS onto granule cells (Masukawa et al., 1992). An
hippocampi, and this is likely to represent the an- example comparing responses to orthodromic and
tidromic activation of recurrent mossy fibers, antidromic stimulation in the same cell, recorded
189

from sclerotic hippocampus, is shown in Fig. 2A. animal model of MTS (Ribak et al., 2000; Dash-
This hippocampus was characterized by robust tipour et al., 2002). However, the data shown in
mossy fiber sprouting, as shown in Fig. 2B. Note Fig. 2C suggest that activation of recurrent mossy
that progressively longer and more complex events fibers is more effective in synchronizing granule
occurred following the antidromic spike, consist- cell activity than activating granule cell basal
ent with the hypothesis that there is a complex dendrites.
network of recurrent fibers that can easily activate An additional clue to the role of mossy fiber
a given granule cell. sprouting in generating hyperexcitability is pro-
A second approach was glutamate application, vided by a distinct group of patients that have cell
using microdrops, to the granule cells located at a loss predominantly in CA1 with less damage in the
distance from the recorded neuron. Presumably dentate gyrus, hilus and CA3, and there is little to
the granule cells would be interconnected by no mossy fiber sprouting (as determined by dynor-
recurrent circuits in the MTS tissue, and the phin immunohistochemistry; (de Lanerolle et al.,
GABAA receptor antagonist bicuculline (30 mM) 1997, 2003). In this group, the ability to generate
was added to ensure that inhibition would bursts with perforant path stimulation was com-
not mask the effects of recurrent excitatory parable to that seen in typical MTS patients; how-
circuits. As shown in Fig. 2C, the microdrop ever, there was significantly less asynchronous
application was followed by a transient increase in synaptic activity and fewer spontaneous baseline
the frequency of spontaneous activity (presumably EPSPs (de Lanerolle et al., 1997). Therefore, only
EPSPs). This type of response was not observed specific aspects of the observed hyperexcitability
when a glutamate was applied to the hilus. can be attributed to robust sprouting. We cannot
These data demonstrate that there are robust rule out a possible contribution of minimal sprout-
polysynaptic connections made by the recurrent ing that was not detected by the immunostaining,
mossy fibers onto granule cells in MTS hippo- however.
campi. In addition to sprouting, a common feature of
We were unable to evoke synaptic activity with the MTS dentate is granule cell dispersion (Hous-
hilar stimulation or with glutamate microdrops er, 1990). The mechanisms for this dispersion have
applied to the granule cell layer in the dentate of included deficiencies in Reelin (Haas et al., 2002;
comparison hippocampi, supporting the hypothe- Heinrich et al., 2006) leading to a broad cell body
sis that this activity is mediated by synaptic reor- layer. However, it appears that the wider granule
ganization. Moreover, they also indicate that the cell layer is linked with MTS, because the width
excitatory mossy cells of the hilus are not able to of the granule cell layers are correlated with the
induce this type of graded excitation, presumably degree of sprouting. However, total cell loss is
because of the longitudinal orientation of their the greatest predictor of the density of sprouting
axonal arbors. However, little is known about (El Bahh et al., 1999).
mossy cells in the human dentate. The enhanced width of the granule cell layer in
An important caveat to the results described MTS hippocampi may reflect an increase in dent-
above is that effects of hilar stimulation, electrical ate granule cell neurogenesis, which occurs after
or chemical, could have been due to activation of seizures in laboratory animals, especially animal
basal dendrites of granule cells. Primates are models of TLE (Scharfman, 2004). Neural pro-
known to have basal dendrites (see chapter by L. genitors have been described in the human dentate
Seress in this volume). Approximately 40% of hu- gyrus (Roy et al., 2000) and thus it is likely that
man granule cells from epileptic tissue have been different populations of granule cells are found in
reported to have basal dendrites, and they extend the human hippocampus that are of different ages.
into the hilus, where they could be activated by How the physiological differences in cells, related
either electrical or chemical stimulation (von to their age and degree of incorporation into the
Campe et al., 1997). Basal dendrites do not exist hippocampal circuitry, has not been addressed in
normally in the rat, and are greatly increased in human material.
190

GABAergic inhibition reduced to the same degree, suggesting that a


decreased afferent input onto the remaining inter-
The possible role of altered inhibition in epileptic neurons does not explain the decreased inhibitory
tissue has been extensively discussed in the liter- efficacy. The decreased evoked responses may be
ature (Bernard et al., 1998, 2000). A significant to due the loss the SS and NPY immunoreactive
problem in studying GABAergic inhibition in the interneurons that target the dendritic arbor rather
dentate gyrus is the enormous complexity of the than the cell body, because there was no change
interneuronal systems found within the hilus and in the density of parvalbumin-immunoreactive
in the dentate gyrus, with different populations of terminals at the cell body of granule cells in
interneurons subserving different functions and tissue removed from patients with MTS compared
contacting distinct regions of the granule cells to controls (Seress et al., 1993).
(Freund and Buzsaki, 1996). Finally, the decrease in evoked inhibition is not
There are numerous changes in interneuronal due to changes in GABA receptor density or re-
networks in the MTS hippocampus compared to sponsiveness. The studies by Shumate et al. (1998)
either comparison tissue or autopsy controls (de and Gibbs et al. (1996) clearly show that acutely
Lanerolle et al., 1989, 1992; Magloczky et al., dissociated cells have an enhanced response to ap-
2000; Wittner et al., 2002). The primary changes plied GABA and that the pharmacology of the
include a loss of hilar interneurons that co-localize GABA responses demonstrate an increased sensi-
specific peptides and GABA. Those cells immuno- tivity to Zn2+, suggesting changes in the g subunit
reactive for neuropeptide Y (NPY) and somatosta- composition of these receptors. These pharmaco-
tin (SST) appear to be especially vulnerable logical experiments are consistent with the mRNA
(de Lanerolle et al., 1992; Mathern et al., 1995), studies from single granule cells showing a signifi-
while cells immunoreactive for parvalbumin and cant increase in the GABA a5 receptor subunit
calbindin are preserved (Magloczky et al., 2000). (Brooks-Kayal et al., 1999). Later studies revealed
However, despite these findings, there is no sig- that there is significant diversity in the patterns of
nificant loss of immunoreactivity for glutamic acid GABAA receptors expressed by different granule
decarboxylase, the GABA synthetic enzyme (Babb cells in pediatric patients with intractable TLE
et al., 1989; Mathern et al., 1995). Adding to (Porter et al., 2005). Thus, the specific character-
the complexity of describing the changes seen in istics of GABA receptor-mediated responses may
inhibitory systems are the alterations in fiber path- be quite heterogeneous between patients.
ways. This includes sprouting of NPY, SS, and SP Taken together, these data indicate that, like the
systems (de Lanerolle et al., 1992; Mathern et al., changes in excitatory systems, there is a complex
1995). Many of these changes were recently interplay between changes in pre- and postsynaptic
reviewed by Magloczky and Freund (2005). elements. The decrease in evoked (polysynaptic
Physiologically, several groups have reported a and monosynaptic) IPSPs could reflect presynap-
decrease in evoked GABAA receptor-mediated in- tic modulation of GABA release, or changes in
hibition in MTS hippocampi relative to compar- the amount of GABA available for release despite
ison material (Isokawa, 1998; Williamson et al., increases in postsynaptic receptor density.
1999; Fig. 3). The degree of reduction is similar to
that seen in the pilocarpine rat model (Kobayashi
and Buckmaster, 2003). As with the membrane Synaptic plasticity
properties, no significant differences in polysynap-
tically IPSP/C conductances have been reported in Several forms of synaptic plasticity are expressed
parallel studies between control rats and the in the dentate gyrus and have been intensively
comparison material in humans (Isokawa, 1996; studied in rodents. These include long-term po-
Williamson et al., 1999). It was interesting that tentiation (LTP), induced by brief periods of high
both polysynaptically and monosynaptically frequency stimulation, and long-term depression
evoked GABAA receptor-mediated events were (LTD), induced by low frequency stimuli. While
191

Fig. 3. Decreased poly- and monosynaptic GABAA receptor-mediated inhibition in MTS hippocampi. We noted significant decreases
in molecular layer-evoked inhibitory responses. A shows events delivered in normal medium; the stimulus intensity was set at twice that
needed to evoke a single action potential. The d in A and B indicate the points at which the fast and slow IPSPs were measured. B
shows monosynaptic responses evoked in presence of APV and CNQX (30 mM each). The stimulus intensity was set to the greatest that
did not evoke an antidromic action potential. C and D shows grouped IPSP conductance data for polysynaptically and monosy-
naptically-evoked IPSPs, respectively. Note the profound and significant decrease in the IPSP conductance in the MTS tissue relative
to comparison material. po0.05; po0.001.

the link between learning and memory and these (SRT) (Bell et al., 2005). For these studies,
forms of synaptic plasticity remains controversial, 200 Hz stimuli were delivered in a theta burst
the availability of human material provides an ad- pattern and the change in the response was meas-
ditional avenue to examine this issue, because ured 30 min following the tetani. This value was
memory deficits are a common clinical feature of then correlated with each patient’s SRT z score on
MTS (Helmstaedter et al., 2003). the selective reminding task, a measure of laterali-
Both LTP and LTD can be reliably obtained in zed hippocampal function. Taken together, these
slices from epileptic human temporal neocortex data support the hypothesis that altered synaptic
with pharmacology comparable to that seen in plasticity is associated with the cognitive dysfunc-
rodents (Chen et al., 1996). However, both forms tion in MTS patients.
of synaptic plasticity are impaired in patients with
MTS relative to patients without this pathology
(Beck et al., 2000). Moreover, as shown in Fig. 4, Neurotransmitter transport
there is a direct correlation between the degree of
LTP impairment and a measure of neuropsycho- Another aspect of neurotransmission that has been
logical function, the selective reminding task investigated in the human dentate is GABA
192

Fig. 4. Altered synaptic plasticity in the hippocampi of patients with MTS. A shows examples of field PSPs recorded in the molecular
layer of a MTS patient. Note that there was pronounced depression of the evoked response following five trains of 100 Hz stimuli (100
stimuli/train delivered at 30 s intervals). B1 shows the group data for seven patients and B2 shows the cumulative probability plot
demonstrating that depression was the predominant response. C shows that those patients that did not express LTD with this
stimulation protocol had normal SRT z scores (control is set to 0) while those that did show LTD had markedly lower SRT scores.

uptake, commonly assessed using focal GABA There are a number of possible consequences of
application (Williamson et al., 1995b; Patrylo altered GABA transport, which could affect the
et al., 2001). These studies showed prolonged overall excitability of the dentate. For example,
responses to GABA in the MTS dentate, which granule cells with the GABAA d receptor subunit
was not evident in the CA2 region (Telfeian et al., could be involved in tonic inhibition (Peng et al.,
1999). Moreover, there was a blunted effect 2004). Moreover, a slowed GABA clearance will
of NO-711, a selective GAT-1 inhibitor, in the also allow for binding to GABAB receptors, many
MTS tissue compared to comparison material. of which are presynaptic, and can limit neuro-
Finally, the responses to NPA, a GAT inhibitor transmitter release (Sperk et al., 2004). Conversely,
that induces heterotransport (Honmou et al., there may be a reduced efficacy of GABA in
1995), were also reduced in MTS hippocampi. this tissue due to GABA receptor desensitization.
These data were interpreted as impaired GABA Finally, under conditions of high interneuronal
uptake. Parallel findings were made in the kainic activity, a reduction in neuronal uptake could
acid rat model of TLE (Patrylo et al., 2001); how- limit the availability of GABA for repackaging
ever, see Frahm et al. (2003). GAT-1 is found both and release. Which of these possibilities dominates
on neurons and on astrocytes, so these data do under normal conditions has yet to be resolved.
not address the role of gliosis in the regulation Glutamate transport is somewhat more com-
of GABA uptake. These physiological data are plex. The bulk of glutamate uptake is mediated
somewhat consistent with anatomical data show- by glial transporters (Danbolt, 2001) and there is
ing a loss of GAT-1 around granule cell somata no real consensus about changes in MTS vs. non-
(where the GABA was applied), but an increase in MTS hippocampi (see Binder and Steinhauser,
the dendrites (Mathern et al., 1999; Lee et al., 2006 for review). For example, there is little
2006). evidence for changes in glial transporter density
193

when studied with Western blot (e.g., Eid et al., neuropeptide expression and numerous changes
2004), however, immunohistochemical studies in peptide distributions have been observed as
show decreases in EAAT-2 (Mathern et al., 1999; described above.
Proper et al., 2002). However, there are no pub- Fewer physiological studies have been per-
lished studies measuring glial glutamate transport formed relative to the anatomical ones. The best
in human astrocytes that can shed light on these studied peptide is NPY which, as in rodent studies,
conflicting anatomical findings. is a potent modulator of excitatory activity
(rodent, Patrylo et al., 1999). While there is a
Neuromodulation loss of NPY-immunopositive cells in MTS hippo-
campi, there is evidence for axonal sprouting of
There is now anatomical and physiological evi- NPY-immunoreactive (IR) fibers that cover the
dence for multiple neuromodulatory systems in the entire molecular layer, instead of being limited to
human dentate gyrus. Three systems will be de- the inner third of the molecular layer (de Lanerolle
scribed here, with the understanding that there are et al., 1989). When NPY was exogenously applied
numerous other modulators that may be impor- to human granule cells, there were no changes
tant in regulating the excitatory tone of the human in the membrane properties, but a significant
dentate. decrease in the evoked response. It was notable
that NPY had a greater effect in cells with greater
excitability: there was a 60% decrease in the
Metabotropic glutamate receptors (mGluRs) evoked response when there were three or more
evoked action potentials, but only a 34% decrease
Several lines of evidence suggest that excitability in in cells where only one action potential could be
epileptic human tissue is constrained by metabo- evoked (Patrylo et al., 1999). This is consistent
tropic glutamate receptors. First, extracellular with the observations that NPY receptors are
glutamate levels are very high in MTS hippocampi expressed on mossy fibers, but not on the perfo-
(Cavus et al., 2005) so that there is ample trans- rant path inputs (Klapstein and Colmers, 1993),
mitter available to bind mGluRs. Second, Dietrich suggesting that there is a greater effect in those
et al. (1999b) demonstrated that there is a reduced hippocampi with robust synaptic reorganization.
efficacy of L-AP4 (a type 2 mGluR agonist) in Another peptide that is reorganized in MTS
MTS tissue vs. comparison material. Similar find- hippocampi is somatostatin. The SS-IR hilar
ings were reported by Patrylo et al. (1999). One interneurons appear to be highly susceptible to
possibility is that there is little additional effect of seizures in both human MTS and animal models
mGluR agonists because the receptors are satu- (Robbins et al., 1991; Buckmaster and Schwa-
rated. As shown in Fig. 5, there is relatively little rtzkroin, 1995; Vezzani and Hoyer, 1999). As with
effect of L-AP4 compared to other presynaptic NPY, the primary effects of SS are on excitatory,
modulators, such as NPY. rather than inhibitory neurotransmission (Tallent
The anatomic distribution of mGluR5 has been and Siggins, 1997). SS receptors are expressed at
studied in the human dentate and a significant specific synapses such as associational-comissural
upregulation of this postsynaptic receptor has fibers, but not on perforant path fibers. As shown
been observed by several investigators (Tang et in Fig. 5A, there was only a modest effect of SS on
al., 2001; Notenboom et al., 2006). However, the orthodromic and evoked hyperexcitability. The
specific physiological effects of mGluR5 receptors ability of SS to reduce hyperexcitability was much
in human tissue have not been investigated. less than that of NPY (Fig. 5B). However, this
may reflect the downregulation of postsynaptic
Neuropeptides SS receptors in the outer dendritic regions MTS
hippocampi (Csaba et al., 2005).
As in rodents, there is tremendous diversity These are two examples of the physiological
within the interneuronal population in terms of effects numerous peptides that are known to exist
194

Fig. 5. Modulation of granule cell synaptic response by the neuropeptides NPY and somatostatin and glutamate receptor antagonists.
A shows that NPY (1 mM) dramatically reduced the orthodromic response in an MTS case. In contrast, somatostatin (SS) had a less
robust effect on the evoked response. Grouped data showing the percent change for NPY and SS are shown in C1. The effect of NPY
was significantly greater than that of SS (po0.05). C2 shows the percent change in area mediated by glutamate receptor antagonists
APV (30 mM) and the mGluR group 2 antagonist L-AP4 (100 mM) in MTS tissue.

in hilar interneurons and which exhibit some re- and NAA (a neuronal mitochondrial marker) with
organization. Much more work will need to be MRS (Casse et al., 2002; Van Paesschen, 2004a, b;
done to determine how these changes in both fiber Pan et al., 2005). However, none of these in vivo
and receptor distributions will affect the overall approaches allows for a regional assessment of
synaptic output. It is also important to note that metabolism.
there are significant differences between humans The cellular basis for this hypometabolism
and rodents in many of these systems, and thus remains somewhat unclear; however, several
data from rodent models may not accurately pre- recent studies have begun to examine the issue.
dict the situation in the human (Magloczky et al., In the elegant studies of Kann et al. (2005),
2000). electrical stimulation was used to metabolically
activate the tissue and the changes in the levels
of NAD(P)H autofluorescence, as well as in mi-
Metabolism tochondrial potential, were studied. These exper-
iments demonstrated that there was a pronounced
A consistent feature of epileptic tissue is hypome- drop in NADH fluorescence and a smaller over-
tabolism. This has been assessed in vivo using a shoot during stimulation in the dentate as well as
variety of methods including glucose or O2 use in other hippocampal areas. Despite this effect,
with PET and SPECT, as well as phosphocreatine fluorescence imaging of mitochondrial potentials
195

indicated that individual mitochondria exhibited into other pathologies with hippocampal involve-
normal voltage changes to these stimuli. Overall, ment, such as Alzheimer’s disease.
these results were interpreted to indicate reduced
mitochondrial function, because small overshoots
are associated with decreased efficacy of mi- References
tochondrial NAD(P)H+ reduction. An alternate
explanation is that these changes reflect the altered Babb, T.L., Pretorius, J.K., Kupfer, W.R. and Crandall, P.H.
neuronal-glial metabolic coupling know to exist in (1989) Glutamate decarboxylase-immunoreactive neurons
human TLE (Petroff et al., 2002). However, it is are preserved in human epileptic hippocampus. J. Neurosci.,
important to note that the dentate is less metabol- 9: 2562–2574.
Beaumont, A. and Whittle, I.R. (2000) The pathogenesis of
ically impaired in MTS than other hippocampal tumour associated epilepsy. Acta Neurochir. (Wien.), 142:
regions, due in part to the differences in neuronal 1–15.
loss between the CA fields and the dentate (Kunz Beck, H., Blumcke, I., Kral, T., Clusmann, H., Schramm, J.,
et al., 2000). Wiestler, O.D., Heinemann, U. and Elger, C.E. (1996) Prop-
Despite these imaging studies, there have been erties of a delayed rectifier potassium current in dentate
granule cells isolated from the hippocampus of patients with
few studies on the interaction between metabolism chronic temporal lobe epilepsy. Epilepsia, 37: 892–901.
and neuronal function in either animals or hu- Beck, H., Steffens, R., Elger, C.E. and Heinemann, U. (1998)
mans. Our initial studies in this field demonstrated Voltage-dependent Ca2+ currents in epilepsy. Epilepsy Res,
that there were strong correlations between the 32: 321–332.
Beck, H., Steffens, R., Heinemann, U. and Elger, C.E. (1999)
levels of phosphocreatine and the extent of neu-
Ca(2+)-dependent inactivation of high-threshold Ca(2+)
ronal bursting as well as with the IPSP conduct- currents in hippocampal granule cells of patients with chronic
ance. For both variables, lower PCr was associated temporal lobe epilepsy. J Neurophysiol, 82: 946–954.
with greater excitability (Williamson et al., 2005). Bell, B.D., Fine, J., Dow, C., Seidenberg, M. and Hermann,
Finally, together with the mitochondrial imaging B.P. (2005) Temporal lobe epilepsy and the selective remind-
data, these results suggest that the intimate rela- ing test: the conventional 30-minute delay suffices. Psychol.
Assess., 17: 103–109.
tionship between excitability and energetics may Bender, R.A., Soleymani, S.V., Brewster, A.L., Nguyen, S.T.,
be altered in specific ways in MTS hippocampi. Beck, H., Mathern, G.W. and Baram, T.Z. (2003) Enhanced
expression of a specific hyperpolarization-activated cyclic
nucleotide-gated cation channel (HCN) in surviving dentate
Conclusions gyrus granule cells of human and experimental epileptic hip-
pocampus. J. Neurosci., 23: 6826–6836.
Bernard, C., Cossart, R., Hirsch, J.C., Esclapez, M. and
While physiological/functional studies of human Ben-Ari, Y. (2000) What is GABAergic inhibition? How is it
tissue are restricted to surgical specimens, the va- modified in epilepsy? Epilepsia, 41(Suppl 6): S90–S95.
rieties in pathology allow for some conclusions to Bernard, C., Esclapez, M., Hirsch, J.C. and Ben-Ari, Y. (1998)
Interneurones are not so dormant in temporal lobe epilepsy:
be made about the changes that are specific to
a critical reappraisal of the dormant basket cell hypothesis.
MTS. These include robust synaptic reorganiza- Epilepsy Res., 32: 93–103.
tion that promotes hyperexcitability, and a de- Binder, D.K. and Steinhauser, C. (2006) Functional changes in
crease in the efficacy of evoked GABA inhibition. astroglial cells in epilepsy. Glia, 54: 358–368.
The alterations in GABAergic and neuromodula- Brooks-Kayal, A.R., Shumate, M.D., Jin, H., Lin, D.D.,
tory systems are quite complex and the specific Rikhter, T.Y., Holloway, K.L. and Coulter, D.A. (1999)
Human neuronal gamma-aminobutyric acid(A) receptors:
consequences of these changes will be difficult to coordinated subunit mRNA expression and functional
separate. However, because there are distinct correlates in individual dentate granule cells. J. Neurosci.,
differences between these systems in humans and 19: 8312–8318.
in animal models (Magloczky and Freund, 2005), Buckmaster, P.S. and Schwartzkroin, P.A. (1995) Interneurons
and inhibition in the dentate gyrus of the rat in vivo.
these studies will require human material. The
J. Neurosci., 15: 774–789.
availability of viable human specimens allows us Casse, R., Rowe, C.C., Newton, M., Berlangieri, S.U.
the unique opportunity to study the physiological and Scott, A.M. (2002) Positron emission tomography and
changes associated with TLE and provide insight epilepsy. Mol. Imag. Biol., 4: 338–351.
196

Cavus, I., Kasoff, W.S., Cassaday, M.P., Jacob, R., Gueor- Frotscher, M., Jonas, P. and Sloviter, R.S. (2006) Synapses
guieva, R., Sherwin, R.S., Krystal, J.H., Spencer, D.D. and formed by normal and abnormal hippocampal mossy fibers.
Abi-Saab, W.M. (2005) Extracellular metabolites in the cortex Cell Tissue Res, 326: 361–367.
and hippocampus of epileptic patients. Ann. Neurol., 57: Gibbs III, J.W., Zhang, Y.F., Kao, C.Q., Holloway, K.L., Oh,
226–235. K.S. and Coulter, D.A. (1996) Characterization of GABAA
Chen, W.R., Lee, S., Kato, K., Spencer, D.D., Shepherd, G.M. receptor function in human temporal cortical neurons. J.
and Williamson, A. (1996) Long-term modifications of Neurophysiol., 75: 1458–1471.
synaptic efficacy in the human inferior and middle temporal Haas, C.A., Dudeck, O., Kirsch, M., Huszka, C., Kann, G.,
cortex. Proc. Natl. Acad. Sci. U.S.A., 93: 8011–8015. Pollak, S., Zentner, J. and Frotscher, M. (2002) Role for
Cohen-Gadol, A.A., Bradley, C.C., Williamson, A., Kim, J.H., reelin in the development of granule cell dispersion in tem-
Westerveld, M., Duckrow, R.B. and Spencer, D.D. (2005) poral lobe epilepsy. J. Neurosci., 22: 5797–5802.
Normal magnetic resonance imaging and medial temporal Harvey, B.D. and Sloviter, R.S. (2005) Hippocampal granule
lobe epilepsy: the clinical syndrome of paradoxical temporal cell activity and c-Fos expression during spontaneous sei-
lobe epilepsy. J. Neurosurg., 102: 902–909. zures in awake, chronically epileptic, pilocarpine-treated rats:
Csaba, Z., Pirker, S., Lelouvier, B., Simon, A., Videau, C., implications for hippocampal epileptogenesis. J. Comp. Ne-
Epelbaum, J., Czech, T., Baumgartner, C., Sperk, G. and urol., 488: 442–463.
Dournaud, P. (2005) Somatostatin receptor type 2 undergoes Heinrich, C., Nitta, N., Flubacher, A., Muller, M., Fahrner, A.,
plastic changes in the human epileptic dentate gyrus. J. Ne- Kirsch, M., Freiman, T., Suzuki, F., Depaulis, A., Frotscher,
uropathol. Exp. Neurol., 64: 956–969. M. and Haas, C.A. (2006) Reelin deficiency and displacement
Danbolt, N.C. (2001) Glutamate uptake. Prog. Neurobiol., 65: of mature neurons, but not neurogenesis, underlie the for-
1–105. mation of granule cell dispersion in the epileptic hippocam-
Dashtipour, K., Yan, X.X., Dinh, T.T., Okazaki, M.M., pus. J. Neurosci., 26: 4701–4713.
Nadler, J.V. and Ribak, C.E. (2002) Quantitative and mor- Helmstaedter, C., Kurthen, M., Lux, S., Reuber, M. and Elger,
phological analysis of dentate granule cells with recurrent C.E. (2003) Chronic epilepsy and cognition: a longitudinal
basal dendrites from normal and epileptic rats. Hippocam- study in temporal lobe epilepsy. Ann. Neurol., 54: 425–432.
pus, 12: 235–244. Honmou, O., Kocsis, J.D. and Richerson, G.B. (1995) Gaba-
Dietrich, D., Clusmann, H., Kral, T., Steinhauser, C., Blumcke, pentin potentiates the conductance increase induced by nip-
I., Heinemann, U. and Schramm, J. (1999a) Two electro- ecotic acid in CA1 pyramidal neurons in vitro. Epilepsy Res.,
physiologically distinct types of granule cells in epileptic hu- 20: 193–202.
man hippocampus. Neuroscience, 90: 1197–1206. Houser, C.R. (1990) Granule cell dispersion in the dentate
Dietrich, D., Kral, T., Clusmann, H., Friedl, M. and Schramm, gyrus of humans with temporal lobe epilepsy. Brain Res.,
J. (1999b) Reduced function of L-AP4-sensitive metabotropic 535: 195–204.
glutamate receptors in human epileptic sclerotic hippocam- Houser, C.R., Miyashiro, J.E., Swartz, B.E., Walsh, G.O.,
pus. Eur. J. Neurosci., 11: 1109–1113. Rich, J.R. and Delgado-Escueta, A.V. (1990) Altered pat-
Eid, T., Thomas, M.J., Spencer, D.D., Runden-Pran, E., Lai, terns of dynorphin immunoreactivity suggest mossy fiber re-
J.C., Malthankar, G.V., Kim, J.H., Danbolt, N.C., Ottersen, organization in human hippocampal epilepsy. J. Neurosci.,
O.P. and de Lanerolle, N.C. (2004) Loss of glutamine syn- 10: 267–282.
thetase in the human epileptogenic hippocampus: possible Isokawa, M. (1996) Decrement of GABAA receptor-mediated
mechanism for raised extracellular glutamate in mesial tem- inhibitory postsynaptic currents in dentate granule cells in
poral lobe epilepsy. Lancet, 363: 28–37. epileptic hippocampus. J. Neurophysiol., 75: 1901–1908.
El Bahh, B., Lespinet, V., Lurton, D., Coussemacq, M., Le Gal Isokawa, M. (1998) Modulation of GABAA receptor-mediated
La Salle, G. and Rougier, A. (1999) Correlations between inhibition by postsynaptic calcium in epileptic hippocampal
granule cell dispersion, mossy fiber sprouting, and hippo- neurons. Brain Res., 810: 241–250.
campal cell loss in temporal lobe epilepsy. Epilepsia, 40: Isokawa, M., Avanzini, G., Finch, D.M., Babb, T.L. and Le-
1393–1401. vesque, M.F. (1991) Physiologic properties of human dentate
Frahm, C., Stief, F., Zuschratter, W. and Draguhn, A. (2003) granule cells in slices prepared from epileptic patients. Ep-
Unaltered control of extracellular GABA-concentration ilepsy Res., 9: 242–250.
through GAT-1 in the hippocampus of rats after pilocar- Isokawa, M. and Levesque, M.F. (1991) Increased NMDA re-
pine-induced status epilepticus. Epilepsy Res., 52: 243–252. sponses and dendritic degeneration in human epileptic hip-
Franck, J.E., Pokorny, J., Kunkel, D.D. and Schwartzkroin, pocampal neurons in slices. Neurosci. Lett., 132: 212–216.
P.A. (1995) Physiologic and morphologic characteristics of Kann, O., Kovacs, R., Njunting, M., Behrens, C.J., Otahal, J.,
granule cell circuitry in human epileptic hippocampus. Epile- Lehmann, T.N., Gabriel, S. and Heinemann, U. (2005) Met-
psia, 36: 543–558. abolic dysfunction during neuronal activation in the ex vivo
Freund, T.F. and Buzsaki, G. (1996) Interneurons of the hip- hippocampus from chronic epileptic rats and humans. Brain,
pocampus. Hippocampus, 6: 347–470. 128: 2396–2407.
197

Kim, J.H., Guimaraes, P.O., Shen, M.Y., Masukawa, L.M. antidromically evoked field responses of the dentate gyrus
and Spencer, D.D. (1990) Hippocampal neuronal density in and mossy fiber reorganization in temporal lobe epileptic
temporal lobe epilepsy with and without gliomas. Acta patients. Brain Res., 579: 119–127.
Neuropathol. (Berl.), 80: 41–45. Mathern, G.W., Babb, T.L., Pretorius, J.K. and Leite, J.P.
Klapstein, G.J. and Colmers, W.F. (1993) On the sites of pre- (1995) Reactive synaptogenesis and neuron densities for ne-
synaptic inhibition by neuropeptide Y in rat hippocampus in uropeptide Y, somatostatin, and glutamate decarboxylase
vitro. Hippocampus, 3: 103–111. immunoreactivity in the epileptogenic human fascia dentata.
Kobayashi, M. and Buckmaster, P.S. (2003) Reduced inhibition J. Neurosci., 15: 3990–4004.
of dentate granule cells in a model of temporal lobe epilepsy. Mathern, G.W., Mendoza, D., Lozada, A., Pretorius, J.K.,
J. Neurosci., 23: 2440–2452. Dehnes, Y., Danbolt, N.C., Nelson, N., Leite, J.P., Chimelli,
Kunz, W.S., Kudin, A.P., Vielhaber, S., Blumcke, I., Zusch- L., Born, D.E., Sakamoto, A.C., Assirati, J.A., Fried, I.,
ratter, W., Schramm, J., Beck, H. and Elger, C.E. (2000) Peacock, W.J., Ojemann, G.A. and Adelson, P.D. (1999)
Mitochondrial complex I deficiency in the epileptic focus of Hippocampal GABA and glutamate transporter immunore-
patients with temporal lobe epilepsy. Ann. Neurol., 48: activity in patients with temporal lobe epilepsy. Neurology,
766–773. 52: 453–472.
de Lanerolle, N.C., Brines, M., Williamson, A., Kim, J.H. and Nagerl, U.V., Mody, I., Jeub, M., Lie, A.A., Elger, C.E. and
Spencer, D.D. (1992) Neurotransmitters and their receptors Beck, H. (2000) Surviving granule cells of the sclerotic human
in human temporal lobe epilepsy. Epilepsy Res. Suppl., 7: hippocampus have reduced Ca(2+) influx because of a loss
235–250. of calbindin-D(28k) in temporal lobe epilepsy. J Neurosci.,
de Lanerolle, N.C., Kim, J.H., Robbins, R.J. and Spencer, 20: 1831–1836.
D.D. (1989) Hippocampal interneuron loss and plasticity in Notenboom, R.G., Hampson, D.R., Jansen, G.H., Van Rijen,
human temporal lobe epilepsy. Brain Res., 495: 387–395. P.C., Van Veelen, C.W., Van Nieuwenhuizen, O. and De
de Lanerolle, N.C., Kim, J.H., Williamson, A., Spencer, S.S., Graan, P.N. (2006) Up-regulation of hippocampal metabo-
Zaveri, H.P., Eid, T. and Spencer, D.D. (2003) A retrospec- tropic glutamate receptor 5 in temporal lobe epilepsy pa-
tive analysis of hippocampal pathology in human temporal tients. Brain, 129: 96–107.
lobe epilepsy: evidence for distinctive patient subcategories. Pan, J.W., Kim, J.H., Cohen-Gadol, A., Pan, C., Spencer, D.D.
Epilepsia, 44: 677–687. and Hetherington, H.P. (2005) Regional energetic dysfunc-
de Lanerolle, N.C., Williamson, A., Meredith, C., Kim, J.H., tion in hippocampal epilepsy. Acta Neurol. Scand., 111:
Tabuteau, H., Spencer, D.D. and Brines, M.L. (1997) Dynor- 218–224.
phin and the kappa 1 ligand [3H]U69,593 binding in the hu- Patrylo, P.R. and Dudek, F.E. (1998) Physiological unmasking
man epileptogenic hippocampus. Epilepsy Res., 28: 189–205. of new glutamatergic pathways in the dentate gyrus of hip-
Lee, T.S., Bjornsen, L.P., Paz, C., Kim, J.H., Spencer, S.S., pocampal slices from kainate-induced epileptic rats. J. Ne-
Spencer, D.D., Eid, T. and de Lanerolle, N.C. (2006) GAT1 urophysiol., 79: 418–429.
and GAT3 expression are differently localized in the human Patrylo, P.R., Spencer, D.D. and Williamson, A. (2001) GABA
epileptogenic hippocampus. Acta Neuropathol. (Berl.), 111: uptake and heterotransport are impaired in the dentate gyrus
351–363. of epileptic rats and humans with temporal lobe sclerosis. J.
Lieberman, D.N. and Mody, I. (1999) Properties of single Neurophysiol., 85: 1533–1542.
NMDA receptor channels in human dentate gyrus granule Patrylo, P.R., Van Den Pol, A.N., Spencer, D.D. and Will-
cells. J. Physiol., 518(Pt 1): 55–70. iamson, A. (1999) NPY inhibits glutamatergic excitation in
Lombardo, A.J., Kuzniecky, R., Powers, R.E. and Brown, G.B. the epileptic human dentate gyrus. J. Neurophysiol., 82:
(1996) Altered brain sodium channel transcript levels in hu- 478–483.
man epilepsy. Brain Res. Mol. Brain Res., 35: 84–90. Peng, Z., Huang, C.S., Stell, B.M., Mody, I. and Houser, C.R.
Magloczky, Z. and Freund, T.F. (2005) Impaired and repaired (2004) Altered expression of the delta subunit of the GABAA
inhibitory circuits in the epileptic human hippocampus. receptor in a mouse model of temporal lobe epilepsy. J. Ne-
Trends Neurosci., 28: 334–340. urosci., 24: 8629–8639.
Magloczky, Z., Halasz, P., Vajda, J., Czirjak, S. and Freund, Petroff, O.A., Errante, L.D., Rothman, D.L., Kim, J.H. and
T.F. (1997) Loss of Calbindin-D28K immunoreactivity from Spencer, D.D. (2002) Glutamate-glutamine cycling in the
dentate granule cells in human temporal lobe epilepsy. Ne- epileptic human hippocampus. Epilepsia, 43: 703–710.
uroscience, 76: 377–385. Porter, B.E., Zhang, G., Celix, J., Hsu, F.C., Raol, Y.H., Tel-
Magloczky, Z., Wittner, L., Borhegyi, Z., Halasz, P., Vajda, J., feian, A., Gallagher, P.R., Coulter, D.A. and Brooks-Kayal,
Czirjak, S. and Freund, T.F. (2000) Changes in the distribu- A.R. (2005) Heterogeneous GABAA receptor subunit ex-
tion and connectivity of interneurons in the epileptic human pression in pediatric epilepsy patients. Neurobiol. Dis., 18:
dentate gyrus. Neuroscience, 96: 7–25. 484–491.
Masukawa, L.M., Uruno, K., Sperling, M., O’connor, M.J. and Proper, E.A., Hoogland, G., Kappen, S.M., Jansen, G.H.,
Burdette, L.J. (1992) The functional relationship between Rensen, M.G., Schrama, L.H., Van Veelen, C.W., Van Rijen,
198

P.C., Van Nieuwenhuizen, O., Gispen, W.H. and De Graan, Tallent, M.K. and Siggins, G.R. (1997) Somatostatin depresses
P.N. (2002) Distribution of glutamate transporters in the excitatory but not inhibitory neurotransmission in rat CA1
hippocampus of patients with pharmaco-resistant temporal hippocampus. J. Neurophysiol., 78: 3008–3018.
lobe epilepsy. Brain, 125: 32–43. Tang, F.R., Lee, W.L. and Yeo, T.T. (2001) Expression of the
Reckziegel, G., Beck, H., Schramm, J., Elger, C.E. and Urban, group I metabotropic glutamate receptor in the hippocampus
B.W. (1998) Electrophysiological characterization of Na+ of patients with mesial temporal lobe epilepsy. J. Ne-
currents in acutely isolated human hippocampal dentate urocytol., 30: 403–411.
granule cells. J. Physiol., 509(Pt 1): 139–150. Telfeian, A.E., Spencer, D.D. and Williamson, A. (1999) Lack
Reckziegel, G., Beck, H., Schramm, J., Urban, B.W. and Elger, of correlation between neuronal hyperexcitability and elect-
C.E. (1999) Carbamazepine effects on Na+ currents in hu- rocorticographic responsiveness in epileptogenic human
man dentate granule cells from epileptogenic tissue. Epile- neocortex. J. Neurosurg., 90: 939–945.
psia, 40: 401–407. Van Paesschen, W. (2004a) Ictal SPECT. Epilepsia, 45(Suppl
Ribak, C.E., Tran, P.H., Spigelman, I., Okazaki, M.M. and 4): 35–40.
Nadler, J.V. (2000) Status epilepticus-induced hilar basal Van Paesschen, W. (2004b) Qualitative and quantitative imag-
dendrites on rodent granule cells contribute to recurrent ex- ing of the hippocampus in mesial temporal lobe epilepsy with
citatory circuitry. J. Comp. Neurol., 428: 240–253. hippocampal sclerosis. Neuroimag. Clin. North Am., 14:
Robbins, R.J., Brines, M.L., Kim, J.H., Adrian, T., de Lane- 373–400.
rolle, N., Welsh, S. and Spencer, D.D. (1991) A selective loss Vezzani, A. and Hoyer, D. (1999) Brain somatostatin: a can-
of somatostatin in the hippocampus of patients with tempo- didate inhibitory role in seizures and epileptogenesis. Eur. J.
ral lobe epilepsy. Ann. Neurol., 29: 325–332. Neurosci., 11: 3767–3776.
Roy, N.S., Wang, S., Jiang, L., Kang, J., Benraiss, A., Harri- Von Campe, G., Spencer, D.D. and de Lanerolle, N.C. (1997)
son-Restelli, C., Fraser, R.A., Couldwell, W.T., Kawaguchi, Morphology of dentate granule cells in the human epilepto-
A., Okano, H., Nedergaard, M. and Goldman, S.A. (2000) In genic hippocampus. Hippocampus, 7: 472–488.
vitro neurogenesis by progenitor cells isolated from the adult Williamson, A., Patrylo, P.R., Pan, J., Spencer, D.D. and He-
human hippocampus. Nat. Med., 6: 271–277. therington, H. (2005) Correlations between granule cell phys-
Scharfman, H.E. (2004) Functional implications of seizure-in- iology and bioenergetics in human temporal lobe epilepsy.
duced neurogenesis. Adv. Exp. Med. Biol., 548: 192–212. Brain, 128: 1199–1208.
Schwartzkroin, P.A. and Prince, D.A. (1976) Microphysiology Williamson, A., Patrylo, P.R. and Spencer, D.D. (1999)
of human cerebral cortex studied in vitro. Brain Res., 115: Decrease in inhibition in dentate granule cells from
497–500. patients with medial temporal lobe epilepsy. Ann. Neurol.,
Seress, L., Gulyas, A.I., Ferrer, I., Tunon, T., Soriano, E. and 45: 92–99.
Freund, T.F. (1993) Distribution, morphological features, Williamson, A. and Spencer, D. (1995) Zinc reduces dentate
and synaptic connections of parvalbumin- and calbindin granule cell hyperexcitability in epileptic humans. Neurore-
D28k-immunoreactive neurons in the human hippocampal port, 6: 1562–1564.
formation. J. Comp. Neurol., 337: 208–230. Williamson, A., Spencer, D.D. and Shepherd, G.M. (1993)
Shumate, M.D., Lin, D.D., Gibbs III, J.W., Holloway, K.L. Comparison between the membrane and synaptic properties
and Coulter, D.A. (1998) GABA(A) receptor function in ep- of human and rodent dentate granule cells. Brain Res., 622:
ileptic human dentate granule cells: comparison to epileptic 194–202.
and control rat. Epilepsy Res., 32: 114–128. Williamson, A., Spencer, S.S. and Spencer, D.D. (1995a) Depth
Sloviter, R.S., Zappone, C.A., Harvey, B.D. and Frotscher, M. electrode studies and intracellular dentate granule cell re-
(2006) Kainic acid-induced recurrent mossy fiber innervation cordings in temporal lobe epilepsy. Ann. Neurol., 38:
of dentate gyrus inhibitory interneurons: possible anatomical 778–787.
substrate of granule cell hyper-inhibition in chronically ep- Williamson, A., Telfeian, A.E. and Spencer, D.D. (1995b) Pro-
ileptic rats. J. Comp. Neurol., 494: 944–960. longed GABA responses in dentate granule cells in slices
Sperk, G., Furtinger, S., Schwarzer, C. and Pirker, S. (2004) isolated from patients with temporal lobe sclerosis. J. Ne-
GABA and its receptors in epilepsy. Adv. Exp. Med. Biol., urophysiol., 74: 378–387.
548: 92–103. Wittner, L., Eross, L., Szabo, Z., Toth, S., Czirjak, S., Halasz,
St John, J.L., Rosene, D.L. and Luebke, J.I. (1997) Morphol- P., Freund, T.F. and Magloczky, Z.S. (2002) Synaptic reor-
ogy and electrophysiology of dentate granule cells in the ganization of calbindin-positive neurons in the human hip-
rhesus monkey: comparison with the rat. J. Comp. Neurol., pocampal CA1 region in temporal lobe epilepsy.
387: 136–147. Neuroscience, 115: 961–978.
Sutula, T., Cascino, G., Cavazos, J., Parada, I. and Ramirez, L. Wuarin, J.P. and Dudek, F.E. (2001) Excitatory synaptic input
(1989) Mossy fiber synaptic reorganization in the epileptic to granule cells increases with time after kainate treatment. J.
human temporal lobe. Ann. Neurol., 26: 321–330. Neurophysiol., 85: 1067–1077.
Plate 11.2. Granule cell hyperexcitability is associated with mossy fiber sprouting. A shows representative traces comparing the
response to molecular layer (orthodromic) and hilar (antidromic) stimulation in a patient with MTS. Note that there is a greater
response to maximal hilar stimulation than to stimulation of outer molecular layer. Moreover, note that hilar stimulation produced a
synaptic response prior to an antidromic spike. These data support the hypothesis that mossy fibers are primarily excitatory in resected
tissue. B shows Timm staining from the region of the dentate where the cell shown in A was recorded. Note the robust supergranular
labeling indicating the presence of sprouted fibers. (C1) Stimulation protocols have the problem of activating multiple fiber systems.
Therefore, in the presence of 30 mM bicuculline, we applied a bolus of glutamate to the granule cell layer at a distance from the
recorded cell and were able to evoke a long lasting barrage of EPSPs. The glutamate was applied at a distance from the recorded cell
and thus we were not observing direct glutamate effects. C2 shows that a similar response was not observed when the glutamate was
applied to the hilus. (For B/W version, see page 188 in the volume.)
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 12

Hilar mossy cells: functional identification and


activity in vivo

Darrell A. Henze1 and György Buzsáki2,

1
Merck Research Laboratories, West Point, PA 19486, USA
2
Center for Molecular and Behavioral Neuroscience, Rutgers, The State University of New Jersey, Newark, NJ, USA

Abstract: Network oscillations are proposed to provide the framework for the ongoing neural compu-
tations of the brain. Thus, an important aspect of understanding the functional roles of various cell classes
in the brain is to understand the relationship of cellular activity to the ongoing oscillations. While many
studies have characterized the firing properties of cells in the hippocampal network including granule cells,
pyramidal cells and interneurons, information about the activity of dentate mossy cells in the intact brain
is scant. Here we review the currently available information and describe biophysical properties and
network-related firing patterns of mossy cells in vivo. These new observations will assist in the extracellular
identification of this unique cell type and help elucidate their functional role in behaving animals.

Keywords: mossy cell; hilus; network patterns; sharp wave; gamma; slow oscillation

Introduction Mossy cells are perhaps the most underinvesti-


gated neurons in the hippocampal formation. Part
The granule cell is the most numerous cell type of the scarcity of information can be explained by
of the dentate gyrus (1 million in one hemisphere the low number of MCs (approximately 10,000 in
in the rat, (Amaral et al., 1990), serves to integrate the rat) (Amaral et al., 1990). Unlike other prin-
entorhinal input and sends its messages to three cipal (excitatory) cell types of the hippocampus,
major target cell populations. These three cell MCs do not form recognizable layers with densely
groups include mossy cells (MCs) of the hilus, packed somata. Instead, they are scattered in
CA3 pyramidal cells and interneurons of both the hilar region under the granule cell layer,
dentate and CA3 regions. Although the MCs are making their in vivo accessibility for physiological
the least numerous cell type in the hilar region, studies difficult. Most of what we know about the
their unique properties suggest that they are functions of MCs comes from the pioneering
likely to be a very important cell type of the dent- in vitro studies of Scharfman (Scharfman and
ate region. This chapter examines the properties Schwartzkroin, 1988; Scharfman et al., 1990, 2001;
of the MCs and their possible role in overall Scharfman, 1991, 1992a, b, 1994a–c, 1995), in vivo
hippocampal function. studies of Schwartzkroin et al. (Buckmaster et al.,
1992, 1993, 1996; Buckmaster and Schwartzkroin
1995; Wenzel et al., 1997) and pathoanatomical
Corresponding author. Tel.: +1 (973) 353-1080 ext. 3131; studies of epilepsy and ischemia by Sloviter et al.
Fax: +1 (973) 353-1820; E-mail: buzsaki@andromeda.rutgers.edu (Sloviter, 1989, 1991a, 1994; Sloviter et al., 1991,

DOI: 10.1016/S0079-6123(07)63012-X 199


200

2003; Goodman et al., 1993; Zappone and MCs containing calretinin (Murakawa and
Sloviter, 2004). Behavioral correlates of MCs are Kosaka, 2001) which is largely absent in MCs of
unknown due to the lack of criteria for the reliable the dorsal hilus. This is in contrast to the rat where
identification of MCs with extracellular methods. no MCs contain calretinin. In addition, calcitonin
A major goal of this chapter is to provide an over- gene related peptide has been reported to selec-
view of the available knowledge about the firing tively label MCs in the rat (Freund et al., 1997).
patterns and biophysical properties of anatomi- Because these peptide markers are conspicuously
cally identified MCs and their network-related absent in the pyramidal cells of the hippocampus
behavior, and discuss how this information can proper, their differential staining can be taken as
facilitate studies on MCs in the intact, behaving clear justification of separating MCs of the dentate
animal. gyrus and pyramidal neurons of the Ammon’s
horn.
The MCs are usually multipolar and have
Mossy cell anatomy tapering dendrites that largely remain restricted
to the hilus proper, although occasional dendrites
Mossy cells of the hilus were first recognized by are observed in the molecular layer of the dentate
Cajal (1911) and Lorente de No (1934) for their gyrus in both rats (Scharfman, 1991), and more of-
dendrites covered with large spines. These cells ten in primates (Frotscher et al., 1991; Buckmaster
were later given the name ‘‘mossy cells’’ by Amaral and Amaral, 2001). These dendrites in the molec-
(1978) due to the ‘‘mossy’’ appearance of their ular layer can receive direct input from the ent-
large spines (see Fig. 1A). It has also been dem- orhinal cortex bypassing the granule cells. This
onstrated that all MCs selectively stain for GluR2/ variability in direct EC input is likely to be im-
3 receptors as opposed to other cells of the hilus portant for physiological function. All CA3 py-
(Petralia and Wenthold, 1992; Leranth et al., 1996; ramidal cells, including those with mostly
Fujise and Kosaka, 1999). Interestingly, there ap- horizontal dendrites residing in zone 3 of Amaral
pears to be differences in the peptide content of the (1978), send at least one dendritic branch to the
mouse and hamster MC population, with ventral stratum lacunosum-moleculare and therefore

A B

Ca3 PC
Layer

C 20 mV
0.6 nA
50 ms

DG granule cell layer

Fig. 1. Hilar mossy cell and associated basic properties. (A) Biocytin labeled mossy cell. Note the large spines covering the soma and
proximal dendrites (arrows). Example current step evoked traces from the mossy cell shown in A. (B) Responses to a series of
hyperpolarizing steps from a resting potential of 57 mV. Note the robust ‘‘sag’’ observed for the larger hyperpolarizing steps and the
long charging curve for the smallest hyperpolarizing step. For this cell, the in vivo input resistance was 52 MO. (C) The response of this
cell to a strong depolarizing step (1.0 nA). Note that there is only very weak accommodation observed.
201

receive inputs from layer 2 neurons of the ent- It is possible that at least some of the input to the
orhinal cortex. In contrast, MCs without dendrites more peripheral dendrites is from CA3 pyramidal
in the molecular layer generally receive only indi- cells which extend axons back into the hilus and
rect information from the entorhinal input relayed granule cell layer (Li et al., 1994). However, direct
by the granule cells (but see Kohler, 1985; Deller, anatomical evidence for a CA3-mossy cell com-
1998). munication is still lacking (but see below for dis-
The proximal dendrites and somata of MCs are cussion of functional data). The dendrites of MCs
covered by large ‘‘thorny excrescences’’ (Fig. 1). are extensive, extending several hundred microm-
While the thorny excrescences at first seem similar eters in both medio-lateral and dorso-caudal di-
to those of the proximal apical dendrite of CA3 rections, suggesting a largely cylindrical dendritic
pyramidal cells, they are qualitatively different in tree arborizing largely in the subgranular zone.
that they resemble ‘‘clusters of spheres’’ (Amaral, The large span of the dendritic arbor suggests
1978) where as CA3 excrescences appear more that MCs are innervated by spatially distributed
irregular and thorny in their shape (Chicurel and granule cells. The 100:1 ratio of granule cells to
Harris, 1992). As in CA3 pyramidal cells, the MCs predicts that a typical MC receives inputs
thorny excrescences of MCs receive synaptic input from as many as a hundred granule cells. Given
from the mossy fibers of the granule cells (Amaral, the relative sparse distribution of mossy fiber
1978; Murakawa and Kosaka, 2001). Recent find- terminals in the hilus, it is also likely that each
ings suggest that the mossy fiber input to CA3 is a granule cell innervates only one or two MCs with
mixed glutamatergic/GABAergic input for the first large mossy fiber boutons. This low divergence
three weeks of development in the rat or the young and convergence should be contrasted to the wide-
guinea pig (Walker et al., 2001). Under normal spread reciprocal innervation of granule cells by
conditions in the adult, there is no detectable the MCs (see below).
GABAergic transmission. However, mossy fiber The MCs make glutamatergic (Soriano and
GABAergic transmission is restored in the adult Frotscher, 1994; Wenzel et al., 1997) asymmetric
after periods of strong repetitive stimulation or synapses with both excitatory and inhibitory
hyperexcitability such as epilepsy (see Gutierrez, postsynaptic targets (Frotscher et al., 1991;
2003, for review). There are also significant GAD- Buckmaster et al., 1996). In general, a single
positive, inhibitory inputs to the somatic region of MC’s synaptic targets can be divided into three
MCs suggesting a strong perisomatic inhibitory classes: local hilar targets; longitudinal targets
input to these cells (Acsady et al., 2000; Murakawa (located >1 mm either septally or temporally); and
and Kosaka, 2001). In fact, the perisomatic inhi- contralateral targets. These three target classes have
bition to MCs is very strong (15–40 times more been observed in both rodents and primates. Excit-
synapses per soma) compared to the inhibitory atory innervation of CA3 pyramidal cells by the
cells of the hilus. The perisomatic inhibitory in- MCs has been repeatedly suggested (Buckmaster
nervation of MCs is primarily from terminals that et al., 1996; Buckmaster and Amaral, 2001) but
contain parvalbumin or cholecystokinin (CCK) conclusive anatomical evidence for this alleged
(Acsady et al., 2000), similar to the pyramidal connection is missing.
cells (Freund and Buzsáki, 1996). There is also a The main local hilar targets are mostly aspiny
perisomatic and proximal dendritic input from dendrites of interneurons and potentially dendrites
cholinergic fibers (Deller et al., 1999) which may of other MCs, although anatomical proof for
serve to provide excitatory tone to the MCs during mutual MC communication is lacking. CA3 re-
oscillations such as theta. The more distal, or current collaterals also innervate aspiny interneu-
so-called peripheral dendrites, are innervated by a rons in the hilus (Buckmaster et al., 1996) but
variety of inputs including more mossy fiber inputs conspicuously avoid spiny interneurons of which
onto regular small spines as well as other anatom- there are many in the hilus (Wittner et al., 2006).
ically uncharacterized inputs making asymmetric The major bulk of the axon cloud of MCs (>90%
and symmetric synapses (Frotscher et al., 1991). of ipsilateral synaptic contacts) target dentate
202

granule cell dendrites in the inner third of the hippocampal function. This can be summarized
molecular layer of the dentate gyrus rostral and most simply as primarily providing distributed
caudal to the soma and dendrites of the parent excitatory feedback to dentate granule cells and
MC (Buckmaster et al., 1992, 1996; Wenzel et al., secondarily providing excitatory drive to local
1997). This longitudinal arrangement is of major inhibitory interneurons of the hilus. The functional
physiological consequence because it eliminates importance of the contralateral projection is
the possibility of a local granule cell–mossy harder to predict since the anatomical targets
cell–granule cell recurrent excitatory loop. Gran- are not yet characterized. A body of work
ule cells, which drive discharge of the MCs, do by Scharfman and colleagues (Scharfman and
not receive excitatory information about the firing Schwartzkroin, 1988; Scharfman et al., 1990,
status of their targets. Instead, the output of 2001; Scharfman, 1991, 1992a, b, 1994a–c, 1995)
the activated MC is distributed over the septal- has provided functional data to support the pre-
temporal extent of the dentate. In these target dictions of the anatomical connectivity to and
areas, MCs innervate both granule cells and inter- from MCs. For example, MCs that have dendrites
neurons, giving rise to potentially interesting phys- that extend into the DG molecular layer have a
iological scenarios. First, a granule cell discharging lower threshold to fire in response to perforant
a target MC can silence competing granule cells path stimulation (Scharfman, 1991). A challenging
over large territories via the feed-forward excita- series of studies in ventral slices used paired
tion of interneurons by the MC. However, if this recordings from anatomically confirmed hilar,
were the sole function of such widespread axon dentate and CA3 pyramidal cells to investigate
arborization, there would be no need for MC the functional connectivity in this region. Paired
excitatory innervation of granule cells. Through recording of granule cells or CA3 pyramidal cells
the latter numerically dominant excitatory path- with MCs showed that single action potentials in
way, the possibility exists that MCs can synchro- either GCs or PCs can evoke EPSPs in MCs
nize spatially distinct granule cells in the (Scharfman, 1994b). Another paired recording
septotemporal axis, although the functional effi- study demonstrated that MCs monosynaptically
cacy of the mossy cell–granule cell synapse is not excite both granule cells and inhibitory interneu-
well characterized (but see Scharfman, 1995). rons. In addition, evidence was observed for
From the latter perspective, the main function of polysynaptic inhibition of GCs in response to
MCs would be to integrate functions of large MC activity (Scharfman, 1995).
numbers of active granules cells in the septotem- The hypothesized physiological function of the
poral axis of the hippocampus. low convergence and divergence of granule cells
The contralateral cellular targets of MCs are onto CA3 pyramidal cells is to disperse (‘‘orthog-
largely undefined. However, in addition to inner- onalize’’) the entorhinal information onto the large
vating granule cells, at least hilar neuropeptide Y recurrent system of CA3 neurons during the en-
(NPY) containing cells may receive input from coding of memories and provide multiple but
contralateral MCs (Deller and Leranth, 1990). sparse representation (Treves and Rolls, 1992,
It will be interesting to learn the extent and iden- 1994). In the retrieval process, the auto-associative
tity of the contralateral targets and how they CA3 recurrent system, in turn, can recover the
compare to the predominant granule cell targets whole memory representation from partial or
ipsilaterally. fragmental information (‘‘pattern completion’’)
(McNaughton and Morris, 1987; Kanerva, 1988).
Given this model of memory encoding and re-
Functional cellular connectivity of MCs trieval, we can ask what role might MCs play in
this system? Although the anatomical connectivity
The anatomical connectivity described above pro- between granule cells and MCs is similar to the
vides a prospective framework for understanding granule cell–CA3 connections, given the small
the functional role of the MCs in normal number of MCs and the lack of their reciprocal
203

excitation with one another makes it unlikely (CA3 ¼ 53.8 mO; n ¼ 8; CA1 ¼ 48.4 MO (Henze
that MCs are part of the pattern completion and Buzsaki, 2001)). The smallest hyperpolarizing
mechanism. Instead, they may assist in increasing step applied (0.2 nA) resulted in a hyperpolari-
the sparseness of the memory representation in the zation of 22 mV and the time constant is best fit
recurrent CA3 system by the hypothesized feed with a double exponential with tau1 ¼ 7.5 ms and
forward suppression of granule cells in the septo- tau2 ¼ 250 ms. A strong inward rectification can
temporal axis or by allowing a sparse but coordi- be seen with a step injection of 0.4 nA that results
nated transfer of information from different ultimately in a maximum hyperpolarization of
segments of the entorhinal cortex onto the CA3 35 mV. The hyperpolarizations of the membrane
system. This may be an important combinatorial potential by small step currents were best fit by a
mechanism because the metric of spatial represen- double exponential in seven of the nine cells where
tation increases in the dorso-caudal axis of the it could be measured; the mean time constant
entorhinal cortex with corresponding projections values were 15.9 and 188.2 ms. The remaining two
to different segments of the septotemporal MCs were best fit with single exponentials with
axis of the hippocampus (Hafting et al., 2005). time-constants of 32 and 26 ms. Figure 1C shows
The synchronizing mechanism of granule cells by that the MCs typically do not show burst firing in
the MCs in the longitudinal axis would secure the response to depolarizing steps (1.0 nA) and only
simultaneous but distributed representation of the show weak accommodation for the duration of
different sampling metric of the entorhinal cortex the step depolarization. This behavior can be
in the associative CA3 network. This hypothesis contrasted to the typical burst pattern and strong
differs from an alternative proposed by Buckmas- spike accommodation in response to current steps
ter and Schwartzkroin (1994) dubbed the ‘‘granule in CA3 pyramidal cells (personal observations,
cell association’’ hypothesis. In that hypothesis, Bilkey and Schwartzkroin, 1990; Buckmaster
the MCs provide the necessary links to form et al., 1993; Scharfman, 1993b).
associative connections within the dentate network The background synaptic activity in MCs has
analogous to the associational collaterals of been reported to be quite high, both in vivo and in
pyramidal cells in the CA3 region. vitro (Strowbridge et al., 1992; Scharfman,
1993a).We also have observed high background
activity that included some very large events
Cellular properties of MCs (>10 mV; Fig. 2). It is likely that these giant PSPs
with fast rise times arise from the mossy fiber
The basic cellular properties of MCs are quite dis- synaptic inputs to the MCs reflecting either
tinctive from other cell types in the hilar region. synchronous multivesicular release from a com-
The MCs tend to have higher input resistance and plex MF bouton or perhaps the release of large
strong inward rectification in response to hyper- individual quanta as has been reported in CA3
polarization. Figure 1B shows a typical MC from a pyramidal cells (e.g. see Henze et al., 2002).
set of 10 cells we have recorded in vivo in response Although the magnitude of the giant PSPs is quite
to a series of hyperpolarizing steps. Each anatom- variable, it is unlikely that it reflects varying
ically verified MC recorded in urethane anestheti- convergence of activity from multiple granule
zed rats had action potentials that crossed 0 mV, a cells. First, the convergence of granule cells
mean resting membrane potential of 5871.8 mV, onto MCs is low (100). Second, the density of
and a mean input resistance of 6178.6 MO mossy boutons in the hilus is quite low. Third,
(means7SEM). The resting potential is similar intracellularly labeled neighboring granule cells
to that of CA3 (63 mV; n ¼ 84) and CA1 never showed spatially clustered boutons that
(65 mV; n ¼ 280) pyramidal cells but different would otherwise suggest common targets (Acsady
from the more hyperpolarized granule cells et al., 2000).
(74 mV; n ¼ 41). Of the principal cell types, One feature of MC activity that we have
MCs had the largest input resistance observed that has not been previously described
204

A
unusual in that the extracellular unit waveform has
a rounded trough that has not been observed in
10 mV other hippocampal unit waveforms. It is possible
that the rounded (wide) pattern derives from the
0.1 s
summed extracellular currents from the soma
B
and the thick dendrites of MCs. In support of this
2 mV interpretation, the extracellular spikes were
0.01 s recorded by three shanks of the silicon probe, a
total distance of 300 mm. We would suggest that
the horizontally wide current distribution of MC
spikes may be used as a distinguishing feature in
extracellular recordings from the granule cells and
other nearby smaller size interneurons.
Fig. 2. (A) Intracellular recording of a hilar mossy cell resting
at 55 mV (solid line). Notice the large spontaneous post-
synaptic potentials that occur from the baseline that sometimes
lead to action potentials. (B) Higher resolution depiction of two
MC cellular activity in a network
large EPSPs from the box in (A).
Although there is a relative lack of information
about the physiological role of MCs under nor-
is their relatively high background firing rate mal conditions, their role has been a frequently
(e.g. see Figs. 3, 5, and 9C). Our findings under discussed topic of debate in the epilepsy literature.
urethane anesthesia are partly consistent with This is because MCs are often observed to be
Scharfman (1993a), who showed that there is reduced in post-mortem tissue taken from people
often a high rate of firing but it waxes and wanes in who have had temporal lobe epilepsy (TLE)
vitro. However, it is not consistent with other studies (Sloviter et al., 1991). Sloviter and colleagues pro-
in vivo (Soltesz et al., 1993) under ketamine/ posed the so-called ‘‘dormant basket cell’’ hypoth-
xylazine where rates of less than 1 Hz were reported. esis of epilepsy (Sloviter, 1991b). The dormant
The source of this high firing rate is not well under- basket cell hypothesis holds that the importance of
stood because granule cells under urethane an- the excitatory tone provided by the MCs is to
esthesia are typically hyperpolarized and fire at a provide excitation of hilar inhibitory interneurons
low rate. One hypothesis is that spontaneous release which in turn then provide strong inhibition of
from MF boutons can lead to MC action potentials granule cells. When MCs are lost due to neurode-
(Henze et al., 2002). If the high firing pattern is generation associated with TLE, the inhibitory
confirmed by investigations in the freely behaving basket cells lose their tonic excitatory drive result-
animal, it can be contrasted with the more clustered ing in a net disinhibition of dentate gyrus granule
firing patterns of CA3 pyramidal cells. Figure 3 il- cells (Sloviter, 1994; Sloviter et al., 2003). A
lustrates the autocorrelograms calculated from our contrasting hypothesis has been called the ‘‘irrita-
MC recordings. These autocorrelograms should fa- ble mossy cell’’ hypothesis as proposed by Soltesz
cilitate the comparison of the spike dynamics of and colleagues. In this view, it is not the loss of
MCs under anesthesia and in the drug-free animal in MCs that leads to a net excitation of granule cells,
future studies. instead it is the remaining MCs that have higher
Another essential piece of information in the firing rates and provide uncontrolled excitatory
process of physiological identification of MCs is feedback to the dentate gyrus granule cells thus
the extracellular features and shape of the extra- exacerbating the epileptic process (Santhakumar
cellular action potential. To this end, we have re- et al., 2000; Ratzliff et al., 2002).
corded from a single MC using simultaneous Both of the ‘‘dormant basket cell’’ and ‘‘irritable
intracellular and extracellular electrodes (Fig. 4). mossy cell’’ hypotheses have their attractions.
The waveform we have observed is somewhat Recent studies by both camps have provided
205

A F

-0.10 -0.05 0.00 0.05 0.10 -0.10 -0.05 0.00 0.05 0.10

B G

-0.10 -0.05 0.00 0.05 0.10 -0.10 -0.05 0.00 0.05 0.10

C H

-0.10 -0.05 0.00 0.05 0.10 -0.10 -0.05 0.00 0.05 0.10

D I

-0.10 -0.05 0.00 0.05 0.10


-0.10 -0.05 0.00 0.05 0.10

E J

-0.10 -0.05 0.00 0.05 0.10 -0.10 -0.05 0.00 0.05 0.10

Fig. 3. Hilar mossy cells show a variety of firing patterns as assessed by autocorrellograms (A–J). The autocorrelogram for each of the
10 MCs in our dataset was calculated from spontaneous spiking from the resting potential (no injected current). The majority of the
cells show a pattern that is more reminiscent of that seen for repetitively firing interneurons than the more bursty firing of CA1 or CA3
pyramidal cells (e.g. Csicsvari et al., 1998).
206

Spontaneous Evoked
20 mV
0.5 mS
IC

EC

50 microV 40 microV
0.5ms 0.5ms

Fig. 4. Extracellular waveforms of mossy cells during spontaneously and evoked spiking. A recording was obtained where the
intracellularly recorded mossy cell was also observed on the extracellular electrode. The extended shape of the mossy cell dendrites
allowed the extracellular signal to be observed on three shanks of the extracellular silicon probe (150 mm between shanks). The
extracellular electrode track is indicated by the white arrows. The soma of the mossy cell is indicated by the yellow arrow. There was a
difference in the shape of the action potentials, both intracellular and extracellular, suggesting differences in the site of spike initiation
for spontaneous and evoked spikes. (See Color Plate 12.4 in Color Plate Section.)

evidence to support the respective theories. Three known, and both low (Soltesz et al., 1993) and
days after kainic acid induced status epilepticus, high firing rates (Figs. 3, 5, and 9C) have been
there is reduction in inhibition and thus increase in observed under anesthesia. Bulk stimulation of
excitability of the dentate gyrus that correlates fiber pathways such as the perforant path input to
with the degree of MC loss (Zappone and Sloviter, the dentate gyrus is inherently not physiological in
2004). However, Ratzliff et al. (2004) showed that that the precise synchrony of synaptic input that
in the hippocampal slice, if they acutely removed results very rarely, if ever, happens in natural
MCs via manual destruction, there was a net processing. As such, the ratio of synchronously
reduction in the excitability of the dentate gyrus activated excitatory to inhibitory inputs may be
presumably due to the loss of direct excitatory very different than that observed during normal
input from MCs. In all likelihood, both hypotheses ongoing hippocampal function.
are at least partially correct and strict interpreta- Although the hilus also contains a large variety
tion of these studies is confounded by the technical of interneurons (Amaral, 1978; Mizumori et al.,
challenges of studying the MCs in vivo under non- 1990; Halasy and Somogyi, 1993; Buhl et al., 1994;
pathological conditions. The intrinsic firing rates Sik et al., 1997), the contribution of these inter-
of MCs in the intact unanesthetized animal is not neurons to the rich variety of dentate area network
207

patterns is not known. Our overriding hypothesis gyrus depend mainly on inputs from the
is that the main goal of both MCs and hilar entorhinal cortex (Bragin et al., 1995a;
interneurons is to control network patterns during Penttonen et al., 1998). In addition to the
various behaviors. This rationale has been suc- classical gamma frequency, slower (12–40 Hz;
cessfully applied to characterize CA1 interneurons beta) oscillations are also observed, often
(Csicsvari et al., 1999; Klausberger et al., 2003, with a larger amplitude in the hilus. It is not
2004, 2005), where the initial network pattern- clear whether the slow oscillation is simply a
based classification (theta phase and sharp waves) slower version of gamma or comprises a
in freely behaving rats was followed by juxtacel- physiologically distinct rhythm.
lular labeling of similarly characterized interneu- 3. The largest amplitude dentate gyrus event is
rons under anesthesia. We suggest that a similar the ‘‘dentate spike’’. This is a short duration
approach in the dentate gyrus can be equally fruit- (o60 ms), large amplitude (>0.5–2.5 mV)
ful. Here we have preliminary yet more advanced field potential characterized by synchronous
knowledge of the network contribution of MCs discharge of granule cells and interneurons
under anesthesia than in the behaving animals. and suppression of CA3 pyramidal cells
In the dentate area, several distinct oscillatory (Bragin et al., 1995b; Penttonen et al.,
patterns are present. 1997). Two types of dentate spikes have been
Network oscillations are known to involve a distinguished. The first type is a short burst of
rhythmic pattern of periods of active inhibition gamma oscillation consisting of 2–5 waves,
that are counterbalanced by periods of permissive one of which of excessively high amplitude
excitability (e.g. theta in CA1) (Buzsaki, 2002). with large sinks in the outer third of the
Perhaps the unique role of the MCs is to provide dentate molecular layer. The second type,
active excitation to granule cells in the periods observed in a subset of animals, has a some-
between the rhythmic inhibitory inputs driven by what different voltage vs. depth profile with a
interneurons. Although all the various patterns are large sink located in the middle third of the
mediated by a limited set of excitatory pathways, it dentate molecular layer. Our unpublished
is expected that MCs and the various interneuron observations suggest that type 2 dentate
groups may differentially participate in these net- spikes occur when thalamocortical high volt-
work patterns because the dynamics of activation age spindles (Buzsaki et al., 1988) invade the
can differentially affect neurons with different dentate area.
properties. The following patterns can be used to 4. Neocortical slow oscillations (Steriade et al.,
characterize the network contribution of MCs and 1990) also exert an impact on the firing
contrast them to granule cells, CA3 pyramidal cells patterns of the hippocampus, likely by way of
and hilar area interneuron types. the entorhinal cortex (Isomura et al., 2006;
Wolansky et al., 2006). During the UP state
1. In the exploring rat and during REM sleep, of slow oscillations, gamma power in the
large amplitude theta oscillations are present dentate gyrus and spiking of neurons in-
in the dentate gyrus. Theta waves in the dent- creases dramatically. In contrast, dentate
ate region are coherent with but phase-shifted gamma activity decreases during the DOWN
by approximately 2701 relative to the theta state (corresponding to delta waves of deep
oscillation in the CA1 pyramidal layer sleep in the neocortex) but CA3 pyramidal
(Buzsaki et al., 1983). cells may increase their firing rates and
2. Concurrent with theta, another prominent generate gamma oscillations (Isomura et al.,
pattern in the dentate area is the gamma fre- 2006)
quency oscillation (40–100 Hz). The power of 5. Sharp waves-ripple complexes (SPW) are truly
gamma oscillations is strongly phase modu- self-organized endogenous hippocampal
lated by the slower theta rhythm and both events that occur during slow-wave sleep,
theta and gamma oscillations in the dentate immobility and consummatory behaviors
208

(Buzsaki et al., 1983). They arise in the CA3 A


recurrent system and can spread to the CA1
region and the dentate region. The transient 1
excitation of CA1 neurons gives rise to a 0.2 mV
short-lived fast oscillation (‘‘ripple’’) 10 mV
0.5 s
(O’Keefe and Nadel, 1978; Buzsáki et al.,
1992; Ylinen et al., 1995). No ripples are as-
sociated with SPW in the dentate area but 2
putative interneurons and, occasionally,
granule cells are depolarized and discharge
in synchrony with CA1 ripples, although
the typical response is hyperpolarization
(Penttonen et al., 1997).
B
We submit that these five unique population
patterns can be used in future studies to function- 1
ally classify dentate region neurons in the freely
behaving animal. In turn, these same patterns can
0.1 mV
be used in anesthetized rats where they can be 10 mV
labeled by intracellular or juxtacellular methods. 0.5 s
The network-based classification should be com-
bined with the spike dynamics and wave-shape
2
features of extracellular spikes. Fortunately, the
repertoire of rat hippocampal network activity
patterns under urethane anesthesia largely reflects
what is observed in the drug-free rat. Using this
approach we have been able to observe how MCs
behave in relation to some of these naturally
occurring network patterns. Fig. 5. (A) Example of mossy cell inhibition by transition to
theta rhythm evoked by tail pinch in urethane anesthetized rat.
Extracellular recording (A1) from area CA1 recorded simulta-
Theta neously with a hilar mossy cell (A2). A tail pinch was applied at
the arrow and the mossy cell responded by slowing its firing
rate. (B) Example of mossy cell excitation by transition to theta
It has been previously reported that MC mem- rhythm evoked by tail pinch in urethane anesthetized rat. Ex-
brane potential shows rhythmic oscillations in the tracellular recording (B1) from area CA3 recorded simultane-
theta band that are phase-locked to the extracel- ously with (B2) a hilar mossy cell that was excited by tail pinch
lular theta oscillation in the contralateral hippo- (arrow) evoked theta.
campus (Soltesz et al., 1993). However, as noted
above, this study reported a very low (o1 Hz) peak of the locally derived theta oscillation
basal firing rate for the MCs. Nevertheless, our (Fig. 6) and coherent with the discharge of some
observations support the involvement of MCs in interneuron types. We predict that MCs in the
theta oscillations. Individual MCs can either be behaving rat will keep a similar relationship to
depolarized (8 of 10) or hyperpolarized (2 of 10) by local hilar/CA3c theta oscillations.
a transition from slow wave sleep to theta evoked
by a tail pinch (Fig. 5). This behavior is similar Beta/gamma oscillations
to pyramidal cells (Kamondi et al., 1998). In
addition, the membrane potential of MCs shows a The power of gamma frequency oscillation in
co-variation with the extracellular field, with peak the hilus is phase modulated by the slower theta
depolarization and discharge slightly after the (Bragin et al., 1995a). This gamma frequency
209

Fig. 6. (A) Continuous theta/delta power ratio from an extracellular electrode located in the hilus. The increases in theta power at 45
and 200 s was induced via tail pinch. (B) rastergrams of isolated units recorded from three extracellular tetrodes in the hilus/area CA3c.
(C) Intracellular membrane potential recorded from a mossy cell during this same recording epoch while passing hyperpolarizing
current to maintain the resting potential near 80 mV. Notice that during the periods of theta power increase, the mossy cell
experiences a depolarization from the 80 mV holding potential. (D) Two examples of average mossy cell membrane potential time
aligned to the two ‘‘theta on’’ putative interneuronal units indicated by the start symbols in (C). The upper trace is the average
membrane potential of the MC. The lower trace is the average wide-band extracellular trace. The autocorrelogram and waveform for
the isolated IN units are also shown as insets.
210

A Gamma cycle

Theta oscillation

20µV
EC unit for MC
200 ms

1mV
B
20 ms

25µV
1ms

Fig. 7. (A) Average of extracellular hilar field potential aligned to the peaks of the intracellular MC spontaneous action potentials
recorded over a 200 sec recording period. This is the same case as in Fig. 4 where the extracellular electrode picked up the same MC as
was recorded intracellularly. The extracellular unit correlate of the intracellular AP is indicated. Notice that the MC fires on the end of
the gamma cycle waveform and there is an overall theta oscillation present that is time aligned to the intracellular spike. (B) Red traces:
MC membrane potential averaged and time aligned on the spike times of isolated interneurons recorded in the hilus/CA3c. Left
column: IN unit autocorrelogram, black traces: average IN unit waveforms. Note similar phase relationship of the interneuron-timed
intracellular gamma frequency oscillation.

modulation is also evident in the relationship be- membrane oscillation of MCs in the gamma fre-
tween the local field and fast spiking interneurons quency band becomes evident (Fig. 7B). Similar
and their effect on the MC (Fig. 7). When the ac- phase-locked behavior has been also observed in
tion potentials of MCs are used as reference for the beta oscillation frequency range as well.
averaging local field potentials, phaselocking of
MC spikes to the gamma oscillation, superimposed Slow oscillations
on the peak of theta waves becomes evident
(Fig. 7A). Furthermore, when nearby fast-spiking Slow oscillations arise in neocortical networks
putative interneurons are used as the reference, (Steriade et al., 1993a–c) and spread to the
211

Fig. 8. (A) CA3 unit activity and MC membrane potential fluctuations. Mossy cells show synaptic activity correlated with unit activity
in CA3 (2/2 cells recorded with extracellular electrode placed in CA3c. Upper traces show extracellular wideband recording middle
trace bandpass filtered between 800 Hz and 3 kHz. Lower traces show intracellular membrane potential. (B) Cross-correlation between
spontaneous MC spikes recorded intracellularly and all extracellular units combined. Lower panels are shuffled controls. (C) Hilar/
CA3c population bursts correlate with spontaneous mossy cell spikes. Hilar/CA3c bursts were defined as five unit spikes with less than
20 ms inter-spike interval. The spikes could originate from any unit. Shuffled correlograms are shown for controls. (D) Cross-
correlation between spontaneous mossy cell spikes and all hilar/CA3 units combined for a second extracellular/intracellular recording
pair.

entorhinal cortex and subiculum from where transition of the entorhinal DOWN–UP state, the
they can invade the dentate gyrus as well (Isomura surge of excitation induces a strong discharge of
et al., 2006; Wolansky et al., 2006). At the single granule cells and also gamma frequency oscilla-
cell level, most cortical neurons show a bimodal tions (Isomura et al., 2006). Often one of these
distribution of the membrane potential and these gamma waves becomes excessively large and has
UP and DOWN states alternate relatively rhyth- been referred to as the dentate spike (Bragin et al.,
mically at 0.5–2 Hz. Slow oscillations are most 1995b). In this case, the MC firing is likely timed
prominent under anesthesia but are also present in by feed forward excitation from dentate gyrus
deep stages of slow wave sleep (Achermann and granule cells that are driven by the entorhinal
Borbely, 1997), where the transient delta waves input. The surge of activity in the input from
correspond to the DOWN (or silent) state. At the entorhinal cortex is also reflected by the sudden
212

Fig. 9. CA1 sharpwave/ripples are associated with inhibition of mossy cells. (A) Example of a CA1 sharpwave/ripple complex
recorded extracellularly in the pyramidal layer of CA1. (B) Intracellular spiking activity of mossy cell is suppressed during SPW. (C, D)
Temporal expansion of region in box in A and B. Mossy cell membrane response during SPW/ripples has a reversal near 60 mV.
(E, F) Average extracellular SPW and mossy cell membrane potential from a hyperpolarized baseline (E; n ¼ 12 SPWs) and a more
depolarized baseline (F; n ¼ 11 SPWs). (G) Plot of the peak change in mossy cell membrane potential during CA1 SPWs from 15
recording epochs at different membrane potentials from four mossy cells (different symbols). The best fit linear regression line is shown
(R ¼ 0.57621, Po0.025).
213

increase of population firing in the dentate. We expected that future studies will clarify whether
exploited the high levels of CA3 activity that occur MCs can become active participants in SPW
during the slow oscillatory pattern for the exam- events. It is worth noting here that in slices treated
ination of the behavior of MCs. The transition with a GABAA receptor antagonist, or slices from
from DOWN to UP state was defined when a an epileptic rat, CA3 and MCs are engaged in
group of dentate/hilar/CA3 neurons fired a burst population bursts while granule cells remain
of activity. This analysis revealed that MCs be- hyperpolarized, suggesting that CA3 pyramidal
come periodically active and silent according to cells may directly excite MCs (Scharfman, 1994a;
the level of activity in the entorhinal–hippocampal Scharfman et al., 2001). The failure of MCs to
network (Fig. 8). discharge during SPWs in the intact brain would
imply that the hypothesized SPW-mediated con-
solidation of synaptic circuits in the CA3–CA1
CA1 sharp waves networks (Buzsáki, 1989) can proceed independent
of the modification of the synapses established by
Although SPW arises in the CA3–CA1 regions, the MCs.
the recurrent axon collaterals of CA3 can directly
excite hilar interneurons and even granule cells
(Li et al., 1994; Penttonen et al., 1997). Typically, Conclusion
feed-forward inhibition prevails by this mechanism
as reflected by the SPW-locked hyperpolarization Although MCs are well-known and critical com-
of granule cells. However, this occasional failure of ponents of the dentate circuitry, their physiological
inhibition and/or a concerted activation of a single function and exact involvement in various hyper-
granule cell by the recurrent CA3 axons can excitable phenomena has remained elusive. A
robustly discharge granule cells during SPWs major technical problem is the lack of reliable
(Penttonen et al., 1997). Our observations, to physiological criteria that may be used in extra-
date, suggest that SPWs are associated with net cellular recordings in freely behaving animal for
inhibition of MCs. Figure 9A–D show the tempo- the positive identification of MCs. Our intracellu-
ral relationship between a spontaneous SPW lar characterization of some of their biophysical
recorded in area CA1 (Fig. 9A, B) while record- features and network-related behavior are the first
ing from a MC in the hilus (Fig. 9C, D). As can be steps in this direction.
appreciated from this example, the ongoing spon-
taneous firing of the MC is inhibited in the period
Abbreviations
overlapping with and following the SPW/ripple
complex. This inhibitory effect is probably medi-
CCK cholecystokinin
ated via activation of chloride flux through
EPSP excitatory postsynaptic potential
GABAA receptors since the change in MC mem-
GABA gamma-aminobutyric acid
brane potential associated with a CA1 SPW has a
GAD glutamic acid decarboxylase
reversal potential near 60 mV similar to GABAA
MC mossy cell
reversal potentials in vivo (Fig. 9E–G). The
NPY neuropeptide Y
potential source of this inhibition is the increased
SPW-related firing of hilar interneurons driven
either directly by the recurrent CA3 collaterals
or the rarely discharging granule cells (Penttonen References
et al., 1997). It appears that, at least under an-
Achermann, P. and Borbely, A.A. (1997) Low-frequency
esthesia, the strong inhibition can prevent MCs
(o1 Hz) oscillations in the human sleep electroencephalo-
from discharging in response to their granule cell gram. Neuroscience, 81: 213–222.
inputs during SPWs. However, the situation might Acsady, L., Katona, I., Martinez-Guijarro, F.J., Buzsaki, G.
be quite different in the drug-free animal and it is and Freund, T.F. (2000) Unusual target selectivity of
214

perisomatic inhibitory cells in the hilar region of the rat hip- Buzsaki, G., Leung, L.W. and Vanderwolf, C.H. (1983) Cellu-
pocampus. J. Neurosci., 20: 6907–6919. lar bases of hippocampal EEG in the behaving rat. Brain
Amaral, D.G. (1978) A Golgi study of cell types in the hilar Res., 287: 139–171.
region of the hippocampus in the rat. J. Comp. Neurol., 182: Cajal, S.R. (1911) Histologie du Systeme Nerveux de l’Homme
851–914. et des Vertebres. Oxford University Press, New York.
Amaral, D.G., Ishizuka, N. and Claiborne, B. (1990) Neurons, Chicurel, M.E. and Harris, K.M. (1992) Three-dimensional
numbers and the hippocampal network. Prog. Brain Res., analysis of the structure and composition of CA3 branched
83(1–11): 1–11. dendritic spines and their synaptic relationships with mossy
Bilkey, D.K. and Schwartzkroin, P.A. (1990) Variation in fiber boutons in the rat hippocampus. J. Comp. Neurol., 325:
electrophysiology and morphology of hippocampal CA3 py- 169–182.
ramidal cells. Brain Res., 514: 77–83. Csicsvari, J., Hirase, H., Czurko, A. and Buzsaki, G. (1998)
Bragin, A., Jando, G., Nadasdy, Z., Hetke, J., Wise, K. Reliability and state dependence of pyramidal cell-interneu-
and Buzsaki, G. (1995a) Gamma (40–100 Hz) oscillation ron synapses in the hippocampus: an ensemble approach in
in the hippocampus of the behaving rat. J. Neurosci., 15: the behaving rat. Neuron, 21: 179–189.
47–60. Csicsvari, J., Hirase, H., Czurko, A., Mamiya, A. and Buzsáki,
Bragin, A., Jando, G., Nadasdy, Z., van Landeghem, M. and G. (1999) Oscillatory coupling of hippocampal pyramidal
Buzsaki, G. (1995b) Dentate EEG spikes and associated cells and interneurons in the behaving rat. J. Neurosci., 19:
interneuronal population bursts in the hippocampal hilar 274–287.
region of the rat. J. Neurophysiol., 73: 1691–1705. Deller, T. (1998) The anatomical organization of the rat fascia
Buckmaster, P.S. and Amaral, D.G. (2001) Intracellular re- dentata: new aspects of laminar organization as revealed by
cording and labeling of mossy cells and proximal CA3 py- anterograde tracing with phaseolus vulgaris-luecoagglutinin
ramidal cells in macaque monkeys. J. Comp. Neurol., 430: (PHAL). Anat. Embryol. (Berl.), 197: 89–103.
264–281. Deller, T., Katona, I., Cozzari, C., Frotscher, M. and Freund,
Buckmaster, P.S. and Schwartzkroin, P.A. (1994) Hippocampal T.F. (1999) Cholinergic innervation of mossy cells in the rat
mossy cell function: a speculative view. Hippocampus, 4: fascia dentata. Hippocampus, 9: 314–320.
393–402. Deller, T. and Leranth, C. (1990) Synaptic connections of ne-
Buckmaster, P.S. and Schwartzkroin, P.A. (1995) Interneurons uropeptide Y (NPY) immunoreactive neurons in the hilar
and inhibition in the dentate gyrus of the rat in vivo. J. area of the rat hippocampus. J. Comp. Neurol., 300:
Neurosci., 15: 774–789. 433–447.
Buckmaster, P.S., Strowbridge, B.W., Kunkel, D.D., Schmiege, Freund, T.F. and Buzsáki, G. (1996) Interneurons of the hip-
D.L. and Schwartzkroin, P.A. (1992) Mossy cell axonal pocampus. Hippocampus, 6: 347–470.
projections to the dentate gyrus molecular layer in the rat Freund, T.F., Hajos, N., Acsady, L., Gorcs, T.J. and Katona, I.
hippocampal slice. Hippocampus, 2: 349–362. (1997) Mossy cells of the rat dentate gyrus are immunore-
Buckmaster, P.S., Strowbridge, B.W. and Schwartzkroin, active for calcitonin gene-related peptide (CGRP). Eur. J.
P.A. (1993) A comparison of rat hippocampal mossy Neurosci., 9: 1815–1830.
cells and CA3c pyramidal cells. J. Neurophysiol., 70: Frotscher, M., Seress, L., Schwerdtfeger, W.K. and Buhl, E.
1281–1299. (1991) The mossy cells of the fascia dentata: a comparative
Buckmaster, P.S., Wenzel, H.J., Kunkel, D.D. and Schwa- study of their fine structure and synaptic connections in ro-
rtzkroin, P.A. (1996) Axon arbors and synaptic connections dents and primates. J. Comp. Neurol., 312: 145–163.
of hippocampal mossy cells in the rat in vivo. J. Comp. Fujise, N. and Kosaka, T. (1999) Mossy cells in the mouse
Neurol., 366: 271–292. dentate gyrus: identification in the dorsal hilus and their dis-
Buhl, E.H., Han, Z.S., Lorinczi, Z., Stezhka, V.V., Karnup, tribution along the dorsoventral axis. Brain Res., 816:
S.V. and Somogyi, P. (1994) Physiological properties of an- 500–511.
atomically identified axo-axonic cells in the rat hippocampus. Goodman, J.H., Wasterlain, C.G., Massarweh, W.F., Dean, E.,
J. Neurophysiol., 71: 1289–1307. Sollas, A.L. and Sloviter, R.S. (1993) Calbindin-D28k
Buzsáki, G. (1989) Two-stage model of memory trace forma- immunoreactivity and selective vulnerability to ischemia in
tion: a role for ‘‘noisy’’ brain states. Neuroscience, 31: the dentate gyrus of the developing rat. Brain Res., 606:
551–570. 309–314.
Buzsaki, G. (2002) Theta oscillations in the hippocampus. Gutierrez, R. (2003) The GABAergic phenotype of the
Neuron, 33: 325–340. ‘‘glutamatergic’’ granule cells of the dentate gyrus. Prog.
Buzsaki, G., Bickford, R.G., Armstrong, D.M., Ponomareff, Neurobiol., 71: 337–358.
G., Chen, K.S., Ruiz, R., Thal, L.J. and Gage, F.H. (1988) Hafting, T., Fyhn, M., Molden, S., Moser, M.B. and Moser,
Electric activity in the neocortex of freely moving young and E.I. (2005) Microstructure of a spatial map in the entorhinal
aged rats. Neuroscience, 26: 735–744. cortex. Nature, 436: 801–806.
Buzsáki, G., Horvath, Z., Urioste, R., Hetke, J. and Wise, K. Halasy, K. and Somogyi, P. (1993) Subdivisions in the multiple
(1992) High-frequency network oscillation in the hippocam- GABAergic innervation of granule cells in the dentate gyrus
pus. Science, 256: 1025–1027. of the rat hippocampus. Eur. J. Neurosci., 5: 411–429.
215

Henze, D.A. and Buzsaki, G. (2001) Action potential threshold Penttonen, M., Kamondi, A., Acsady, L. and Buzsaki, G.
of hippocampal pyramidal cells in vivo is increased by recent (1998) Gamma frequency oscillation in the hippocampus of
spiking activity. Neuroscience, 105: 121–130. the rat: intracellular analysis in vivo. Eur. J. Neurosci., 10:
Henze, D.A., McMahon, D.B., Harris, K.M. and Barrionuevo, 718–728.
G. (2002) Giant miniature EPSCs at the hippocampal mossy Penttonen, M., Kamondi, A., Sik, A., Acsady, L. and Buzsaki,
fiber to CA3 pyramidal cell synapse are monoquantal. J. G. (1997) Feed-forward and feed-back activation of the
Neurophysiol., 87: 15–29. dentate gyrus in vivo during dentate spikes and sharp wave
Isomura, Y., Sirota, A., Montgomery, S., Mizuseki, K., Özen, bursts. Hippocampus, 7: 437–450.
S., Henze, D.A. and Buzsáki, G. (2006) Integration and Seg- Petralia, R.S. and Wenthold, R.J. (1992) Light and electron
regation of Activity in Entorhinal-hippocampal Subregions immunocytochemical localization of AMPA-selective gluta-
by Neocortical Slow Oscillations. Neuron, 52: 871–882. mate receptors in the rat brain. J. Comp. Neurol., 318:
Kamondi, A., Acsady, L., Wang, X.J. and Buzsáki, G. (1998) 329–354.
Theta oscillations in somata and dendrites of hippocampal Ratzliff, A.H., Howard, A.L., Santhakumar, V., Osapay, I. and
pyramidal cells in vivo: activity-dependent phase-precession Soltesz, I. (2004) Rapid deletion of mossy cells does not result
of action potentials. Hippocampus, 8: 244–261. in a hyperexcitable dentate gyrus: implications for epilepto-
Kanerva, P. (1988) Sparse Distributed Memory. Bradford/MIT genesis. J. Neurosci., 24: 2259–2269.
Press, Cambridge, MA. Ratzliff, A.H., Santhakumar, V., Howard, A. and Soltesz, I.
Klausberger, T., Magill, P.J., Marton, L.F., Roberts, J.D., (2002) Mossy cells in epilepsy: rigor mortis or vigor mortis?
Cobden, P.M., Buzsaki, G. and Somogyi, P. (2003) Brain- Trends Neurosci., 25: 140–144.
state- and cell-type-specific firing of hippocampal interneu- Santhakumar, V., Bender, R., Frotscher, M., Ross, S.T.,
rons in vivo. Nature, 421: 844–848. Hollrigel, G.S., Toth, Z. and Soltesz, I. (2000) Granule cell
Klausberger, T., Marton, L.F., Baude, A., Roberts, J.D., hyperexcitability in the early post-traumatic rat dentate
Magill, P.J. and Somogyi, P. (2004) Spike timing of dendrite- gyrus: the ‘irritable mossy cell’ hypothesis. J. Physiol.,
targeting bistratified cells during hippocampal network os- 524(Pt 1): 117–134.
cillations in vivo. Nat. Neurosci., 7: 41–47. Scharfman, H.E. (1991) Dentate hilar cells with dendrites in the
Klausberger, T., Marton, L.F., O’Neill, J., Huck, J.H., molecular layer have lower thresholds for synaptic activation
Dalezios, Y., Fuentealba, P., Suen, W.Y., Papp, E., Kaneko, by perforant path than granule cells. J. Neurosci., 11:
T., Watanabe, M., Csicsvari, J. and Somogyi, P. (2005) 1660–1673.
Complementary roles of cholecystokinin- and parvalbumin- Scharfman, H.E. (1992a) Blockade of excitation reveals inhi-
expressing GABAergic neurons in hippocampal network os- bition of dentate spiny hilar neurons recorded in rat hippo-
cillations. J. Neurosci., 25: 9782–9793. campal slices. J. Neurophysiol., 68: 978–984.
Kohler, C. (1985) A projection from the deep layers of the Scharfman, H.E. (1992b) Differentiation of rat dentate neurons
entorhinal area to the hippocampal formation in the rat by morphology and electrophysiology in hippocampal slices:
brain. Neurosci. Lett., 56: 13–19. granule cells, spiny hilar cells and aspiny ‘fast-spiking’ cells.
Leranth, C., Szeidemann, Z., Hsu, M. and Buzsaki, G. (1996) Epilepsy Res. Suppl., 7: 93–109.
AMPA receptors in the rat and primate hippocampus: a Scharfman, H.E. (1993a) Characteristics of spontaneous and
possible absence of GluR2/3 subunits in most interneurons. evoked EPSPs recorded from dentate spiny hilar cells in rat
Neuroscience, 70: 631–652. hippocampal slices. J. Neurophysiol., 70: 742–757.
Li, X.G., Somogyi, P., Ylinen, A. and Buzsaki, G. (1994) The Scharfman, H.E. (1993b) Spiny neurons of area CA3c in
hippocampal CA3 network: an in vivo intracellular labeling rat hippocampal slices have similar electrophysiological
study. J. Comp. Neurol., 339: 181–208. characteristics and synaptic responses despite morphological
Lorente de No, R. (1934) Studies on the structure of the cer- variation. Hippocampus, 3: 9–28.
ebral cortex. II. Continuation of the study of the ammonic Scharfman, H.E. (1994a) EPSPs of dentate gyrus granule cells
system. J. Psychol. Neurol. (Leipzig), 46: 113–177. during epileptiform bursts of dentate hilar ‘‘mossy’’ cells and
McNaughton, B.L. and Morris, R.G.M. (1987) Hippocampal area CA3 pyramidal cells in disinhibited rat hippocampal
synaptic enhancement and information storage within a dis- slices. J. Neurosci., 14: 6041–6057.
tributed memory system. TINS, 10(10): 408–415. Scharfman, H.E. (1994b) Evidence from simultaneous intracel-
Mizumori, S.J., Barnes, C.A. and McNaughton, B.L. (1990) lular recordings in rat hippocampal slices that area CA3
Behavioral correlates of theta-on and theta-off cells recorded pyramidal cells innervate dentate hilar mossy cells. J. Ne-
from hippocampal formation of mature young and aged rats. urophysiol., 72: 2167–2180.
Exp. Brain Res., 80: 365–373. Scharfman, H.E. (1994c) Synchronization of area CA3 hippo-
Murakawa, R. and Kosaka, T. (2001) Structural features of campal pyramidal cells and non-granule cells of the dentate
mossy cells in the hamster dentate gyrus, with special refer- gyrus in bicuculline-treated rat hippocampal slices. Neuro-
ence to somatic thorny excrescences. J. Comp. Neurol., 429: science, 59: 245–257.
113–126. Scharfman, H.E. (1995) Electrophysiological evidence that
O’Keefe, J. and Nadel, L. (1978) Hippocampus as a cognitive dentate hilar mossy cells are excitatory and innervate both
map. Clarendon, Oxford. granule cells and interneurons. J. Neurophysiol., 74: 179–194.
216

Scharfman, H.E., Kunkel, D.D. and Schwartzkroin, P.A. Steriade, M., Contreras, D., Curro, D.R. and Nunez, A.
(1990) Synaptic connections of dentate granule cells and (1993a) The slow (o1 Hz) oscillation in reticular thalamic
hilar neurons: results of paired intracellular recordings and and thalamocortical neurons: scenario of sleep rhythm
intracellular horseradish peroxidase injections. Neuroscience, generation in interacting thalamic and neocortical networks.
37: 693–707. J. Neurosci., 13: 3284–3299.
Scharfman, H.E. and Schwartzkroin, P.A. (1988) Electrophys- Steriade, M., Gloor, P., Llinas, R.R., Lopes de Silva, F.H. and
iology of morphologically identified mossy cells of the dent- Mesulam, M.M. (1990) Report of IFCN committee on basic
ate hilus recorded in guinea pig hippocampal slices. J. mechanisms. Basic mechanisms of cerebral rhythmic activi-
Neurosci., 8: 3812–3821. ties. Electroencephalogr. Clin. Neurophysiol., 76: 481–508.
Scharfman, H.E., Smith, K.L., Goodman, J.H. and Sollas, A.L. Steriade, M., Nunez, A. and Amzica, F. (1993b) A novel slow
(2001) Survival of dentate hilar mossy cells after pilocarpine- (o1 Hz) oscillation of neocortical neurons in vivo: depolar-
induced seizures and their synchronized burst discharges with izing and hyperpolarizing components. J. Neurosci., 13:
area CA3 pyramidal cells. Neuroscience, 104: 741–759. 3252–3265.
Sik, A., Penttonen, M. and Buzsáki, G. (1997) Interneurons in Steriade, M., Nunez, A. and Amzica, F. (1993c) Intracellular
the hippocampal dentate gyrus: an in vivo intracellular study. analysis of relations between the slow (o1 Hz) neocortical
Eur. J. Neurosci., 9: 573–588. oscillation and other sleep rhythms of the electroencephalo-
Sloviter, R.S. (1989) Calcium-binding protein (calbindin-D28k) gram. J. Neurosci., 13: 3266–3283.
and parvalbumin immunocytochemistry: localization in the Strowbridge, B.W., Buckmaster, P.S. and Schwartzkroin, P.A.
rat hippocampus with specific reference to the selective (1992) Potentiation of spontaneous synaptic activity in rat
vulnerability of hippocampal neurons to seizure activity. mossy cells. Neurosci. Lett., 142: 205–210.
J. Comp. Neurol., 280: 183–196. Treves, A. and Rolls, E.T. (1992) Computational constraints
Sloviter, R.S. (1991a) Feedforward and feedback inhibition of suggest the need for two distinct input systems to the hip-
hippocampal principal cell activity evoked by perforant path pocampal CA3 network. Hippocampus, 2: 189–199.
stimulation: GABA-mediated mechanisms that regulate ex- Treves, A. and Rolls, E.T. (1994) Computational analysis of the
citability in vivo. Hippocampus, 1: 31–40. role of the hippocampus in memory. Hippocampus, 4:
Sloviter, R.S. (1991b) Permanently altered hippocampal struc- 374–391.
ture, excitability, and inhibition after experimental status ep- Walker, M.C., Ruiz, A. and Kullmann, D.M. (2001) Monosy-
ilepticus in the rat: the ‘‘dormant basket cell’’ hypothesis and naptic GABAergic signaling from dentate to CA3 with a
its possible relevance to temporal lobe epilepsy. Hippocam- pharmacological and physiological profile typical of mossy
pus, 1: 41–66. fiber synapses. Neuron, 29: 703–715.
Sloviter, R.S. (1994) The functional organization of the hippo- Wenzel, H.J., Buckmaster, P.S., Anderson, N.L., Wenzel, M.E.
campal dentate gyrus and its relevance to the pathogenesis of and Schwartzkroin, P.A. (1997) Ultrastructural localization
temporal lobe epilepsy. Ann. Neurol., 35: 640–654. of neurotransmitter immunoreactivity in mossy cell axons
Sloviter, R.S., Sollas, A.L., Barbaro, N.M. and Laxer, K.D. and their synaptic targets in the rat dentate gyrus. Hippo-
(1991) Calcium-binding protein (calbindin-D28k) and campus, 7: 559–570.
parvalbumin immunocytochemistry in the normal and Wittner, L., Henze, D.A., Zaborszky, L. and Buzsaki, G. (2006)
epileptic human hippocampus. J. Comp. Neurol., 308: Hippocampal CA3 pyramidal cells selectively innervate
381–396. aspiny interneurons. Eur. J. Neurosci., 24: 1286–1298.
Sloviter, R.S., Zappone, C.A., Harvey, B.D., Bumanglag, A.V., Wolansky, T., Clement, E.A., Peters, S.R., Palczak, M.A. and
Bender, R.A. and Frotscher, M. (2003) ‘‘Dormant basket Dickson, C.T. (2006) Hippocampal slow oscillation: a novel
cell’’ hypothesis revisited: relative vulnerabilities of dentate EEG state and its coordination with ongoing neocortical ac-
gyrus mossy cells and inhibitory interneurons after hippo- tivity. J. Neurosci., 26: 6213–6229.
campal status epilepticus in the rat. J. Comp. Neurol., 459: Ylinen, A., Bragin, A., Nadasdy, Z., Jando, G., Szabo, I., Sik,
44–76. A. and Buzsáki, G. (1995) Sharp wave-associated high-fre-
Soltesz, I., Bourassa, J. and Deschenes, M. (1993) The behavior quency oscillation (200 Hz) in the intact hippocampus: net-
of mossy cells of the rat dentate gyrus during theta oscilla- work and intracellular mechanisms. J. Neurosci., 15: 30–46.
tions in vivo. Neuroscience, 57: 555–564. Zappone, C.A. and Sloviter, R.S. (2004) Translamellar disinhi-
Soriano, E. and Frotscher, M. (1994) Mossy cells of the rat bition in the rat hippocampal dentate gyrus after seizure-in-
fascia dentata are glutamate-immunoreactive. Hippocampus, duced degeneration of vulnerable hilar neurons. J. Neurosci.,
4: 65–69. 24: 853–864.
Spontaneous Evoked
20 mV
0.5 mS
IC

EC

50 microV 40 microV
0.5ms 0.5ms

Plate 12.4. Extracellular waveforms of mossy cells during spontaneously and evoked spiking. A recording was obtained where the
intracellularly recorded mossy cell was also observed on the extracellular electrode. The extended shape of the mossy cell dendrites
allowed the extracellular signal to be observed on three shanks of the extracellular silicon probe (150 mm between shanks). The
extracellular electrode track is indicated by the white arrows. The soma of the mossy cell is indicated by the yellow arrow. There was a
difference in the shape of the action potentials, both intracellular and extracellular, suggesting differences in the site of spike initiation
for spontaneous and evoked spikes. (For B/W version, see page 206 in the volume.)
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 13

Interneurons of the dentate gyrus: an overview of cell


types, terminal fields and neurochemical identity

Carolyn R. Houser1,2,

1
Department of Neurobiology and Brain Research Institute, David Geffen School of Medicine at UCLA,
Los Angeles, CA 90095, USA
2
Research Service, VA Greater Los Angeles Healthcare System, West Los Angeles, Los Angeles, CA 90073, USA

Abstract: Interneurons of the dentate gyrus are a diverse group of neurons that use GABA as their primary
neurotransmitter. Morphological studies of these neurons have been challenging since no single neuro-
anatomical method provides a complete view of these interneurons. However, through the integration of
findings obtained from multiple methods, an interesting picture of this complex group of neurons is
emerging, and this review focuses on studies in rats and mice. In situ hybridization of mRNAs for the two
isoforms of the GABA synthesizing enzyme, glutamate decarboxylase (GAD65 and GAD67), demonstrates
the abundance of GABA neurons in the dentate gyrus and their high concentration in the hilus and along
the base of the granule cell layer. Likewise, immunohistochemical studies, particularly of GAD65, dem-
onstrate the rich fields of GABA terminals not only around the somata of granule cells but also in the
dendritic regions of the molecular layer. This broad group of GABA neurons and their terminals can be
subdivided according to their morphological characteristics, including the distribution of their axonal
plexus, and their neurochemical identity. Intracellular labeling of single interneurons has been instrumental
in demonstrating the extensiveness of their axonal plexus and the relatively specific spatial distribution of
their axonal fields. These findings have led to the broad classification of interneurons into those that
terminate primarily at perisomatic regions and those that innervate the dendrites of granule cells. The
interneurons also can be classified according to their neuropeptide and calcium-binding protein content.
These and other molecules contribute to the rich diversity of dentate interneurons and may provide
opportunities for selectively regulating specific groups of GABA neurons in the dentate gyrus in order to
enhance their function or protect vulnerable neurons from damage.

Keywords: GABA; glutamate decarboxylase (GAD); interneurons; hilus; parvalbumin; somatostatin;


calretinin

Introduction early morphological studies identified at least 21


cell types in the dentate hilus alone (Amaral,
Interneurons of the dentate gyrus are a particu- 1978). Despite such diversity, a unifying feature of
larly fascinating and diverse group of neurons, and the dentate interneurons is their use of g-amino-
butyric acid (GABA) as their major neurotrans-
Corresponding author. Tel.: +1 (310) 206-1567; mitter. However, as in other cortical and
Fax: +1 (310) 825-2224; E-mail: houser@mednet.ucla.edu
hippocampal regions, multiple subtypes of GABA

DOI: 10.1016/S0079-6123(07)63013-1 217


218

neurons can be distinguished, and several classifi- both isoforms appear to be present in essentially
cation systems have emerged. These include cate- all GABAergic interneurons of the dentate gyrus
gorization according to the axonal distributions (Houser and Esclapez, 1994). Thus clear visuali-
and postsynaptic targets of the interneurons, their zation of the entire population of GABA neuron
neuropeptide or calcium-binding protein content, somata is possible, and the abundance of these
and their physiological characteristics (Parra et al., neurons within the relatively small region of the
1998; Maccaferri and Lacaille, 2003; Somogyi and dentate gyrus is striking (Fig. 1). Such in situ
Klausberger, 2005; Jinno and Kosaka, 2006). Each hybridization methods for identifying GABA
system highlights special features of the interneu- neurons have the additional advantage of labeling
ron population, and, while there is not a perfect the cell bodies of these GABA neurons without
overlap between categories in the different sys- occlusion by axon terminals.
tems, some relationships can be established. This GAD mRNA-labeled neurons are distributed in
review will focus first on the broad distribution of all three relatively distinct regions of the dentate
GABA neuron cell bodies and axon terminals gyrus. First, GAD mRNA-labeled neurons are
within the dentate gyrus, and then consider several highly concentrated in the hilus, and their distri-
major categories of dentate interneurons and their bution often helps define the hilar region (Fig. 1A,
relationships to the overall distribution of GABA B). While labeled interneurons are abundant in the
neurons and their processes. deep parts of the hilus, between the tip of the hilus
and CA3, they are also numerous in the hilar
regions that are located beneath the granule cell
Distribution of GABA neurons layer (Fig. 1A, C). Second, labeled neurons asso-
ciated with the granule cell layer are frequently
Demonstrating the entire population of GABA aligned along the base of this layer and may be
neurons in the dentate gyrus in histological prep- either highly concentrated or more dispersed along
arations has proven to be an interesting challenge. this zone (Fig. 1A–C). Some GAD mRNA-labeled
Immunohistochemical studies utilizing various an- cell bodies are also located within the granule cell
tisera to GABA and its synthesizing enzyme, layer and along its outer border. Finally, GABA
glutamate decarboxylase (GAD), have yielded neurons are distributed diffusely within the mo-
different and, occasionally, contradictory results lecular layer, and a few are found along the hip-
(Ribak et al., 1978; Sloviter and Nilaver, 1987; pocampal fissure at the outer border of the
Babb et al., 1988; Woodson et al., 1989; Jinno et molecular layer (Fig. 1A, C). These basic patterns
al., 1998). Although cell bodies in many regions of of GABA neurons are present in both the dorsal
the hippocampal formation are labeled with these (septal) and ventral (temporal) regions of the dent-
methods, several groups of neurons in the dentate ate gyrus (Fig. 1B).
gyrus, particularly those in the hilus, are often in- The numbers of GAD mRNA-labeled neurons
consistently labeled. Alternate methods of identi- in all three regions (hilus, granule cell layer and
fying the cell bodies of GABA neurons in the molecular layer) generally appear to be larger than
dentate gyrus are thus needed. those identified with other labeling methods, and
Currently, one of the most effective methods for the numerous labeled neurons in the hilus are par-
demonstrating the entire population of GABA ticularly notable. Their abundance could give the
neurons in the dentate gyrus is in situ hybridiza- impression that most hilar neurons are GABA
tion of either GAD65 or GAD67 mRNA (Houser neurons. However, non-GABA neurons are also
et al., 2000, for review). Both forms of GAD can present in the hilus (Fig. 2A), in approximately
synthesize GABA, but they have functional differ- equal numbers, and the vast majority of these are
ences that include different interactions with the mossy cells that use glutamate as their transmitter.
cofactor pyridoxal phosphate (Erlander et al., Double labeling for GAD mRNA and a general
1991; Kaufman et al., 1991; Martin et al., 1991; neuronal marker, such as the neuron-specific nu-
Esclapez et al., 1994). However, the mRNAs for clear protein NeuN, demonstrates the extensive
219

Fig. 1. In situ hybridization of GAD65 in the rat dentate gyrus. (A) In a coronal section through the rostral dentate gyrus, labeled
neurons are numerous in the deep regions of the hilus (H), between the hilar tip and CA3, and in the region beneath the granule cell
layer (arrow). Labeled neurons are aligned along the base of the granule cell layer (G) and are dispersed in the molecular layer (M). (B)
In a coronal section through a more caudal level of the dentate gyrus, labeled neurons are concentrated in the dorsal and ventral hilus
and are numerous along the base of the granule cell layer (G). (C) At higher magnification, labeled neurons in the hilus (H) can be
distinguished from those at the base of the granule cell layer (G). Labeled neurons are evident at all laminar levels of the molecular
layer (M).

intermingling of GABA and non-GABA neurons neurons are being identified as the different sub-
throughout the hilus (Fig. 2A). (Mossy cells are groups of GABA neurons are labeled. As one
generally considered to be principal cells and are example, NADPH-diaphorase labels exclusively
discussed elsewhere in this volume; see reviews by GABAergic interneurons in the dentate gyrus,
Seress, Henze and Buszaki, and Scharfman.) Iden- and the labeled neurons are relatively abundant
tifying the relatively large numbers of GAD (Valtschanoff et al., 1993; Czeh et al., 2005). Yet,
mRNA-labeled neurons in the dentate gyrus pro- even though NADPH-diaphorase-labeled neurons
vides a base on which to consider subpopulations are present in all regions of the dentate gyrus
of GABA neurons. Visualizing the entire group of and are not restricted to a specific subtype of
GABA neurons in the region also makes it clear interneuron, they represent only a portion of the
that only limited numbers of dentate GABA GABA neurons in each region (Fig. 2B).
220

Fig. 2. Double labeling of GAD65 mRNA and either NeuN (A) or NADPH-diaphorase (B) in the rat dentate gyrus. (A) GAD65
mRNA-labeled neurons (black) are numerous throughout the dentate hilus (H) where they are intermingled with NeuN single-labeled
(non-GABA) neurons (red; examples at arrows). Some GAD65 mRNA-labeled neurons are present in the granule cell layer (G) but are
most abundant at the base of this layer. Most neurons in the molecular layer (M) are labeled for GAD65 mRNA. Granule cells and
CA3 pyramidal cells are single-labeled for NeuN. (B) NADPH-diaphorase (dark blue) is present in only subgroups of the GAD
mRNA-labeled neurons (red) in the hilus (H), granule cell layer (G) and molecular layer (M) (examples at arrows). See Plate 13.2 in
Color Plate Section.

While it is expected that many peptides and cal- influences the extent of cell body labeling. The
cium-binding proteins will define specific sub- varying somal content could be related to physi-
groups of GABA neurons, as will be discussed in ological conditions, including the levels of activity
later sections, it is less clear why immunohisto- of the neurons, as well as transport of GAD and
chemical labeling of GABA, as well as GAD65 GABA from the soma to axon terminals.
and GAD67 proteins, does not label all GABA As additional antibodies that recognize GAD
neurons as reliably as the in situ hybridization and GABA are obtained and new labeling meth-
labeling of GAD mRNAs. One possibility is that ods are developed, the most effective methods for
the actual levels of GAD and GABA in the cell labeling GABA neurons are likely to change.
bodies of GABA neurons vary, and that this However, currently in situ hybridization of either
221

GAD65 or GAD67 mRNA yields the most com- Several subgroups of GABA neurons contribute
plete and reliable labeling of cell bodies of the en- to the rich fields of GABAergic terminals in the
tire population of GABA neurons in the dentate dentate gyrus, and many of these interneurons
gyrus. Immunohistochemical localization of the have extensive axonal arborizations not only
GAD proteins generally provides labeling of some within a slice but also along the septotemporal
but not all cell bodies, but these immunohisto- length of the dentate gyrus (Han et al., 1993;
chemical methods are essential for demonstrating Buckmaster and Schwartzkroin, 1995; Sik et al.,
the extensive plexus of GABAergic terminals 1997). Indeed the axonal patterns and postsynaptic
within the dentate gyrus. The following descrip- targets of the interneurons prove to be the most
tions of immunohistochemical labeling will con- distinguishing morphological features of the differ-
sider GABA neurons in both rat and mouse. The ent subclasses of interneurons. The current classi-
most detailed information on the distributions and fication of dentate interneurons is based primarily
connections of dentate interneurons has been ob- on the studies of Somogyi and colleagues in which
tained from rat, but there is considerable interest interneurons were labeled intracellularly, and their
in the characteristics of these neurons in mice, due axonal plexus and targets were identified with light
to the rapid increase in studies of genetically mod- and electron microscopic methods (Halasy and
ified animals. Thus the illustrations of immuno- Somogyi, 1993b; Han et al., 1993).
labeling will be from mouse (C57/BL6) primarily. Interneurons in the dentate gyrus, as in other
Current findings suggest, however, that most mor- cortical regions, can be divided into those that
phological and biochemical aspects of dentate form synapses at either perisomatic or dendritic
interneurons are quite similar in rat and mouse locations. Interneurons with synapses at periso-
(personal observations; Mátyás et al., 2004). matic sites can be divided further into those that
synapse on either the cell bodies and proximal
dendrites (basket cells) or the axon initial segments
Axonal arborizations and postsynaptic targets of (axoaxonic cells) of granule cells.
GABA neurons The cell bodies of many basket cells are located
along the base of the granule cell layer but are also
GABAergic terminals are abundant throughout the found within the granule cell layer and along its
dentate gyrus, but reach their highest densities in outer border. While basket cells have a variety of
the outer third of the molecular layer (Fig. 3A). shapes and dendritic patterns, the pyramidal basket
Slightly lower concentrations of GABAergic termi- cells are the most readily identified because of their
nals are evident in the inner two-thirds of this layer. large cell bodies and long apical dendrites (Ribak
These high concentrations of GABAergic axon ter- and Seress, 1983). But regardless of the dendritic
minals in the dendritic regions of the granule cells patterns, the axons are concentrated in the granule
are consistent with electron microscopic evidence cell layer, where each basket cell forms synaptic
that, in quantitative terms, GABAergic synapses in contacts with the cell bodies of multiple granule cells
dendritic regions significantly outnumber those in (Han et al., 1993). The distributions of the axon
somatic regions of the dentate gyrus (75% vs. 25%) terminals of basket cells on granule cell somata are
(Halasy and Somogyi, 1993a). Within the granule reflected in the numerous GAD-labeled terminals
cell layer, GABAergic terminals surround the gran- throughout the granule cell layer (Fig. 3A). In ad-
ule cell somata, and a narrow band of increased dition, basket cells form synaptic contacts with
labeling is present at the outer border of the layer other GABA neurons, including other basket cells
(Fig. 3A). The lowest concentration of GABAergic (Freund and Buzsáki, 1996).
terminals is found in the hilus where labeled termi- A second class of GABA neurons with periso-
nals are dispersed throughout the neuropil. While matic targets is the axoaxonic cell. These inter-
the cell bodies of some hilar neurons are sur- neurons are amazingly selective in their targets
rounded by GABAergic terminals, others exhibit and, in the dentate gyrus, form synapses exclu-
little perisomatic labeling. sively on axon initial segments of granule cells
222

Fig. 3. GAD65-labeled axon terminals in the mouse dentate gyrus (A) and hypothetical cell types of origin (B). (A & B) Dense terminal
fields in the outer molecular layer (Mo) originate from hilar (H) and molecular layer GABA neurons that project to the perforant path
zone (HIPP and MOPP cells, respectively). Moderately dense terminal fields in the inner molecular layer (Mi) originate from hilar
interneurons that innervate the commissural-association projection region (HICAP cells). GABAergic terminals in the granule cell
layer (G) orginate from basket and axoaxonic cells that provide perisomatic innervation.

(Soriano and Frotscher, 1989; Soriano et al., 1990; granule cells through their contacts with multiple
Han et al., 1993; DeFelipe, 1999). Their axons granule cells (Miles et al., 1996).
branch extensively, and each branch forms a series The interneurons that terminate on granule cell
of synaptic contacts along an axon initial segment. dendrites, and form the GABAergic plexus in the
The numerous descending branches resemble the molecular layer, can be divided into multiple
candles or lights on a chandelier, and this led to classes based on the location of their cell bodies
their early identification as chandelier cells in the and the laminar distribution of their axon termi-
cerebral cortex (Szentágothai, 1974; Somogyi, nals in the molecular layer (Halasy and Somogyi,
1977). The cell bodies of these neurons are often 1993b; Han et al., 1993). Interestingly, the cell
located near the granule cell layer, including loca- bodies of neurons that innervate granule cell dend-
tions along the outer border of the layer, but can rites in the molecular layer are located in both the
also be found within the hilus. Because of the hilus and the molecular layer itself. GABA neu-
proximal location of the axon terminals of basket rons with their cell bodies in the hilus and axon
cells and axoaxonic cells, these neurons can exert projections to the outer two-thirds of the molec-
powerful control over the output of granule cells ular layer are identified as hilar perforant path-
and potentially contribute to synchronization of associated (HIPP) cells. Likewise, neurons with
223

cell bodies in the hilus that project to the inner their terminals in the perforant path zone (MOPP
third of the molecular layer are identified as cells).
hilar commissural-association pathway-related Two other subgroups of GABA neurons in the
(HICAP) cells. Because of the location of the cell dentate gyrus have unique patterns of axonal ter-
bodies of these neurons in the hilus, they receive minations but cannot be readily distinguished by
considerable input from the granule cells and are their cell body locations, dendritic morphology, or
thus positioned to provide feedback inhibition to laminar distributions of their axons. The first is a
granule cells near the location of either the exci- group of interneurons that selectively contact
tatory input from the perforant path or commis- other interneurons. Many of these GABA neu-
sural-association neurons (primarily mossy cells) rons are located in the hilus and contain the cal-
in the hilus (Han et al., 1993). cium-binding proteins calretinin and calbindin in
In contrast, GABA neurons with cell bodies in the rat (Gulyás et al., 1996). However, the chem-
the molecular layer are likely to receive excitatory ical identity of interneuron-specific GABA neu-
input directly from the perforant path or, poten- rons in the mouse is not yet known.
tially, commissural-association fibers and thus Another subgroup of interneurons in the hilus
could provide feed-forward inhibition to granule projects beyond the dentate to the medial septum.
cell dendrites (Freund and Buzsáki, 1996). Molec- The hippocampo-septal neurons have been de-
ular layer neurons that innervate the outer molec- scribed most extensively in CA1 and CA3 of the
ular layer have been named molecular layer rat (Alonso and Kohler, 1982; Toth and Freund,
perforant path-associated (MOPP) cells (Halasy 1992). However, they are also found in the hilus,
and Somogyi, 1993b). Influences of these neurons and many express somatostatin (Zappone and
on granule cells could be critical for regulating the Sloviter, 2001; Jinno and Kosaka, 2002a; Gulyás
response of the granule cells to their major exci- et al., 2003). In the septal region, these neurons
tatory inputs. form synapses with parvalbumin (PV)-containing
As descriptions of specific cell types are ob- GABA neurons that, in turn, project back to the
tained, it is possible to relate the general locations hippocampus and dentate gyrus and innervate
of the interneurons and their terminal fields to other GABA neurons (Freund and Antal, 1988;
the broad distribution of axon terminals in the Acsády et al., 2000b). This recurrent GABAergic
dentate gyrus (Fig. 3). Many of the cell bodies circuit is viewed as a potential regulator of hippo-
along the base of the granule cell layer, and campal theta activity although the precise mech-
within the layer, are likely to be basket cells and anism of action remains to be determined (Dragoi
axoaxonic cells that contribute to the GABAergic et al., 1999).
terminals within the granule cell layer (Fig. 3). In
contrast, many GABA neurons in the hilus are
dendrite-innervating neurons and are a major Neurochemical identity
source of the GABAergic innervation of the mo-
lecular layer. The dense innervation in the outer Interneurons of the dentate gyrus can also be clas-
molecular layer is derived in part from the HIPP sified according to their content of specific cal-
cells whereas the more moderate densities of ter- cium-binding proteins and neuropeptides, and a
minals in the inner molecular region include ax- number of distinct groups of interneurons have
onal projections from the HICAP cells (Fig. 3). been identified, even though overlap is found
However, it is not possible to distinguish different among some subgroups. By studying the cell bod-
classes of hilar neurons, such as the HICAP and ies and axon terminals of these chemically defined
HIPP cells, on the basis of their locations in the interneurons, it has been possible to ‘‘dissect’’ the
hilus or the patterns of their proximal dendrites. broad, complex pattern of GABAergic cell bodies
Most GABA neurons in the molecular layer pre- and terminals and relate several of the groups to
sumably contribute to the GABAergic plexus the morphologically identified cell types discussed
within the layer, with the highest concentration of previously.
224

Parvalbumin dendrites that extend in the subgranular region


(Fig. 4A).
PV-containing interneurons have a restricted dis- The PV-labeled neurons receive substantial in-
tribution of their axon terminals and form put from axon collaterals of the mossy fibers
synapses predominantly on cell bodies and axon through synaptic contacts on their cell bodies and
initial segments of granule cells, consistent with proximal dendrites within the granule cell layer
their identity as basket and axo-axonic cells and also on their basal dendrites (Blasco-Ibánez
(Ribak et al., 1990). PV-labeled cell bodies are lo- et al., 2000). These interneurons can thus provide
cated primarily near the granule cell layer and are strong feedback inhibition of the granule cells.
most prominent at the base of the granule cell In addition, the ascending dendrites of many
layer (Fig. 4A). However, some PV-labeled inter- PV-labeled interneurons are positioned to receive
neurons are also located near the junction of the input from the major afferents to the dentate
granule cell and molecular layers (Fig. 4A, C), and gyrus, including the entorhinal cortex (Zipp et al.,
smaller numbers of PV-labeled neurons are found 1989). Thus the PV-labeled interneurons can also
in the hilus and molecular layer. Many labeled provide feed-forward inhibition to the granule
neurons along the base of the granule cell layer cells, in response to excitatory inputs to their
resemble pyramidal basket cells, with labeled api- dendrites in the molecular layer.
cal dendrites that often branch and ascend through Although PV-containing neurons are considered
the full extent of the molecular layer and basal to be a prominent class of interneurons in the

Fig. 4. Parvalbumin (PV)-labeled neurons in the mouse dentate gyrus. (A) Cell bodies of PV-labeled neurons are located at the base of
the granule cell layer (G), and their apical dendrites ascend through the granule cell layer and often branch as they extend through the
molecular layer (M). Basal dendrites are evident along the inner border of the granule cell layer (arrows). (B) A large PV-labeled
interneuron has the characteristics of a pyramidal basket cell and is located at the base of the granule cell layer (G). (C) A PV-labeled
multipolar interneuron in the inner molecular layer extends several dendrites (arrows) into the molecular layer (M). The axon is likely
to descend and contribute to the axonal plexus in the granule cell layer (G). (D) PV-labeled terminals are concentrated in the granule
cell layer where they surround the cell bodies and axon initial segments of granule cells.
225

dentate gyrus, the number of labeled cell bodies in resemble the patterns of PV-containing terminals
the dentate is often relatively low (Ribak et al., within the granule cell layer, including a slightly
1990). The percentage of GAD67-labeled inter- increased density of fibers and terminals along the
neurons that express PV in the granule cell layer is outer border of the granule cell layer. In contrast
estimated to be 20–25% in the mouse (Jinno and to the PV-containing group, CCK-labeled neurons
Kosaka, 2002b). This contrasts with the consider- are restricted to the basket cell population and
ably higher 40% of interneurons that express PV apparently do not contribute to the axoaxonic sub-
in the pyramidal cell layer of CA3 and CA1 of the group, at least in the rat.
hippocampus (Jinno and Kosaka, 2002b). The
limited number of PV-labeled cell bodies could
reflect generally low numbers of PV-containing Somatostatin
dentate interneurons, or, possibly, a low level of
the protein in the cell bodies despite the presence Somatostatin neurons constitute one of the largest
of PV in the axon terminals. Such differential in- chemically defined subgroups of GABA neurons in
tracellular labeling has been suggested in some ex- the dentate gyrus, and virtually all somatostatin
perimental and pathological conditions, including neurons of this region are located within the
human temporal lobe epilepsy, in which cell body hilus (Fig. 5A). Furthermore, approximately 55%
labeling of PV is substantially decreased despite of hilar GABA neurons express somatostatin
persistent terminal labeling (Sloviter et al., 1991; (Esclapez and Houser, 1995; Buckmaster and Jon-
Scotti et al., 1997; Wittner et al., 2001). gen-Rêlo, 1999; Jinno and Kosaka, 2003).
The precise function of PV in the interneurons In both mouse and rat, somatostatin-containing
remains uncertain, but the presence of this cal- neurons are more abundant, and constitute a
cium-binding protein could be a physiological slightly higher percentage of total GABA neurons
adaptation related to the high levels of activity of in the ventral than in the dorsal dentate gyrus
this group of neurons and their identification (Buckmaster et al., 1994; Esclapez and Houser,
as fast-spiking neurons in several brain regions 1995; Jinno and Kosaka, 2003).
(Freund and Buzsáki, 1996). Interestingly, PV- While somatostatin-labeled cell bodies are con-
labeled interneurons express particularly high fined primarily to the hilus in the dentate gyrus,
levels of cytochrome c in the rat, and this, along their terminals fields are most prominent in the
with high numbers of mitochondrial profiles in outer molecular layer where the perforant path
their axon terminals, is consistent with high levels fibers terminate (Bakst et al., 1986; Katona et al.,
of metabolic activity in this group of GABA neu- 1999). Thus many of the somatostatin neurons in
rons (Gulyás et al., 2006). the hilus are considered to be HIPP cells, based on
the location of their cell bodies and axon termi-
nals. The dendrites of somatostatin neurons ex-
Cholecystokinin tend for considerable distances across the hilus but
generally remain within the region (Fig. 5B).
The peptide cholecystokinin (CCK) is also located Although the proximal dendrites of many of
in interneurons near the granule cell layer, and these the somatostatin interneurons are smooth and
neurons constitute a second group of GABAergic sparsely spinous, as is characteristic of many
basket cells. There appears to be little overlap GABAergic interneurons (Ribak, 1978), the more
between the CCK- and PV-labeled neurons in the distal dendrites have a considerable number of
hippocampus, and this suggests that there are differ- spines (Fig. 5C). These extensive dendrites and their
ent groups of basket cells with potentially different spines are likely targets of the granule cell mossy
but complementary functions, as has been demon- fibers, and the mossy fiber contacts on GABA neu-
strated for basket cells in CA1 (Klausberger et al., rons outnumber those on mossy cells by a ratio of
2005). Relatively few CCK-labeled cell bodies are approximately 5:1 (Acsády et al., 1998). The som-
found in the dentate gyrus, but their terminal fields atostatin neurons, in turn, form a moderately dense
226

Fig. 5. Somatostatin-containing neurons in the ventral dentate gyrus of the mouse. (A) Numerous somatostatin-immunoreactive cell
bodies are present in the hilus (H), and a labeled axonal plexus is evident in the outer molecular layer (M). Little labeling is found in the
granule cell layer (G) except for axons from hilar somatostatin neurons that project through this layer. (B) Localization of green
fluorescent protein (GFP) in somatostatin neurons of a transgenic mouse reveals the extensive dendrites of these neurons. Several
dendrites extend across large extents of the hilus (arrows). (C) GFP-labeled dendrites extend from a multipolar hilar neuron, and small
spines can be detected on some dendritic segments (arrows).

axonal plexus in the outer two-thirds of the molec- high percentage of hilar somatostatin neurons that
ular layer (Bakst et al., 1986). The somatostatin innervate the septum suggests that somatostatin as
terminals synapse primarily with granule cell dend- well as GABA could have important functional
rites and are often located adjacent to asymmetric roles in the hippocampal-septal circuitry.
excitatory synapses (Katona et al., 1999). Som- Somatostatin neurons in the dentate gyrus are
atostatin neurons are thus ideally positioned to highly vulnerable to excitatory and ischemic dam-
modulate the influence of the perforant path on age, and loss of these neurons is commonly ob-
dentate granule cells in response to ongoing granule served in models of temporal lobe epilepsy and
cell activity (see chapter by Tallent in this volume). forebrain ischemia (e.g. Johansen et al., 1987;
Some hilar somatostatin neurons project beyond Sloviter, 1987; Buckmaster and Dudek, 1997).
the dentate gyrus to the medial septum. In both Characteristics of these neurons that may contrib-
mouse and rat, a high proportion of the hippo- ute to their vulnerability include the high levels
campo-septal neurons in the hilus express som- of excitatory mossy fiber contacts (Acsády
atostatin (Zappone and Sloviter, 2001; Jinno and et al., 1998) and the relatively low GABAergic
Kosaka, 2002a). Furthermore, retrograde labeling innervation on the soma and proximal dendrites of
studies in the mouse indicate that a surprisingly these interneurons (Acsády et al., 2000a).
high percentage (44%) of somatostatin-labeled
neurons in the hilus project to the septal region
(Jinno and Kosaka, 2002a). Currently, it is unclear Calretinin
whether separate groups of GABA neurons
project to either the dentate molecular layer or Most classes of GABAergic interneurons in the
the medial septum, or whether single somatostatin dentate gyrus appear to be quite similar in their
neurons project to both regions. Regardless, the patterns of neuropeptide and calcium-binding
227

Fig. 6. Comparisons of calretinin labeling in the dorsal dentate gyrus of mouse (A) and rat (B). (A) Calretinin labeling in the mouse
includes darkly-labeled (arrows) and moderately-labeled (arrowheads) neurons in the hilus (H) that correspond to GABAergic in-
terneurons and mossy cells, respectively. In addition, small labeled cell bodies at the base of the granule cell layer (G) are likely to be
newly-generated granule cells. A dense band of labeled axon terminals (asterisk) from the mossy cells is prominent in the inner
molecular layer (M). (B) In the rat, darkly-labeled neurons (arrows) in the hilus (H) are presumed to be GABA neurons. A narrow
band of labeled terminals is evident along the outer border of the granule cell layer (G), and these may originate in the sup-
ramammillary region.

protein expression in rats and mice. However, sev- (Fig. 6A). Another difference between the two
eral species differences are found in the neurons species is the lack of spiny calretinin-containing
that express the calcium-binding protein calretinin neurons in the mouse (Blasco-Ibáñez and Freund,
(Liu et al., 1996; Blasco-Ibáñez and Freund, 1997). 1997; Fujise et al., 1998; Mátyás et al., 2004). Such
Most notably, in the mouse, calretinin is expressed neurons are distinctive hilar neurons in the rat
in mossy cells of the hilus, and such mossy cells are and belong to the subgroup of GABA neurons
particularly numerous in the ventral hilus (Blasco- that contacts primarily other interneurons (Gulyás
Ibáñez and Freund, 1997; Fujise et al., 1998). et al., 1996). Finally, in the mouse, newborn granule
Calretinin is present in axon terminals of the cells along the base of the granule cell layer
mossy cells, and this creates a dense band of labe- express calretinin (Fig. 6A) (Liu et al., 1996;
ling in the inner third of the molecular layer Blasco-Ibáñez and Freund, 1997). Likewise, Cajal-
(the major terminal field of the mossy cells) in the Retzius cells that are presumed to use glutamate as
mouse (Fig. 6A). In contrast, labeling is primarily their neurotransmitter are labeled for calretinin in
restricted to GABAergic interneurons of the dent- the mouse.
ate gyrus in the rat, and only a narrow band of Thus, calretinin labeling of cell bodies in the
calretinin labeling is present in the supragranular dentate gyrus is largely restricted to GABA neu-
region in the rat (Fig. 6B). These terminals are rons in the rat but is present in several groups
not from mossy cells but may originate from the of glutamatergic, as well as some GABAergic,
supramammillary region and are considered to be neurons, in the mouse.
glutamatergic (Maglóczky et al., 1994; Kiss et al.,
2000).
Despite calretinin labeling of the mossy cells in Neuropeptide Y
the mouse, some GABAergic interneurons that co-
express calretinin can be detected (Gulyás et al., Within the dentate gyrus, neuropeptide Y (NPY)-
1996). The GABA neurons are located primarily in labeled cell bodies are located primarily in the
the hilus, near the granule cell layer, and some- hilus (Fig. 7) and constitute one of the larger
times can be distinguished by their darker labeling groups of hilar GABA neurons. It has been esti-
compared to that of mossy cells in the region mated that NPY-containing neurons account for
228

Fig. 7. NPY-labeled neurons in the dorsal (A) and ventral (B) hilus of the mouse dentate gyrus. (A & B) Labeled neurons are largely
confined to the hilus (H) at all levels of the dentate gyrus. Labeled neurons are relatively numerous in the hilus but are sparse in the
granule cell (G) and molecular (M) layers as well as the pyramidal cell layer of CA3.

approximately 57–82% of GAD67-labeled neu- GABA neurons in two locations within the dent-
rons in the dorsal and ventral hilus, respectively ate gyrus appear to be particularly underrepre-
(Jinno and Kosaka, 2003). These relatively high sented by the currently available markers for
percentages suggest that there is considerable co- subgroups of GABA neurons. These include the
localization of NPY and somatostatin in GABA numerous GABA neurons in the hilar or poly-
neurons in the hilus (Kohler et al., 1987; Deller morph region that is located below the base of the
and Leranth, 1990). Small numbers of NPY-labe- granule cell layer and GABA neurons in the mo-
led cell bodies are also located at the base of the lecular layer.
granule cell layer and in the molecular layer. The Other molecules however call attention to some
axon terminal labeling of NPY-expressing neurons of these GABA neurons. For example, virtually all
is generally limited with current immunohisto- of the molecular layer interneurons express high
chemical methods, but some terminals are evident levels of the a1 subunit of the GABAA receptor, as
in do many, but not all, groups of interneurons in the
the outer molecular layer, suggesting that many of dentate gyrus (Fig. 8) (Gao and Fritschy, 1994;
the NPY-containing neurons are HIPP cells and Esclapez et al., 1996; Sperk et al., 1997; Peng et al.,
resemble somatostatin neurons. Some NPY- 2004). In addition, these molecular layer interneu-
labeled axon terminals are also present in the rons express the d subunit of the GABAA receptor
hilus, and many of their targets are likely to be and, consistent with such expression, exhibit tonic
mossy cells (Deller and Leranth, 1990; Acsády inhibition (Glykys et al., 2007). Activation of
et al., 2000a). GABAA receptors on the cell bodies of GABA
neurons could be an additional mechanism
through which the activity of GABA neurons is
Further characterization of dentate interneurons modulated or limited. Inhibition of even small
numbers of GABA neurons that have extensive
The presence of specific neuropeptides and cal- axonal projections, such as those in the molecular
cium-binding proteins in selected interneurons has layer, could reduce inhibition and thus enhance
been extremely useful for identifying major classes excitation of the granule cells.
of GABA neurons in the hippocampal formation. Numerous other peptides and proteins are local-
Yet, in some regions, the number of these neurons ized in GABA neurons in the dentate gyrus, and
appears small when compared to the total number their identities and distributions highlight addi-
of GABA neuron cell bodies and terminals. tional, unique characteristics of these interneurons.
229

1992; Buckmaster and Dudek, 1997). Continued


characterization of the vulnerable groups of
GABA neurons could eventually lead to specific
neuroprotective strategies.
Other molecules are localized primarily in inter-
neurons but are found in several subtypes of
GABA neurons in the hippocampal formation.
These includes Reelin (Pesold et al., 1998) and
nerve growth factor (Lauterborn et al., 1993;
Pascual et al., 1998). The preferential localization
of these particular substances in interneurons
suggests roles for these GABA neurons in basic
developmental processes that extend beyond their
essential roles in regulating information process-
ing, rhythm generation and excitability of the hip-
pocampal formation.

Conclusions
Fig. 8. GABAA receptor a1 subunit labeling in the mouse
dentate gyrus. The a1 subunit is highly expressed on the surface Despite years of study, GABA neurons in the dent-
of cell bodies and dendritic processes of interneurons in the ate gyrus continue to hold surprises and challenges.
granule cell (G) and molecular (M) layers (examples at arrows).
Labeled neurons at the base of the granule cell layer closely
No single morphological method provides an ade-
resemble parvalbumin-labeled pyramidal basket cells (compare quate view of the richness and complexity of these
with Fig. 4B). neurons. Thus the findings obtained from different
methods are pieces of a puzzle that require assem-
Some expression patterns mirror the major subdi- bly. Key features of the emerging picture are the
visions of interneurons, based on their innervation extensive and yet highly defined axonal plexus
of perisomatic or dendritic regions, for which PV of single interneurons in the dentate gyrus; the rich
and somatostatin serve as prototypic markers, repertoire of molecules that each interneuron pos-
respectively. For example, mu-opioid receptors are sesses; and the remarkable functional regulation
preferentially localized on PV-labeled neurons of levels of GABA, GAD and many neuropeptides
whereas delta-opioid receptors are more abundant in these interneurons. While such complexity can
on somatostatin neurons (Commons and Milner, be daunting, it also provides opportunities for dis-
1996; Stumm et al., 2004; Drake and Milner, 2006). covering new ways to enhance normal GABAergic
Likewise, the a1 subunit of the GABAA receptor is function and possibly protects these critical neurons
highly expressed on PV-labeled neurons, as well as from damage in pathological conditions.
other interneurons in the dentate gyrus, but shows
very little or no expression on somatostatin neurons
in the hilus (Gao and Fritschy, 1994; Esclapez et al., Acknowledgments
1996).
Differences between these major groups of dent- I thank Christine Huang and Siroun Tahtakran
ate interneurons are also reflected in their vulner- for excellent histologic and photographic work,
ability to ischemic and seizure-induced damage. and in situ hybridization labeling, respectively, and
Somatostatin interneurons are some of the most gratefully acknowledge the numerous contribu-
vulnerable neurons in the dentate gyrus, whereas tions of Drs. Monique Esclapez, Zechun Peng and
PV-containing neurons are more resistant to dam- Nianhui Zhang to our studies of GABA neurons
age in some models (Sloviter, 1987; Sperk et al., in the hippocampus. These studies were supported
230

by Veterans Administration Medical Research Buckmaster, P.S. and Schwartzkroin, P.A. (1995) Interneurons
Funds and National Institutes of Health grants and inhibition in the dentate gyrus of the rat in vivo. J.
Neurosci., 15(1): 774–789.
NS046524 and NS051311 to C.R.H.
Commons, K.G. and Milner, T.A. (1996) Cellular and subcel-
lular localization of delta opioid receptor immunoreactivity
in the rat dentate gyrus. Brain Res., 738: 181–195.
Czeh, B., Hajnal, A. and Seress, L. (2005) NADPH-diaphorase
References positive neurons of the rat hippocampal formation: regional
distribution, total number and colocalization with calcium
Acsády, L., Kamondi, A., Sik, A., Freund, T. and Buzsáki, G. binding proteins. Prague Med. Rep., 106: 261–274.
DeFelipe, J. (1999) Chandelier cells and epilepsy. Brain,
(1998) GABAergic cells are the major postsynaptic targets
of mossy fibers in the rat hippocampus. J. Neurosci., 18: 122(Pt 10): 1807–1822.
3386–3403. Deller, T. and Leranth, C. (1990) Synaptic connections of ne-
uropeptide Y (NPY) immunoreactive neurons in the hilar area
Acsády, L., Katona, I., Martinez-Guijarro, F.J., Buzsáki, G.
of the rat hippocampus. J. Comp. Neurol., 300: 433–447.
and Freund, T.F. (2000a) Unusual target selectivity of
perisomatic inhibitory cells in the hilar region of the rat hip- Dragoi, G., Carpi, D., Recce, M., Csicsvari, J. and Buzsaki, G.
pocampus. J. Neurosci., 20: 6907–6919. (1999) Interactions between hippocampus and medial septum
Acsády, L., Pascual, M., Rocamora, N., Soriano, E. and during sharp waves and theta oscillation in the behaving rat.
Freund, T.F. (2000b) Nerve growth factor but not neurotro- J. Neurosci., 19: 6191–6199.
phin-3 is synthesized by hippocampal GABAergic neurons Drake, C.T. and Milner, T.A. (2006) Mu opioid receptors are
that project to the medial septum. Neuroscience, 98: 23–31. extensively co-localized with parvalbumin, but not som-
Alonso, A. and Kohler, C. (1982) Evidence for separate atostatin, in the dentate gyrus. Neurosci. Lett., 403: 176–180.
projections of hippocampal pyramidal and non-pyramidal Erlander, M.G., Tillakaratne, N.J.K., Feldblum, S., Patel, N.
neurons to different parts of the septum in the rat brain. and Tobin, A.J. (1991) Two genes encode distinct glutamate
Neurosci. Lett., 31: 209–214. decarboxylases. Neuron, 7: 91–100.
Amaral, D.G. (1978) A Golgi study of cell types in the hilar Esclapez, M., Chang, D.K. and Houser, C.R. (1996) Sub-
region of the hippocampus in the rat. J. Comp. Neurol., 182: populations of GABA neurons in the dentate gyrus express
851–914. high levels of the a1 subunit of the GABAA receptor.
Babb, T.L., Pretorius, J.K., Kupfer, W.R. and Brown, W.J. Hippocampus, 6: 225–238.
(1988) Distribution of glutamate-decarboxylase-immunore- Esclapez, M. and Houser, C.R. (1995) Somatostatin neurons
active neurons and synapses in the rat and monkey hippo- are a subpopulation of GABA neurons in the rat dentate
campus: light and electron microscopy. J. Comp. Neurol., gyrus: evidence from co-localization of pre-prosomatostatin
278: 121–138. and glutamate decarboxylase mRNAs. Neuroscience, 64:
Bakst, I., Avendano, C., Morrison, J.H. and Amaral, D.G. 339–355.
(1986) An experimental analysis of the origins of somatosta- Esclapez, M., Tillakaratne, N.J.K., Kaufman, D.L., Tobin, A.J.
tin-like immunoreactivity in the dentate gyrus of the rat. J. and Houser, C.R. (1994) Comparative localization of two
Neurosci., 6: 1452–1462. forms of glutamic acid decarboxylase and their mRNAs in
Blasco-Ibáñez, J.M. and Freund, T.F. (1997) Distribution, ul- rat brain supports the concept of functional differences be-
trastructure, and connectivity of calretinin-immunoreactive tween the forms. J. Neurosci., 14: 1834–1855.
mossy cells of the mouse dentate gyrus. Hippocampus, 7: Freund, T.F. and Antal, M. (1988) GABA-containing neurons
307–320. in the septum control inhibitory interneurons in the hippo-
Blasco-Ibánez, J.M., Martı́nez-Guijarro, F.J. and Freund, T.F. campus. Nature, 336: 170–173.
(2000) Recurrent mossy fibers preferentially innervate parv- Freund, T.F. and Buzsáki, G. (1996) Interneurons of the hip-
albumin-immunoreactive interneurons in the granule cell pocampus. Hippocampus, 6: 347–470.
layer of the rat dentate gyrus. Neuroreport, 11: 3219–3225. Fujise, N., Liu, Y., Hori, N. and Kosaka, T. (1998) Distribution
Buckmaster, P.S. and Dudek, F.E. (1997) Neuron loss, granule of calretinin immunoreactivity in the mouse dentate gyrus: II.
cell axon reorganization, and functional changes in the dent- Mossy cells, with special reference to their dorsoventral
ate gyrus of epileptic kainate-treated rats. J. Comp. Neurol., difference in calretinin immunoreactivity. Neuroscience,
385: 385–404. 82: 181–200.
Buckmaster, P.S. and Jongen-Rêlo, A.L. (1999) Highly specific Gao, B. and Fritschy, J.M. (1994) Selective allocation of
neuron loss preserves lateral inhibitory circuits in the dentate GABAA receptors containing the a1 subunit to neurochem-
gyrus of kainate-induced epileptic rats. J. Neurosci., 19: ically distinct subpopulations of rat hippocampal interneu-
9519–9529. rons. Eur. J. Neurosci., 6: 837–853.
Buckmaster, P.S., Kunkel, D.D., Robbins, R.J. and Schwa- Glykys, J., Peng, Z., Chandra, D., Homanics, G.E., Houser,
rtzkroin, P.A. (1994) Somatostatin-immunoreactivity in the C.R. and Mody, I. (2007) A new naturally occurring GABAA
hippocampus of mouse, rat, guinea pig, and rabbit. Hippo- receptor subunit partnership with high sensitivity to ethanol.
campus, 4: 167–180. Nat. Neurosci., 10: 40–48.
231

Gulyás, A.I., Hajos, N., Katona, I. and Freund, T.F. (2003) Katona, I., Acsády, L. and Freund, T. (1999) Postsynaptic tar-
Interneurons are the local targets of hippocampal inhibitory gets of somatostatin-immunoreactive interneurons in the rat
cells which project to the medial septum. Eur. J. Neurosci., hippocampus. Neuroscience, 88: 37–55.
17: 1861–1872. Kaufman, D.L., Houser, C.R. and Tobin, A.J. (1991) Two
Gulyás, A.I., Buzsáki, G., Freund, T.F. and Hirase, H. (2006) forms of the gamma-aminobutyric acid synthetic enzyme
Populations of hippocampal inhibitory neurons express glutamate decarboxylase have distinct intraneuronal distri-
different levels of cytochrome c. Eur. J. Neurosci., 23: butions and cofactor interactions. J. Neurochem., 56:
2581–2594. 720–723.
Gulyás, A.I., Hájos, N. and Freund, T.F. (1996) Interneurons Kiss, J., Csaki, A., Bokor, H., Shanabrough, M. and Leranth,
containing calretinin are specialized to control other inter- C. (2000) The supramammillo-hippocampal and supramam-
neurons in the rat hippocampus. J. Neurosci., 16(10): millo-septal glutamatergic/aspartatergic projections in the
3397–3411. rat: a combined [3H]D-aspartate autoradiographic and
Halasy, K. and Somogyi, P. (1993a) Distribution of immunohistochemical study. Neuroscience, 97: 657–669.
GABAergic synapses and their targets in the dentate gyrus Klausberger, T., Marton, L.F., O’Neill, J., Huck, J.H., Dale-
of rat: a quantitative immunoelectron microscopic analysis. zios, Y., Fuentealba, P., Suen, W.Y., Papp, E., Kaneko, T.,
J. Hirnforsch, 34: 299–308. Watanabe, M., Csicsvari, J. and Somogyi, P. (2005) Com-
Halasy, K. and Somogyi, P. (1993b) Subdivision in the multiple plementary roles of cholecystokinin and parvalbumin-ex-
GABAergic innervation of granule cells in the dentate gyrus of pressing GABAergic neurons in hippocampal network
the rat hippocampus. Eur. J. Neurosci., 5: 411–429. oscillations. J. Neurosci., 25: 9782–9793.
Han, Z.-S., Buhl, E.H., Lorinczi, Z. and Somogyi, P. (1993) Kohler, C., Eriksson, L.G., Davies, S. and Chan-Palay, V.
A high degree of spatial selectivity in the axonal and dendritic (1987) Co-localization of neuropeptide tyrosine and som-
domains of physiologically identified local-circuit neurons in atostatin immunoreactivity in neurons of individual subfields
the dentate gyrus of the rat hippocampus. Eur. J. Neurosci., of the rat hippocampal region. Neurosci. Lett., 78: 1–6.
5: 395–410. Lauterborn, J.C., Tran, T.M., Isackson, P.J. and Gall, C.M.
Houser, C.R. and Esclapez, M. (1994) Localization of (1993) Nerve growth factor mRNA is expressed by GAB-
mRNAs encoding two forms of glutamic acid decarboxy- Aergic neurons in rat hippocampus. Neuroreport, 5:
lase in the rat hippocampal formation. Hippocampus, 5: 273–276.
530–545. Liu, Y., Fujise, N. and Kosaka, T. (1996) Distribution of ca-
Houser, C.R., Esclapez, M. and Zhang, N. (2000) GABA Neu- lretinin immunoreactivity in the mouse dentate gyrus. I.
rons of the hippocampal formation: localization, vulnerability General description. Exp. Brain Res., 108: 389–403.
and plasticity. In: Martin D.L. and Olsen R.W. (Eds.), GABA Maccaferri, G. and Lacaille, J.C. (2003) Interneuron diversity
in the Nervous System: The View at Fifty Years. Lippincott series: hippocampal interneuron classifications — making
Williams & Wilkins, Philadelphia, PA, pp. 337–355. things as simple as possible, not simpler. Trends Neurosci.,
Jinno, S., Aika, Y., Fukuda, T. and Kosaka, T. (1998) 26: 564–571.
Quantitative analysis of GABAergic neurons in the mouse Maglóczky, Z., Acsády, L. and Freund, T. (1994) Principal
hippocampus, with optical disector using confocal laser scan- cells are the postsynaptic targets of supramammillary
ning microscope. Brain Res., 814(1–2): 55–70. afferents in the hippocampus of the rat. Hippocampus, 4:
Jinno, S. and Kosaka, T. (2002a) Immunocytochemical char- 322–334.
acterization of hippocamposeptal projecting GABAergic Martin, D.L., Martin, S.B., Wu, S.J. and Espina, N. (1991)
nonprincipal neurons in the mouse brain: a retrograde labe- Cofactor interactions and the regulation of glutamate dec-
ling study. Brain Res., 945: 219–231. arboxylase activity. Neurochem. Res., 16: 243–249.
Jinno, S. and Kosaka, T. (2002b) Patterns of expression of Mátyás, F., Freund, T.F. and Gulyás, A.I. (2004) Immunocyto-
calcium binding proteins and neuronal nitric oxide synthase chemically defined interneuron populations in the hippocam-
in different populations of hippocampal GABAergic neurons pus of mouse strains used in transgenic technology.
in mice. J. Comp. Neurol., 449: 1–25. Hippocampus, 14: 460–481.
Jinno, S. and Kosaka, T. (2003) Patterns of expression of Miles, R., Tóth, K., Gulyás, A.I., Hájos, N. and Freund, T.F.
neuropeptides in GABAergic nonprincipal neurons in the (1996) Differences between somatic and dendritic inhibition
mouse hippocampus: quantitative analysis with optical in the hippocampus. Neuron, 16: 815–823.
disector. J. Comp. Neurol., 461: 333–349. Parra, P., Gulyas, A.I. and Miles, R. (1998) How many sub-
Jinno, S. and Kosaka, T. (2006) Cellular architecture of the types of inhibitory cells in the hippocampus? Neuron, 20:
mouse hippocampus: a quantitative aspect of chemically 983–993.
defined GABAergic neurons with stereology. Neurosci. Res., Pascual, M., Rocamora, N., Acsády, L., Freund, T.F. and
56: 229–245. Soriano, E. (1998) Expression of nerve growth factor and
Johansen, F.F., Zimmer, J. and Diemer, N.H. (1987) Early loss neurotrophin-3 mRNAs in hippocampal interneurons: mor-
of somatostatin neurons in dentate hilus after cerebral phological characterization, levels of expression, and colo-
ischemia in the rat precedes CA-1 pyramidal cell loss. Acta calization of nerve growth factor and neurotrophin-3. J.
Neuropathol. (Berl.), 73: 110–114. Comp. Neurol., 395: 73–90.
232

Peng, Z., Huang, C.S., Stell, B.M., Mody, I. and Houser, C.R. Soriano, E. and Frotscher, M. (1989) A GABAergic axo-axonic
(2004) Altered expression of the d subunit of the GABAA cell in the fascia dentata controls the main excitatory hippo-
receptor in a mouse model of temporal lobe epilepsy. J. Ne- campal pathway. Brain Res., 503: 170–174.
urosci., 24: 8629–8639. Soriano, E., Nitsch, R. and Frotscher, M. (1990) Axo-axonic
Pesold, C., Impagnatiello, F., Pisu, M.G., Uzunov, D.P., Costa, chandelier cells in the rat fascia dentata: Golgi-electron
E., Guidotti, A. and Caruncho, H.J. (1998) Reelin is pref- microscopy and immunocytochemical studies. J. Comp.
erentially expressed in neurons synthesizing gamma-amino- Neurol., 293: 1–25.
butyric acid in cortex and hippocampus of adult rats. Proc. Sperk, G., Marksteiner, J., Gruber, B., Bellmann, R., Mahata,
Natl. Acad. Sci. U.S.A., 95: 3221–3226. M. and Ortler, M. (1992) Functional changes in neuropeptide
Ribak, C.E. (1978) Aspinous and sparsely-spinous stellate neu- Y- and somatostatin-containing neurons induced by limbic
rons in the visual cortex of rats contain glutamic acid dec- seizures in the rat. Neuroscience, 50: 831–846.
arboxylase. J. Neurocytol., 7: 461–478. Sperk, G., Schwarzer, C., Tsunashima, K., Fuchs, J. and
Ribak, C.E., Nitsch, R. and Seress, L. (1990) Proportion of Sieghart, W. (1997) GABAA receptor subunits in the rat
parvalbumin-positive basket cells in the GABAergic inner- hippocampus I: immunocytochemical distribution of 13 sub-
vation of pyramidal and granule cells of the rat hippocampal units. Neuroscience, 80: 987–1000.
formation. J. Comp. Neurol., 300: 449–461. Stumm, R.K., Zhou, C., Schulz, S. and Hollt, V. (2004)
Ribak, C.E. and Seress, L. (1983) Five types of basket cell in the Neuronal types expressing mu- and delta-opioid receptor
hippocampal dentate gyrus: a combined Golgi and electron mRNA in the rat hippocampal formation. J. Comp. Neurol.,
microscopic study. J. Neurocytol., 12: 577–597. 469: 107–118.
Ribak, C.E., Vaughn, J.E. and Saito, K. (1978) Immunocyto- Szentágothai, J. (1974) Conceptual models of neural organiza-
chemical localization of glutamic acid decarboxylase in neu- tion. Neurosci. Res. Prog. Bull., 12: 307–510.
ronal somata following colchicine inhibition of axonal Toth, K. and Freund, T.F. (1992) Calbindin D28k-containing
transport. Brain Res., 140: 315–332. nonpyramidal cells in the rat hippocampus: their immuno-
Scotti, A.L., Bollag, O., Kalt, G. and Nitsch, C. (1997) Loss of reactivity for GABA and projection to the medial septum.
perikaryal parvalbumin immunoreactivity from surviving Neuroscience, 49: 793–805.
GABAergic neurons in the CA1 field of epileptic gerbils. Valtschanoff, J.G., Weinberg, R.J., Kharazia, V.N.,
Hippocampus, 7: 524–535. Nakane, M. and Schmidt, H.H. (1993) Neurons in rat hip-
Sik, A., Penttonen, M. and Buzsaki, G. (1997) Interneurons in pocampus that synthesize nitric oxide. J. Comp. Neurol., 331:
the hippocampal dentate gyrus: an in vivo intracellular study. 111–121.
Eur. J. Neurosci., 9: 573–588. Wittner, L., Maglóczky, Z., Borhegyi, Z., Halász, P., Tóth, S.,
Sloviter, R.S. (1987) Decreased hippocampal inhibition and Eröss, L., Szabó, Z. and Freund, T.F. (2001) Preserva-
a selective loss of interneurons in experimental epilepsy. tion of perisomatic inhibitory input of granule cells in the
Science, 235: 73–76. epileptic human dentate gyrus. Neuroscience, 108:
Sloviter, R.S. and Nilaver, G. (1987) Immunocytochemical lo- 587–600.
calization of GABA-, cholecystokinin-, vasoactive intestinal Woodson, W., Nitecka, L. and Ben-Ari, Y. (1989) Organization
polypeptide-, and somatostatin-like immunoreactivity in the of the GABAergic system in the rat hippocampal formation:
area dentata and hippocampus of the rat. J. Comp. Neurol., a quantitative immunocytochemical study. J. Comp. Neurol.,
256: 42–60. 280: 254–271.
Sloviter, R.S., Sollas, A.L., Barbaro, N.M. and Laxer, K.D. Zappone, C.A. and Sloviter, R.S. (2001) Commissurally
(1991) Calcium-binding protein (calbindin-D28K) and parv- projecting inhibitory interneurons of the rat hippocampal
albumin immunocytochemistry in the normal and epileptic dentate gyrus: a colocalization study of neuronal markers
human hippocampus. J. Comp. Neurol., 308: 381–396. and the retrograde tracer Fluoro-gold. J. Comp. Neurol.,
Somogyi, P. (1977) A specific ‘axo-axonal’ interneuron in the 441: 324–344.
visual cortex of the rat. Brain Res., 136: 345–350. Zipp, F., Nitsch, R., Soriano, E. and Frotscher, M. (1989)
Somogyi, P. and Klausberger, T. (2005) Defined types of cor- Entorhinal fibers form synaptic contacts on parvalbumin-
tical interneurone structure space and spike timing in the immunoreactive neurons in the rat fascia dentata. Brain Res.,
hippocampus. J. Physiol., 562: 9–26. 495: 161–166.
Plate 13.2. Double labeling of GAD65 mRNA and either NeuN (A) or NADPH-diaphorase (B) in the rat dentate gyrus. (A) GAD65
mRNA-labeled neurons (black) are numerous throughout the dentate hilus (H) where they are intermingled with NeuN single-labeled
(non-GABA) neurons (red; examples at arrows). Some GAD65 mRNA-labeled neurons are present in the granule cell layer (G) but are
most abundant at the base of this layer. Most neurons in the molecular layer (M) are labeled for GAD65 mRNA. Granule cells and
CA3 pyramidal cells are single-labeled for NeuN. (B) NADPH-diaphorase (dark blue) is present in only subgroups of the GAD
mRNA-labeled neurons (red) in the hilus (H), granule cell layer (G) and molecular layer (M) (examples at arrows). (For B/W version,
see page 220 in the volume.)
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 14

Functional regulation of the dentate gyrus


by GABA-mediated inhibition

Douglas A. Coulter1,2, and Gregory C. Carlson1

1
The Children’s Hospital of Philadelphia, Abramson Pediatric Research Center, Room 410D, 3516 Civic Center
Boulevard, Philadelphia, PA 19104-4318, USA
2
Departments of Pediatrics, Neurology, and Neuroscience, University of Pennsylvania School of Medicine, Philadelphia,
PA, USA

Abstract: Dentate granule cells are characterized by their low levels of excitability, an important aspect of
hippocampal function, which distinguishes them from other principal cells of the hippocampus. This low
excitability derives in large part from the degree and nature of GABAergic inhibition evident in the dentate
gyrus. Granule cells express a unique and complex assortment of GABAA receptor subunits, found in few
areas of the brain. Associated with this receptor complexity, granule cells are endowed with both synaptic
and extrasynaptic GABAA receptors with distinctive properties. In particular, extrasynaptic GABAA re-
ceptors in granule cells exhibit high affinity for GABA and do not desensitize. This results in activation of a
tonic current by ambient levels of GABA present in the extracellular space. This tonic current contributes
significantly to the circuit properties of the dentate gyrus. Both synaptic and extrasynaptic GABAA re-
ceptors exhibit profound dysregulation in animal models of temporal lobe epilepsy, which may contribute
to the hippocampal hyperexcitability that defines this disorder.

Keywords: hippocampus; dentate gyrus; GABA receptor; inhibition; epilepsy

Introduction 1976; Wilson and McNaughton, 1993). This ‘sparse


coding’ of dentate granule cells is theorized to be
As a part of its role in memory formation, the hip- important in information processing and memory
pocampus encodes a cognitive map of the space in formation of the hippocampus (McNaughton and
which an animal (or human) navigates. During en- Morris, 1987). It also reflects a general trend evi-
vironmental exploration, studies recording place dent in most studies of dentate granule cell excit-
fields in rat hippocampal neurons have typically ability: these cells exhibit a fundamental reluctance
demonstrated that information coding of spatial to fire, particularly synchronously in network
position (i.e. degree of firing and firing patterns) is bursts. This is due, in part, to a combination of
specific but sparse in dentate granule cells (Jung and intrinsic factors (including hyperpolarized resting
McNaughton, 1993; Chawla et al., 2005) compared membrane potential, lack of conductances which
to dorsal hippocampal pyramidal cells (O’Keefe, permit phasic firing or ‘‘bursting’’, and marked
spike frequency adaptation), but is principally a
consequence of the powerful feedforward and feed-
Corresponding author. Tel.: +1 215-590-1937; back GABAergic inhibition characteristic of dent-
Fax: +1 215-590-4121; E-mail: coulterd@email.chop.edu ate gyrus circuit function.

DOI: 10.1016/S0079-6123(07)63014-3 235


236

In addition to its important role in cognitive possible receptor composition by alternative splic-
processing, the low excitability of the dentate gyrus ing, which is found in several subunit families (re-
may serve to filter or ‘gate’ synchronous excitatory viewed in Farrant and Nusser, 2005, and in
activity in entorhinal cortex, preventing this type of references therein). Combinatorial calculations
activity from hyperactivating and damaging the based on random assembly of pentameric recep-
relatively fragile hippocampal structures down- tors would result in a million or more possible
stream, and from triggering seizure activity (Heine- GABAA receptors. However, this complexity is
mann et al., 1992; Lothman et al., 1992). Like simplified by the fact that there appears to be a
sparse coding, this gating function of the dentate preferred receptor stoichiometry (usually two a,
gyrus is also due to powerful feedforward and feed- two b, and a g subunit as the prevalent receptor,
back GABAergic inhibition, as well as intrinsic with the g subunit replaced by d, e, or p subunit in
properties of granule cells. Furthermore, the filter rare subsets of receptors). In GABAA receptors
function of the dentate gyrus may be compromised isolated from brain, there may be as few as 10–20
in animals with epilepsy, or in animals in the proc- preferred receptor subunit configurations (Wisden
ess of developing epilepsy, and this loss of function et al. 1992; McKernan and Whiting, 1996; Pirker
may reflect alterations in GABAA receptor expres- et al. 2000). This heterogeneity in subunit compo-
sion and function in dentate granule cells. The in- sition of GABAA receptors has significant func-
tent of the present chapter is to discuss dentate tional consequences. The subunit composition of a
gyrus circuit excitability in context of the GABA GABAA receptor dictates the kinetic properties,
receptors, which granule cells express, and also to cellular and subcellular localization, and pharma-
extend this discussion to how alterations in expres- cology of the receptor when expressed in neurons.
sion and function of inhibitory synaptic receptors in
epileptic animals may disrupt normal operation of
the dentate gyrus. GABAA receptors expressed by dentate granule
cells

Heterogeneous composition and function of GABAA In situ hybridization (Wisden et al., 1992),
receptors in the CNS immunocytochemical (Sperk et al., 1997; Pirker
et al., 2000), and single-cell combined antisense
GABAA receptors are members of the cysteine- RNA amplification profiling-functional studies
loop ligand-gated ion channel family, which, like (Brooks-Kayal et al., 1998) have demonstrated
other members of this family, are pentameric as- that granule cells express as many as 10 or more
semblies of subunits that surround a central ion GABAA receptor subunits, even when only a sin-
selective pore. The properties of the pore in GABA gle cell is measured (Brooks-Kayal et al., 1998). Of
receptors make it primarily permeable to chloride 19 possible subunits, technological and experimen-
and bicarbonate ions. The net result of activation tal issues have limited profiling to 13 subunits or
of GABA receptors is therefore flow of chloride less in most cases. In these studies, granule cells
into cells down its concentration gradient, con- express five of six a subunits (with no expression of
comitant hyperpolarization of the membrane po- a6, and a3 found in some studies, and not in oth-
tential, and a large conductance increase. Both of ers), three of three b subunits, one of three g sub-
these events (hyperpolarization and activation of a units (only g2), as well as d subunits (Wisden et al.,
shunting conductance) usually result in a de- 1992; Sperk et al., 1997; Brooks-Kayal et al., 1998;
creased propensity to fire action potentials in a Pirker et al., 2000).
neuron following GABA receptor activation, i.e. Pharmacological studies examining the actions
inhibition. To date, 19 GABAA receptor subunits of subunit selective agonists and antagonists on
have been cloned from the mammalian CNS, en- synaptic responses in dentate granule cells have
compassing 8 families (a1–6, b1–3, g1–3, d, e, y, p, provided some insight into which of the above de-
and r1–3). Additional complexity is conferred on scribed set of 10 subunits may comprise
237

subsynaptic GABAA receptors. In normal, control composition may play a critical role in discrete
rodents, zolpidem (a hypnotic that preferentially coding of information, and in opposing synchro-
augments responses in a1-containing GABAA nous firing.
receptors markedly enhances inhibitory synaptic What can the properties of a4bd receptors tell us
responses in dentate granule cells, while zinc has about how they may accomplish this role in de-
little or no antagonist effect, even at very high fining circuit function in the dentate gyrus? Re-
concentrations (Buhl et al., 1996; Cohen et al., ceptors lacking the g2 subunit are not clustered in
2003). This suggests that, in addition to g2 sub- synapses (Essrich et al., 1998), and so a4bd recep-
units required to anchor GABAA receptors in tors are primarily localized extra- or perisynaptic-
synapses (Essrich et al., 1998) and an undefined set ally (Wei et al., 2003; Sun et al., 2004). Receptors
of b subunits, a1 subunits are highly expressed in of this composition are endowed with a set of
synaptic GABAA receptors in dentate granule pharmacological properties which make them per-
cells, since GABAA receptors containing these fectly suited to constitute extrasynaptic receptors:
subunits respond preferentially to zolpidem, and they have extremely high affinity for GABA (in the
are very resistant to blockade by zinc. Other a nM compared to the low mM range for synaptic
subunits, such as a2 and a4 may also make a con- receptors), and they exhibit little or no desensiti-
tribution to synaptic responses. The degree of zation (Stell and Mody, 2002; Mtchedlishvili and
contribution of these other subunits has not as yet Kapur, 2006; reviewed in Farrant and Nusser,
been defined experimentally in granule cells. In 2005). Therefore, these receptors can respond to
addition, this ensemble of synaptic GABAA re- ambient or synaptic spillover levels of GABA, and
ceptors changes significantly in animals with epi- can maintain a tonic current for sustained periods
lepsy, which will be discussed in detail below. without desensitizing, both of which define tonic
GABAA receptors (Farrant and Nusser, 2005).
Gene targeting studies in mice support this hy-
Tonic and phasic inhibition in dentate granule cells pothesis, because it was found that deletion of
either the a4 or d subunit significantly reduced
In addition to synaptic GABAA receptors, which tonic current as well as accelerated decay of IPSCs
demonstrate a significant contribution of a1 sub- in dentate granule cells, both of which were ac-
units, dentate granule cells express a relatively companied by alterations in neuronal excitability
unique set of GABAA receptors comprised of a4bd (Spiegelman et al., 2003; Wei et al., 2003; Chandra
subunits. Receptors with this composition are only et al., 2006). Therefore, tonic and spillover-
found to any significant extent in one other region mediated GABAA currents derived from activity
of the brain, the thalamus, where they are ex- of a4bd-containing receptors are a significant
pressed in thalamocortical relay neurons (McKe- and prevalent feature of dentate granule cell ne-
rnan and Whiting, 1996). These two brain regions, urophysiological characteristics, and are impor-
the dentate gyrus and the thalamus, share a func- tant contributors to the function of the dentate
tional requirement to fire action potentials in dis- gyrus.
crete, extremely small populations of neurons, and
to resist synchronous firing during information
transfer. These requirements are evident as ‘sparse The ‘gatekeeper’ function of the dentate gyrus is
coding’ properties discussed above for the dentate maintained by GABAergic inhibition
gyrus, and as somatosensory and visual represen-
tations evident in thalamic projections to cortex, Assessing the role of GABAergic inhibition in reg-
where synchronous firing would disrupt informa- ulating function of the dentate gyrus requires si-
tion transfer. Given the shared circuit require- multaneous recording in afferents to the dentate
ments in these two brain regions, and the selective gyrus, the dentate gyrus itself, and efferents of the
expression of a4bd GABA receptors in only these dentate gyrus during synaptic activation. This is
regions, this suggests that receptors of this necessary to ascertain how the dentate gyrus
238

circuitry may filter and constrain the amplitude of hippocampal activation despite robust, syn-
and duration of afferent inputs as it passes infor- chronous excitation of the dentate gyrus by ent-
mation on to area CA3 of the hippocampus. In orhinal cortical afferents; Fig. 1A–C; see also Ang
addition, these recordings need to be conducted et al., 2006). In addition, these studies also dem-
under conditions where the extracellular milieu onstrate that this dentate filtering is accomplished
can be easily manipulated. by activation of GABAA receptors. Even modest
Several studies have been conducted examining disinhibition (by picrotoxin concentrations suffi-
filtering or ‘gatekeeper’ function of the dentate cient to block 20–25% of synaptic responses) re-
gyrus under control and pathophysiological con- sults in collapse of the gate of the dentate gyrus
ditions (animal models of epilepsy). Behr et al. (Fig. 1D–F), a finding which was also evident in
(1998), using multiple field potential recordings in earlier imaging studies of dentate gyrus function
entorhinal cortex, dentate gyrus, and area CA3, (Iijima et al., 1996).
found that, in control animals, epileptiform bursts In the Carlson et al. (2002a) studies, picrotoxin
originating in entorhinal cortex did not propagate was used in modest concentrations to trigger dis-
through the dentate gyrus to activate CA3 with a inhibition. Picrotoxin is a non-competitive GABAA
similar burst. Rather, the input was filtered, and antagonist, and has equivalent efficacy in blocking
only moderate activity was recorded in CA3. This both lower affinity, synaptic and higher affinity,
contrasted with kindled animals, where the ent- extrasynaptic GABAA receptors (Stell and Mody,
orhinal cortical epileptiform burst activity was 2002). In contrast, gabazine, a competitive GABAA
able to propagate through the dentate gyrus and antagonist, is much more effective in blocking lower
trigger similar epileptiform bursts in area CA3. No affinity, synaptic GABAA receptors than tonic, ex-
attempt was made in the Behr et al. (1998) study to trasynaptic receptors, since it can compete more
determine what aspect of dentate gyrus circuit effectively for the GABA binding site when recep-
physiology was responsible for the filtering func- tors have relatively low affinity for GABA. When
tion in control animals, nor what aspect of circuit applied in nanomolar concentrations, sufficient to
function was perturbed in kindled animals. block 70% of synaptic GABA responses, but only a
We (Carlson et al., 2002a; Ang et al., 2006) have few percent of tonic GABAA receptor responses
utilized voltage sensitive dye imaging techniques to (Stell and Mody, 2002), gabazine had no effect on
monitor function of the dentate gyrus in control gatekeeper function of the dentate gyrus (Carlson
and epileptic animals, as entorhinal cortical affer- et al., 2002b; Carlson and Coulter, unpublished
ents to dentate gyrus are activated. Voltage sen- observations). This implicates tonic GABAA recep-
sitive dyes, when combined with state-of-the-art tors as critical mediators of dentate gyrus filter
CCD cameras, allow monitoring of multiple sites function.
within a brain slice (in our case, 6400 sites), at This concept of a critical role of tonic GABAA
heretofore unprecedented temporal resolution receptors as important regulators of dentate gyrus
(1–5 kHz frame rates). This allows circuit activity function was further supported by a recent study,
to be monitored with high temporal and spatial which described ovarian cycle-linked reduction in
resolution, which can facilitate recording of synap- expression of the d subunit of GABAA receptors in
tic integration within multiple circuits. This per- dentate granule cells, together with a reduction in
mits monitoring of afferent activation of the tonic current, which were accompanied by an in-
dentate gyrus, processing within the dentate, and crease in seizure susceptibility (Maguire et al.,
propagation of dentate signals to efferent struc- 2005). Increased seizure susceptibility could be
tures to be monitored simultaneously, while re- mimicked by experimental downregulation of the d
cording from individual neurons with a patch subunit. This demonstration that reduced expres-
electrode (Ang et al., 2006; see Fig. 1). sion of a single subunit (d) critical for tonic current
Data derived from these studies clearly illustrate in dentate granule cells altered dentate circuit
both the ‘gatekeeper’ function of the dentate (lack function in whole animals further corroborates the
239

Fig. 1. ‘Gatekeeper’ function of the dentate gyrus is maintained by GABAergic inhibition. Simultaneous voltage sensitive dye (A
snapshot taken at the peak of the response, B trace illustrating the VSD response over time), patch clamp (C), and field potential (C)
recording of dentate gyrus response to perforant path activation in control ACSF. Note robust activation of dentate gyrus molecular
layer (red color in A, corresponding to a 10–15 mV EPSP in B), which does not result in activation of downstream structures (note lack
of response in area CA3 in A and B). This lack of CA3 activation is because dentate granule cells do not fire action potentials in
response to perforant path activation under these conditions. This is evident in both the patch (C, top trace, the neuron depolarized to
Vm of 50 mV) and field potential recording (C, bottom trace), due to powerful feedforward inhibition activated by perforant path
stimulation (C, note large IPSP in patch recording). The importance of inhibition in mediating this ‘gatekeeper’ function is illustrated
in responses in panels D, E, and F, following perfusion with 5 mM picrotoxin, a non-competitive GABAA receptor antagonist. This
concentration blocks 20–25% of inhibition (see inset [located above panel E] depicting an averaged spontaneous IPSC [sIPSC] before
and after perfusion with 5 mM picrotoxin). During 25% GABAergic blockade, perforant path activation resulted in powerful ac-
tivation of both the dentate gyrus and downstream structures (CA3 and hilus; D, E). It also triggered action potential firing in dentate
granule cells (see patch and field potential recordings in F, both of which exhibit action potential firing). A grayscale image of the slice,
with patch and field potential recording electrode location is depicted in the inset above A. From Carlson and Coulter (unpublished).
(See Color Plate 14.1 in color plate section.)
240

important role played by these small amplitude, Alterations in pharmacology and subunit expression
sustained currents in regulating function of the of dentate granule cell GABAA receptors in epileptic
dentate gyrus. animals

This upregulation of GABAA receptors appears to


Alterations in GABAA receptor expression in be inconsistent with the hypothesis that compro-
dentate granule cells, and function of the dentate mised inhibition in the dentate gyrus might con-
gyrus in animal models of epilepsy tribute to seizure generation in epileptic animals.
However, experimental evidence provides addi-
Temporal lobe epilepsy is defined by seizures dis- tional clues as to how alterations in inhibition may
charges which activate the temporal lobe, includ- contribute to seizure susceptibility. Not only is
ing the hippocampus. Because the dentate gyrus is there an upregulation in density of GABAA re-
hypothesized to be a critical checkpoint regulating ceptors (discussed above), but the nature (subunit
excitability of the limbic system (Heinemann et al., composition) of the receptors themselves is altered.
1992; Lothman et al., 1992), it has been a focus of This is evident as large-scale changes in the phar-
multiple studies in animal models of epilepsy, as- macological properties of the receptors, and
sessing whether cellular, synaptic, and circuit changes in the expression patterns of subunits en-
properties are altered in a manner consistent with coding GABAA receptors in dentate granule cells.
seizure susceptibility. A primary focus of this work Synaptic GABAA receptors exhibit the de novo
has been GABAA receptor expression and func- appearance of sensitivity to zinc blockade (Buhl
tion, as well as inhibitory synaptic function. et al., 1996; Gibbs et al., 1997; Brooks-Kayal et al.,
1998; Cohen et al., 2003), as well as loss of sensi-
tivity to benzodiazepine site agonists (Gibbs et al.,
Upregulation of synaptic GABAA receptors in 1997; Mtchedlishvili et al., 2001; Leroy et al., 2004).
granule cells of epileptic animals Both of these findings (elevated zinc sensitivity, re-
duced benzodiazepine sensitivity) suggest that dent-
If levels of GABAergic inhibition are critical in ate granule cell GABAA receptors exhibit altered
mediating gatekeeper function of the dentate gyrus, subunit composition following the development of
and compromised filter function of the dentate is a epilepsy. This has been demonstrated in individual
primary contributor to seizure generation in ani- dentate granule cells with this pharmacological
mals with temporal lobe epilepsy, then one might phenotype, with a net downregulation in expression
expect that overall levels of expression of GABAA of a1 subunits, and an upregulation in expression of
receptors might be reduced in epileptic animals. In a4 subunits (Brooks-Kayal et al., 1998). This a
both kindling and post-status epilepticus models of subunit switch, if manifest as an upregulation in a4
temporal lobe epilepsy, studies examining GABAA subunits in synapses, could explain both the en-
receptor function in dentate granule cells have de- hanced sensitivity to zinc blockade and diminished
scribed a paradoxical upregulation in expression of sensitivity to benzodiazepine site agonists, particu-
GABAA receptors, both synaptically (Otis et al., larly the a1-preferring GABA receptor modulator,
1994; Buhl et al., 1996; Nusser et al., 1998; Cohen zolpidem.
et al., 2003), and in whole cell studies (Gibbs et al.,
1997; Brooks-Kayal et al., 1998; Mtchedlishvili et
al., 2001; Leroy et al., 2004). There is a consistent Possible consequences of epilepsy-associated altered
finding of a virtual doubling in the amplitude of subunit composition of GABAA receptors: zinc-
quantal inhibitory synaptic responses, accompanied induced collapse of augmented inhibition
by a doubling in density of GABAA receptors re-
corded in whole-cell GABA responses, evident This altered zinc sensitivity of GABAA receptors
across multiple, divergent models of temporal lobe in animals with epilepsy may have functional sig-
epilepsy. nificance. In addition to alterations in inhibitory
241

synaptic responses, the dentate gyrus of animals appear critical in ‘gatekeeper’ function of the dent-
with epilepsy frequently demonstrates aberrant ate gyrus, this downregulation in d subunit expres-
sprouting of the output axons of granule cells, the sion could compromise function of the dentate
mossy fibers. These axons, perhaps in response to gyrus, despite the concomitant upregulation of
loss of targets in the hilus and area CA3, sprout synaptic GABAA receptors. However, recent cir-
and reinnervate the proximal dendritic tree of cuit studies have failed to identify compromised
other dentate granule cells, creating a reentrant ‘gatekeeper’ function of the dentate gyrus in epi-
excitatory circuit. In addition to releasing gluta- leptic animals (Ang et al., 2006), under stimulus
mate, these mossy fibers release large quantities of parameter conditions where blockade of tonic
zinc (reviewed in Coulter, 2000). Therefore, in the GABAA receptors induced collapse of the dentate
epileptic brain, there is an apposition of the de gate in normal animals (Carlson et al., 2002b). This
novo appearance of a zinc delivery system, suggests that possible dentate gate compromise me-
sprouted mossy fibers, as well as the de novo ap- diated by downregulation in expression of d sub-
pearance of zinc-sensitive synaptic GABAA recep- units (and reduction in tonic GABAA currents) in
tors. The simultaneous manifestation of these two epileptic animals may require certain patterns of
alterations in the epileptic dentate gyrus has led to afferent activation to become operative. Perhaps
the hypothesis that, during periods of repetitive repetitive activation, as during seizure initiation,
afferent activation of the dentate gyrus, which will elevate extrasynaptic GABA concentrations
triggers zinc release, there may be a zinc-induced and enhance tonic current in normal animals
collapse of augmented GABAergic inhibition, fa- (suppressing seizures), and this major check on
cilitating seizure initiation (Buhl et al., 1996; Gibbs excitability will be compromised in animals with
et al., 1997; Coulter, 2000). Evidence supporting epilepsy.
this hypothesis includes the de novo appearance of
zinc-sensitivity to ‘gatekeeper’ function of the
dentate gyrus evident in voltage-sensitive dye im- Conclusions
aging studies (Carlson et al., 2002a).
However, all of the above studies demonstrating The dentate gyrus is a structure characterized by
compromised circuit function and enhanced zinc low cellular excitability of its output neurons, dent-
responses of GABAA receptors have examined the ate granule cells. This low excitability is important
effects of exogenously applied zinc. For the above in cognitive function of the hippocampus, and re-
hypothesis to be operative, zinc released at exci- sults predominantly from the high degree of inhib-
tatory synapses needs to be sufficiently mobile to itory synaptic regulation, as well as the unique
diffuse to and block neighboring GABAergic GABAA receptor properties of these cells, including
synapses. In vitro studies examining the effects of expression of tonic GABAA receptors rarely ex-
endogenously released zinc have as yet failed to pressed in other brain regions. Altered expression
identify effects of zinc release on inhibitory synap- and function of GABAA receptors in dentate gran-
tic responses in slices prepared from epileptic an- ule cells is a prominent feature of temporal lobe
imals (Molnar and Nadler, 2001). epilepsy, a disorder characterized by hypersynchro-
nous discharges involving the hippocampus. This
association of aberrant GABAA receptor properties
Reductions in tonic GABAA current in granule cells in granule cells from hyperexcitable hippocampi
of epileptic animals further demonstrates the critical role of inhibition
in controlling dentate gyrus function.
A second finding has recently been described in
animal models of temporal lobe epilepsy: reduction References
in tonic GABAA receptor current accompanied by
a downregulation in expression of d subunits (Peng Ang, C.W., Carlson, G.C. and Coulter, D.A. (2006) Massive
et al., 2004). Given that tonic GABAA receptors and specific dysregulation of direct cortical input to the
242

hippocampus in temporal lobe epilepsy. J. Neurosci., 26: Jung, M.W. and McNaughton, B.L. (1993) Spatial selectivity of
11850–11856. unit activity in the hippocampal granular cell layer. Hippo-
Behr, J., Lyson, K.J. and Mody, I. (1998) Enhanced propaga- campus, 3: 165–182.
tion of epileptiform activity through the kindled dentate O’Keefe, J. (1976) Place units in the hippocampus of the freely
gyrus. J. Neurophysiol., 79: 1726–1732. moving rat. Exp. Neurol., 51: 78–109.
Brooks-Kayal, A.R., Shumate, M.D., Jin, H., Lin, D.D., Ri- Leroy, C., Poisbeau, P., Keller, A.F. and Nehlig, A. (2004)
khter, T.Y. and Coulter, D.A. (1998) Selective changes in Pharmacological plasticity of GABAA receptors at dentate
single cell GABAA receptor subunit expression and function gyrus synapses in a rat model of temporal lobe epilepsy.
in temporal lobe epilepsy. Nat. Med., 4: 1166–1172. J. Physiol., 557: 473–487.
Buhl, E.H., Otis, T.S. and Mody, I. (1996) Zinc-induced col- Lothman, E.W., Stringer, J.L. and Bertram, E.H. (1992) The
lapse of augmented inhibition by GABA in a temporal lobe dentate gyrus as a control point for seizures in the hippo-
epilepsy model. Science, 271: 369–373. campus and beyond. Epilepsy Res. Suppl., 7: 301–313.
Carlson, G., Lee, C.J. and Coulter, D.A. (2002a) Zinc facilitates Maguire, J.L., Stell, B.M., Rafizadeh, M. and Mody, I. (2005)
hyperexcitability in the hippocampus of epileptic rats. Epile- Ovarian cycle-linked changes in GABA(A) receptors medi-
psia, 43(Suppl 7): 2.031. ating tonic inhibition alter seizure susceptibility and anxiety.
Carlson, G.C., Lee, C.J. and Coulter, D.A. (2002b) The role of Nat. Neurosci., 8(6): 797–804.
inhibition and dentate gatekeeper function: voltage-sensitive McKernan, R.M. and Whiting, P.J. (1996) Which GABAA-
dye imaging study. Soc. Neurosci. Abstr., 28: 145.9. receptor subtypes really occur in the brain. Trends Neurosci.,
Chandra, D., Jia, F., Peng, Z., Suryanarayanan, A., Werner, 19: 139–143.
D.F., Spigelman, I., Houser, C.R., Olsen, R.W., Harrison, McNaughton, B.L. and Morris, R.G.M. (1987) Hippocampal
N.L. and Homanics, G.E. (2006) GABAA receptor a4 sub- synaptic enhancement and information storage within a dis-
units mediate extrasynaptic inhibition in thalamus and dent- tributed memory system. Trends Neurosci., 10: 408–415.
ate gyrus and the action of gaboxadol. PNAS, 103: Molnar, P. and Nadler, J.V. (2001) Lack of effect of mossy
15230–15235. fiber-released zinc on granule cell GABAA receptors in the
Chawla, K.M., Guzowski, J.F., Ramirez-Amaya, V., Lipa, P., pilocarpine model of epilepsy. J. Neurophysiol., 85:
Hoffman, K.L., Marriott, L.K., Worley, P.F., McNaughton, 1932–1940.
B.L. and Barnes, C.A. (2005) Sparse, environmentally selec- Mtchedlishvili, Z., Bertram, E.H. and Kapur, J. (2001) Dimin-
tive expression of Arc RNA in the upper blade of the rodent ished allopregnanolone enhancement of GABAA receptor
fascia dentate by brief spatial experience. Hippocampus, 15: currents in a rat model of chronic temporal lobe epilepsy.
579–586. J. Physiol., 537: 453–465.
Cohen, A.S., Lin, D.D., Quirk, G.L. and Coulter, D.A. (2003) Mtchedlishvili, Z. and Kapur, J. (2006) High-affinity, slowly
Dentate granule cell GABAA receptors in epileptic hippo- desensitizing GABAA receptors mediate tonic inhibition in
campus: enhanced synaptic efficacy and altered pharmacol- hippocampal dentate granule cells. Mol. Pharmacol., 69:
ogy. Eur. J. Neurosci., 17: 1607–1616. 564–575.
Coulter, D.A. (2000) Mossy fiber zinc and temporal lobe ep- Nusser, Z., Hajos, N., Somogyi, P. and Mody, I. (1998) In-
ilepsy: pathological association with altered epileptic creased number of synaptic GABAA receptors underlies po-
GABAA receptors in dentate granule cells. Epilepsia, 41(Sup- tentiation at hippocampal inhibitory synapses. Nature, 395:
pl 6): S96–S99. 172–177.
Essrich, C., Lorez, M., Benson, J.A., Fritschy, J.M. and Lusc- Otis, T.S., De Koninck, Y. and Mody, I. (1994) Lasting po-
her, B. (1998) Postsynaptic clustering of major GABAA re- tentiation of inhibition is associated with an increased
ceptor subtypes requires the gamma 2 subunit and gephyrin. number of gamma-aminobutyric acid type A receptors acti-
Nat. Neurosci., 1: 563–571. vated during miniature inhibitory postsynaptic currents.
Farrant, M. and Nusser, Z. (2005) Variations on an inhibitory PNAS, 91: 7698–7702.
theme: phasic and tonic activation of GABAA receptors. Nat. Peng, Z., Huang, C.S., Stell, B.S., Mody, I. and Houser, C.R.
Rev. Neurosci., 6: 215–229. (2004) Altered expression of the d subunit of the GABAA
Gibbs III, J.W., Shumate, M.D. and Coulter, D.A. (1997) receptor in a mouse model of temporal lobe epilepsy. J. Ne-
Differential epilepsy-associated alterations in postsynaptic urosci., 24: 8629–8639.
GABAA receptor function in dentate granule and CA1 neu- Pirker, S., Schwarzer, C., Wieselthaler, A., Sieghart, W. and
rons. J. Neurophysiol., 77: 1924–1938. Sperk, G. (2000) GABAA receptors: immunocytochemical
Heinemann, U., Beck, H., Dreier, J.P., Ficker, E., Stabel, J. and distribution of 13 subunits in the adult rat brain. Neurosci-
Zhang, C.L. (1992) The dentate gyrus as a regulated gate for ence, 101: 815–850.
the propagation of epileptiform activity. Epilepsy Res. Sup- Sperk, G., Schwarzer, C., Tsunaskima, K., Fuchs, K. and Si-
pl., 7: 273–280. eghart, W. (1997) GABA(A) receptor subunits in the rat
Iijima, T., Witter, M.P., Ichikawa, M., Tominage, T., Kajiwara, hippocampus. I: immunocytochemical distribution of 13 sub-
R. and Matsumoto, G. (1996) Entorhinal-hippocampal in- units. Neuroscience, 80: 987–1000.
teractions revealed by real-time imaging. Science, 272: Spiegelman, I., Li, Z., Liang, J., Cagetti, E., Sanzadeh, S., Mi-
1176–1179. halek, R.M., Homanics, G.E. and Olsen, R.W. (2003)
243

Reduced inhibition and sensitivity to neurosteroids in hip- Wei, W., Zhang, N., Peng, Z., Houser, C.R. and Mody, I.
pocampus of mice lacking the GABA(A) receptor delta sub- (2003) Perisynaptic localization of d-containing GABAA re-
unit. J. Neurophysiol., 90: 903–910. ceptors and their activation by GABA spillover in the mouse
Stell, B.M. and Mody, I. (2002) Receptors with different dentate gyrus. J. Neurosci., 23: 10650–10661.
affinities mediate phasic and tonic GABAA conduc- Wilson, M.A. and McNaughton, B.L. (1993) Dynamics of the
tances in hippocampal neurons. J. Neurosci., 22(1–5): hippocampal ensemble code for space. Science, 261(5124):
RC223. 1055–1058.
Sun, C., Sieghart, W. and Kapur, J. (2004) Distribution of al- Wisden, W., Laurie, D.J., Monyer, H. and Seeburg, P.H. (1992)
pha1, alpha4, gamma2, and delta subunits of GABAA re- The distribution of 13 GABAA receptor subunit mRNAs in
ceptors in hippocampal dentate granule cells. Brain Res., the rat brain. I. Telencephalon, diencephalon, mesencepha-
1029: 207–216. lon. J. Neurosci., 12: 1040–1062.
Plate 14.1. ‘Gatekeeper’ function of the dentate gyrus is maintained by GABAergic inhibition. Simultaneous voltage sensitive dye (A
snapshot taken at the peak of the response, B trace illustrating the VSD response over time), patch clamp (C), and field potential (C)
recording of dentate gyrus response to perforant path activation in control ACSF. Note robust activation of dentate gyrus molecular
layer (red color in A, corresponding to a 10–15 mV EPSP in B), which does not result in activation of downstream structures (note lack
of response in area CA3 in A and B). This lack of CA3 activation is because dentate granule cells do not fire action potentials in
response to perforant path activation under these conditions. This is evident in both the patch (C, top trace, the neuron depolarized to
Vm of 50 mV) and field potential recording (C, bottom trace), due to powerful feedforward inhibition activated by perforant path
stimulation (C, note large IPSP in patch recording). The importance of inhibition in mediating this ‘gatekeeper’ function is illustrated
in responses in panels D, E, and F, following perfusion with 5 mM picrotoxin, a non-competitive GABAA receptor antagonist. This
concentration blocks 20–25% of inhibition (see inset [located above panel E] depicting an averaged spontaneous IPSC [sIPSC] before
and after perfusion with 5 mM picrotoxin). During 25% GABAergic blockade, perforant path activation resulted in powerful ac-
tivation of both the dentate gyrus and downstream structures (CA3 and hilus; D, E). It also triggered action potential firing in dentate
granule cells (see patch and field potential recordings in F, both of which exhibit action potential firing). A grayscale image of the slice,
with patch and field potential recording electrode location is depicted in the inset above A. From Carlson and Coulter (unpublished).
(For B/W version, see page 239 in the volume.)
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 15

Opioid systems in the dentate gyrus

Carrie T. Drake1, Charles Chavkin2 and Teresa A. Milner1,

1
Division of Neurobiology, Department of Neurology and Neuroscience, Weill-Cornell Medical College,
411 East 69th Street, New York, NY 10021, USA
2
Department of Pharmacology, University of Washington, Seattle, WA 98195, USA

Abstract: Opiate drugs alter cognitive performance and influence hippocampal excitability, including long-
term potentiation (LTP) and seizure activity. The dentate gyrus (DG) contains two major opioid peptides,
enkephalins and dynorphins, which have opposing effects on excitability. Enkephalins preferentially bind
to delta- and mu-opioid receptors (DORs and MORs) while dynorphins preferentially bind to kappa-
opioid receptors (KORs). Opioid receptors can also be activated by exogenous opiate drugs such as the
MOR agonist morphine. Enkephalins are contained in the mossy fiber pathway, in the lateral perforant
path (PP) and in scattered GABAergic interneurons. MORs and DORs are predominantly in distinct
subpopulations of GABAergic interneurons known to inhibit granule cells, and are present at low levels
within granule cells. MOR and DOR agonists increase excitability and facilitate LTP in the molecular layer.
Anatomical and physiological evidence is consistent with somatodendritic and axon terminal targeting of
both MORs and DORs. Dynorphins are in the granule cells, most abundantly in mossy fibers but also in
dendrites. KORs have been localized to granule cell mossy fibers, supramammillary afferents to granule
cells, and PP terminals. KOR agonists, including endogenous dynorphins, diminish the induction of LTP.
Recent evidence indicates that opiates and opioids also modulate other processes in the hippocampal
formation, including adult neurogenesis, the actions of gonadal hormones, and development of neonatal
transmitter systems.

Keywords: enkephalin; dynorphin; opioid receptors; GABA interneuron; disinhibition; neurogenesis;


seizure; long-term potentiation

Introduction Endogenous opioid peptides also modulate hippo-


campal excitability (reviewed by Simmons and
Application of opiate drugs alters cognitive per- Chavkin, 1996) and the endogenous hippocampal
formance and influences hippocampal excitability, opioid systems are implicated in learning (Sandin
including long-term potentiation (LTP) (Guerra et al., 1998), including that associated with drug
et al., 1987; Bramham, 1992; Wagner and Chavkin, use (Nestler, 2001). Recent evidence indicates that
1992; Pu et al., 2002; Robbins and Everitt, opiates and opioids also modulate other processes
2002) and seizure activity (Hong et al., 1993). in the hippocampal formation, including adult
neurogenesis, the actions of gonadal hormones,
and development of neonatal transmitter systems
Corresponding author. Tel.: +1 (212) 570 2900; (Eisch et al., 2000; Slamberová et al., 2003;
Fax: +1 (212) 988 3672; E-mail: tmilner@med.cornell.edu Schindler et al., 2004). The dentate gyrus (DG)

DOI: 10.1016/S0079-6123(07)63015-5 245


246

contains several types of opioid peptides, which Opioid receptors


have varying degrees of receptor selectivity, and
opioid receptors, which can be activated by both Three major types of opioid receptors have been
endogenous opioid peptides and exogenous opiate characterized; MORs, DORs, and KORs (reviewed
drugs. Most studies of opioid systems have been by Martin, 1983). These receptor types have vary-
conducted on rat, and although other species stud- ing preferences for endogenous opioid peptides and
ied are generally similar, there are some notable exogenous ligands. MORs have a high affinity for
species differences. endomorphin and the opiate alkaloid morphine
(Martin, 1983; Zadina et al., 1997). MORs and
DORs preferentially interact with enkephalins over
dynorphins, and DORs have a 10-fold higher affin-
Opioid peptides ity than MORs for enkephalins (Corbett et al.,
1993). KORs have a high affinity for dynorphins
The enkephalins and dynorphins are the most but not enkephalins (Chavkin et al., 1982, 1985b;
abundant opioids in the DG. They are derived Corbett et al., 1982) and also bind tightly to the
from two genes with distinct but overlapping cel- exogenous agent ethylketocyclazocine (Corbett
lular distributions. Proenkephalin-derived opioids et al., 1993). As described below, there is also ev-
include [Met5]-enkephalin, [Leu5]-enkephalin, and idence that dynorphins, although primarily KOR
other C-terminally extended forms of [Met5]- agonists, have some affinity for MORS and can
enkephalin (Gubler et al., 1982; Noda et al., modulate N-methyl-D-aspartic acid (NMDA) re-
1982). Prodynorphin-derived opioids include ceptors (Chavkin et al., 1985a; Caudle et al., 1994;
[Leu5]-enkephalin, dynorphin A(1–17), dynorphin Caudle and Dubner, 1998).
A(1–8), dynorphin B(1–13), leumorphin [a.k.a. The complexity of opioid receptors goes far
dynorphin B(1–29)], alpha-neo-endorphin, and beyond three straightforward types. Additional
beta-neo-endorphin (Kakidani et al., 1982). The opioid receptor types and subtypes have been pro-
first five amino acids of dynorphin peptides posed based on receptor-binding studies, receptor
are identical to the (entire) sequence of [Leu5]- autoradiography, and behavioral studies. These
enkephalin. At times this has complicated the include mu1 and mu2 (Wolozin and Pasternak,
localization and study of the individual opioids, 1981), mu/delta (Rothman et al., 1983, 1987),
however, fairly selective agents have been delta1 and delta2 (Jiang et al., 1991; Mattia et al.,
developed. 1991), and kappa1, kappa2, and kappa3 (Nock
The recently discovered opioid endomorphin et al., 1988; Zukin et al., 1988; Clark et al., 1989).
is extremely rare in the rat DG (Pierce and MOR, DOR, and KOR subtypes may arise from
Wessendorf, 2000). Another opioid, beta-endorphin several sources. The MOR, DOR, and KOR have
(a product of the pro-opiomelanocortin precur- been cloned (Evans et al., 1992; Kieffer et al., 1992;
sor), was reported in the hippocampal formation Chen et al., 1993; Minami et al., 1993; Yasuda
(Zakarian and Smyth, 1979, 1982) based on radio- et al., 1993), and although to date only one gene
immunoassay and immunofluorescence; however has been found for each receptor type (Slowe et al.,
later chromatographic work showed that the 1999), other genes may exist. Another potential
immunoreactivity did not correspond to authentic source of subtypes is alternatively spliced opioid
beta-endorphin (Chavkin et al., 1985b). Finally, receptor mRNA (Zimprich et al., 1995; Pan et al.,
nociceptin/orphanin FQ is a hectadecapeptide sim- 1999). In both rat and mouse DG, splice variants
ilar in structure to dynorphin A (Meunier et al., derived from the cloned MOR1-gene are found,
1995; Reinscheid et al., 1995). Although it shares and have slightly different distributions, implying
structural similarity to opioids, its actions and region-specific mRNA processing (Abbadie et al.,
pharmacology are distinct in that it acts through 2000). Yet another source of opioid receptor
the naloxone-insensitive ORL1 (a.k.a NOP) re- diversity is functional association with other pro-
ceptor (Heinricher, 2003). teins. Like other G protein-coupled receptors,
247

opioid receptors form stable associations with one and 2 and discussed in detail below. The present
another (termed ‘‘dimers’’) and with accessory description of opioid peptides is restricted to
proteins. These interactions may serve to increase enkephalins and dynorphins, since the endogenous
pharmacological complexity, to aid in transport or MOR ligand endomorphin is present only in
maturation of the receptors, and to regulate signal rare scattered fibers in the rat DG (Pierce and
transduction. Both homodimers (e.g., DOR–DOR) Wessendorf, 2000).
and heterodimers (e.g., KOR–DOR) have been
demonstrated (see Bouvier, 2001; Rios et al., 2001).
A novel opioid-like receptor, termed ORL1, was Enkephalins
identified based on homology to opioid receptors,
particularly KORs. This receptor has a high affin- Both [met5]-enkephalin and [leu5]-enkephalin are
ity for OFQ/nociceptin, but not opioids (Bunzow present in the DG (Hong et al., 1980; Gall et al.,
et al., 1994). The ORL1 and its endogenous ligand 1981). Enkephalins are present in afferents to the
OFQ/nociceptin are present in the DG and modify DG and in DG neurons in a wide range of species
many of the same functions as do opioids; the including mice, rats, guinea pigs, hamsters, and
history and details of this system are reviewed in humans (Gall, 1988b; Herkenham and McLean,
detail elsewhere (Mogil and Pasternak, 2001). 1988; Rees et al., 1994). Enkephalin is present in
As we describe in the following review of the a subset of mossy fibers (Gall et al., 1981; Gall,
anatomy and physiology of opioid systems in the 1988a), and is expressed by 15% of granule cells
DG, the different opioid systems share at least two (Johnston and Morris, 1994). Leu-enkephalin
features. One, in the hippocampal opioid system immunoreactivity is present in both types of
there is frequently a mismatch between opioid re- terminals formed by mossy fibers, that is, small
ceptors and endogenous opioid peptides. Mis- terminals in the hilus and the large complex mossy
matches at the regional level (e.g., tens or hundreds terminals (Commons and Milner, 1995).
of micrometers) raise the question of how much of Enkephalins also are found in a portion of the
a given pool of opioid peptides reaches opioid lateral perforant path (PP) (Gall et al., 1981;
receptors (see McLean et al., 1987) and predict dis- Fredens et al., 1984; Gall, 1988a), particularly the
tinct effects of endogenously released opioids com- temporal portion of this pathway. At the subcel-
pared to exogenous opiates. Mismatches at the lular level, leu-enkephalin immunoreactivity in the
ultrastructural level indicate that opioids often are rat is primarily associated with distinctive organ-
not active at synapses, but instead function through elles known as large dense-core vesicles (described
nonsynaptic diffusion to nearby cellular targets in more detail in Opioid Peptide Release) in both
(e.g., Drake et al., 1994, 2002). Two, although dif- the small hilar and large mossy terminals. Addi-
ferent opioids have differing effects on excitability, tionally, a few interneuron-like cells in the hilus are
they all share the property of being modulators of leu-enkephalin-immunoreactive (-ir) (Gall et al.,
fast neurotransmission. Their modulatory influ- 1981; Commons and Milner, 1995). Analysis of
ences have regionally selective effects on plasticity proenkephalin mRNA expression indicates addi-
and sometimes exert complex or subtle influences tional proenkephalin-expressing interneurons in
on dentate circuits. Furthermore, physiologically the granule cell layer (GCL) and molecular layer
significant functional effects may occur in regions (Stumm et al., 2004).
with relatively few receptors (Weisskopf et al., 1993). There are some relevant species differences in
dentate enkephalins. Compared to the rat, the
mouse appears to have sparse enkephalin
Anatomical distribution of opioids and immunoreactivity in the hilus (Khalid et al., per-
opioid receptors sonal observation). In the guinea pig there is a
prominent bulge of enkephalin immunoreactivity
The distribution of opioid peptides and their in stratum lucidum (SLu) at the CA3/CA2 border.
receptors is summarized schematically in Figs. 1 Enkephalin-ir fibers can been seen to course along
248

SO CA1
p
SP
SR
SUB

SLM
TAT
PP

p HIL m
SLu g
CA3
DG

GCL
ML

ENK

DYN

Fig. 1. Location of enkephalins and dynorphins in the hippocampal formation. Opioid peptides are found in all regions of the
hippocampal formation: the hippocampus proper (fields CA1 and CA3), the dentate gyrus (DG), and the subiculum (SUB). The
glutamatergic neurons [pyramidal cells (p), granule cells (g), mossy cells (m)] and their projections are shown. GABAergic interneurons
are scattered in all laminae. The perforant path (PP) from entorhinal cortex provides the major excitatory innervation. Subcortical
afferents and contralateral hippocampal projections enter via the fimbria/fornix (not shown) adjacent to CA3. Enkephalins (green) are
in some granule cells, particularly their mossy fiber terminals in hilus and stratum lucidum (SLu). Mossy fibers extensively innervate
hilar mossy cells and CA3 pyramidal cells. Enkephalins are in a portion of the projection from entorhinal cortex; the lateral portion of
the perforant path (PP) in the DG and the temporal-ammonic tract (TAT) in CA1. Enkephalins are in a few scattered interneurons,
which are relatively abundant on the border of stratum radiatum (SR) and stratum lacunosum-moleculare (SLM). Dynorphins (yellow)
are in the granule cells. Dynorphin peptides are most striking in the mossy fiber terminals, and are also in the mossy fiber axons,
granule cell somata, and granule cell dendrites. Abbreviations: SO, stratum oriens; SP, stratum pyramidale; GCL, granule cell layer.
(See Color Plate 15.1 in color plate section.)

the longitudinal axis of the hippocampal forma- studies described below, this observation indicates
tion at this point and are almost certainly in the that prodynorphin-derived peptides play a larger
thick end-bulb formed by the septotemporal pro- role in the DG of the guinea pig than the rat.
jections of mossy fibers (Tielen et al., 1982; Anatomical data are consistent with the sugges-
McLean et al., 1987). In the rat DG, the tion that enkephalins nonsynaptically modulate
enkephalin-containing PP fibers are restricted to inhibitory transmission. In the dentate molecular
the outer molecular layer, that is, the terminal layer, leu-enkephalin-ir terminals occasionally
zone of the lateral PP (Gall et al., 1981). In con- appose GABA-ir terminals, and in the hilus leu-
trast, two bands of enkephalin immunoreactivity enkephalin-ir terminals often contact gamma
are seen in molecular layer of guinea pig, one in amino butyric acid (GABA)-containing perikarya
the lateral PP terminal zone and the other in the and dendrites (Commons and Milner, 1996b).
medial PP terminal zone (Tielen et al., 1982). Ad- These data suggest that leu-enkephalin may have
ditionally, while the rat contains dense enkephalin an important role in regulating inhibition of
immunoreactivity in the PP terminal zone, in the GABA-containing neurons.
guinea pig this immunoreactivity is considerably One interesting aspect of enkephalin biology is
lighter (McLean et al., 1987). Consistent with that the production may be regulated by highly
249

+
Molecular ENK
Layer PP
_ GABA

?
?

+
SubP ACh

Granule Granule ?
Cell Cell
Layer

GABA
Hilus PARV

+
ENK/DYN
+

ACh GABA
NPY
SOM

MOR

to CA3 DOR

KOR
+

Fig. 2. Subcellular locations of opioid peptides and receptors in the dentate gyrus. Enkephalins (green shading) are in: (1) som-
atodendritic, axon, and terminal portions of granule cells; (2) lateral perforant path axons and terminals in the outer molecular layer;
and (3) occasional GABAergic interneurons (not shown). Mu-opioid receptors (MORs; blue) are most frequently in all portions of
parvalbumin (PARV)-containing basket cells that innervate granule cell somata and proximal dendrites. MORs are also in cholinergic
and GABAergic afferents from the lateral septum/diagonal band. Delta-opioid receptors (DORs; pink) are in many hilar interneurons
that contain somatostatin or neuropeptide Y (SOM/NPY) and inhibit the distal dendrites of granule cells. PARV-containing neurons
express DOR mRNA. DORs, and to a lesser extent MORs, are occasionally present in granule cell dendrites. Kappa-opioid receptors
(KORs) in guinea pigs are in substance P containing afferents to granule cells, in granule cell mossy fibers, and most likely on perforant
path terminals in the outer molecular layer. (See Color Plate 15.2 in color plate section.)

region-specific mechanisms. For example, the Dynorphin


biosynthesis and posttranslational proteolytic
processing of proenkephalin in the mossy fiber Dynorphin has a similar distribution in rats, mice,
system is distinct from other brain regions, sug- guinea pigs, squirrels, hamsters, and humans
gesting that particular functions of the enkephalins (Chavkin et al., 1985b; McLean et al., 1987; Gall,
may be elicited depending on which met-enkepha- 1988a; Houser et al., 1990). The vast majority of
lin-containing peptides are produced in a given dynorphin in the hippocampal formation is made
region or subregion (White et al., 1986). by the dentate granule cells, and dynorphin
250

immunoreactivity is abundant in the mossy fiber immunoreactivity than do dendritic shafts, and
pathway in the hilus and CA3 SLu (McGinty dynorphin-containing dense-core vesicles in spines
et al., 1983; Khachaturian et al., 1993). In the hi- are usually near the extrasynaptic plasma mem-
lus, both the large mossy terminals and the smaller brane and seldom near synapses (Drake et al.,
mossy fiber collateral terminals contain dynorphin 1994). In contrast, dynorphin-containing dense-
immunoreactivity (Pierce et al., 1999), suggesting core vesicles in dendritic shafts are often in the
nonselective release near all types of synaptic tar- center of the profile, appearing to be in transit.
gets of these terminals, that is, CA3 pyramidal This localization may be relevant to synaptic plas-
cells, mossy cells, and interneurons. Released ticity. In the outer two-thirds of the molecular
dynorphin is more stable than enkephalin layer, spines receive virtually all of the excitatory
(Wagner et al., 1991) and diffusion may play a input to granule cells. Individual spines in some
large role in the actions of dynorphin. ways act as discrete functional units, with different
Some dynorphin may be contributed by affer- calcium dynamics and chemical events than the
ents. For example, in situ hybridization histo- adjacent dendritic shaft (reviewed by Yuste and
chemistry in rat identified preprodynorphin Tank, 1996). Such compartmentalization is
mRNA-containing perikarya in hippocampally thought to be critical for synapse-specific plastic-
projecting regions of the septum (Merchenthaler ity (e.g., LTP) (Yuste and Tank, 1996). Only 26%
et al., 1997). Also, at the extreme ventral pole of of dynorphin-labeled dense-core vesicles in the
the guinea pig hippocampal formation, a diffuse guinea pig molecular layer are in axons or axon
band of varicose dynorphin-ir processes is found terminals; of these, most are in small terminals
in the molecular layer (Drake et al., 1994). that may have originated in entorhinal cortex or
The granule cell dendrites may also be a relevant are rare collaterals of granule cells. Consistent with
source of dynorphin. As discussed in more detail the latter possibility, the inner molecular layer
below, dynorphin, like enkephalins and other pep- (which is almost completely devoid of entorhinal
tides, is stored in dense-core vesicles that are read- afferents) contains the largest proportion of ax-
ily visible by electron microscopy (Drake et al., onal dense-core vesicles (Drake et al., 1994). In the
1994; Pierce et al., 1999). In the molecular layer of inner molecular layer, dynorphin-containing ter-
the guinea pig, midway along the septotemporal minals form asymmetric synapses on granule cell
axis, electron microscopic analysis showed that al- dendrites. These may be mossy fiber collaterals,
though some dynorphin-immunoreactivity is pre- consistent with physiological evidence for direct
sent in unmyelinated axons, the majority (74%) of granule cell-to-granule cell connections (Terman
dynorphin-labeled dense-core vesicles are found in et al., 2000).
granule cell dendrites (Drake et al., 1994). Den- Interestingly, at the extreme ventral pole of the
dritic dynorphin has also been observed in samples guinea pig hippocampal formation, the number of
from epileptic human DG (Zhang and Houser, axonal dynorphin-labeled dense-core vesicles was
1999). The functional relevance of dynorphin re- higher than at other levels, with no change in the
lease in guinea pig outer molecular layer was sug- number of dendritic dense-core vesicles (Drake
gested by physiological data showing that the time et al., 1994), suggesting that some innervation
course of released dynorphin actions was more differences are present. The most likely possibility
consistent with local release than with release from seems to be the mossy fiber collaterals, which as
hilar mossy fibers (Drake et al., 1994). Release described above are more common at the ventral
of dendritic dynorphin in the molecular layer pole. This dorsal–ventral difference in dynorphin
involves calcium channel types and mechanisms distribution has not been further studied, but is
distinct from axonal dynorphin release (Simmons intriguing in light of other functional and struc-
et al., 1995). tural differences in dynorphin systems of the
There is an interesting compartmentalization of dorsal compared to the ventral hippocampal
dynorphin within guinea pig granule cell dendrites. formation (McDaniel et al., 1990; Pierce et al.,
The dendritic spines contain more dynorphin 1999).
251

Opioid peptide release the DG of mice and rats (Goodman et al., 1980;
Crain et al., 1986; Gulya et al., 1986; Mansour et al.,
Like many other neuropeptides, dynorphin and 1987; McLean et al., 1987; Tempel and Zukin, 1987;
enkephalins are stored in large dense-core vesicles. Mansour et al., 1988). Immunocytochemical labe-
These are larger (80–120 nm diameter) than small ling of DORs in the rat (Mansour et al., 1993) and
synaptic vesicles, the storage vesicles for fast trans- mouse (Bausch et al., 1995a) revealed that DOR-ir
mitters such as GABA or glutamate. Dense-core neurons and processes are scattered in the DG, in a
vesicles are synthesized, transported, and released distribution that overlaps with, but is not restricted
differently from small synaptic vesicles (reviewed to, regions known to contain enkephalin. Interest-
in De Camilli and Jahn, 1990). In particular, ingly, neurons with intense DOR immunoreactivity
dense-core vesicles require a more prolonged, in- (Commons and Milner, 1996a) or high levels of
tense stimulation of the neuron to fuse with the DOR mRNA (Stumm et al., 2004) are found in the
membrane and release their contents (Seward central hilus, while neurons with light DOR-labeling
et al., 1995). This is linked to differential calcium are in the GCL. Close examination with electron
coupling: dense-core vesicle release is facilitated by microscopy (Commons and Milner, 1996a) further
a small, widespread elevation of calcium while small identified DOR immunoreactivity in dendritic spines
synaptic vesicle release requires robust, localized of granule cells, as well as in perikarya and dendrites
calcium elevation (Verhage et al., 1991). Consistent of granule cells and nongranule cells. In DOR-ir
with this, endogenous opioid release in the hippo- dendrites, DOR labeling is most often affiliated with
campal formation requires specific forms of high- the plasmalemmal surface near excitatory-type
intensity stimulation (Wagner et al., 1990, 1991, synapses, consistent with a role in modulating ex-
1993). Dense-core vesicles often fuse with, and re- citatory neurotransmission. Additionally, presynap-
lease their contents from, portions of the membrane tic modulation by DORs is suggested by DOR
that are distant from classical synaptic specializa- immunoreactivity in axon terminals of both mouse
tions (Thureson-Klein and Klein, 1990). Such non- (Bausch et al., 1995a) and rat (Commons and
synaptic or extrasynaptic sites of release can be the Milner, 1996a). DOR-labeled terminals exhibit a
preferential location of dense-core vesicle fusion, range of excitatory and inhibitory-type morph-
as in Aplysia californica (Karhunen et al., 2001). ologies and are in all dentate lamina, consistent
In rat and guinea pig mossy fiber terminals, with a variety of neuronal sources (Commons and
dense-core vesicles containing enkephalin- or Milner, 1996a). However, DOR immunoreactivity is
dynorphin-immunoreactivities are distributed found on the plasmalemmal surface in axon termi-
along both extrasynaptic and synaptic portions nals less frequently than in dendrites.
of the membrane (Drake et al., 1994; Commons Most DOR-labeled perikarya contain GABA
and Milner, 1995; Pierce et al., 1999). The majority (Bausch et al., 1995a; Commons and Milner,
of these dense-core vesicles are extrasynaptic, and 1996a; Stumm et al., 2004). DOR has been identi-
changes in the distribution of membrane-associ- fied within several interneuron subtypes, although
ated dense-core vesicles following seizure suggest variations in prevalence have been reported de-
that dynorphin-containing dense-core vesicles are pending on the technique used. For example, with
specifically targeted to particular portions of the dual-labeling immunocytochemistry 75% of DOR-
terminals (Pierce et al., 1999). ir neurons sampled in the hilus contained neuro-
peptide Y (NPY) immunoreactivity and 39% con-
tained somatostatin (SOM) immunoreactivity
Opioid receptors (Commons and Milner, 1996a). Conversely, 75%
of NPY neurons and 93% of SOM neurons con-
Delta-opioid receptors (DORs) tained DOR immunoreactivity (Commons and
Milner, 1996a). This suggests that almost all of
Functional DOR binding, as indicated by receptor the somatostatinergic cells in the hilus, which are a
autoradiography, is diffusely distributed throughout subpopulation of NPY neurons, contain DOR. In
252

this study, few DOR-ir neurons were found in the binding, although both are similarly distributed in
infragranular zone (home to many of the parval- a diffuse pattern (Mansour et al., 1988). By light
bumin (PARV)-containing basket cells). On the microscopy, MOR immunoreactivity is present in
other hand, using a combination of immunocyto- scattered neurons in the rat and mouse DG
chemistry and in situ hybridization, DOR mRNA (Arvidsson et al., 1995; Bausch et al., 1995b;
was observed in more than 90% of the PARV- Mansour et al., 1995; Ding et al., 1996; Bausch
positive neurons sampled in the GCL and hilus, in and Chavkin, 1997; Kalyuzhny and Wessendorf,
11% of hilar SOM-ir neurons, and in none of the 1997; Drake and Milner, 1999). Dual-labeling
calretinin-ir neurons (Stumm et al., 2004). The rea- immunocytochemistry and ultrastructural mor-
son for the discrepancy in prevalence with the ear- phological characterization showed that MOR-
lier study (Commons and Milner, 1996a) is not labeling is in both the somatodendritic and axonal
clear. Possibilities include translation-induced compartments of GABAergic interneurons (Drake
differences (e.g., low production of the DOR pro- and Milner, 1999). Notably, MOR is common in
tein in PARV-containing neurons or low levels of presumed basket cell terminals that form inhibi-
SOM production in some of the neurons containing tory-type synapses with granule cell somata and
SOM mRNA), or perhaps a dimerization of dendrites (Drake and Milner, 1999). Analysis of
MORs and DORs (George et al., 2000; Gomes the neurochemical content of MOR-labeled cells
et al., 2000) in PARV-containing neurons that leads indicates that MOR mRNA (Stumm et al., 2004)
to lower immunocytochemical recognition of DOR and immunoreactivity (Drake and Milner, 2006)
protein. are most frequent in PARV-containing interneu-
None of the DOR mRNA-containing neurons rons, which comprise one group of basket cells
in the dentate were observed to contain pro- (see Freund and Buzsáki, 1996 for review). MOR
enkephalin mRNA (Stumm et al., 2004), suggest- immunoreactivity (Drake and Milner, 2006) and
ing little autoregulation of DOR-bearing interneu- mRNA (Stumm et al., 2004) are also a modest
rons. However, since enkephalins and modest number of the SOM-containing hilar interneurons
DOR levels have both been observed in granule (the ‘‘HIPP’’ neurons that project to the outer
cells, autocrine regulation of mossy fibers or gran- molecular layer). In the outer molecular layer,
ule cell dendrites remains a possibility. MOR-ir axons and terminals form inhibitory-type
In summary, DORs are primarily in several in- synapses with dendrites containing NMDA recep-
terneuron populations that inhibit granule cells, tor immunoreactivity (Milner and Drake, 2001).
are present at low levels in granule cells, and are These results are consistent with physiological
absent from the calretinin-containing interneurons studies described below, and suggest that opioids
that inhibit other interneurons. Since the targets of could inhibit GABAergic terminals that modulate
the identified DOR-containing interneuron popu- granule cell dendrites, boosting depolarizing
lations are relatively well understood (for review events in granule cells and facilitating the activa-
see (Freund and Buzsáki, 1996), a functional link tion of NMDA receptors. Finally, MOR mRNA is
can be postulated. The presence of DORs in SOM/ in a small proportion (10%) of calretinin-contain-
NPY-containing interneurons suggests a role in ing interneurons in the GCL (Stumm et al., 2004).
the inhibition of granule cell distal dendrites, and This is in contrast to DOR mRNA, which is
in PARV neurons (if the DOR protein is indeed absent from calretinin-ir neurons (Stumm et al.,
made by these cells) suggests a role in inhibiting 2004). Together the above morphological and
granule cell output. dual-labeling studies suggest that MORs are most
frequent in interneurons specialized to inhibit
granule cell output (i.e., PARV-containing basket
Mu-opioid receptors (MORs) cells), are in a limited number of interneurons
that inhibit granule cell distal dendrites, and are
Receptor autoradiography has shown that MOR occasionally in interneurons that innervate other
binding in the DG is more prominent than DOR interneurons.
253

The subcellular location of MOR immunoreac- MORs are in a few other interesting locations in
tivity is consistent with extrasynaptic activation by the DG. First, MOR immunoreactivity is seen in
diffusing endogenous opioids. In somata and occasional dendrites of granule cells (Drake and
dendrites of GABAergic neurons, MOR immuno- Milner, 1999). This location is consistent with
reactivity is often along the plasma membrane but other evidence that MOR agonists can act directly
rather distant from synapses (Drake and Milner, on granule cells (Piguet and North, 1993; Stumm
1999). Moreover, in the rat hilus, profiles bearing et al., 2004), and contrasts with the hippocampus
MORs are not contacted by many leu-enkephalin- proper, where MORs are absent from principal
labeled profiles, but about one-third of MOR-ir cells (Drake and Milner, 1999). Second, MORs are
axons, terminals, and dendrites are within 3 mm of on some afferents to the DG. MOR labeling is in
leu-enkephalin-ir profiles. An estimation of the cholinergic septal terminals in the hilus (Kaplan
effective range of one leu-enkephalin-containing et al., 2004) and in a subset of GABAergic septal
vesicle can be made by calculating how far leu- afferents (Alreja et al., 2000). MOR-ir terminals
enkephalin released from a single dense-core ves- also contact the same hilar neuronal targets as do
icle would diffuse to reach a concentration that cholinergic terminals (Kaplan et al., 2004). These
would activate half of the available MORs. By findings suggest the existence of numerous inter-
knowing the estimated affinity of leu-enkephalin actions between opioid systems and septal projec-
for MORs, the concentration of leu-enkephalin in tions in the DG.
a single dense-core vesicle, and the extracellular
volume fraction of the hilus, we have calculated
that the approximate volume occupied by relevant Kappa-opioid receptors (KORs)
concentration of peptide from one dense-core ves-
icle is 35–104 fl, with a radius of 2.03–2.92 mm (for The rat DG contains little KOR binding and vir-
more detail see Drake et al., 2002). Thus, many tually no KOR immunoreactivity (Mansour et al.,
MORs are likely to be within a functionally rel- 1988, 1996; Unterwald et al., 1991). [However, see
evant distance of released enkephalin (Drake et al., Maderspach et al., 1991, 1995)] Considerably more
2002). The lack of synaptic associations suggests KOR binding is found in the guinea pig DG
that activation of MORs by leu-enkephalin re- (McLean et al., 1987; Wagner et al., 1991). In
quires volume transmission, and furthermore that guinea pig, KOR binding is most intense in the
dendritic, axonal, and terminal MORs are equally dentate outer molecular layer, with a thin band in
likely to be activated by leu-enkephalin released the inner molecular layer (McLean et al., 1987).
from mossy fibers. There is also a possibility that The outer molecular layer KORs disappear fol-
some MORs are activated in an autocrine fashion lowing entorhinal cortex lesions, consistent with a
by enkephalins. One group observed that 75% of presynaptic localization in PP afferents (Simmons
hilar cells with MOR mRNA also label for pro- et al., 1994). At least two subtypes of pharmaco-
enkephalin mRNA (Stumm et al., 2004). However, logically identified KORs (K1 and K2) have been
these neurons may not translate both MORs and identified in the guinea pig DG (Zukin et al., 1988;
enkephalin, as proenkephalin- and MOR- mRNAs Unterwald et al., 1991). Two groups, using differ-
are not seen in the same transmitter-identified hilar ent carefully characterized antibodies, have local-
populations (Stumm et al., 2004) and an electron ized KOR immunoreactivity in the guinea pig
microsopic study of the hilus showed little colo- brain, and found a pattern that largely overlaps
calization of immunoreactivities for MOR and leu- with KOR binding but has some notable differ-
enkephalin in dendrite or axon terminal profiles ences (Arvidsson et al., 1994; Drake et al., 1996).
(Drake et al., 2002). This suggests that, at least in The light microscopic distribution of KOR
the hilus, autoregulation of neurotransmitter re- immunoreactivity overlaps the distribution of
lease by MORs may be minor. Rather, enkephalins KOR binding in the inner molecular layer (Drake
released from the mossy fibers or other interneu- et al., 1996), but is much less intense in the outer
rons appear more likely to activate hilar MORs. molecular layer. This initially surprising result
254

suggests that in these latter regions KOR-binding hypothalamus, translocation of KORs to the
may reflect receptors that are less accessible to the plasma membrane is observed following the stim-
antibodies (due to alternative conformations or ulation of peptide release (Shuster et al., 1999),
associations with other proteins), or are different providing direct evidence for activity-dependent
subtypes. KOR immunoreactivity in the inner presentation of KORs to the membrane.
molecular layer is near (within 100 mm) large stores
of dynorphin, but these immunoreactivities do not
overlap, suggesting diffusion must occur if the
KORs are to be activated (Drake et al., 1996). The Opioid physiology
reliance on diffusion is consistent with other
strong evidence that released dynorphin can Mu- and delta-opioid actions in the hippocampus
diffuse a sufficient distance to activate dentate are largely disinhibitory (Zieglgänsberger et al.,
KORs (Wagner et al., 1991, 1993; Drake et al., 1979). For example, enkephalin was found to
1994). hyperpolarize GABAergic interneurons in the rat
KOR immunoreactivity in the granule cell and hippocampus (Madison and Nicoll, 1988), and
inner molecular layers is restricted to unmyelin- MOR and DOR activation blocked GABAergic
ated axons and axon terminals. Many KOR input and induced disinhibitory effects in the rat
immunolabeled terminals colocalize substance P DG (Neumaier et al., 1988). Whole-cell voltage
(SubP) and form asymmetric synapses with gran- clamp recordings of granule cells also showed that
ule cell perikarya and large unlabeled dendrites mu-receptor activation reduced monosynaptic
concentrated in the dentate inner molecular layer stimulation-evoked inhibitory post-synaptic cur-
(Drake et al., 1997). Following fornix lesions, rents (IPSCs) (Xie et al., 1992). The reduction in
KOR- and SubP-ir processes dramatically de- GABAergic tone generally increases the excitabil-
crease in the molecular layer. These data suggest ity of the hippocampus in ways that affect seizure
a role for KORs in presynaptically modulating susceptibility and synaptic plasticity [i.e., LTP and
supramammillary afferents to granule cells, thus long-term depression (LTD)] (Cohen et al., 1992;
helping to regulate information flow into the DG. Morris and Johnston, 1995). LTP induction by
KORs in the DG have an interesting subcellular high-frequency stimulation of the PP input to the
location. There is an association with the inner DG in rat hippocampal slices was facilitated by
surfaces of plasma membranes (Drake et al., MOR and DOR activation (Bramham and Sarvey,
1996), as expected for the cytoplasmic portion of 1996) and this was found to be caused by disin-
functional cell-surface receptors. KOR immuno- hibition.
reactivity is also found in large dense-core vesicles The inhibition of GABAergic neurons by
(Drake et al., 1996). This localization suggests the MORs and DORs is due to membrane hyperpo-
possibility that KORs are incorporated into the larization, and the ionic basis has been character-
presynaptic membrane when the dense-core vesi- ized by several groups using voltage clamp
cle’s contents are released. Since strong stimula- electrophysiology. Mu opioids activate both in-
tion is required to release dense-core vesicles ward-rectifying and voltage-gated potassium chan-
(Thureson-Klein and Klein, 1990), the implication nels in hippocampal interneurons (Wimpey and
is that these KORs may be a reserve pool of Chavkin, 1991). Opioids can also increase the
receptors that undergo excitation-dependent pres- M-type potassium channel conductance in hippo-
entation to the plasma membrane. Such a mech- campus (Moore et al., 1994). G protein-coupled
anism is of obvious relevance for a role in LTP or inwardly rectifying potassium channels (GIRK or
in response to hyperstimulation. A presence in Kir3) can be activated by the MOR agonist DAM-
dense-core vesicles may be a common property GO in slices from mice (Luscher et al., 1997).
of KORs, since KORs have been localized to These multiple hyperpolarizing mechanisms re-
dense-core vesicles in other brain regions (Shuster duce inhibitory input within the DG and indirectly
et al., 1999; Svingos et al., 1999). Moreover, in increase granule cell excitability.
255

Kappa opioids can directly affect presynaptic dentate opioids has been reviewed elsewhere
release of excitatory amino acids in the guinea pig (Tortella et al., 1988; Simmons and Chavkin,
DG, leading to inhibition of excitatory transmis- 1996; Simonato and Romualdi, 1996), and will be
sion. (However, in rat, disinhibition resembling discussed only briefly here. Levels of enkephalins
MOR and DOR activation has been reported and dynorphins in the hippocampal formation are
(Neumaier et al., 1988).) Intracellular recordings altered in humans with temporal lobe epilepsy
of dentate granule cells demonstrated that KOR (Houser et al., 1990; Rees et al., 1994), as well as in
activation significantly reduced the amplitude of rodent seizure models (Hong et al., 1993). Dentate
glutamatergic excitatory post-synaptic potentials levels of enkephalin and dynorphin are altered in
(EPSPs) while having no direct effects on mem- rodents by seizure-inducing treatments, including
brane conductance (Wagner et al., 1992). Provid- electroconvulsive shock, amygdala kindling
ing further evidence for presynaptic inhibition, a (Vindrola et al., 1981; McGinty et al., 1986), and
KOR-selective agonist inhibited excitatory re- administration of excitatory chemical agents
sponses to PP stimulation but not to applied (Hong et al., 1980, 1988; Pierce et al., 1999; Pierce
glutamate, and potentiated paired-pulse facilita- and Milner, 2001). The time course and direction
tion (Simmons et al., 1994). Kappa receptor acti- of peptide changes varies; in general enkephalin
vation blocked excitatory amino acid release from mRNA and immunoreactivity increase within
PP and recurrent hilar input to granule cells in the hours of seizure while dynorphin decreases rap-
DG, and this inhibition reduced LTP induction of idly but then recovers (for reviews, see Hong et al.,
these two pathways (Terman et al., 1994, 2000). 1993; Pierce et al., 1999). Opioid receptors in the
Much of the evidence for KOR agonist inhibition DG also change following seizures, with both
of LTP in the DG (and in region CA3) has been MORs and DORs showing a shift in distribution
reviewed by (Morris and Johnston, 1995). that is in line with anatomical reorganization of
Opioids cause presynaptic inhibition of certain particular neuronal circuits, such as mossy fiber
subcortical afferents to the DG. Neurochemical sprouting (Bausch and Chavkin, 1997; Skyers et
measures of transmitter release showed that the al., 2003). Electrophysiological responses to MOR
stimulated release of serotonin (Yoshioka et al., and DOR agonists also are altered after such sei-
1993), norepinephrine (Jackisch et al., 1984; zure-associated circuit reorganization (Bausch and
Matsumoto et al., 1994), and acetylcholine Chavkin, 1997), consistent with a persistent func-
(Jackisch et al., 1986) were each directly inhibited tional change in the dentate opioid systems.
by opioid receptor activation. The ionic mechanisms Opioid regulation of excitability within the DG
of these inhibitory presynaptic actions have not and hippocampus also has profound effects on
been directly defined because it is difficult to electro- seizure susceptibility. Iontophoretic application of
physiologically record from hippocampal nerve ter- opioids produced epileptiform increases in excita-
minals, but molecular mechanisms that have been bility (French and Siggins, 1980). Morphine-
suggested include activation of delayed rectifying pretreated rats show enhanced susceptibility to
potassium channels (Wimpey and Chavkin, 1991), amydala kindling and morphological changes in
opioid inhibition of voltage-sensitive calcium chan- the DG (Rocha et al., 1996). Seizure development
nels by Gbg binding (Herlitze et al., 1996; Ikeda, is thought to involve MOR activation, since MOR
1996), or direct inhibition of the vesicular fusion agonists, but not DOR agonists, produce convul-
machinery by opioid receptor activation (Scholz sions and wet-dog shakes (Lee et al., 1989; Hong
and Miller, 1992; Capogna et al., 1996). et al., 1993). However, in some reports, MOR
agonists administered systemically had anti-
convulsant actions that may reflect their inhibi-
Opioids and seizures tory actions on other brain regions (reviewed by
Simmons and Chavkin, 1996). KOR activation
Seizures modulate, and are modulated by, hippo- has been consistently reported to dampen seizure
campal opioid systems. The topic of seizures and (reviewed by Tortella et al., 1988). For example,
256

KOR activation in rats reduces excitability of the CA3 compared to rats in diestrus. Rats with 24 h
dentate and reduces pilocarpine-induced seizures (but not 6 or 72 h) of estrogen replacement showed
(Bausch et al., 1998). This anti-convulsant action is increased levels of leu-enkephalin. Dynorphin-
consistent with the inhibition of excitatory neuro- immunoreactivity was elevated only in the apex
transmission in the DG and CA3 produced by of the hilus in proestrus rats, but showed signifi-
exogenous and endogenous KOR agonists. cant increases in the entire hilus and CA3 SLu 24 h
after estrogen replacement in ovariectomized rats.
These data suggest that during the estrogen peaks
Opioids and gonadal steroids of the ovarian cycle, a greater amount of
enkephalin is available for release from mossy
There is ample precedent for interactions between fibers and perhaps enkephalinergic interneurons,
opioids and gonadal steroids in other brain regions suggesting greater excitability in hippocampal cir-
(e.g., hypothalamus). Because of this, and because cuits. We also found that immunoreactivity for
both opioids (for details see Opioid Physiology) the estrogen receptor ERb was colocalized with
and gonadal steroids (for review see Woolley and enkephalin- or dynorphin-labeling in mossy fiber
Schwartzkroin, 1998) modulate hippocampal plas- terminals and smaller terminals. Thus, estrogen
ticity, it has been hypothesized that these two sys- may directly influence enkephalin and/or dynor-
tems may interact in the hippocampal formation. phin release through activation of estrogen recep-
A few studies have examined gonadal steroid tors on opioidergic neurons.
effects on opioid receptors. Estradiol treatment
has been reported to increase MOR binding in
hippocampal homogenates (Piva et al., 1995). Opioids and neurogenesis
Since estrogens do not appear to alter levels of
hippocampal MOR mRNA (Quinones-Jenab The DG is one of the few brain regions that un-
et al., 1997), the increased MOR binding presum- dergoes neurogenesis throughout adulthood (see
ably involves another mechanism, such as un- Gould and Gross, 2002). New granule cells, and
masking, receptor dimerization, or increased interneurons (Liu et al., 2003), are generated in the
translation efficiency. subgranular zone of the hilus by progenitor cells
Evidence also suggests that sex hormones may (Seri et al., 2001; Turlejski and Djavadian, 2002).
affect levels of dentate opioid peptides. One group As noted in earlier sections, granule cells express
(Roman et al., 2006) used radioimmunoassay to low levels of MORs and DORs (Commons and
determine whether opioid peptide levels changed Milner, 1996a; Drake and Milner, 1999), and con-
across the estrous cycle. They found no significant tain enkephalin (Gall et al., 1981) and dynorphin
changes in the levels of dynorphins or pro- (Chavkin, 2000). Some hilar interneurons likewise
enkephalin-derived peptides in homogenates of contain opioid receptors and/or enkephalin
the hippocampal formation, however, the ratio of (Gall et al., 1981; Commons and Milner, 1996a).
dynorphins to their conversion products (including Neurogenesis has been linked to certain forms
perhaps leu-enkephalin) fluctuated. This suggests of hippocampal-dependent associative learning
that enzymatic processing of dynorphin peptides (Shors et al., 2001, 2002) and is influenced by
changes across the estrous cycle, and may affect many factors; it is enhanced by exercise and anti-
the availability of dynorphin and its derivatives. depressant drugs, inhibited by stress, and fluctu-
Using anatomical methods and focusing on par- ates across the female estrous cycle (Tanapat et al.,
ticular subregions, we recently found evidence that 1999; Eisch et al., 2000; Gould and Gross, 2002).
estrogen affects the levels of leu-enkephalin and Since cognitive impairment is one of the problems
dynorphin immunoreactivity in the DG and the associated with chronic opiate use, and opiates can
CA3 SLu (Torres-Reveron et al., 2006a, b). Rats impair neurogenesis in the prenatal brain, Eisch
in proestrus (high estrogen) have significantly et al. (2000) examined whether opiates also may
higher levels of leu-enkephalin in the hilus and in affect neurogenesis in adult rats. Although an
257

acute exposure to morphine did not affect neuro- Prenatal morphine exposure produces functional
genesis, neurogenesis was inhibited by chronic consequences for the hippocampal formation. Al-
morphine treatment or heroin self-administration. though rats that are exposed to morphine prenatally
The mechanism requires opioid-receptor activa- did not have a reduced seizure threshold to MOR
tion, as evidenced by sensitivity to the opioid re- agonists, they showed an opioid-receptor dependent
ceptor antagonist naltrexone, and does not involve decrease in sensitivity to bicuculline-induced seizures
stress hormones (Eisch et al., 2000). Newly born (Schindler et al., 2004). Similarly, in slices from
neurons contain MORs (Eisch and Harburg, adult male rats, prenatal morphine exposure was
2006), consistent with the possibility that mor- reported to enhance susceptibility to epileptiform
phine may directly act on newly born neurons. activity in the outer molecular layer, but to shift
These findings suggest that inhibition of neuro- long-term plasticity of PP-granule cell synapses in
genesis is one mechanism by which opiates may favor of LTD (Velı́sek et al., 2000). Clearly the
exert long-lasting effects on dentate circuits. effects of prenatal morphine are complex, and
involve both afferents and intrinsic dentate circuits.

Prenatal morphine and opioid system development


Summary
Prenatal exposure to opiates has many conse-
quences for the developing brain, including alter- Opiate drugs alter cognitive performance and in-
ation of the dentate opioid system and of fluence hippocampal excitability, including LTP
phenomena known to be influenced by opioids, and seizure activity. In the DG, the two major
such as LTP and seizure threshold. In a study of opioid peptides, enkephalins, and dynorphins,
the DG of adult male rats exposed to morphine have opposing effects on excitability. Enkephalins
prenatally, proenkephalin mRNA and met-enke- preferentially bind to DORs and MORs while
phalin peptides were decreased, while prodynor- dynorphins preferentially bind to KORs. Opioid
phin mRNA and dynorphin B peptides were receptors can also be activated by exogenous opi-
increased (Schindler et al., 2004). Increases in the ate drugs such as morphine. Enkephalins are con-
density of MOR, but not DOR, binding were re- tained in the mossy fiber pathway, in the lateral PP
ported in some strata of hippocampal subfields in and in scattered GABAergic interneurons. MORs
morphine-exposed males (Schindler et al., 2004). and DORs are predominantly in distinct sub-
In an investigation of the interactions between populations of GABAergic interneurons known to
prenatal morphine exposure and adult hormone inhibit granule cells, and are present at low levels
status (Slamberová et al., 2003), MOR density in within granule cells. MOR and DOR agonists
the DG was reduced in prenatal, morphine- increase excitability and facilitate LTP in the mo-
exposed males that received gonadal hormone re- lecular layer. Anatomical and physiological evi-
placement compared to gonadectomized males. dence is consistent with somatodendritic and axon
Controls exposed to prenatal saline did not show terminal targeting of both MORs and DORs.
this difference. There was a lower density of Dynorphins are most abundant in mossy fibers but
MORs in the DG of females after hormone re- also are in the granule cells and their dendrites.
placement, but this occurred with either prenatal KORs have been localized to granule cell mossy
morphine treatment or with prenatal saline expo- fibers, supramammillary afferents to granule cells,
sure, suggesting that interactions between prenatal and PP terminals. KOR agonists, including en-
morphine and adult hormone status were less rel- dogenous dynorphins, diminish the induction of
evant in the female. Interestingly, the hippocam- LTP. Opiates and opioids also modulate other
pus proper was affected differently than the DG, processes in the hippocampal formation, including
with prenatal exposure to morphine altering the adult neurogenesis, the actions of gonadal hor-
density of MORs in adult female but not adult mones, and development of neonatal transmitter
male rats (Slamberová et al., 2003). systems.
258

Abbreviations Arvidsson, U., Riedl, M., Chakrabarti, S., Lee, J.H., Nakano,
A.H., Dado, R.J., Loh, H.H., Law, P.Y., Wessendorf, M.W.
DG dentate gyrus and Elde, R. (1995) Distribution and targeting of a mu-
opioid receptor (MOR1) in brain and spinal cord. J. Neuro-
DOR delta-opioid receptor sci., 15: 3328–3341.
EPSPs excitatory post-synaptic potentials Arvidsson, U., Risling, M., Frisén, J., Piehl, F., Fried, K.,
GABA gamma amino butyric acid Hökfelt, T. and Cullheim, S. (1994) trkC-like immunoreac-
GCL granule cell layer tivity in the primate descending serotoninergic system. Eur. J.
GIRK G protein-coupled inwardly recti- Neurosci., 6: 230–236.
Bausch, S.B. and Chavkin, C. (1997) Changes in hippocampal
fying potassium channels circuitry after pilocarpine-induced seizures as revealed by
ir immunoreactive opioid receptor distribution and activation. J. Neurosci., 17:
KOR kappa-opioid receptor 477–492.
LTD long-term depression Bausch, S.B., Esteb, T.M., Terman, G.W. and Chavkin, C.
LTP long-term potentiation (1998) Administered and endogenously released kappa op-
ioids decrease pilocarpine-induced seizures and seizure-
MOR mu-opioid receptor
induced histopathology. J. Pharmacol. Exp. Ther., 284:
ORL1 opioid-like receptor 1147–1155.
PARV parvalbumin Bausch, S.B., Patterson, T.A., Appleyard, S.M. and Chavkin,
PP perforant path C. (1995a) Immunocytochemical localization of delta
SLM stratum lacunosum-moleculare opioid receptors in mouse brain. J. Chem. Neuroanat., 8:
175–189.
SLu stratum lucidum
Bausch, S.B., Patterson, T.A., Ehrengruber, M., Lester, H.A.,
SO stratum oriens Davidson, N. and Chavkin, C. (1995b) Colocalization of mu
SOM/NPY somatostatin/neuropeptide Y opioid receptors with GIRK1 potassium channels in rat
SP stratum pyramidale brain: an immunocytochemical study. Receptors Channels, 3:
SubP substance P 221–241.
Bouvier, M. (2001) Oligomerization of G-protein-coupled
SR stratum radiatum
transmitter receptors. Nat. Rev. Neurosci., 2: 274–286.
SUB subiculum Bramham, C.R. (1992) Opioid receptor dependent long-term
TAT temporal-ammonic tract potentiation: peptidergic regulation of synaptic plasticity in
the hippocampus. Neurochem. Int., 20: 441–455.
Bramham, C.R. and Sarvey, J.M. (1996) Endogenous activa-
tion of m and d-1 opioid receptors is required for long-
Acknowledgments term potentiation induction in the lateral perforant path:
Dependence on GABAergic inhibition. J. Neurosci., 16:
We thank Ms. Katherine Mitterling and Mr. 8123–8131.
Bunzow, J.R., Saez, C., Mortrud, M., Bouvier, C., Williams,
Bradley Graustein for technical assistance. Sup-
J.T., Low, M. and Grandy, D.K. (1994) Molecular cloning
ported by NIH grants: DA 08259 (T.A.M; and tissue distribution of a putative member of the rat opioid
C.T.D.; C.C.) and HL 18974 (T.A.M.; CTD) and receptor gene family that is not a mu, delta or kappa opioid
DA04123 (C.C.). receptor type. FEBS Lett., 347: 284–288.
Capogna, M., Gahwiler, B.H. and Thompson, S.M. (1996)
Presynaptic inhibition of calcium-dependent and -independ-
ent release elicited with ionomycin, gadolinium, and alpha-
References latrotoxin in the hippocampus. J. Neurophysiol., 75:
2017–2028.
Abbadie, C., Pan, Y.X., Drake, C.T. and Pasternak, G.W. Caudle, R.M., Chavkin, C. and Dubner, R. (1994) Kappa2
(2000) Comparative immunohistochemical distributions of opioid receptors inhibit NMDA receptor-mediated synaptic
carboxy terminus epitopes from the mu-opioid receptor currents in guinea pig CA3 pyramidal cells. J. Neurosci., 14:
splice variants MOR-1D, MOR-1 and MOR-1C in the mouse 5580–5589.
and rat CNS. Neuroscience, 100: 141–153. Caudle, R.M. and Dubner, R. (1998) Ifenprodil blocks the ex-
Alreja, M., Shanabrough, M., Liu, W.M. and Leranth, C. citatory effects of the opioid peptide dynorphin 1–17 on
(2000) Opioids suppress IPSCs in neurons of the rat medial NMDA receptor mediated currents in the CA3 region of the
septum/diagonal band of Broca: Involvement of m-opioid guinea pig hippocampus. Neuropeptides, 32: 87–95.
receptors and septohippocampal GABAergic neurons. Chavkin, C. (2000) Dynorphins are endogenous opioid peptides
J. Neurosci., 20: 1179–1189. released from granule cells to act neurohumorly and inhibit
259

excitatory neurotransmission in the hippocampus. Prog. P-containing subcortical afferents in guinea pig dentate
Brain Res., 125: 363–367. gyrus. Hippocampus, 7: 36–47.
Chavkin, C., Henriksen, S.J., Siggins, G.R. and Bloom, F.E. Drake, C.T. and Milner, T.A. (1999) Mu opioid receptors are in
(1985a) Selective inactivation of opioid receptors in rat hip- somatodendritic and axonal compartments of GABAergic
pocampus demonstrates that dynorphin-A and -B may act on neurons in rat hippocampal formation. Brain Res., 849:
mu-receptors in the CA1 region. Brain Res., 331: 366–370. 203–215.
Chavkin, C., James, I.F. and Goldstein, A. (1982) Dynorphin is Drake, C.T. and Milner, T.A. (2006) Mu opioid receptors are
a specific endogenous ligand of the kappa opioid receptor. extensively co-localized with parvalbumin, but not som-
Science, 215: 413–415. atostatin, in the dentate gyrus. Neurosci. Lett., 403: 176–180.
Chavkin, C., Shoemaker, W.J., McGinty, J.F., Bayon, A. and Drake, C.T., Patterson, T.A., Simmons, M.L., Chavkin, C. and
Bloom, F.E. (1985b) Characterization of the prodynorphin Milner, T.A. (1996) Kappa opioid receptor-like immunore-
and proenkephalin neuropeptide systems in rat hippocam- activity in guinea pig brain: ultrastructural localization in
pus. J. Neurosci., 5: 808–816. presynaptic terminals in hippocampal formation. J. Comp.
Chen, Y., Mestek, A., Liu, J., Hurley, J.A. and Yu, L. (1993) Neurol., 370: 377–395.
Molecular cloning and functional expression of a mu-opioid Drake, C.T., Terman, G.W., Simmons, M.L., Milner, T.A.,
receptor from rat brain. Mol. Pharmacol., 44: 8–12. Kunkel, D.D., Schwartzkroin, P.A. and Chavkin, C. (1994)
Clark, J.A., Liu, L., Price, M., Hersh, B., Edelson, M. and Dynorphin opioids present in dentate granule cells may func-
Pasternak, G.W. (1989) Kappa opiate receptor multiplicity: tion as retrograde inhibitory neurotransmitters. J. Neurosci.,
evidence for two U50,488-sensitive kappa 1 subtypes and a 14: 3736–3750.
novel kappa 3 subtype. J. Pharmacol. Exp. Ther., 251: Eisch, A.J., Barrot, M., Schad, C.A., Self, D.W. and Nestler,
461–468. E.J. (2000) Opiates inhibit neurogenesis in the adult rat hip-
Cohen, G.A., Doze, V.A. and Madison, D.V. (1992) Opioid pocampus. Proc. Natl. Acad. Sci. U.S.A., 97: 7579–7584.
inhibition of GABA release from presynaptic terminals of rat Eisch, A.J. and Harburg, G.C. (2006) Opiates, psychostimu-
hippocampal interneurons. Neuron, 9: 325–335. lants, and adult hippocampal neurogenesis: Insights for
Commons, K.G. and Milner, T.A. (1995) Ultrastructural het- addiction and stem cell biology. Hippocampus, 16: 271–286.
erogeneity of enkephalin-containing neurons in the rat hip- Evans, C.J., Keith, D.E., Morrison, H., Magendzo, K. and
pocampal formation. J. Comp. Neurol., 358: 324–342. Edwards, R.H. (1992) Cloning of a delta opioid receptor by
Commons, K.G. and Milner, T.A. (1996a) Cellular and sub- functional expression. Science, 258: 1952–1955.
cellular localization of delta opiate receptor immunoreactiv- Fredens, K., Stengaard Pedersen, K. and Larsson, L.I. (1984)
ity in the rat dentate gyrus. Brain Res., 738: 181–195. Localization of enkephalin and cholecystokinin immunore-
Commons, K.G. and Milner, T.A. (1996b) The ultrastructural activities in the perforant path terminal fields of the rat hip-
relationships between leu-enkephalin and GABA containing pocampal formation. Brain Res., 304: 255–263.
neurons differ within the hippocampal formation. Brain Res., French, E.D. and Siggins, G.R. (1980) An iontophoretic
724: 1–15. survey of opioid peptide actions in the rat limbic system: in
Corbett, A.D., Paterson, S.J. and Kosterlitz, H.W. (1993) Se- search of opiate epileptogenic mechanisms. Regul. Pept., 1:
lectivity of ligands for opioid receptors. In: Herz A. (Ed.), 127–146.
Opioids I. Springer-Verlag, Berlin, pp. 645–680. Freund, T.F. and Buzsáki, G. (1996) Interneurons of the hip-
Corbett, A.D., Paterson, S.J., McKnight, A.T., Magnan, J. and pocampus. Hippocampus, 6: 347–470.
Kosterlitz, H.W. (1982) Dynorphin and dynorphin are lig- Gall, C. (1988a) Seizures induce dramatic and distinctly differ-
ands for the kappa-subtype of opiate receptor. Nature, 299: ent changes in enkephalin, dynorphin, and CCK immunore-
79–81. activities in mouse hippocampal mossy fibers. J. Neurosci., 8:
Crain, B.J., Chang, K.J. and McNamara, J.O. (1986) Quanti- 1852–1862.
tative autoradiographic analysis of mu and delta opioid Gall, C.M. (1988b) Localization and seizure-induced alterations
binding sites in the rat hippocampal formation. J. Comp. of opioid peptides and CCK in the hippocampus. Natl. Inst.
Neurol., 246: 170–180. Drug Abuse Res. Monogr. Ser., 82: 12–32.
De Camilli, P. and Jahn, R. (1990) Pathways to regulated ex- Gall, C., Brecha, N., Karten, H.J. and Chang, K.-J. (1981)
ocytosis in neurons. Annu. Rev. Physiol., 52: 625–645. Localization of enkephalin-like immunoreactivity to identi-
Ding, Y.Q., Kaneko, T., Nomura, S. and Mizuno, N. (1996) fied axonal and neuronal populations of the rat hippocam-
Immunohistochemical localization of m-opioid receptors in pus. J. Comp. Neurol., 198: 335–350.
the central nervous system of the rat. J. Comp. Neurol., 367: George, S.R., Fan, T., Xie, Z., Tse, R., Tam, V., Varghese, G.
375–402. and O’Dowd, B.F. (2000) Oligomerization of mu- and
Drake, C.T., Chang, P.C., Harris, J.A. and Milner, T.A. (2002) delta-opioid receptors, generation of novel functional prop-
Neurons with mu opioid receptors interact indirectly with erties. J. Biol. Chem., 275: 26128–26135.
enkephalin-containing neurons in the rat dentate gyrus. Exp. Gomes, I., Jordan, B.A., Gupta, A., Trapaidze, N., Nagy, V.
Neurol., 176: 254–261. and Devi, L.A. (2000) Heterodimerization of m and d opioid
Drake, C.T., Chavkin, C. and Milner, T.A. (1997) Kappa op- receptors: a role in opiate synergy. J. Neurosci., 20:
ioid receptor-like immunoreactivity is present in substance NIL22–NIL26.
260

Goodman, R.R., Snyder, S.H., Kuhar, M.J. and Young III, Differential antagonism of opioid delta antinociception by
W.S. (1980) Differentiation of delta and mu opiate receptor [D-Ala2,Leu5,Cys6]enkephalin and naltrindole 50 -isothiocyan-
localizations by light microscopic autoradiography. Proc. ate: evidence for delta receptor subtypes. J. Pharmacol. Exp.
Natl. Acad. Sci. U.S.A., 77: 6239–6243. Ther., 257: 1069–1075.
Gould, E. and Gross, C.G. (2002) Neurogenesis in adult Johnston, H.M. and Morris, B.J. (1994) Nitric oxide alters
mammals: some progress and problems. J. Neurosci., 22: proenkephalin and prodynorphin gene expression in hippo-
619–623. campal granule cells. Neuroscience, 61: 435–439.
Gubler, U., Seeburg, P., Hoffman, B.J., Gage, L.P. and Uden- Kakidani, H., Furutani, Y., Takahashi, H., Noda, M., Mori-
friend, S. (1982) Molecular cloning establishes proenkephalin moto, Y., Hirose, T., Asai, M., Inayama, S., Nakanishi, S.
as precursor of enkephalin-containing peptides. Nature, 295: and Numa, S. (1982) Cloning and sequence analysis of
206–208. cDNA for porcine beta-neo-endorphin/dynorphin precursor.
Guerra, D., Sole, A., Cami, J. and Tobena, A. (1987) Neuro- Nature, 298: 245–249.
physiological performance in opiate addicts after rapid Kalyuzhny, A.E. and Wessendorf, M.W. (1997) CNS GABA
detoxification. Drug Alcohol Depend., 20: 261–270. neurons express the mu-opioid receptor: immunocytochem-
Gulya, K., Gehlert, D.R., Wamsley, J.K., Mosberg, H., Hruby, ical studies. Neuroreport, 20(8): 3367–3372.
V.J. and Yamamura, H.I. (1986) Light microscopic autora- Kaplan, T.J., Skyers, P.R., Tabori, N.E., Drake, C.T. and
diographic localization of delta opioid receptors in the rat Milner, T.A. (2004) Ultrastructural evidence for mu-opioid
brain using a highly selective bis-penicillamine cyclic enke- modulation of cholinergic pathways in rat dentate gyrus.
phalin analog. J. Pharmacol. Exp. Ther., 238: 720–726. Brain Res., 1019: 28–38.
Heinricher, M.M. (2003) Orphanin FQ/nociceptin: from neural Karhunen, T., Vilim, F.S., Alexeeva, V., Weiss, K.R. and
circuitry to behavior. Life Sci., 73: 813–822. Church, P.J. (2001) Targeting of peptidergic vesicles in
Herkenham, M. and McLean, S. (1988) The anatomical rela- cotransmitting terminals. J. Neurosci., 21: RC127.
tionship of opioid peptides and opiate receptors in the hip- Khachaturian, H., Schaefer, M.K.H. and Lewis, M.E. (1993)
pocampi of four rodent species. Natl. Inst. Drug Abuse Res. Anatomy and function of the endogenous opioid system.
Monogr. Ser., 82: 33–47. In: Herz A. (Ed.), Opioid I. Springer-Verlag, Berlin,
Herlitze, S., Garcia, D.E., Mackie, K., Hille, B., Scheuer, T. pp. 471–497.
and Catterall, W.A. (1996) Modulation of Ca2+ channels by Kieffer, B.L., Befort, K., Gaveriaux-Ruff, C. and Hirth, C.G.
G-protein beta gamma subunits. Nature, 380: 258–262. (1992) The delta opioid receptor: isolation of a cDNA by
Hong, J.S., McGinty, J.F., Grimes, L., Kanamatsu, T., Obie, J. expression cloning and pharmacological characterization.
and Mitchell, C.L. (1988) Seizure-induced alterations in the Proc. Natl. Acad. Sci. U.S.A., 89: 12048–12052.
metabolism of hippocampal opioid peptides suggest opioid Lee, P.H., Obie, J. and Hong, J.S. (1989) Opioids induce con-
modulation of seizure-related behaviors. Natl. Inst. Drug vulsions and wet dog shakes in rats: mediation by hippo-
Abuse Res. Monogr. Ser., 82: 48–66. campal mu, but not delta or kappa opioid receptors. J.
Hong, J.S., McGinty, J.F., Lee, P.H.K., Xie, C.W. and Mitc- Neurosci., 9: 692–697.
hell, C.L. (1993) Relationship between hippocampal opioid Liu, S., Wang, J., Zhu, D.Y., Fu, Y., Lukowiak, K. and Lu,
peptides and seizures. Prog. Neurobiol., 40: 507–528. Y.M. (2003) Generation of functional inhibitory neurons in
Hong, J.S., Wood, P.L., Gillin, J.C., Yang, H.Y. and Costa, E. the adult rat hippocampus. J. Neurosci., 23: 732–736.
(1980) Changes of hippocampal Met-enkephalin content af- Luscher, C., Jan, L.Y., Stoffel, M., Malenka, R.C. and Nicoll,
ter recurrent motor seizures. Nature, 285: 231–232. R.A. (1997) G protein-coupled inwardly rectifying K+ chan-
Houser, C.R., Miyashiro, J.E., Swartz, B.E., Walsh, G.O., nels (GIRKs) mediate postsynaptic but not presynaptic trans-
Rich, J.R. and Delgado-Escueta, A.V. (1990) Altered pat- mitter actions in hippocampal neurons. Neuron, 19: 687–695.
terns of dynorphin immunoreactivity suggest mossy fiber re- Maderspach, K., Nemeth, K., Simon, J., Benyhe, S., Szucs, M.
organization in human hippocampal epilepsy. J. Neurosci., and Wollemann, M. (1991) A monoclonal antibody recog-
10: 267–282. nizing kappa—but not mu—and delta-opioid receptors.
Ikeda, S.R. (1996) Voltage-dependent modulation of N-type J. Neurochem., 56: 1897–1904.
calcium channels by G-protein beta gamma subunits. Nature, Maderspach, K., Takacs, J., Niewiadomska, G. and Csillag, A.
380: 255–258. (1995) Postsynaptic and extrasynaptic localization of kappa-
Jackisch, R., Geppert, M., Brenner, A.S. and Illes, P. (1986) opioid receptor in selected brain areas of young rat and chick
Presynaptic opioid receptors modulating acetylcholine re- using an anti-receptor monoclonal antibody. J. Neurocytol.,
lease in the hippocampus of the rabbit. Naunyn Schmiedeb- 24: 478–486.
ergs Arch. Pharmacol., 332: 156–162. Madison, D.V. and Nicoll, R.A. (1988) Enkephalin hyperpo-
Jackisch, R., Werle, E. and Hertting, G. (1984) Identification of larizes interneurons in the rat hippocampus. J. Physiol., 398:
mechanisms involved in the modulation of release of nor- 123–130.
adrenaline in the hippocampus of the rabbit in vitro. Mansour, A., Burke, S., Pavlic, R.J., Akil, H. and Watson, S.J.
Neuropharmacology, 23: 1363–1371. (1996) Immunohistochemical localization of the cloned
Jiang, Q., Takemori, A.E., Sultana, M., Portoghese, P.S., kappa 1, receptor in the rat CNS and pituitary. Neuro-
Bowen, W.D., Mosberg, H.I. and Porreca, F. (1991) science, 71: 671–690.
261

Mansour, A., Fox, C.A., Burke, S., Akil, H. and Watson, S.J. Milner, T.A. and Drake, C.T. (2001) Ultrastructural evidence
(1995) Immunohistochemical localization of the cloned mu for presynaptic m opioid receptor modulation of synaptic
opioid receptor in the rat CNS. J. Chem. Neuroanat., 8: plasticity in NMDA-receptor-containing dendrites in the
283–305. dentate gyrus. Brain Res. Bull., 54: 131–140.
Mansour, A., Khachaturian, H., Lewis, M.E., Akil, H. and Minami, M., Hosoi, Y., Toya, T., Katao, Y., Maekawa, K.,
Watson, S.J. (1987) Autoradiographic differentiation of mu, Katsumata, S., Yabuuchi, K., Onogi, T. and Satoh, M.
delta, and kappa opioid receptors in the rat forebrain and (1993) In situ hybridization study of kappa-opioid receptor
midbrain. J. Neurosci., 7: 2445–2464. mRNA in the rat brain. Neurosci. Lett., 162: 161–164.
Mansour, A., Khachaturian, H., Lewis, M.E., Akil, H. and Mogil, J.S. and Pasternak, G.W. (2001) The molecular
Watson, S.J. (1988) Anatomy of CNS opioid receptors. and behavioral pharmacology of the orphanin FQ/nocicep-
Trends Neurosci., 11: 308–314. tin peptide and receptor family. Pharmacol. Rev., 53:
Mansour, A., Thompson, R.C., Akil, H. and Watson, S.J. 381–415.
(1993) Delta opioid receptor mRNA distribution in the brain: Moore, S.D., Madamba, S.G., Schweitzer, P. and Siggins, G.R.
comparison to delta receptor binding and proenkephalin (1994) Voltage-dependent effects of opioid peptides on hip-
mRNA. J. Chem. Neuroanat., 6: 351–362. pocampal CA3 pyramidal neurons in vitro. J. Neurosci., 14:
Martin, W.R. (1983) Pharmacology of opioids. Pharmacol. 809–820.
Rev., 35: 283–323. Morris, B.J. and Johnston, H.M. (1995) A role for hippocampal
Matsumoto, M., Yoshioka, M., Togashi, H., Hirokami, M., opioids in long-term functional plasticity. Trends Neurosci.,
Tochihara, M., Ikeda, T., Smith, C.B. and Saito, H. (1994) 18: 350–355.
mu-Opioid receptors modulate noradrenaline release from Nestler, E.J. (2001) Molecular basis of long-term plasticity un-
the rat hippocampus as measured by brain microdialysis. derlying addiction. Nat. Rev. Neurosci., 2: 119–128.
Brain Res., 636: 1–8. Neumaier, J.F., Mailheau, S. and Chavkin, C. (1988) Opioid
Mattia, A., Vanderah, T., Mosberg, H.I. and Porreca, F. (1991) receptor-mediated responses in the dentate gyrus and CA1
Lack of antinociceptive cross-tolerance between [D-Pen2, region of the rat hippocampus. J. Pharmacol. Exp. Ther.,
D-Pen5]enkephalin and [D-Ala2]deltorphin II in mice: evi- 244: 564–570.
dence for delta receptor subtypes. J. Pharmacol. Exp. Ther., Nock, B., Rajpara, A., O’Connor, L.H. and Cicero, T.J. (1988)
258: 583–587. Autoradiography of [3 H]U-69593 binding sites in rat brain:
McDaniel, K.L., Mundy, W.R. and Tilsom, H.A. (1990) evidence for kappa opioid receptor subtypes. Eur. J. Pharma-
Microinjection of dynorphin into the hippocampal forma- col., 154: 27–34.
tion impairs spatial learning in rats. Pharmacol. Biochem. Noda, M., Furutani, Y., Takahashi, H., Toyosato, M., Hirose,
Behav., 35: 429–435. T., Inayama, S., Nakanishi, S. and Numa, S. (1982) Cloning
McGinty, J.F., Henriksen, S.J., Goldstein, A., Terenius, L. and sequence analysis of cDNA for bovine adrenal prep-
and Bloom, F.E. (1983) Dynorphin is contained within roenkephalin. Nature, 295: 202–206.
hippocampal mossy fibers: immunochemical altera- Pan, Y.X., Xu, J., Bolan, E., Abbadie, C., Chang, A., Zucker-
tions after Kainic acid administration and colchicine- man, A., Rossi, G. and Pasternak, G.W. (1999) Identification
induced neurotoxicity. Proc. Natl. Acad. Sci. U.S.A., 80: and characterization of three new alternatively spliced mu-
589–593. opioid receptor isoforms. Mol. Pharmacol., 56: 396–403.
McGinty, J.F., Kanamatsu, T., Obie, J. and Hong, J.S. (1986) Pierce, J.P., Kurucz, O. and Milner, T.A. (1999) The morph-
Modulation of opioid peptide metabolism by seizures: differ- ometry of a peptidergic transmitter system before and after
entiation of opioid subclasses. Natl. Inst. Drug Abuse Res. seizure. I. Dynorphin B-like immunoreactivity in the hippo-
Monogr. Ser., 71: 89–101. campal mossy fiber system. Hippocampus, 9: 255–276.
McLean, S., Rothman, R.B., Jacobson, A.E., Rice, K.C. and Pierce, J.P. and Milner, T.A. (2001) Parallel increases in the
Herkenham, M. (1987) Distribution of opiate receptor sub- synaptic and surface areas of mossy fiber terminals following
types and enkephalin and dynorphin immunoreactivity in the seizure induction. Synapse, 39: 249–256.
hippocampus of squirrel, guinea pig, rat, and hamster. Pierce, T.L. and Wessendorf, M.W. (2000) Immunocytochem-
J. Comp. Neurol., 255: 497–510. ical mapping of endomorphin-2-immunoreactivity in rat
Merchenthaler, I., Maderdrut, J.L., Cianchetta, P., Shughrue, brain. J. Chem. Neuroanat., 18: 181–207.
P.J. and Bronstein, D. (1997) In situ hybridization histo- Piguet, P. and North, R.A. (1993) Opioid actions at mu and
chemical localization of prodynorphin messenger RNA in the delta receptors in the rat dentate gyrus in vitro. J. Pharmacol.
central nervous system of the rat. J. Comp. Neurol., 384: Exp. Ther., 266: 1139–1146.
211–232. Piva, F., Limonta, P., Dondi, D., Pimpinelli, F., Martini, L.
Meunier, J.-C., Mollereau, C., Toll, L., Suaudeau, C., Moisand, and Maggi, R. (1995) Effects of steroids on the brain opioid
C., Alvinerie, P., Butour, J.-L., Guillemot, J.-C., Ferrara, P., system. J. Steroid Biochem. Mol. Biol., 53: 343–348.
Monsarrat, B., Mazarguil, H., Vassart, G., Parmentier, M. Pu, L., Bao, G.B., Xu, N.J., Ma, L. and Pei, G. (2002)
and Costentin, J. (1995) Isolation and structure of the en- Hippocampal long-term potentiation is reduced by chronic
dogenous agonist of opioid receptor-like ORL1 receptor. opiate treatment and can be restored by re-exposure to opi-
Nature, 377: 532–536. ates. J. Neurosci., 22: 1914–1921.
262

Quinones-Jenab, V., Jenab, S., Ogawa, S., Inturrisi, C. and Shors, T.J., Townsend, D.A., Zhao, M., Kozorovitskiy, Y. and
Pfaff, D.W. (1997) Estrogen regulation of mu-opioid receptor Gould, E. (2002) Neurogenesis may relate to some but not all
mRNA in the forebrain of female rats. Brain Res., 47: 1. types of hippocampal-dependent learning. Hippocampus, 12:
Rees, H., Ang, L.C., Shul, D.D., George, D.H., Begley, H. and 578–584.
McConnell, T. (1994) Increase in enkephalin-like immuno- Shuster, S.J., Riedl, M., Li, X.R., Vulchanova, L. and Elde, R.
reactivity in hippocampi of adults with generalized epilepsy. (1999) Stimulus-dependent translocation of kappa opioid re-
Brain Res., 652: 113–119. ceptors to the plasma membrane. J. Neurosci., 19: 2658–2664.
Reinscheid, R.K., Nothacker, H.P., Bourson, A., Ardati, A., Simmons, M.L. and Chavkin, C. (1996) Endogenous opioid
Henningsen, R.A., Bunzow, J.R., Grandy, D.K., Langen, H., regulation of hippocampal function. Int. Rev. Neurobiol., 39:
Monsma Jr., F.J. and Civelli, O. (1995) Orphanin FQ: a ne- 145–196.
uropeptide that activates an opioid-like G protein-coupled Simmons, M.L., Terman, G.W., Drake, C.T. and Chavkin, C.
receptor. Science, 270: 792–794. (1994) Inhibition of glutamate release by presynaptic kappa
Rios, C.D., Jordan, B.A., Gomes, I. and Devi, L.A. (2001) 1—opioid receptors in the guinea pig dentate gyrus. J. Ne-
G-protein-coupled receptor dimerization: modulation of re- urophysiol., 72: 1697–1705.
ceptor function. Pharmacol. Ther., 92: 71–87. Simmons, M.L., Terman, G.W., Gibbs, S.M. and Chavkin, C.
Robbins, T.W. and Everitt, B.J. (2002) Limbic-striatal memory (1995) L-type calcium channels mediate dynorphin neuro-
systems and drug addiction. Neurobiol. Learn. Mem., 78: peptide release from dendrites but not axons of hippocampal
625–636. granule cells. Neuron, 14: 1265–1272.
Rocha, L., Ackermann, R.F. and Engel Jr., J. (1996) Effects of Simonato, M. and Romualdi, P. (1996) Dynorphin and epi-
chronic morphine pretreatment on amygdaloid kindling de- lepsy. Prog. Neurobiol., 50: 557–583.
velopment, postictal seizure and suppression and benzodiaze- Skyers, P.S., Einheber, S., Pierce, J.P. and Milner, T.A. (2003)
pine receptor binding in rats. Epilepsy Res., 23: 225–233. Increased mu-opioid receptor labeling is found on inner mo-
Roman, E., Ploj, K., Gustafsson, L., Meyerson, B.J. and lecular layer terminals of the dentate gyrus following seizures.
Nylander, I. (2006) Variations in opioid peptide levels during Exp. Neurol., 179: 200–209.
the estrous cycle in Sprague-Dawley rats. Neuropeptides, 40: Slamberová, R., Rimanoczy, A., Bar, N., Schindler, C.J. and
195–206. Vathy, I. (2003) Density of mu-opioid receptors in the hip-
Rothman, R.B., Bowen, W.D., Schumacher, U.K. and Pert, pocampus of adult male and female rats is altered by prenatal
C.B. (1983) Effect of beta-FNA on opiate receptor binding: morphine exposure and gonadal hormone treatment. Hippo-
preliminary evidence for two types of mu receptors. Eur. J. campus, 13: 461–471.
Pharmacol., 95: 147–148. Slowe, S.J., Simonin, F., Kieffer, B. and Kitchen, I. (1999)
Rothman, R.B., Jacobson, A.E., Rice, K.C. and Herkenham, Quantitative autoradiography of mu-, delta- and kappa 1
M. (1987) Autoradiographic evidence for two classes of mu opioid receptors in kappa-opioid receptor knockout mice.
opioid binding sites in rat brain using [125I]FK33824. Pep- Brain Res., 818: 335–345.
tides, 8: 1015–1021. Stumm, R.K., Zhou, C., Schulz, S. and Hollt, V. (2004) Neuronal
Sandin, J., Nylander, I., Georgieva, J., Schött, P.A., Ögren, types expressing mu- and delta-opioid receptor mRNA in the
S.O. and Terenius, L. (1998) Hippocampal dynorphin B in- rat hippocampal formation. J. Comp. Neurol., 469: 107–118.
jections impair spatial learning in rats: a kappa-opioid re- Svingos, A.L., Colago, E.E.O. and Pickel, V.M. (1999) Cellular
ceptor-mediated effect. Neuroscience, 85: 375–382. sites for dynorphin activation of kappa-opioid receptors in
Schindler, C.J., Slamberova, R., Rimanoczy, A., Hnactzuk, the rat nucleus accumbens shell. J. Neurosci., 19: 1804–1813.
O.C., Riley, M.A. and Vathy, I. (2004) Field-specific changes Tanapat, P., Hastings, N.B., Reeves, A.J. and Gould, E. (1999)
in hippocampal opioid mRNA, peptides, and receptors due Estrogen stimulates a transient increase in the number of new
to prenatal morphine exposure in adult male rats. Neurosci- neurons in the dentate gyrus of the adult female rat. J.
ence, 126: 355–364. Neurosci., 19: 5792–5801.
Scholz, K.P. and Miller, R.J. (1992) Inhibition of quantal Tempel, A. and Zukin, R.S. (1987) Neuroanatomical patterns
transmitter release in the absence of calcium influx by a G of the mu, delta, and kappa opioid receptors of rat brain as
protein-linked adenosine receptor at hippocampal synapses. determined by quantitative in vitro autoradiography. Proc.
Neuron, 8: 1139–1150. Natl. Acad. Sci. U.S.A., 84: 4308–4312.
Seri, B., Garcia-Verdugo, J.M., McEwen, B.S. and Alvarez- Terman, G.W., Drake, C.T., Simmons, M.L., Milner, T.A. and
Buylla, A. (2001) Astrocytes give rise to new neurons in the Chavkin, C. (2000) Opioid modulation of recurrent excitation
adult mammalian hippocampus. J. Neurosci., 21: 7153–7160. in the hippocampal dentate gyrus. J. Neurosci., 20: 4379–4388.
Seward, E.P., Chernevskaya, N.I. and Nowycky, M.C. (1995) Terman, G.W., Wagner, J.J. and Chavkin, C. (1994) Kappa
Exocytosis in peptidergic nerve terminals exhibits two cal- opioids inhibit induction of long-term potentiation in the
cium-sensitive phases during pulsatile calcium entry. J. Ne- dentate gyrus of the guinea pig hippocampus. J. Neurosci.,
urosci., 15: 3390–3399. 14: 4740–4747.
Shors, T.J., Miesegaes, G., Beylin, A., Zhao, M., Rydel, T. and Thureson-Klein, A.K. and Klein, R.L. (1990) Exocytosis from
Gould, E. (2001) Neurogenesis in the adult is involved in the neuronal large dense-core vesicles. Int. Rev. Cytol., 121:
formation of trace memories. Nature, 410: 372–376. 67–126.
263

Tielen, A.M., van Leeuwen, F.W. and Lopes da Silva, F.H. Weisskopf, M.G., Zalutsky, R.A. and Nicoll, R.A. (1993) The
(1982) The localization of leucine-enkephalin immunoreac- opioid peptide dynorphin mediates heterosynaptic depression
tivity within the guinea pig hippocampus. Exp. Brain Res., of hippocampal mossy fibre synapses and modulates long-
48: 288–295. term potentiation. Nature, 362: 423–427.
Torres-Reveron, A., Khalid, S., Drake, C.T. and Milner, T.A. White, J.D., Gall, C.M. and McKelvy, J.F. (1986) Pro-
(2006a) Enkephalin and dynorphin immunoreactivities in- enkephalin is processed in a projection-specific manner in
crease in the dorsal hippocampus in response to estrogens the rat central nervous system. Proc. Natl. Acad. Sci. U.S.A.,
and colocalize with estrogen receptor beta in mossy fiber 83: 7099–7103.
terminals. Program No. 724.6 2006. Neuroscience Meeting Wimpey, T.L. and Chavkin, C. (1991) Opioids activate both an
Planner Atlanta GA: Society for Neuroscience, 2006. Online. inward rectifier and a novel voltage-gated potassium con-
Torres-Reveron, A., Khalid, S., Drake, C.T. and Milner, T.A. ductance in the hippocampal formation. Neuron, 6: 281–289.
(2006b) Enkephalin-immunoreactivity colocalizes with estro- Wolozin, B.L. and Pasternak, G.W. (1981) Classification of
gen receptor immunoreactivity in terminals and increases in multiple morphine and enkephalin binding sites in the central
proestrus and estrogen replaced female rats in the dorsal nervous system. Proc. Natl. Acad. Sci. U.S.A., 78:
hippocampus. Program No. Tn-26 2006. Poster Session Ab- 6181–6185.
stracts St. Paul Minnesota. International Narcotics Research Woolley, C.S. and Schwartzkroin, P.A. (1998) Hormonal
Conference. effects on the brain. Epilepsia, 39(Suppl 8): S2–S8.
Tortella, F.C., Echevarria, E., Robles, L., Mosberg, H.I. and Xie, C.W., Morrisett, R.A. and Lewis, D.V. (1992) Mu opioid
Holaday, J.W. (1988) Anticonvulsant effects of mu (DAGO) receptor-mediated modulation of synaptic currents in dentate
and delta (DPDPE) enkephalins in rats. Peptides, 9: granule cells of rat hippocampus. J. Neurophysiol., 68:
1177–1181. 1113–1120.
Turlejski, K. and Djavadian, R. (2002) Life-long stability of Yasuda, K., Raynor, H., Kong, C.D., Breder, J., Takeda, J.,
neurons: a century of research on neurogenesis, neuronal Reisine, T. and Bell, G.I. (1993) Cloning and functional
death and neuron quantification in adult CNS. Prog. Brain comparison of kappa and delta opioid receptors from mouse
Res., 136: 39–65. brain. Proc. Natl. Acad. Sci. U.S.A., 90: 6736–6740.
Unterwald, E.M., Knapp, C. and Zukin, R.S. (1991) Neuro- Yoshioka, M., Matsumoto, M., Togashi, H., Smith, C.B. and
anatomical localization of kappa 1 and kappa 2 opioid Saito, H. (1993) Opioid receptor regulation of 5-hydroxytry-
receptors in rat and guinea pig brain. Brain Res., 562: ptamine release from the rat hippocampus measured by in
57–65. vivo microdialysis. Brain Res., 613: 74–79.
Velı́sek, L., Stanton, P.K., Moshé, S.L. and Vathy, I. (2000) Yuste, R. and Tank, D.W. (1996) Dendritic integration in
Prenatal morphine exposure enhances seizure susceptibility mammalian neurons, a century after Cajal. Neuron, 16:
but suppresses long-term potentiation in the limbic system of 701–716.
adult male rats. Brain Res., 869: 186–193. Zadina, J.E., Hackler, L., Ge, L.J. and Kastin, A.J. (1997) A
Verhage, M., McMahon, H.T., Ghijsen, W.E.J.M., Boomsma, potent and selective endogenous agonist for the m-opiate re-
F., Scholten, G., Wiegant, V.M. and Nicholls, D.G. (1991) ceptor. Nature, 386: 499–502.
Differential release of amino acids, neuropeptides, and cat- Zakarian, S. and Smyth, D. (1979) Distribution of active and
echolamines from isolated nerve terminals. Neuron, 6: inactive forms of endorphins in rat pituitary and brain. Proc.
517–524. Natl. Acad. Sci. U.S.A., 76: 5972–5976.
Vindrola, O., Briones, R., Asai, M. and Fernandez Guardiola, Zakarian, S. and Smyth, D.G. (1982) Distribution of beta-
A. (1981) Amygdaloid kindling enhances the enkephalin endorphin-related peptides in rat pituitary and brain. Bio-
content in rat brain. Neurosci. Lett., 21: 39–44. chem. J., 202: 561–571.
Wagner, J.J., Caudle, R.M. and Chavkin, C. (1992) Kappa- Zieglgänsberger, W., French, E.D., Siggins, G.R. and Bloom,
opioids decrease excitatory transmission in the dentate gyrus F.E. (1979) Opioid peptides may excite hippocampal pyram-
of the guinea pig hippocampus. J. Neurosci., 12: 132–141. idal neurons by inhibiting adjacent inhibitory interneurons.
Wagner, J.J., Caudle, R.M., Neumaier, J.F. and Chavkin, C. Science, 205: 415–417.
(1990) Stimulation of endogenous opioid release displaces mu Zimprich, A., Simon, T. and Hollt, V. (1995) Cloning and ex-
receptor binding in rat hippocampus. Neuroscience, 37: 45–53. pression of an isoform of the rat mu opioid receptor
Wagner, J.J. and Chavkin, C. (1992) Endogenous opioids and (rMOR1B) which differs in agonist induced desensitization
LTP: another view. Neurochem. Int., 20: 457–460. from rMOR1. FEBS Lett., 359: 142–146.
Wagner, J.J., Evans, C.J. and Chavkin, C. (1991) Focal stim- Zukin, R.S., Eghbali, M., Olive, D., Unterwald, E.M. and
ulation of the mossy fibers releases endogenous dynorphins Tempel, A. (1988) Characterization and visualization of rat
that bind kappa 1-opioid receptors in guinea pig hippocam- and guinea pig brain kappa opioid receptors: evidence for
pus. J. Neurochem., 57: 333–343. kappa 1 and kappa 2 opioid receptors. Proc. Natl. Acad. Sci.
Wagner, J.J., Terman, G.W. and Chavkin, C. (1993) Endog- U.S.A., 85: 4061–4065. RU:1 do you mean KORs too? above
enous dynorphins inhibit excitatory neurotransmission and you mention KORs do opposite things( and see below com-
block LTP induction in the hippocampus. Nature, 363: ment 8. RU:2 could you provide a reference - to a review -
451–454. because yours is the only opioids chapter in the book.
SO CA1
p
SP
SR
SUB

SLM
TAT
PP

p HIL m
SLu g
CA3
DG

GCL
ML

ENK

DYN

Plate 15.1. Location of enkephalins and dynorphins in the hippocampal formation. Opioid peptides are found in all regions of the
hippocampal formation: the hippocampus proper (fields CA1 and CA3), the dentate gyrus (DG), and the subiculum (SUB). The
glutamatergic neurons [pyramidal cells (p), granule cells (g), mossy cells (m)] and their projections are shown. GABAergic interneurons
are scattered in all laminae. The perforant path (PP) from entorhinal cortex provides the major excitatory innervation. Subcortical
afferents and contralateral hippocampal projections enter via the fimbria/fornix (not shown) adjacent to CA3. Enkephalins (green) are
in some granule cells, particularly their mossy fiber terminals in hilus and stratum lucidum (SLu). Mossy fibers extensively innervate
hilar mossy cells and CA3 pyramidal cells. Enkephalins are in a portion of the projection from entorhinal cortex; the lateral portion of
the perforant path (PP) in the DG and the temporal-ammonic tract (TAT) in CA1. Enkephalins are in a few scattered interneurons,
which are relatively abundant on the border of stratum radiatum (SR) and stratum lacunosum-moleculare (SLM). Dynorphins (yellow)
are in the granule cells. Dynorphin peptides are most striking in the mossy fiber terminals, and are also in the mossy fiber axons,
granule cell somata, and granule cell dendrites. Abbreviations: SO, stratum oriens; SP, stratum pyramidale; GCL, granule cell layer.
(For B/W version, see page 248 in the volume.)
+
Molecular ENK
Layer PP
_ GABA

?
?

+
SubP ACh

Granule Granule ?
Cell Cell
Layer

GABA
Hilus PARV

+
ENK/DYN
+

ACh GABA
NPY
SOM

MOR

to CA3 DOR

KOR
+

Plate 15.2. Subcellular locations of opioid peptides and receptors in the dentate gyrus. Enkephalins (green shading) are in: (1)
somatodendritic, axon, and terminal portions of granule cells; (2) lateral perforant path axons and terminals in the outer molecular
layer; and (3) occasional GABAergic interneurons (not shown). Mu-opioid receptors (MORs; blue) are most frequently in all portions
of parvalbumin (PARV)-containing basket cells that innervate granule cell somata and proximal dendrites. MORs are also in
cholinergic and GABAergic afferents from the lateral septum/diagonal band. Delta-opioid receptors (DORs; pink) are in many hilar
interneurons that contain somatostatin or neuropeptide Y (SOM/NPY) and inhibit the distal dendrites of granule cells. PARV-
containing neurons express DOR mRNA. DORs, and to a lesser extent MORs, are occasionally present in granule cell dendrites.
Kappa-opioid receptors (KORs) in guinea pigs are in substance P containing afferents to granule cells, in granule cell mossy fibers, and
most likely on perforant path terminals in the outer molecular layer. (For B/W version, see page 249 in the volume.)
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 16

Somatostatin in the dentate gyrus

Melanie K. Tallent

Department of Pharmacology and Physiology, Drexel University College of Medicine, 245 N. 15 St., Philadelphia,
PA 19102, USA

Abstract: The neuropeptide somatostatin (SST) is expressed in a discrete population of interneurons in the
dentate gyrus. These interneurons have their soma in the hilus and project to the outer molecular layer onto
dendrites of dentate granule cells, adjacent to perforant path input. SST-containing interneurons are very
sensitive to excitotoxicty, and thus are vulnerable to a variety of neurological diseases and insults, including
epilepsy, Alzheimer’s disease, traumatic brain injury, and ischemia. The SST gene contains a prototypical
cyclic AMP response element (CRE) site. Such a regulatory site confers activity-dependence to the gene,
such that it is turned on when neuronal activity is high. Thus SST expression is increased by pathological
conditions such as seizures and by natural stimulation such as environmental enrichment. SST may play
an important role in cognition by modulating the response of neurons to synaptic input. In the dentate,
SST and the related peptide cortistatin (CST) reduce the likelihood of generating long-term potentiation,
a cellular process involved in learning and memory. Thus these neuropeptides would increase the threshold
of input required for acquisition of new memories, increasing ‘‘signal to noise’’ to filter out irrelevant
environmental cues. The major mechanism through which SST inhibits LTP is likely through inhibition
of voltage-gated Ca2+ channels on dentate granule cell dendrites. Transgenic overexpression of CST in the
dentate leads to profound deficits in spatial learning and memory, validating its role in cognitive process-
ing. A reduction of synaptic potentiation by SST and CST in dentate may also contribute to the
well-characterized antiepileptic properties of these neuropeptides. Thus SST and CST are important
neuromodulators in the dentate gyrus, and disruption of this signaling system may have major impact on
hippocampal function.

Keywords: somatostatin; cortistatin; neuropeptide; electrophysiology; long-term potentiation; interneuron;


cognition; epilepsy

Introduction neurons or HIPP cells), however, its physiological


role here is relatively unknown. The selective tar-
Somatostatin (SST) is a neuropeptide abundantly geting of SST-containing interneuron terminals to
expressed throughout the brain and periphery. In distal dendrites of principal neurons in both dent-
the dentate gyrus, SST has been extensively used as ate and cornu ammonis suggest that this peptide
a chemical marker of a subset of interneurons lo- may play a similar role throughout the hippocam-
cated in the hilus (hilar perforant path associated pus. Interest in the role of SST in the dentate has
been heightened by studies showing it is upregu-
Corresponding author. Tel.: +1 215 762 8208; lated by mild seizures, but SST-containing inter-
Fax: +1 215 762 2299; E-mail: tallent@drexel.edu neurons are vulnerable to severe seizures. What

DOI: 10.1016/S0079-6123(07)63016-7 265


266

role, if any, the loss of the SSTergic system plays in Gai3, or Gao1 (Straiker et al., 2002). Since Ga
postseizure excitability in the dentate is still under subunits show preference in coupling to distinct bg
investigation in our lab and others. Recent evi- subunits (Robishaw and Berlot, 2004), specificity
dence suggests that SST may be a critical regulator in Ga coupling likely confers specificity in bg
of synaptic plasticity that underlies learning and subunits that that interact with SNAP-25.
memory in the dentate (Baratta et al., 2002). SST and CST mediate their actions via the same
SST was first discovered in 1972 by Vale and col- family of receptors. Five SST receptors have been
leagues as a hypothalamic inhibitor of growth hor- cloned, SST1–SST5. These are prototypical Gi/Go
mone release from the anterior pituitary (Brazeau G-protein coupled receptors most closely related
et al., 1972). In the periphery, SST generally acts as to the opioid family of receptors. Only SST2 is
an inhibitor of hormone release from the gut and alternatively spliced, with two described variants,
pancreas (Reisine, 1995). In the brain, SST also SST2a and SST2b (Schindler et al., 1999). Expres-
generally has inhibitory actions at the cellular sion of SST1–SST4 in brain has been confirmed;
level, but can increase neurotransmitter release in SST5 expression in brain is still somewhat contro-
some instances, likely through disinhibitory mech- versial (Stroh et al., 1999; Schulz et al., 2000).
anisms [i.e., inhibition of inhibitory interneurons, SST2, SST3, and SST4 are all expressed in hippo-
Chesselet and Reisine (1983)]. SST distribution in campus (Dournaud et al., 1996a; Handel et al.,
the brain is extensive, with virtually every region 1999; Schreff et al., 2000; Schulz et al., 2000); SST1
showing some localization at some point in devel- expression in forebrain has not been validated
opment. SST expression in forebrain is prominent, using antibodies that have had their specificity
containing perhaps the highest brain levels outside confirmed in SST1 receptor knockout mice
the hypothalamus. More recently, a related (Hervieu and Emson, 1998; Schulz et al., 2000).
peptide from a distinct gene was discovered by Interesting distinct subcellular localization of the
de Lecea and colleagues, cortistatin (CST). The different SST receptor subtypes has been reported
distribution of CST is more regionally limited on the principal neurons of hippocampus. SST2
than SST, being largely restricted to cortex and expression is largely restricted to soma and prox-
hippocampus (de Lecea et al., 1996). imal dendrites (Dournaud et al., 1996a), SST4 to
SST activates at least three types of K+ chan- medial and distal dendrites (Schreff et al., 2000),
nels, KIR, KM, and KLK (Moore et al., 1988; and SST3 has a pattern of expression unique from
Velimirovic et al., 1995; Schweitzer et al., 1998). any other G protein-coupled receptor in that it is
These actions of SST hyperpolarize neurons, exclusively expressed on neuronal cilia (Handel
moving them away from their threshold for firing. et al., 1999). Thus, although these three receptors
Activation of K+ channels also decreases input are expressed in the same neurons, there is little
resistance of neurons, lessening the depolarizing spatial overlap (Schreff et al., 2000). Excepting
actions of excitatory synaptic inputs. SST inhibits axons (which may contain SST1, Hervieu and
voltage-sensitive Ca2+ channels in neurons as well Emson, 1998), the principal projecting neurons in
(Wang et al., 1990b; Boehm and Betz, 1997). This hippocampus therefore are virtually coated with
action could also limit excitability, either through SST receptors, suggesting they are exquisitely
postsynaptic decrease of Ca2+-induced depolari- sensitive to this neuropeptide, and that SST is
zations and/or signaling, or through presynaptic an extremely important signaling molecule in this
inhibition of neurotransmitter release. However, region.
increasing evidence suggests that the major mech-
anism for inhibition of neurotransmitter release by
Gi/Go coupled receptors is by direct inhibition of SST receptors activate multiple signaling systems
the synaptic protein SNAP-25 by bg subunits
(Blackmer et al., 2005; Gerachshenko et al., 2005). Studies in expression systems have shown SST to
SST inhibition of glutamate release in cultured couple to multiple signaling systems, including
hippocampal neurons was shown to require Gai2, adenylyl cyclase, protein kinase C, and MAP
267

kinases. Signaling of these receptors in brain CREB-regulated immediate early gene c-fos,
has been less studied, besides their regulation of and MAP kinase activity, an activator of CREB
voltage-dependent ion channels. Most of the stud- (Yoshitomi et al., 1997). This suggests SST recep-
ies examining signaling pathways activated by tor activation can inhibit CREB independently of
SST receptors have used recombinant receptors in adenylyl cyclase, although this was not directly
expression systems. Such studies have shown that assessed in this study. Interestingly, although
SST receptors can activate or inhibit a multitude SST inhibition of c-fos was blocked by pertussis
of effectors, most of which are typical for the toxin (PTX), activation of c-fos expression was
Gi/Go family of G proteins (Neves et al., 2002). PTX-insensitive, suggesting a non-Gi/Go mecha-
For example, activation of all cloned SST recep- nism (Yoshitomi et al., 1997). In the pituitary
tors leads to the inhibition of adenylyl cyclase adenoma cell line GH3, SST inhibited c-fos
(Moller et al., 2003). Activation of tyrosine phos- mRNA expression through a PTX-sensitive mech-
phates and phospholipase C (PLC) has also been anism that appeared to involve inhibition of ERK
reported for all of the cloned receptors (Lahlou activity and Elk-1 phosphorylation (Todisco et al.,
et al., 2004), although one study in which all 1994). This same group also found SST inhibited
five human receptors were expressed in the same binding of the AP-1 transcription complex to its
cell line showed only SST3 and SST5 significantly target DNA, which is necessary for c-fos activity.
activated PLC and increased intracellular Ca2+ This action of SST was blocked by PTX and the
(Siehler and Hoyer, 1999). Activation of MAP phosphatase inhibitors sodium orthovanadate
kinases has been observed for the majority of the and okadaic acid (Todisco et al., 1994, 1995).
receptors (Florio et al., 1999), specifically ERK In the pancreatic cell line AR42J and in human
(Lahlou et al., 2003), and p38 (Moller et al., 2003), leukemia cell lines, SST also inhibited c-fos
although inhibition of ERK has also been reported expression in a PTX-sensitive fashion (Ishihara
(Todisco et al., 1994; Pola et al., 2003; Lahlou et al., 1999; Cowles et al., 2002). Thus most studies
et al., 2004). SST has also been shown to activate suggest that SST inhibits cAMP-mediated gene
PI3 kinase (Lahlou et al., 2003). expression, which is in keeping with its antiprolif-
Several studies have addressed signaling mech- erative actions and its Gi/Go coupling. However,
anisms of endogenous SST receptors in cell lines robust activation of c-fos expression by SST3 has
and in situ; interest has been raised because of the been reported (Yoshitomi et al., 1997). This action
finding that SST has antiproliferative actions on may be mediated by activation of MAP kinases
cancer cells. Since activation of SST receptors in- via the small GTPase Ras (Florio et al., 1999;
hibits adenylyl cyclase activity, inhibition of Lahlou et al., 2003), which is a prototypical Gi-
cAMP-dependent signaling pathways would be a mediated signaling pathway (Neves et al., 2002).
likely mechanism for the antiproliferative effects of SST has also been found to stimulate adenylyl
SST. One such molecule is the cAMP response el- cyclase in somatotrophic cells (Ramirez et al.,
ement binding protein (CREB), a transcription 2002), suggesting atypical G-protein coupling can
factor that regulates protein expression by binding occur in some cell types.
to CRE sites in their promoter region. Indeed, the It is interesting that SST receptors inhibit aden-
SST gene itself has such a CRE site (as does SST2, ylyl cyclase, leading to a decrease in cAMP levels
Kimura et al., 2001), suggesting a mechanism of and decreased activity of PKA, while these same
reciprocal regulation (Montminy and Bilezikjian, receptors can activate the MAP kinase ERK.
1987). SST was shown to inhibit CREB phos- Thus, although most studies suggest that SST in-
phorylation in GH4 somatotrophic cells via inhi- hibits CREB-mediated transcription, activation
bition of PKA (Tentler et al., 1997). SST has been has also been reported. These contrasting actions
shown to regulate other CREB-regulated signaling may be receptor subtype and tissue specific,
molecules in several cell lines. In the mouse insu- although this issue has not been adequately
linoma cell line MIN6, that expresses only SST3, addressed in situ or in vivo. This complexity in
SST transiently increases and then decreases the signaling is of course not unique to SST receptors,
268

and simply because SST receptors have been correlation was found between levels of SST in
shown that they can couple to a pathway in an cerebrospinal fluid (CSF) and degree of dementia
artificial system, it does not necessarily follow that in patients with Alzheimer’s disease (Edvinsson et
they do couple to these pathways to mediate their al., 1993; Minthon et al., 1997). Thus a reduction
important physiological effects. in the CNS SST system is consistently observed in
Surprisingly few studies have addressed modu- brains from Alzheimer’s patients; however, it is
lation of signaling pathways by SST in the brain, currently unknown how or whether this loss con-
aside from its well-known actions on K+ and tributes to Alzheimer’s dementia, as the function
Ca2+ channels (Mihara et al., 1987; Inoue et al., of SST in cognition is still unclear. Interestingly,
1988; Ikeda and Schofield, 1989; Wang et al., SST receptors appear to be largely preserved in
1990a, b; Meriney et al., 1994; Connor et al., 1997; Alzheimer’s disease (Kumar, 2005), suggesting
Sodickson and Bean, 1998). Inhibition of cAMP them as potential targets for therapeutics. In
by SST in membranes from several brain regions mouse models of Alzheimer’s, decreases in brain
has been observed, including cortex and hippo- SST levels have also been observed (Tomidokoro
campus (Raynor and Reisine, 1992; Blake, 2001). et al., 2000; Ramos et al., 2006), suggesting they
Inhibition of c-fos mRNA expression has been could be appropriate models to study the role of
reported in the trigeminal nucleus and arcuate SST in this disease.
nucleus (Bereiter, 1997; Dickson et al., 1997; A novel role for SST with direct implications
Zheng et al., 1997), however, it is unclear whether on Alzheimer’s disease has recently been re-
this action is through activation of signaling path- ported. SST was shown to increase degradation
ways or through inhibition of neuronal firing, a of b-amyloid by upregulating the activity of ne-
widespread action of SST due to its augmentation prilysin, an Ab degrading peptidase (Saito et al.,
of K+ currents. 2005). Interestingly, in SST knockout mice,
immunoreactive neprilysin was reduced by 78%
in the molecular layer of the dentate gyrus. A
Decrease in SST levels are found in Alzheimer’s corresponding 1.5-fold increase in Ab 42, the b
brain amyloid fragment most closely associated with
Alzheimer’s disease and amyloid plaques, was
Davies et al. (1980) first reported reduced SST found in brains of SST knockout mice (Saito
levels in postmortem brains of Alzheimer’s disease et al., 2005). Like SST, neprilysin decreases with
patients. Since then there have been dozens of aging and in Alzheimer’s disease (Yasojima et al.,
studies confirming and expanding on this initial 2001; Iwata et al., 2002). An Alzheimer’s trans-
observation. SST-containing interneurons were genic mouse model (APP23) crossed with a ne-
shown to be selectively associated with neuronal prilysin knockout mouse showed increased
tangles in Alzheimer’s brain, suggesting a mecha- oligomeric forms of Ab at synapses, and cognitive
nism by which these neurons are vulnerable impairment, including deficits in spatial learning
(Roberts et al., 1985). However, studies in rats and object recognition memory at young ages
suggest that lesioning cholinergic neurons also de- prior to accumulation of amyloid plaques (Huang
creases SST neurons in cortex, suggesting the SST et al., 2006). Thus one functional consequence
reduction is downstream of the well-characterized of reduced SST levels in brain may be increased
cholinergic deficits (Zhang et al., 1998). Other accumulation of Ab, and upregulation of SST
studies have examined regional changes in SST could be protective.
reductions, showing a loss of SST-containing neu- Changes in CNS SST levels have been reported
rons in frontal and parahippocampal cortex and in other diseases as well, in particular those asso-
hippocampus (Grouselle et al., 1998). A reduction ciated with cognitive dysfunction. Whether these
in SST mRNA levels per surviving neuron was changes influence the disease state is unknown.
also observed in hippocampus but not cortex Reduced CSF SST levels were found in patients
(Dournaud et al., 1994). A strong negative with depression (Agren and Lundqvist, 1984) or
269

multiple sclerosis (Vecsei et al., 1990), although Approximately 30% of SST interneurons co-
regional changes in brain have not been explored express neuropeptide Y. This is a common theme
in these diseases. In bipolar disorder, transcription throughout the brain, and co-localization can be
analysis using microarrays found that SST was one as common among SST and neuropeptide Y cells
of eight genes that were significantly correlated as 100% (see Chronwall et al., 1984; McDonald,
with the severity of the disease in postmortem 1989; Figueredo-Cardenas et al., 1996). The func-
tissue. In this study, SST was upregulated, and tional implication of this co-expression has
was the only gene to show significant association not been examined in the dentate or elsewhere,
with bipolarism when genetic polymorphisms were but since both of these neuropeptides have gener-
analyzed (Nakatani et al., 2006). Increases in CSF ally inhibitory actions, they are likely to act
SST was also found in the related disorder, mania synergistically when co-released.
(Sharma et al., 1995). Elevated SST levels are also In rat and primate, including humans, immuno-
detected in basal ganglia of patients with Hunt- staining reveals a dense plexus of SST immunore-
ington’s disease (Aronin et al., 1983; Nemeroff activity in the outer molecular layer, in agreement
et al., 1983). In Parkinson’s disease, a reduction of with the projection pattern of HIPP cells (Amaral
SST in hippocampus and cortex was found only in et al., 1988; Milner and Bacon, 1989; Leranth
patients showing cognitive deficits, suggesting that et al., 1990; Austin and Buckmaster, 2004;
changes occur in the later stages of this disease Csaba et al., 2005). Similar findings have been
(Epelbaum et al., 1983). Reduction in cortical SST reported in other species, including pig (Holm
is also found in schizophrenia (Gabriel et al., et al., 1992), guinea pig, and rabbit (Buckmaster
1996); interestingly, in rat chronic haloperidol et al., 1994). In mice, however, this SST staining
treatment increases cortical SST levels (Sakai profile is absent (Buckmaster et al., 1994; Jinno
et al., 1995). and Kosaka, 2000). This does not appear to be due
to a difference in the projection pattern of HIPP
cells, since enhanced green fluorescent protein
SST is expressed in a subset of hilar interneurons in (EGFP) staining is present in the outer molecular
dentate gyrus layer in mice engineered to express this marker in
SST-containing neurons (see below Oliva et al.,
In rat dentate gyrus, 16% of GAD-containing in- 2000). Whether this difference in SST density in
terneurons express SST (Kosaka et al., 1988). This the outer molecular layer reflects a functional
neuropeptide has a distinct pattern of expression; difference of SST between rats and mice is
it is co-localized with gamma amino butyric acid unknown.
(GABA) in a specific subset of hilar neurons, the Electron microscopy has shown that HIPP cell
HIPP cells that project to the outer molecular axons terminate on spines of dentate granule
layer. These neurons also have an axonal plexus cell (DGC) dendrites, forming symmetrical (inhib-
within the hilus, suggesting they can modulate itory) synapses. Interestingly, asymmetrical (exci-
hilar neurons in addition to their projection to tatory) synapses are frequently observed on the
the outer molecular layer, where they presumably same spines, suggesting entorhinal afferents and
influence granule cells (Leranth et al., 1990; Lubke HIPP axons form synapses on the same spines
et al., 1998). Some SST neurons have axons that (Milner and Bacon, 1989; Leranth et al., 1990).
cross the hippocampal fissure into CA1, suggesting Thus, based on the anatomical data, one would
interactions of SST terminals from the dentate predict that HIPP interneurons would mediate
with SST terminals from the analogs population of feedback inhibition of granule cells, with their
SST neurons in CA1, which terminate nearby in dendrites remaining in the hilus and receiving in-
the stratum lanunosum moleculare. The dendrites put from mossy fibers, and their axons terminating
of HIPP cells are spiny and appear to be restricted in the outer molecular layer adjacent to excitatory
to the hilus. They express mGluR1 and substance perforant path input onto granule cell dendrites.
P receptors (Freund and Buzsaki, 1996). Definitive proof of this feedback role is lacking,
270

in fact, enhanced feedback inhibition has been (Vela et al., 2003). In aged primates, a dramatic
consistently observed after seizure-induced death decline in hippocampal SST mRNA is also
of HIPP neurons (Swanson et al., 1998; Sokal and observed, although no regional analysis was
Large, 2001; Kobayashi and Buckmaster, 2003). performed in this study (Hayashi et al., 1997).
HIPP-like neurons have been characterized Interestingly, a significant negative correlation
electrophysiologically in rat (Scharfman, 1992; has been found in cognitive decline in aged rats
Mott et al., 1997; Lubke et al., 1998), however, and SST levels (Dournaud et al., 1996b). That a
identification in these studies were based on lim- similar negative relationship has been observed
ited morphological analysis and there are some between SST levels and dementia in both Parkin-
discrepancies in the findings. An interesting model son’s (Epelbaum et al., 1983) and Alzheimer’s
to study HIPP neuron properties has been devel- disease in humans (Davies et al., 1980) suggests an
oped. A subset of interneurons was engineered intriguing potential role for SST in cognition.
to express EGFP in a transgenic mouse, the However, in most cases it is unclear whether this
GIN mice (GFP-expressing inhibitory neurons), decrease in SST expression is due to changes in
using the GAD67 promoter (Oliva et al., 2000). SST gene or transcript regulation, or due to the
SST-containing neurons account for 90–98% of death of SST-containing interneurons (see below).
the EGFP expressing cells in hippocampus and If the latter is true, the relationship between SST
cortex. In the hilus, 95% of the EGFP expressing levels and cognitive decline becomes more indirect.
neurons were found to contain SST, although only CST expression in the dentate gyrus is postna-
16% of the SST-containing neurons expressed tally regulated as well, with a pattern of expression
EGFP (Oliva et al., 2000). Anatomically, the hilar that is distinct from SST. At P0 in rat,
EGFP interneurons were similar to the HIPP cells CST mRNA is apparent in hippocampal stratum
described by Freund and Buszaki (1996) in rat, oriens but not the dentate. By P10, expression
with terminals in the outer two-thirds of the throughout the hippocampus is increasing, and
molecular layer. These neurons also were positive SST-positive neurons appear in the hilus. Expres-
for mGluR1, also a characteristic of HIPP cells sion of CST mRNA in hilar neurons peaks at
(Freund and Buzsaki, 1996). These mice have been approximately postnatal day 16 and then declines,
used to characterize electrophysiological proper- with very few labeled neurons present in adult
ties of SST interneurons in sensorimotor cortex, hilus (de Lecea et al., 1997a). Whether CST has
which revealed distinct subpopulations of these a specific function during development is un-
neurons (Halabisky et al., 2006). These mice pro- known, although developmental differences in its
vide an excellent model system for a rigorous study electrophysiological actions have been reported
of electrophysiological and network properties of (de Lecea et al., 1997a). Interestingly, in aged mice
SST-containing HIPP cells in mice. (24 months), CST levels have increased in hilar
interneurons (Winsky-Sommerer et al., 2004).

SST expression in the hippocampus is regulated


during development SST expression is activity-dependent

Expression of SST in hippocampus changes during The SST gene promoter has a prototypical CRE
development. A surge in SST immunoreactivity site that has been studied extensively to under-
occurs between days E12 and E15, with expression stand mechanisms of action of CREB (Montminy
beginning in the subiculum and progressing ‘‘back- and Bilezikjian, 1987; Powers et al., 1989; Mont-
ward’’ though the hippocampus, so that hilar ex- miny et al., 1996). The transcription factor CREB
pression appears last (Rapp and Amaral, 1988). is a major transducer of activity-dependent gene
After maturity, a decline in SST levels occurs expression and thus mediates transformation of
throughout rat hippocampus, with hilar expres- electrical to genetic signaling. CREB mediates
sion showing the largest decrease in aged animals its actions by interacting with CRE sites on
271

promoters. Thus, SST mRNA and ultimately pep- frequencies, it seems logical that during seizures,
tide levels are upregulated by neuronal activity. SST release would be robust. This has proven to be
SST in hilar interneurons appears to be especially the case. In vivo voltammetry showed detectable
sensitive to upregulation by seizure-like activity. SST release in hippocampus 48 min following a
This could be due to dense innervation of these seizure, with increased release continuing for the
neurons by mossy fibers, the axons of DGCs, be- 3 h recording period after the seizure (Manfridi
cause granule cells and other principal neurons et al., 1991). In hippocampal slices, both baseline
of hippocampus are active during many types of and K+-induced SST release was significantly
seizures, particularly limbic seizures. Studies using higher in kindled rats than in control rats (Vezzani
c-fos, a downstream target of CREB, indicate that et al., 1992). An in vivo study showed similar
hilar interneurons are ‘‘turned on’’ transcription- findings using microdialysis, that is, an increase
ally shortly after DGCs and before CA3 or CA1 in both basal and K+-stimulated release in
pyramidal neurons (Le Gal La Salle, 1988; Peng hippocampus after kindling (Marti et al., 2000).
and Houser, 2005). That SST release is enhanced after kindling
Thus, it is not surprising that the intense suggests that upregulation of SST after seizures
neuronal activation underlying seizures results in has functional implications.
upregulation of SST expression. Hippocampal Interestingly, SST has antiepileptic properties in
kindling upregulates hilar SST mRNA and pep- hippocampus. This was first shown using in vivo
tide once stage II seizures (mild seizures) are models in rats, where SST was injected intrahip-
reached (Bendotti et al., 1993; Vezzani et al., pocampally or ICV. For example, Vezzani and
1996). This effect was transient, lasting 1 week. colleagues demonstrated that SST or SST2 selec-
Repeated, but not single electroconvulsive shock tive ligands could decrease the severity of seizures
(ECS)-induced seizures increases SST mRNA in evoked by pentylenetetrazol (Perez et al., 1995),
the hilus (Passarelli and Orzi, 1993; Mikkelsen and kainate (Vezzani et al., 2000), or quinolic acid
Woldbye, 2006). This regulation is specific to dent- (Vezzani et al., 1991). This group also showed that
ate, as no significant changes were detected in CA3 continuous infusion of anti-SST antibody into the
and CA1 (Passarelli and Orzi, 1993). Another hippocampus decreased the latency, and dramat-
study showed repeated ECS increased the density ically reduced the number of kindling stimulations
of SST immunoreactivity in the outer molecular required to reach stage V seizures, suggesting
layer as well as in hilar cell bodies. This effect was endogenous SST plays an important role in main-
additive over 36 days of repeated ECS. Interest- taining a seizure-free state. This was confirmed by
ingly, lesioning the perforant path input to the a different group, who also demonstrated that
dentate did not alter ECS-induced upregulation of anti-SST antibody injected into the hippocampus
SST (Kragh et al., 1994). Kainate-induced seizures blocked the protective effects of SST, administered
also induce expression of SST mRNA in the hilus, ICV, on picrotoxin-kindled rats (Mazarati and
and elsewhere in hippocampus, beginning as early Telegdy, 1992). This study supports the hypothesis
as 3 h after seizures (Hashimoto and Obata, 1991). that hippocampal SST is critical in controlling
Ectopic expression in DGCs and pyramidal cells epileptogenesis and seizures. We used an in vitro
has also been reported after kainate-induced sei- model system to examine whether SST could
zures, suggesting SST expression can be activated suppress epileptiform activity in the rat hippo-
in neurons that normally do not express the ne- campal slice (Tallent and Siggins, 1999). We found
uropeptide (Hashimoto and Obata, 1991; Calbet a robust effect of SST to decrease epileptiform
et al., 1999). Expression of CST, with a promoter events generated either by increasing excitability
distinct from SST, is not induced by kainate, (by removing extracellular Mg++) or by decreas-
demonstrating that these two related peptides are ing inhibition (by blocking GABAA receptors).
independently regulated (Calbet et al., 1999). Although this study focused on epileptiform
Considering that neuropeptides are released events in CA1 and CA3, our later studies in
when peptidergic neurons are activated at high mouse hippocampal slices showed that SST has
272

robust effects on synaptic potentiation in dentate dentate at E18, validating the presence of func-
gyrus (Baratta et al., 2002, discussed below), a tional SST receptors (Thoss et al., 1996). At P5,
mechanism that likely contributes to epileptogen- SST2 and SST3 expression levels remain flat,
esis in this region (Sutula and Steward, 1986, 1987; while SST4 significantly increases, so that mRNA
Wasterlain et al., 1999). Thus, SST has potential for SST2, SST3, and SST4 are expressed at similar
sites of action throughout the hippocampus to levels. Likewise, an increase in labeling of SST
control seizure activity. receptors using radioligands is apparent in dentate
SST expression in the dentate is also regulated from E18 to P5 (Thoss et al., 1996). From P5 to
by pathological conditions that do not involve adult (no intermediate ages have been reported),
seizures. For example, SST expression increases SST1 levels have increased so that it is significantly
72 h following traumatic brain injury (Cook et al., expressed for the first time, SST2 levels modestly
1998). Interestingly, this upregulation of endog- increase, SST4 levels modestly decrease, while
enous SST could be protective, since ICV admin- SST3 levels robustly increase (Thoss et al., 1996).
istration of SST or CST protected against Regulation of SST receptor subtypes during aging
ischemia-induced neuronal death (Rauca et al., has not been reported, although one study showed
1999). Stress also increases expression and release expression of SST1–SST5 in cortex of aged humans
of SST in the hilus (Arancibia et al., 2001). (Kumar, 2005).

Expression of SST receptors in dentate gyrus Plasticity in expression of SST receptors

Prior to the cloning of the SST receptors, autora- Activity-dependent changes in expression of SST
diographic studies showed dense, high-affinity receptors in the dentate following seizures have
binding of radiolabeled SST analogs throughout been observed, but do not appear to involve
the molecular layer, with much less binding in the changes in gene expression. After kainate-induced
granule cell layer and diffuse binding in the hilus seizures, no change was observed in mRNA for
(Leroux et al., 1993). These data suggested that SST1–SST4 in dentate gyrus, although a decrease
SST released from terminals could bind to recep- for SST3 and SST4 mRNA was found in CA1,
tors throughout the dendritic layer, if it could corresponding to the death of pyramidal cells in
diffuse far enough without proteolysis. Character- this region (Perez et al., 1995). Kindling also did
ization of mRNA expression of the different re- not cause changes in mRNA for SST receptors
ceptor subtypes extended these earlier studies by in the dentate, although a 40% reduction in SST
providing evidence for SST1–SST4 in DGCs receptor binding was observed in the molecular
(Bruno et al., 1992; Kong et al., 1994). layer immediately, and 1 week following, kindling-
Immunohistochemical detection has confirmed induced seizures. Binding returned to normal
the presence of SST2, SST3, and SST4 on DGCs 30 days following seizures. Decreased SST bind-
(Dournaud et al., 1996a; Handel et al., 1999; ing without changes in mRNA expression could
Schreff et al., 2000). As described above for all have been due to receptor downregulation as a
hippocampal principal neurons, SST2 is expressed consequence of increased release of SST, although
predominately on DGCs soma and proximal it is unclear why reductions in binding were still
dendrites, SST3 on neuronal cilia, and SST4 is apparent 1 week following seizures unless recep-
almost exclusively dendritic. tor-containing neurons were damaged or were lost
Developmental changes in SST receptor mRNA (Piwko et al., 1996). Kindling also reduced
expression have been reported (Thoss et al., 1995, immunoreactivity for SST2 in the outer molecular
1996). Only SST2 and SST3 mRNA are present in layer; this was associated with an increase in SST
significant amounts at E18, suggesting these two immunoreactivity, again suggesting downregula-
receptors could play a role in development. Robust tion on binding of the receptor to its ligand (Csaba
high-affinity binding of SST ligands is present in et al., 2004). In dentate gyrus from humans with
273

epilepsy, SST2 receptor is strongly regulated Interestingly, changes in the distribution of SST-
(Csaba et al., 2005). In hippocampal tissue re- immunoreactive axons are also observed in the
sected from patients with temporal lobe seizures, dentate of hippocampal tissue resected from
SST2 expression was decreased in CA1 and CA3, patients with temporal lobe epilepsy (Mathern
likely reflecting the neuronal loss found in this re- et al., 1995; Csaba et al., 2005). In control post-
gions in patients exhibiting hippocampal sclerosis mortem tissue or in hippocampus of patients
(De Lanerolle et al., 2003). In contrast, in DGCs, without hippocampal sclerosis, a fine, diffuse
which are more resistant to seizure-induced death, network of SST-containing processes was
SST2 mRNA was strongly upregulated. However, observed in the outer two-thirds of the molecular
SST2 receptor binding and immunoreactivity layer. However, in patients with hippocampal
showed a strong decrease in the outer molecular sclerosis, accompanying the dramatic loss of
layer, suggesting SST2 may be redistributed. SST-containing hilar neurons was a change in
the SST-immunoreactive fibers in the dentate.
SST-containing axons in the molecular layer were
Hilar SST interneurons are vulnerable to more numerous, thicker, longer, and more
excitotoxicity strongly immunoreactive for SST. Large terminal
varicosities were also present. Similar, though less
A hallmark of epileptic hippocampus is the death dramatic changes were observed in fibers in the
of the SST-containing hilar interneurons. This was hilus (Csaba et al., 2005). These results suggest
first shown by Sloviter (1987) in a rat model after changes in the remaining SST-containing neurons
induction of seizures by perforant path stimula- that are consistent with axon sprouting (Mathern
tion. In hippocampal tissue surgically removed et al., 1995).
from humans with intractable temporal lobe Thus, SST interneurons are activated by
seizures, a similar pattern of loss was seen seizures, and SST expression initially increases.
(De Lanerolle et al., 1989; Robbins et al., 1991). This is followed by death of the SST neurons,
Of all subsets of GABAergic neurons, this class of presumably via excitotoxic mechanisms. Interest-
interneurons appears to be selectively vulnerable, ingly, this group of neurons is selectively vulner-
since other interneuronal markers, such as chole- able to other insults as well. Ischemia, for example,
cystokinin and vasoactive intestinal peptide, also causes a reduction in SST-containing hilar
appear unchanged (Robbins et al., 1991). This neurons (Johansen et al., 1987; Bering et al., 1997),
vulnerability of SST interneurons to seizure- as does traumatic brain injury (Lowenstein et al.,
induced death has since been confirmed in other 1992). Both ethanol ingestion and withdrawal
animal models, including kainate- and pilocarpine- induces the death of these neurons (Andrade
induced seizures (Buckmaster and Dudek, 1997; et al., 1992). Neurological diseases in which a
Kobayashi and Buckmaster, 2003). Interestingly, reduction in hilar SST has been found to include
in electrically kindled rats, a model in which spon- Alzheimer’s disease (Chan-Palay, 1987) and an
taneous seizures do not generally develop, most animal model of HIV dementia (Mitchell et al.,
studies show that the hilar SST neurons are spared 1999).
(Bendotti et al., 1993; Schwarzer et al., 1996; What causes this subset of interneurons to be
Vezzani et al., 1996; Simonato et al., 1998). These selectively vulnerable to multiple types of insults?
studies suggest that a functional consequence of Two major theories have been advanced; that
this interneuronal loss is the development of spon- these interneurons are ‘‘overstimulated’’ during
taneous seizures, and that the hilar SST neurons or hyperexcitability because they receive input from
SST itself is important in preventing epileptogen- both mossy fibers and perforant path, or that
esis. However, rate of epileptogenesis is not strictly they lack appropriate Ca2+ binding proteins so
correlated with degree of hilar neuron loss (e.g., that they are vulnerable to Ca2+-induced excito-
Lahteinen et al., 2003), suggesting other factors are toxicity. The first hypothesis assumes that HIPP
likely to contribute. interneurons have a unique innervation pattern
274

compared with other, less vulnerable, hilar inter- Electrophysiological effects of SST in dentate gyrus
neurons. However, the majority of anatomical
data suggests that their main excitatory input On the basis of projection profile of SST-contain-
comes exclusively from mossy fibers of DGCs (see ing HIPP cells, in that their major target is the
above). In fact, when activity-dependent c-fos distal dendrites of DGCs, and that SST-containing
activation is measured following seizures, HIPP terminals are often found adjacent to presumed
neurons are activated after DGCs (Le Gal La lateral perforant path synapses, it seems likely that
Salle, 1988; Peng and Houser, 2005), supporting a major role for SST would be to regulate perfo-
the anatomical data. rant path input onto DGCs. Unlike CA1, and to a
As for Ca2+ binding proteins, SST containing lesser extent CA3, the electrophysiological effects
interneurons in the hilus do not contain parval- of SST in the dentate gyrus have not been well
bumin, calretinin, or calbindin, although most characterized, especially in rat. In rat CA1 pyram-
CA1 and CA3 SST interneurons contain one of idal neurons, the major effects of SST appear to be
these Ca2+ binding proteins (Bouilleret et al., postsynaptic. SST and CST robustly enhance the
2000). Thus, it is an appealing hypothesis that the M-current (Moore et al., 1988; Schweitzer et al.,
underlying vulnerability of hilar SST interneurons 1990; de Lecea et al., 1996), a subthreshold non-
vs. SST interneurons in the rest of the hippocam- inactivating K+ current that is important in set-
pus is due to their lack of Ca2+ binding proteins. ting the excitatory tone of neurons. SST has sim-
Indeed, a study using Ca2+ binding protein ilar actions on the M-current in mouse CA1
knockout mice suggests that this may be a con- pyramidal neurons (Qiu and Tallent, unpublished
tributing factor. This study found that in wild observations). SST and CST also augment another
type mice, intrahippocampal injection of kainate K+ current in these neurons, a voltage-independ-
resulted in a reduction of SST containing inter- ent leak current (Schweitzer et al., 1998, 2003). By
neurons in CA1 to 43 and 29% of control 1 and acting on these two K+ currents, application of
30 days after injection, respectively. However, a SST causes a several millivolt hyperpolarization of
dramatic decrease in SST-containing interneurons the resting membrane potential and a decrease in
after kainate injection was observed in mice with input resistance. Thus, incoming excitatory synap-
both parvalbumin and calbindin knocked out, tic input would be less likely to trigger action po-
but not with a knockout of parvalbumin alone. tentials. SST also has presynaptic actions in rat
In the double knockout, only 12 and 6% of the CA1 pyramidal neurons, inhibiting EPSCs but not
CA1 SST-containing neurons remained 1 and IPSCs (Tallent and Siggins, 1997). Interestingly,
30 days after kainate injection, respectively. In we have observed no presynaptic effect of SST in
mice with both parvalbumin and calretinin mouse CA1 pyramidal neurons, even though in
knocked out, a larger decrease in SST immuno- side-by-side studies SST reduced EPSC amplitudes
reactivity in CA1 was observed 30 days but not 1 in rat CA1 pyramidal neurons (Qiu and Tallent,
day following kainate injections (Bouilleret et al., unpublished observations). In rat CA3 pyramidal
2000). Although changes in network excitability neurons, SST also causes membrane hyperpolari-
cannot be ruled out (although no differences in zation and reduces associational/commissural
behavioral seizures was observed between the EPSCs presynaptically (Tallent and Siggins,
groups), this study suggests that knockout of cal- 1999). In cultured immature rat hippocampal py-
bindin in particular leads to increased suscepti- ramidal neurons, SST presynaptically inhibits
bility of SST interneurons to seizure-induced glutamate but not GABA release, and inhibits
death. Thus, the lack of any of these Ca2+ bind- Ca2+ currents (Boehm and Betz, 1997).
ing proteins in hilar SST interneurons may con- In vivo studies by de Lecea and colleagues (de
tribute to their vulnerability to seizures and other Lecea et al., 1997a) have shown that CST de-
insults. presses population spike amplitude in DGCs in
275

immature (P15) but not adult rats. Interestingly, Ca2+ channels activated during the NMDA-
CST is significantly expressed in dentate gyrus (in receptor mediated depolarization. We next re-
hilar interneurons) only at early stages of devel- corded Ca2+ spikes generated in DGCs by
opment but not in adults (see above; de Lecea et depolarization. SST inhibited the Ca2+ spikes;
al., 1997a). These results suggest developmental this inhibition remained in the presence of the
changes in some SST receptor that mediates this L-type Ca2+ blocker nifedipine and the T-type
CST effect. The developmental regulation of SST Ca2+ channel blocker nickel. However, SST did
receptors in the dentate gyrus has not been exten- not inhibit Ca2+ spikes generated in the presence
sively examined, although only SST4 mRNA ex- of o-contotoxin, indicating that the peptide
pression is higher in immature than adult dentate was acting specifically on N-type Ca2+ channels.
gyrus (Thoss et al., 1995). In the presence of o-conotoxin, no LTP could be
The only detailed characterization of the elect- generated at lateral perforant path synapses,
rophysiological effects of SST in dentate was done suggesting this mechanism alone could account
by our group in adult mouse (Baratta et al., 2002). for SST inhibition of LTP (Baratta et al., 2002).
In mouse DGCs, no action of SST on postsynaptic We have also demonstrated that transgenic
K+ currents was detected, and the peptide did not overexpression of CST prevents generation of
alter membrane potential. SST also did not affect LTP at lateral perforant path synapses (Tallent
the firing properties of DGCs. We also examined et al., 2005a). In adult wild type mice, CST is not
whether SST modulated excitatory synaptic input expressed in dentate (de Lecea et al., 1997a, b). We
onto DGCs. We found that SST did not affect generated a CST overexpressing transgenic mouse
field EPSPs recorded in the outer molecular in which the transgene was driven by the neuron-
layer and evoked by low frequency stimulation specific enolase promoter. CST transgene expres-
the lateral perforant path (entorhinal cortical sion was detected in the hippocampus, cortex, and
input). However, in spite of these negative find- reticular thalamus, with especially high expression
ings on pre- and postsynaptic actions, SST in the dentate. These mice had normal baseline
robustly depressed long-term potentiation (LTP) synaptic transmission at lateral perforant path
at lateral perforant path synapses. Thus, although synapses, with the exception of reduced paired-
not acting on low-frequency input, SST depressed pulse potentiation at interstimulus intervals less
synaptic plasticity generated by high frequency than 100 ms. However, no LTP could be generated
input. This finding corroborates the speculated at these synapses, both in vivo and in vitro (Fig. 1).
role of SST based on the close association between CA1 LTP in the mice was not significantly differ-
SST-containing terminals and lateral perforant ent from wild type 60 min following induction.
path input. We further examined the mechanism A correlated deficit in spatial learning was present
by which SST inhibited LTP. Since LTP at this in the CST transgenic mice (Fig. 2), suggesting
synapse is n-methyl-D-aspartate (NMDA)-receptor synaptic plasticity in dentate gyrus may be critical
dependent, we determined whether SST could re- for this type of learning, as previous studies had
duce pharmacologically isolated NMDA receptor- suggested (Kauer et al., 1988; Moser et al., 1998;
mediated synaptic responses. SST did not act on Nakao et al., 2002).
NMDA receptor-mediated EPSCs recorded in These results and our results with exogenous
voltage-clamp, but modestly depressed NMDA SST (Baratta et al., 2002) indicate SST receptors
receptor-mediated field EPSPs recorded extra- may mediate important regulatory control over
cellularly. Interestingly, the SST effect was blocked synaptic plasticity in dentate gyrus. It is interesting
when recordings were made in the presence of o- that SST appears to have a somewhat different
conotoxin, a specific blocker of N-type Ca2+ physiological role in dentate gyrus than in
channels. These results indicated that SST does CA1 and CA3 hippocampus. In cornu ammonis,
not directly inhibit NMDA receptors but acts on high frequency activation of somatostatinergic
276

A 260 B
Control CST Transgenic 220
Baseline Post-HFT
220 0.25 mV
% Control fEPSP slope

5ms
180

180

140
140

100
100

CST
60 60
-10 0 10 20 30 40 50 60 -10 0 10 20 30 40 50 60
Time (min) Time (min)

C
260
Baseline CST/Post-HFT

220

180

140

CST
100

60
-10 0 10 20 30 40 50 60
Time (min)

Fig. 1. Transgenic overexpression or exogenous application of CST prevents induction of LTP at lateral perforant path/dentate
granule cell synapses. (A) Mean data from control and CST transgenic slices. Initial slope is plotted over time 10 min prior to 60 min
following high frequency trains (double arrows). Reduced short-term potentiation and no long-term potentiation is seen in CST
transgenics (open circles) relative to controls (filled triangles). Inset: Representative fEPSPs from a control and a transgenic mouse
before (gray) and 60 min following two 1 s 100 Hz trains (black). No potentiation of the fEPSP is seen in slices from dentate of the CST
transgenic mouse. Scale bar labels are the same for all subsequent traces. (B) Exogenous CST also reduces LTP. CST (1 mM; open
circles) was superfused beginning 7 min prior to HFTs for a total of 8 min (black bar). A small decrease in the fEPSP slope was
observed prior to HFTs (maximal inhibition was 1174% 7 min following the beginning of the superfusion). No significant potent-
iation is present 60 min following the trains when CST was superfused during the induction protocol. Inset: Representative fEPSPs
before and 60 min following HFTs with CST superfusion. (C) Exogenous CST does not affect LTP maintenance. 1 mM CST was
applied from 30 to 45 min (black bar) following HFTs (open circles). No significant effect on magnitude of LTP was observed. Inset:
Representative baseline and 60 min post-train fEPSP from experiment where CST was applied 30–45 min following train. (Reprinted
from Tallent et al., 2005a, copyright 2005, with permission from Elsevier.)
277

neurons, would be less effective mechanism in


DGCs, with their already quite hyperpolarized
membrane potential (78 to 80 mV). Certainly
regulation of the M-current in particular would be
an ineffective mechanism to control output of
granule cells, since, unlike in CA1, this current
would not be active at the resting membrane
potential of DGCs (in fact we do not observe this
current in granule cells, even though immunohisto-
chemical studies suggest its component subunits
are present: Cooper et al., 2000; Yus-Najera et al.,
2003). We also did not observe presynaptic effects
of SST or CST in mouse dentate gyrus (Baratta
et al., 2002; Tallent et al., 2005b). Thus in spite of
the rather dense localization of SST receptors on
granule cell dendrites, they do not appear to be
functionally important during standard synaptic
transmission even when SST or CST is exogenou-
sly applied. However, the effect of these peptides
Fig. 2. CST transgenic mice show a deficit in spatial learning in
on synaptic plasticity in dentate gyrus is profound.
the Barnes platform maze. In this mouse strain, overexpression No LTP is generated at lateral perforant path
of CST was largely restricted to dentate and reticular thalamus. synapses when nanomolar concentrations of SST
The percent of mice in each group of males (top) and females is applied prior to induction. Likewise, transgenic
(bottom) meeting the criterion for learning in this task (three overexpression of CST prevents generation of any
errors or less on seven of eight consecutive trials) across the 40
days of testing. Both male and female CST transgenic mice
perforant path LTP both in vivo and in vitro
showed impairment in this task as determined by significantly (Tallent et al., 2005b). Therefore, in spite of the
fewer mice in these groups learning this task. (Reprinted from similar projection pattern of somatostatinergic in-
Tallent et al., 2005a, copyright 2005, with permission from terneurons to distal dendrites of principle neurons,
Elsevier.) and of the similar localization of SST receptors on
principle cell soma and dendrites in both regions,
SST may have different physiological roles in
interneurons would release SST to hyperpolarize the cornu ammonis vs. dentate gyrus. In dentate,
pyramidal neurons, decreasing their likelihood of release of SST during high frequency events would
firing during intense activation. Presynaptic inhi- act to increase the threshold for inducing LTP.
bition of glutamate release by SST (in rat but not This mechanism may be important in increasing
mouse) would act synergistically to inhibit high the signal to noise properties of synaptic input into
frequency activity. Thus, SST has robust effects on the dentate, as well as in preventing invasion of
epileptiform activity in both CA1 and CA3 hip- seizures into the hippocampus.
pocampus, in both rats (Tallent and Siggins, 1999)
and mice (Qiu et al., 2005). Interestingly, though,
SST and CST has less robust actions on synaptic Conclusions
plasticity in CA1, reducing but not blocking LTP
(Qiu et al., 2003; Tallent et al., 2005b). In dentate SST in the dentate gyrus is expressed primarily
gyrus, SST has no observable actions on mem- in a subset of interneurons of the hilus which co-
brane potential or firing rate, thus this peptide express GABA and project to the dendritic layer of
would not regulate responsivity of DGCs to nor- DGCs. This places them in a position to modulate
mal, low-frequency synaptic input. Perhaps this granule cell activity, and the striking plasticity of
regulatory mechanism, important in pyramidal these neurons, as well as their vulnerability,
278

suggests that they could play an important role in density and inhibitory postsynaptic current frequency in the
normal and pathological conditions. Indeed, SST dentate gyrus of macaque monkeys. J. Comp. Neurol.,
476(3): 205–218.
levels in different brain regions are altered in many
Baratta, M.V., Lamp, T. and Tallent, M.K. (2002) Som-
diseases associated with cognitive deficits. It is atostatin depresses long-term potentiation and Ca2+ signa-
tempting to speculate that these peptides are im- ling in mouse dentate gyrus. J. Neurophysiol., 88(6):
portant in regulation of synaptic changes that un- 3078–3086.
derlie learning and memory. Although our studies Bendotti, C., Vezzani, A., Tarizzo, G. and Samanin, R. (1993)
in CST overexpressing mice support this hypoth- Increased expression of GAP-43, somatostatin and neuro-
peptide Y mRNA in the hippocampus during development of
esis, further characterization of the role of SST hippocampal kindling in rats. Eur. J. Neurosci., 5(10):
and CST in cognitive processing is needed. 1312–1320.
Bereiter, D.A. (1997) Morphine and somatostatin analogue
List of abbreviations reduce c-fos expression in trigeminal subnucleus caudalis
produced by corneal stimulation in the rat. Neuroscience,
77(3): 863–874.
CREB cAMP response element binding Bering, R., Draguhn, A., Diemer, N.H. and Johansen, F.F.
protein (1997) Ischemia changes the coexpression of somatostatin
CSF cerebrospinal fluid and neuropeptide Y in hippocampal interneurons. Exp.
CST cortistatin Brain. Res., 115(3): 423–429.
Blackmer, T., Larsen, E.C., Bartleson, C., Kowalchyk, J.A.,
DGC dentate granule cell Yoon, E.-J., Preininger, A.M., Alford, S., Hamm, H.E.
EGFP enhanced green fluorescent pro- and Martin, T.F.J. (2005) G protein [beta][gamma] directly
tein regulates SNARE protein fusion machinery for secretory
GABA gamma amino butyric acid granule exocytosis. Nat. Neurosci., 8(4): 421–425.
HIPP hilar perforant path associated Blake, A.D. (2001) Somatostatin receptor subtype 1 (sst(1))
regulates intracellular 30 ,50 -cyclic adenosine monophosphate
LTP long-term potentiation accumulation in rat embryonic cortical neurons: evidence
NMDA n-methyl-D-aspartate with L-797,591, an sst(1)-subtype-selective nonpeptidyl
PTX pertussis toxin agonist. Neuropharmacology, 40(4): 590–596.
SST somatostatin Boehm, S. and Betz, H. (1997) Somatostatin inhibits excitatory
transmission at rat hippocampal synapses via presynaptic
receptors. J. Neurosci., 17(11): 4066–4075.
Bouilleret, V., Schwaller, B., Schurmans, S., Celio, M.R. and
References Fritschy, J.M. (2000) Neurodegenerative and morphogenic
changes in a mouse model of temporal lobe epilepsy do not
Agren, H. and Lundqvist, G. (1984) Low levels of somatostatin depend on the expression of the calcium-binding proteins
in human CSF mark depressive episodes. Psycho- parvalbumin, calbindin, or calretinin. Neuroscience, 97(1):
neuroendocrinology, 9(3): 233–248. 47–58.
Amaral, D.G., Insausti, R. and Campbell, M.J. (1988) Distri- Brazeau, P., Vale, W., Burgus, R., Ling, N., Rivier, J. and
bution of somatostatin immunoreactivity in the human dent- Guillemin, R. (1972) Hypothalamic polypeptide that inhibits
ate gyrus. J. Neurosci., 8(9): 3306–3316. the secretion of immunoreactive pituitary growth hormone.
Andrade, J.P., Fernando, P.M., Madeira, M.D., Paula-Barb- Science, 129: 77–79.
osa, M.M., Cadete-Leite, A. and Zimmer, J. (1992) Effects of Bruno, J., Yu, Y., Song, J. and Berlowitz, M. (1992) Molecular
chronic alcohol consumption and withdrawal on the som- cloning and functional expression of a novel brain som-
atostatin-immunoreactive neurons of the rat hippocampal atostatin receptor. PNAS, 89: 11151–11155.
dentate hilus. Hippocampus, 2(1): 65–71. Buckmaster, P.S. and Dudek, F.E. (1997) Neuron loss, granule
Arancibia, S., Payet, O., Givalois, L. and Tapia-Arancibia, L. cell axon reorganization, and functional changes in the dent-
(2001) Acute stress and dexamethasone rapidly increase hip- ate gyrus of epileptic kainate-treated rats. J. Comp. Neurol.,
pocampal somatostatin synthesis and release from the dent- 385(3): 385–404.
ate gyrus hilus. Hippocampus, 11(4): 469–477. Buckmaster, P.S., Kunkel, D.D., Robbins, R.J. and Schwa-
Aronin, N., Cooper, P.E., Lorenz, L.J., Bird, E.D., Sagar, rtzkroin, P.A. (1994) Somatostatin-immunoreactivity in the
S.M., Leeman, S.E. and Martin, J.B. (1983) Somatostatin is hippocampus of mouse, rat, guinea pig, and rabbit. Hippo-
increased in the basal ganglia in Huntington disease. Ann. campus, 4(2): 167–180.
Neurol., 13(5): 519–526. Calbet, M., Guadano-Ferraz, A., Spier, A.D., Maj, M.,
Austin, J.E. and Buckmaster, P.S. (2004) Recurrent excitation Sutcliffe, J.G., Przewlocki, R. and de Lecea, L. (1999) Co-
of granule cells with basal dendrites and low interneuron rtistatin and somatostatin mRNAs are differentially
279

regulated in response to kainate. Brain Res. Mol. Brain Res., de Lecea, L., Del Rio, J.A., Criado, J.R., Alcantara, S., Mo-
72(1): 55–64. rales, M., Danielson, P.E., Henriksen, S.J., Soriano, E. and
Chan-Palay, V. (1987) Somatostatin immunoreactive neurons Sutcliffe, J.G. (1997a) Cortistatin is expressed in a distinct
in the human hippocampus and cortex shown by immuno- subset of cortical interneurons. J. Neurosci., 17(15):
gold/silver intensification on vibratome sections: coexistence 5868–5880.
with neuropeptide Y neurons, and effects in Alzheimer-type de Lecea, L., Ruiz-Lozano, P., Danielson, P.E., Peelle-Kirley,
dementia. J. Comp. Neurol., 260(2): 201–223. J., Foye, P.E., Frankel, W.N. and Sutcliffe, J.G. (1997b)
Chesselet, M.F. and Reisine, T.D. (1983) Somatostatin regu- Cloning, mRNA expression, and chromosomal mapping of
lates dopamine release in rat striatal slices and cat caudate mouse and human preprocortistatin. Genomics, 42(3):
nuclei. J. Neurosci., 3(1): 232–236. 499–506.
Chronwall, B.M., Chase, T.N. and O’donohue, T.L. (1984) Dickson, S.L., Viltart, O., Bailey, A.R. and Leng, G. (1997)
Coexistence of neuropeptide Y and somatostatin in rat and Attenuation of the growth hormone secretagogue induction
human cortical and rat hypothalamic neurons. Neurosci. of Fos protein in the rat arcuate nucleus by central som-
Lett., 52(3): 213–217. atostatin action. Neuroendocrinology, 66(3): 188–194.
Connor, M., Ingram, S.L. and Christie, M.J. (1997) Cortistatin Dournaud, P., Cervera-Pierot, P., Hirsch, E., Javoy-Agid, F.,
increase of a potassium conductance in rat locus coeruleus in Kordon, C., Agid, Y. and Epelbaum, J. (1994) Somatostatin
vitro. Br. J. Pharmacol., 122(8): 1567–1572. messenger RNA-containing neurons in Alzheimer’s disease:
Cook, J.L., Marcheselli, V., Alam, J., Deininger, P.L. and an in situ hybridization study in hippocampus, parahippo-
Bazan, N.G. (1998) Temporal changes in gene expression campal cortex and frontal cortex. Neuroscience, 61(4):
following cryogenic rat brain injury. Brain Res. Mol. Brain 755–764.
Res., 55(1): 9–19. Dournaud, P., Gu, Y.Z., Schonbrunn, A., Mazella, J., Tan-
Cooper, E.C., Aldape, K.D., Abosch, A., Barbaro, N.M., nenbaum, G.S. and Beaudet, A. (1996a) Localization of the
Berger, M.S., Peacock, W.S., Jan, Y.N. and Jan, L.Y. (2000) somatostatin receptor SST2A in rat brain using a specific anti-
Colocalization and coassembly of two human brain M-type peptide antibody. J. Neurosci., 16(14): 4468–4478.
potassium channel subunits that are mutated in epilepsy. Dournaud, P., Jazat-Poindessous, F., Slama, A., Lamour, Y.
Proc. Natl. Acad. Sci. U.S.A., 97(9): 4914–4919. and Epelbaum, J. (1996b) Correlations between water maze
Cowles, R.A., Segura, B.J. and Mulholland, M.W. (2002) performance and cortical somatostatin mRNA and high-
Regulation of carbachol-induced c-fos mRNA expression in affinity binding sites during ageing in rats. Eur. J. Neurosci.,
AR42J cells by somatostatin receptor subtypes 1, 2, and 3. 8(3): 476–485.
Pancreas, 25(3): 239–244. Edvinsson, L., Minthon, L., Ekman, R. and Gustafson, L.
Csaba, Z., Pirker, S., Lelouvier, B., Simon, A., Videau, C., (1993) Neuropeptides in cerebrospinal fluid of patients with
Epelbaum, J., Czech, T., Baumgartner, C., Sperk, G. and Alzheimer’s disease and dementia with frontotemporal lobe
Dournaud, P. (2005) Somatostatin receptor type 2 undergoes degeneration. Dementia, 4(3–4): 167–171.
plastic changes in the human epileptic dentate gyrus. J. Epelbaum, J., Ruberg, M., Moyse, E., Javoy-Agid, F., Dubois,
Neuropathol. Exp. Neurol., 64(11): 956–969. B. and Agid, Y. (1983) Somatostatin and dementia in Par-
Csaba, Z., Richichi, C., Bernard, V., Epelbaum, J., Vezzani, A. kinson’s disease. Brain Res., 278(1–2): 376–379.
and Dournaud, P. (2004) Plasticity of somatostatin and Figueredo-Cardenas, G., Morello, M., Sancesario, G.,
somatostatin sst2A receptors in the rat dentate gyrus during Bernardi, G. and Reiner, A. (1996) Colocalization of som-
kindling epileptogenesis. Eur. J. Neurosci., 19(9): 2531–2538. atostatin, neuropeptide Y, neuronal nitric oxide synthase and
Davies, P., Katzman, R. and Terry, R.D. (1980) Reduced som- NADPH-diaphorase in striatal interneurons in rats. Brain
atostatin-like immunoreactivity in cerebral cortex from cases Res., 735(2): 317–324.
of Alzheimer disease and Alzheimer senile dementia. Nature, Florio, T., Yao, H., Carey, K.D., Dillon, T.J. and Stork, P.J.
288(5788): 279–280. (1999) Somatostatin activation of mitogen-activated
De Lanerolle, N.C., Kim, J.H., Robbins, R.J. and Spencer, protein kinase via somatostatin receptor 1 (SSTR1). Mol.
D.D. (1989) Hippocampal interneuron loss and plasticity Endocrinol., 13(1): 24–37.
in human temporal lobe epilepsy. Brain Res., 495(2): Freund, T.F. and Buzsaki, G. (1996) Interneurons of the
387–395. hippocampus. Hippocampus, 6(4): 347–470.
De Lanerolle, N.C., Kim, J.H., Williamson, A., Spencer, S.S., Gabriel, S.M., Davidson, M., Haroutunian, V., Powchik, P.,
Zaveri, H.P., Eid, T. and Spencer, D.D. (2003) A retrospec- Bierer, L.M., Purohit, D.P., Perl, D.P. and Davis, K.L.
tive analysis of hippocampal pathology in human temporal (1996) Neuropeptide deficits in schizophrenia vs. Alzheimer’s
lobe epilepsy: evidence for distinctive patient subcategories. disease cerebral cortex. Biol. Psychiatry, 39(2): 82–91.
Epilepsia, 44(5): 677–687. Gerachshenko, T., Blackmer, T., Yoon, E.-J., Bartleson, C.,
de Lecea, L., Criado, J.R., Prospero-Garcia, O., Gautivik, Hamm, H.E. and Alford, S. (2005) G[beta][gamma] acts at
K.M., Schweitzer, P., Danielson, P.E., Dunlop, C.L.M., Sig- the C terminus of SNAP-25 to mediate presynaptic inhibi-
gins, G.R., Henriksen, S.J. and Sutcliffe, J.G. (1996) A cor- tion. Nat. Neurosci., 8(5): 597–605.
tical neuropeptide with neuronal depressant and sleep- Grouselle, D., Winsky-Sommerer, R., David, J.P., Delacourte,
modulating properties. Nature, 381: 242–245. A., Dournaud, P. and Epelbaum, J. (1998) Loss of
280

somatostatin-like immunoreactivity in the frontal cortex Kauer, J.A., Malenka, R.C. and Nicoll, R.A. (1988) A persist-
of Alzheimer patients carrying the apolipoprotein epsilon 4 ent postsynaptic modification mediates long-term potentiat-
allele. Neurosci. Lett., 255(1): 21–24. ion in the hippocampus. Neuron, 1(10): 911–917.
Halabisky, B., Shen, F., Huguenard, J.R. and Prince, D.A. Kimura, N., Tomizawa, S., Arai, K.N. and Osamura, R.Y.
(2006) Electrophysiological classification of somatostatin- (2001) Characterization of 50 -flanking region of rat
positive interneurons in mouse sensorimotor cortex. J. Ne- somatostatin receptor sst2 gene: transcriptional regulatory
urophysiol., 96(2): 834–845. elements and activation by Pitx1 and estrogen. Endocrino-
Handel, M., Schulz, S., Stanarius, A., Schreff, M., Erdtmann- logy, 142(4): 1427–1441.
Vourliotis, M., Schmidt, H., Wolf, G. and Hollt, V. (1999) Kobayashi, M. and Buckmaster, P.S. (2003) Reduced inhibition
Selective targeting of somatostatin receptor 3 to neuronal of dentate granule cells in a model of temporal lobe epilepsy.
cilia. Neuroscience, 89(3): 909–926. J. Neurosci., 23(6): 2440–2452.
Hashimoto, T. and Obata, K. (1991) Induction of somatostatin Kong, H., Depaoli, A.M., Breder, C.D., Yasuda, K., Bell, G.I.
by kainic acid in pyramidal and granule cells of the rat hip- and Reisine, T. (1994) Differential expression of messenger
pocampus. Neurosci. Res., 12(4): 514–527. RNAs for somatostatin receptor subtypes SSTR1, SSTR2,
Hayashi, M., Yamashita, A. and Shimizu, K. (1997) Som- and SSTR3 in adult rat brain: analysis by RNA blotting and
atostatin and brain-derived neurotrophic factor mRNA in situ hybridization histochemistry. Neuroscience, 59(1):
expression in the primate brain: decreased levels of mRNAs 175–184.
during aging. Brain Res., 749(2): 283–289. Kosaka, T., Wu, J.-Y. and Benoit, R. (1988) GABAergic
Hervieu, G. and Emson, P.C. (1998) The localization of som- neurons containing somatostatin-like immunoreactivity in
atostatin receptor 1 (sst1) immunoreactivity in the rat brain the rat hippocampus and dentate gyrus. Exp. Brain Res., 71:
using an N-terminal specific antibody. Neuroscience, 85(4): 388–398.
1263–1284. Kragh, J., Tonder, N., Finsen, B.R., Zimmer, J. and Bolwig,
Holm, I.E., Geneser, F.A. and Zimmer, J. (1992) Somatostatin- T.G. (1994) Repeated electroconvulsive shocks cause tran-
and neuropeptide Y-like immunoreactivity in the dentate sient changes in rat hippocampal somatostatin and neuro-
area, hippocampus, and subiculum of the domestic pig. peptide Y immunoreactivity and mRNA in situ hybridization
J. Comp. Neurol., 322(3): 390–408. signals. Exp. Brain Res., 98(2): 305–313.
Huang, S.M., Mouri, A., Kokubo, H., Nakajima, R., Suemoto, Kumar, U. (2005) Expression of somatostatin receptor subtypes
T., Higuchi, M., Staufenbiel, M., Noda, Y., Yamaguchi, H., (SSTR1-5) in Alzheimer’s disease brain: an immunohisto-
Nabeshima, T., Saido, T.C. and Iwata, N. (2006) Neprilysin- chemical analysis. Neuroscience, 134(2): 525–538.
sensitive synapse-associated Abeta oligomers impair neuron- Lahlou, H., Guillermet, J., Hortala, M., Vernejoul, F., Pyron-
al plasticity and cognitive function. J. Biol. Chem., 281(26): net, S., Bousquet, C. and Susini, C. (2004) Molecular signa-
17941–17951. ling of somatostatin receptors. Ann. N.Y. Acad. Sci.,
Ikeda, S.R. and Schofield, G.G. (1989) Somatostatin blocks 1014(1): 121–131.
a calcium current in rat sympathetic ganglion neurons. Lahlou, H., Saint-Laurent, N., Esteve, J.P., Eychene, A.,
J. Physiol., 409: 221–240. Pradayrol, L., Pyronnet, S. and Susini, C. (2003) sst2
Inoue, M., Nakajima, S. and Nakajima, Y. (1988) Somatostatin Somatostatin receptor inhibits cell proliferation through
induces an inward rectification in rat locus coeruleus neurons Ras-, Rap1-, and B-Raf-dependent ERK2 activation. J. Biol.
through a pertussis toxin sensitive mechanism. J. Physiol., Chem., 278(41): 39356–39371.
407: 177–198. Lahteinen, S., Pitkanen, A., Koponen, E., Saarelainen, T. and
Ishihara, S., Hassan, S., Kinoshita, Y., Moriyama, N., Castren, E. (2003) Exacerbated status epilepticus and acute
Fukuda, R., Maekawa, T., Okada, A. and Chiba, T. (1999) cell loss, but no changes in epileptogenesis, in mice
Growth inhibitory effects of somatostatin on human le- with increased brain-derived neurotrophic factor signaling.
ukemia cell lines mediated by somatostatin receptor subtype Neuroscience, 122(4): 1081–1092.
1. Peptides, 20(3): 313–318. Le Gal La Salle, G. (1988) Long-lasting and sequential increase
Iwata, N., Takaki, Y., Fukami, S., Tsubuki, S. and Saido, T.C. of c-fos oncoprotein expression in kainic acid-induced status
(2002) Region-specific reduction of A beta-degrading end- epilepticus. Neurosci. Lett., 88(2): 127–130.
opeptidase, neprilysin, in mouse hippocampus upon aging. Leranth, C., Malcolm, A.J. and Frotscher, M. (1990) Affer-
J. Neurosci. Res., 70(3): 493–500. ent and efferent synaptic connections of somatostatin-
Jinno, S. and Kosaka, T. (2000) Colocalization of parvalbumin immunoreactive neurons in the rat fascia dentata. J. Comp.
and somatostatin-like immunoreactivity in the mouse hippo- Neurol., 295(1): 111–122.
campus: quantitative analysis with optical dissector. Leroux, P., Weissmann, D., Pujol, J.-F. and Vaudry, H. (1993)
J. Comp. Neurol., 428(3): 377–388. Quantitative autoradiography of somatostatin receptors in
Johansen, F.F., Zimmer, J. and Diemer, N.H. (1987) Early loss the rat limbic system. J. Comp. Neurol., 331: 389–401.
of somatostatin neurons in dentate hilus after cerebral is- Lowenstein, D.H., Thomas, M.J., Smith, D.H. and Mcintosh,
chemia in the rat precedes CA-1 pyramidal cell loss. Acta T.K. (1992) Selective vulnerability of dentate hilar neurons
Neuropathol., 73(2): 110–114. following traumatic brain injury: a potential mechanistic link
281

between head trauma and disorders of the hippocampus. Montminy, M.R. and Bilezikjian, L.M. (1987) Binding of a
J. Neurosci., 12(12): 4846–4853. nuclear protein to the cyclic-AMP response element of the
Lubke, J., Frotscher, M. and Spruston, N. (1998) Specialized somatostatin gene. Nature, 328(6126): 175–178.
electrophysiological properties of anatomically identified Moore, S.D., Madamba, S.G., Joels, M. and Siggins, G.R.
neurons in the hilar region of the rat fascia dentata. J. (1988) Somatostatin augments the M-current in hippocampal
Neurophysiol., 79(3): 1518–1534. neurons. Science, 239(4837): 278–280.
Manfridi, A., Forloni, G.L., Vezzani, A., Fodritto, F. and De Moser, E.I., Krobert, K.A., Moser, M.B. and Morris, R.G.
Simoni, M.G. (1991) Functional and histological consequen- (1998) Impaired spatial learning after saturation of long-term
ces of quinolinic and kainic acid-induced seizures on hippo- potentiation. Science, 281(5385): 2038–2042.
campal somatostatin neurons. Neuroscience, 41(1): 127–135. Mott, D.D., Turner, D.A., Okazaki, M.M. and Lewis, D.V.
Marti, M., Bregola, G., Morari, M., Gemignani, A. and (1997) Interneurons of the dentate-hilus border of the rat
Simonato, M. (2000) Somatostatin release in the hippocam- dentate gyrus: morphological and electrophysiological heter-
pus in the kindling model of epilepsy: a microdialysis study. ogeneity. J. Neurosci., 17(11): 3990–4005.
J. Neurochem., 74(6): 2497–2503. Nakao, K., Ikegaya, Y., Yamada, M.K., Nishiyama, N. and
Mathern, G.W., Babb, T.L., Pretorius, J.K. and Leite, J.P. Matsuki, N. (2002) Hippocampal long-term depression as an
(1995) Reactive synaptogenesis and neuron densities for index of spatial working memory. Eur. J. Neurosci., 16(5):
neuropeptide Y, somatostatin, and glutamate decarboxylase 970–974.
immunoreactivity in the epileptogenic human fascia dentata. Nakatani, N., Hattori, E., Ohnishi, T., Dean, B., Iwayama, Y.,
J. Neurosci., 15(5 pt 2): 3990–4004. Matsumoto, I., Kato, T., Osumi, N., Higuchi, T., Niwa, S.I.
Mazarati, A.M. and Telegdy, G. (1992) Effects of somatostatin and Yoshikawa, T. (2006) Genome-wide expression analysis
and anti-somatostatin serum on picrotoxin-kindled seizures. detects eight genes with robust alterations specific to bipolar I
Neuropharmacology, 31(8): 793–797. disorder: relevance to neuronal network perturbation. Hum.
Mcdonald, A.J. (1989) Coexistence of somatostatin with Mol. Genet., 15(12): 1949–1962.
neuropeptide Y, but not with cholecystokinin or vasoactive Nemeroff, C.B., Youngblood, W.W., Manberg, P.J., Prange
intestinal peptide, in neurons of the rat amygdala. Brain Res., Jr., A.J. and Kizer, J.S. (1983) Regional brain concentrations
500(1–2): 37–45. of neuropeptides in Huntington’s chorea and schizophrenia.
Meriney, S.D., Gray, D.B. and Pilar, G.R. (1994) Somatostatin- Science, 221(4614): 972–975.
induced inhibition of neuronal calcium current modulated by Neves, S.R., Ram, P.T. and Iyengar, R. (2002) G protein path-
cGMP-dependent protein kinase. Nature, 369(6478): 336–339. ways. Science, 296(5573): 1636–1639.
Mihara, S., North, R.A. and Surprenant, A. (1987) Somatosta- Oliva Jr., A.A., Jiang, M., Lam, T., Smith, K.L. and Swann,
tin increases an inwardly rectifying potassium conductance in J.W. (2000) Novel hippocampal interneuronal subtypes iden-
guinea pig submucosa plexus neurones. J. Physiol., 390: tified using transgenic mice that express green fluorescent
335–355. protein in GABAergic interneurons. J. Neurosci., 20(9):
Mikkelsen, J.D. and Woldbye, D.P. (2006) Accumulated in- 3354–3368.
crease in neuropeptide Y and somatostatin gene expression of Passarelli, F. and Orzi, F. (1993) Somatostatin mRNA in the
the rat in response to repeated electroconvulsive stimulation. hippocampal formation following electroconvulsive shock in
J. Psychiatr. Res., 40(2): 153–159. the rat. Neurosci. Lett., 153(2): 197–201.
Milner, T.A. and Bacon, C.E. (1989) Ultrastructural localiza- Peng, Z. and Houser, C.R. (2005) Temporal patterns of fos
tion of somatostatin-like immunoreactivity in the rat dentate expression in the dentate gyrus after spontaneous seizures in
gyrus. J. Comp. Neurol., 290(4): 544–560. a mouse model of temporal lobe epilepsy. J. Neurosci.,
Minthon, L., Edvinsson, L. and Gustafson, L. (1997) Som- 25(31): 7210–7220.
atostatin and neuropeptide Y in cerebrospinal fluid: correla- Perez, J., Vezzani, A., Civenni, G., Tutka, P., Rizzi, M.,
tions with severity of disease and clinical signs in Alzheimer’s Schuepbach, E. and Hoyer, D. (1995) Functional effects of D-
disease and frontotemporal dementia. Dement. Geriatr. Phe-c[Cys-Tyr-D-Trp-Lys-Val-Cys]-Trp-NH2 and differential
Cogn. Disord., 8(4): 232–239. changes in somatostatin receptor messenger RNAs, binding
Mitchell, T.W., Buckmaster, P.S., Hoover, E.A., Whalen, L.R. sites and somatostatin release in kainic acid-treated rats.
and Dudek, F.E. (1999) Neuron loss and axon reorganization Neuroscience, 65(4): 1087–1097.
in the dentate gyrus of cats infected with the feline immuno- Piwko, C., Thoss, V.S., Samanin, R., Hoyer, D. and Vezzani,
deficiency virus. J. Comp. Neurol., 411(4): 563–577. A. (1996) Status of somatostatin receptor messenger RNAs
Moller, L.N., Stidsen, C.E., Hartmann, B. and Holst, J.J. and binding sites in rat brain during kindling epileptogenesis.
(2003) Somatostatin receptors. Biochim. Biophys. Acta, Neuroscience, 75(3): 857–868.
1616(1): 1–84. Pola, S., Cattaneo, M.G. and Vicentini, L.M. (2003) Anti-
Montminy, M., Brindle, P., Arias, J., Ferreri, K. and Arm- migratory and anti-invasive effect of somatostatin in
strong, R. (1996) Regulation of somatostatin gene transcrip- human neuroblastoma cells: involvement of Rac and
tion by cyclic adenosine monophosphate. Metabolism, 45(8 MAP kinase activity. J. Biol. Chem., 278(42):
Suppl 1): 4–7. 40601–40606.
282

Powers, A.C., Tedeschi, F., Wright, K.E., Chan, J.S. and Ha- concentrations following haloperidol-depot treatment in rats.
bener, J.F. (1989) Somatostatin gene expression in pancreatic Neuropeptides, 29(3): 157–161.
islet cells is directed by cell-specific DNA control elements and Scharfman, H.E. (1992) Differentiation of rat dentate neurons
DNA-binding proteins. J. Biol. Chem., 264(17): 10048–10056. by morphology and electrophysiology in hippocampal slices:
Qiu, C., Suntheimer, R.J. and Tallent, M.K. (2003) Somatosta- granule cells, spiny hilar cells and aspiny ‘fast-spiking’ cells.
tin depresses CA1 long term potentiation. Abstract Viewer/ Epilepsy Res. Suppl., 7: 93–109.
Itinerary Planner. Washington, DC: Society for Neurosci- Schindler, M., Humphrey, P.P., Lohrke, S. and Friauf, E.
ence, Program No. 889.1. (1999) Immunohistochemical localization of the somatostatin
Qiu, C., Suzuki, C., Zeyda, T., Hochgeschwender, U., de Lecea, sst2(b) receptor splice variant in the rat central nervous sys-
L. and Tallent, M.K. (2005) Increased seizure severity and tem. Neuroscience, 90(3): 859–874.
reduced somatostatin effects in hippocampus of somatostatin Schreff, M., Schulz, S., Handel, M., Keilhoff, G., Braun, H.,
receptor subtype 4 (SST4) knockout mice. 2005 Abstract Pereira, G., Klutzny, M., Schmidt, H., Wolf, G. and Hollt, V.
Viewer/Itinerary Planner. Washington, DC: Society for (2000) Distribution, targeting, and internalization of the sst4
Neuroscience, Program No. 607.11. somatostatin receptor in rat brain. J. Neurosci., 20(10):
Ramirez, J.L., Gracia-Navarro, F., Garcia-Navarro, S., 3785–3797.
Torronteras, R., Malagon, M.M. and Castano, J.P. (2002) Schulz, S., Handel, M., Schreff, M., Schmidt, H. and Hollt, V.
Somatostatin stimulates GH secretion in two porcine soma- (2000) Localization of five somatostatin receptors in the rat
totrope subpopulations through a cAMP-dependent path- central nervous system using subtype-specific antibodies. J.
way. Endocrinology, 143(3): 889–897. Physiol. Paris, 94(3–4): 259–264.
Ramos, B., Baglietto-Vargas, D., Del Rio, J.C., Moreno- Schwarzer, C., Sperk, G., Samanin, R., Rizzi, M., Gariboldi,
Gonzalez, I., Santa-Maria, C., Jimenez, S., Caballero, C., M. and Vezzani, A. (1996) Neuropeptides-immunoreactivity
Lopez-Tellez, J.F., Khan, Z.U., Ruano, D., Gutierrez, A. and their mRNA expression in kindling: functional implica-
and Vitorica, J. (2006) Early neuropathology of somatosta- tions for limbic epileptogenesis. Brain Res. Brain Res. Rev.,
tin/NPY GABAergic cells in the hippocampus of a 22(1): 27–50.
PS1  APP transgenic model of Alzheimer’s disease. Schweitzer, P., Madamba, S. and Siggins, G.R. (1990)
Neurobiol. Aging, 27(11): 1658–1672. Arachidonic acid metabolites as mediators of somatostatin-
Rapp, P.R. and Amaral, D.G. (1988) The time of origin of induced increase of neuronal M-current. Nature, 346(6283):
somatostatin-immunoreactive neurons in the rat hippocam- 464–467.
pal formation. Brain Res., 469(1–2): 231–239. Schweitzer, P., Madamba, S.G. and Siggins, G.R. (1998) Som-
Rauca, C., Schafer, K. and Hollt, V. (1999) Effects of som- atostatin increases a voltage-insensitive K+ conductance in
atostatin, octreotide and cortistatin on ischaemic neuronal rat CA1 hippocampal neurons. J. Neurophysiol., 79(3):
damage following permanent middle cerebral artery occlu- 1230–1238.
sion in the rat. Naunyn. Schmiedebergs Arch. Pharmacol., Schweitzer, P., Madamba, S.G. and Siggins, G.R. (2003) The
360(6): 633–638. sleep-modulating peptide cortistatin augments the h-current
Raynor, K. and Reisine, T. (1992) Differential coupling of in hippocampal neurons. J. Neurosci., 23(34): 10884–10891.
somatostatin1 receptors to adenylyl cyclase in the rat stria- Sharma, R.P., Bissette, G., Janicak, P.G., Davis, J.M. and Ne-
tum vs. the pituitary and other regions of the rat brain. meroff, C.B. (1995) Elevation of CSF somatostatin concen-
J. Pharmacol. Exp. Ther., 260: 841–848. trations in mania. Am. J. Psychiatry, 152(12): 1807–1809.
Reisine, T. (1995) Somatostatin. Cell Mol. Neurobiol., 15(6): Siehler, S. and Hoyer, D. (1999) Characterisation of human
597–614. recombinant somatostatin receptors. 4. Modulation of
Robbins, R.J., Brines, M.L., Kim, J.H., Adrian, T., De Lane- phospholipase C activity. Naunyn Schmiedebergs Arch.
rolle, N., Welsh, S. and Spencer, D.D. (1991) A selective loss Pharmacol., 360(5): 522–532.
of somatostatin in the hippocampus of patients with tempo- Simonato, M., Bregola, G., Beani, L., Vezzani, A., Sala, R.,
ral lobe epilepsy. Ann. Neurol., 29(3): 325–332. Raiteri, M. and Bonanno, G. (1998) Time- and region-spe-
Roberts, G.W., Crow, T.J. and Polak, J.M. (1985) Location cific variations in somatostatin release following amygdala
of neuronal tangles in somatostatin neurons in Alzheimer’s kindling in the rat. J. Neurochem., 70(1): 252–259.
disease. Nature, 314: 92–94. Sloviter, R.S. (1987) Decreased hippocampal inhibition and a se-
Robishaw, J.D. and Berlot, C.H. (2004) Translating G protein lective loss of interneurons in experimental epilepsy. Science,
subunit diversity into functional specificity. Curr. Opin. Cell 235(4784): 73–76.
Biol., 16(2): 206–209. Sodickson, D.L. and Bean, B.P. (1998) Neurotransmitter acti-
Saito, T., Iwata, N., Tsubuki, S., Takaki, Y., Takano, J., vation of inwardly rectifying potassium current in dissociated
Huang, S.M., Suemoto, T., Higuchi, M. and Saido, T.C. hippocampal CA3 neurons: interactions among multiple re-
(2005) Somatostatin regulates brain amyloid beta peptide ceptors. J. Neurosci., 18(20): 8153–8162.
Abeta42 through modulation of proteolytic degradation. Sokal, D.M. and Large, C.H. (2001) The effects of GABA(B)
Nat. Med., 11(4): 434–439. receptor activation on spontaneous and evoked activity in
Sakai, K., Maeda, K., Chihara, K. and Kaneda, H. (1995) In- the dentate gyrus of kainic acid-treated rats. Neuropharma-
creases in cortical neuropeptide Y and somatostatin cology, 40(2): 193–202.
283

Straiker, A.J., Borden, C.R. and Sullivan, J.M. (2002) G-pro- Tomidokoro, Y., Harigaya, Y., Matsubara, E., Ikeda, M.,
tein alpha subunit isoforms couple differentially to receptors Kawarabayashi, T., Okamoto, K. and Shoji, M. (2000) Im-
that mediate presynaptic inhibition at rat hippocampal paired neurotransmitter systems by Abeta amyloidosis in
synapses. J. Neurosci., 22(7): 2460–2468. APPsw transgenic mice overexpressing amyloid beta protein
Stroh, T., Kreienkamp, H.J. and Beaudet, A. (1999) precursor. Neurosci. Lett., 292(3): 155–158.
Immunohistochemical distribution of the somatostatin re- Vecsei, L., Csala, B., Widerlov, E., Ekman, R., Czopf, J. and
ceptor subtype 5 in the adult rat brain: predominant expres- Palffy, G. (1990) Lumbar cerebrospinal fluid concentrations
sion in the basal forebrain. J. Comp. Neurol., 412(1): 69–82. of somatostatin and neuropeptide Y in multiple sclerosis.
Sutula, T. and Steward, O. (1986) Quantitative analysis of Brain Res. Bull., 25(3): 411–413.
synaptic potentiation during kindling of the perforant path. Vela, J., Gutierrez, A., Vitorica, J. and Ruano, D. (2003) Rat
J. Neurophysiol., 56(3): 732–746. hippocampal GABAergic molecular markers are differen-
Sutula, T. and Steward, O. (1987) Facilitation of kindling by tially affected by ageing. J. Neurochem., 85(2): 368–377.
prior induction of long-term potentiation in the perforant Velimirovic, B.M., Koyano, K., Nakajima, S. and Nakajima,
path. Brain Res., 420(1): 109–117. Y. (1995) Opposing mechanisms of regulation of a G-pro-
Swanson, T.H., Sperling, M.R. and O’connor, M.J. (1998) tein-coupled inward rectifier K+ channel in rat brain neu-
Strong paired pulse depression of dentate granule cells in rons. Proc. Natl. Acad. Sci. U.S.A., 92(5): 1590–1594.
slices from patients with temporal lobe epilepsy. J. Neural Vezzani, A., Monno, A., Rizzi, M., Galli, A., Barrios, M. and
Transm., 105(6–7): 613–625. Samanin, R. (1992) Somatostatin release is enhanced in the
Tallent, M.K., Fabre, V., Qiu, C., Calbet, M., Lamp, T., Bar- hippocampus of partially and fully kindled rats. Neurosci-
atta, M.V., Suzuki, C., Levy, C.L., Siggins, G.R. and Hen- ence, 51(1): 41–46.
riksen, S.J. (2005a) Cortistatin overexpression in transgenic Vezzani, A., Rizzi, M., Conti, M. and Samanin, R. (2000)
mice produces deficits in synaptic plasticity and learning. Modulatory role of neuropeptides in seizures induced in
Mol. Cell. Neurosci., 30(3): 465–475. rats by stimulation of glutamate receptors. J. Nutr., 130(4S
Tallent, M.K., Patterson, C., Hon, B., Qiu, C., Zeyda, T., Ho- Suppl): 1046S–1048S.
chgeschwender, U. and de Lecea, L. (2005b) Somatostatin Vezzani, A., Schwarzer, C., Lothman, E.W., Williamson, J. and
receptor subtype 3 is critical for hippocampal-dependent Sperk, G. (1996) Functional changes in somatostatin and
memory: studies in SST3 receptor knockout mice. Abstract neuropeptide Y containing neurons in the rat hippocampus
Viewer/Itinerary Planner. Washington, DC: Society for in chronic models of limbic seizures. Epilepsy Res., 26(1):
Neuroscience, Program No. 607.10. 267–279.
Tallent, M.K. and Siggins, G.R. (1997) Somatostatin depresses Vezzani, A., Serafini, R., Stasi, M.A., Vigano, G., Rizzi, M. and
excitatory but not inhibitory neurotransmission in rat CA1 Samanin, R. (1991) A peptidase-resistant cyclic octapeptide
hippocampus. J. Neurophysiol., 78(6): 3008–3018. analogue of somatostatin (SMS 201-995) modulates seizures
Tallent, M.K. and Siggins, G.R. (1999) Somatostatin acts in induced by quinolic acid and kainic acid differently in the rat
CA1 and CA3 to reduce hippocampal epileptiform activity. hippocampus. Neuropharmacology, 30(4): 345–352.
J. Neurophysiol., 81(4): 1626–1635. Wang, H.-L., Dichter, M.A. and Reisine, T. (1990a) Lack of
Tentler, J.J., Hadcock, J.R. and Gutierrez-Hartmann, A. (1997) cross-desensitization of somatostatin-14 and somatostatin-28
Somatostatin acts by inhibiting the cyclic 30 ,50 -adenosine receptors coupled to potassium channels in rat neocortical
monophosphate (cAMP)/protein kinase A pathway, cAMP neurons. Mol. Pharm., 38(3): 357–361.
response element-binding protein (CREB) phosphorylation, Wang, H.-L., Reisine, T. and Dichter, M.A. (1990b) Som-
and CREB transcription potency. Mol. Endocrinol., 11(7): atostatin-14 and somatostatin-28 inhibit calcium currents in
859–866. rat neocortical neurons. Neuroscience, 38(2): 335–342.
Thoss, V.S., Duc, D. and Hoyer, D. (1996) Somatostatin re- Wasterlain, C.G., Mazarati, A.M., Shirasaka, Y., Thompson,
ceptors in the developing rat brain. Eur. J. Pharmacol., K.W., Penix, L., Liu, H. and Katsumori, H. (1999) Seizure-
297(1–2): 145–155. induced hippocampal damage and chronic epilepsy: a he-
Thoss, V.S., Perez, J., Duc, D. and Hoyer, D. (1995) Embryonic bbian theory of epileptogenesis. Adv. Neurol., 79: 829–843.
and postnatal mRNA distribution of five somatostatin re- Winsky-Sommerer, R., Spier, A.D., Fabre, V., de Lecea, L. and
ceptor subtypes in the rat brain. Neuropharmacology, 34(12): Criado, J.R. (2004) Overexpression of the human beta-am-
1673–1688. yloid precursor protein downregulates cortistatin mRNA in
Todisco, A., Campbell, V., Dickinson, C.J., Delvalle, J. PDAPP mice. Brain Res., 1023(1): 157–162.
and Yamada, T. (1994) Molecular basis for somatostatin Yoshitomi, H., Fujii, Y., Miyazaki, M., Nakajima, N., Inagaki,
action: inhibition of c-fos expression and AP-1 binding. N. and Seino, S. (1997) Involvement of MAP kinase and
Am. J. Physiol. Gastrointest. Liver Physiol., 267(2): c-fos signaling in the inhibition of cell growth by somatosta-
G245–G253. tin. Am. J. Physiol., 272(5 Pt 1): E769–E774.
Todisco, A., Seva, C., Takeuchi, Y., Dickinson, C.J. and Ya- Yasojima, K., Akiyama, H., Mcgeer, E.G. and Mcgeer, P.L.
mada, T. (1995) Somatostatin inhibits AP-1 function via (2001) Reduced neprilysin in high plaque areas of Alzheimer
multiple protein phosphatases. Am. J. Physiol., 269(1 Pt 1): brain: a possible relationship to deficient degradation of beta-
G160–G166. amyloid peptide. Neurosci. Lett., 297(2): 97–100.
284

Yus-Najera, E., Munoz, A., Salvador, N., Jensen, B.S., neuropeptide Y and somatostatin neurons. Brain Res.,
Rasmussen, H.B., Defelipe, J. and Villarroel, A. (2003) 800(2): 198–206.
Localization of KCNQ5 in the normal and epileptic human Zheng, H., Bailey, A., Jiang, M.H., Honda, K., Chen, H.Y.,
temporal neocortex and hippocampal formation. Neurosci- Trumbauer, M.E., Van Der Ploeg, L.H., Schaeffer, J.M.,
ence, 120(2): 353–364. Leng, G. and Smith, R.G. (1997) Somatostatin receptor sub-
Zhang, Z.J., Lappi, D.A., Wrenn, C.C., Milner, T.A. type 2 knockout mice are refractory to growth hormone-
and Wiley, R.G. (1998) Selective lesion of the negative feedback on arcuate neurons. Mol. Endocrinol.,
cholinergic basal forebrain causes a loss of cortical 11(11): 1709–1717.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 17

Neuropeptide Y in the dentate gyrus

Günther Sperk1,, Trevor Hamilton2 and William F. Colmers2,

1
Department of Pharmacology, Medical University Innsbruck, Peter-Mayr-Str. 1a, 6020 Innsbruck, Austria
2
Department of Pharmacology, University of Alberta, 9– 36 Medical Sciences Building, Edmonton, AB, T6G 2H7, Canada

Abstract: Neuropeptide Y (NPY) is contained in at least four types of GABAergic interneurons in the
dentate gyrus, many of which also contain somatostatin and give rise to the dense NPY innervation of the
dentate outer molecular layer. In humans but not rats, minute amounts of NPY are also normally expressed
in dentate granule cells, while seizure activity in rats induces robust NPY expression in granule cells. Y1 and
Y2 receptors are the most abundant NPY receptors expressed in the dentate gyrus. Y1 receptors are
postsynaptic receptors, primarily located on granule cell dendrites in the molecular layer and some in-
terneurons, while Y2 receptors are presynaptic receptors mediating inhibition of glutamate release, and
potentially that of NPY and GABA depending on their presynaptic localization, and may also be expressed
on some hilar interneurons. In humans, monkeys and mice, Y2 receptors are also present on mossy fibers,
but not in most rat species, though functional evidence suggests their presence. Hilar interneurons con-
taining NPY degenerate in temporal lobe epilepsy and in Alzheimer’s disease and reduced levels of NPY in
dentate hilus are associated with depression. By activating Y1 receptors, NPY also exerts powerful ne-
uroproliferative effects on subgranular zone progenitor cells, increasing the number of newly born granule
cells in the adult dentate gyrus. Functionally, NPY exerts anticonvulsive actions mediated by Y2 receptors
at mossy fiber terminals, but there are no presynaptic responses to NPY at perforant path inputs to dentate
granule cells in rats or mice. NPY also has potentially complicated actions on NPY-containing interneu-
rons. Elevated expression of NPY in mossy fibers of the rat, sprouting of NPY interneurons in the human
dentate, and over-expression of Y2 receptors in mossy fibers indicate an anticonvulsive role of endogenous
NPY in epilepsy. However, the physiological role of NPY in the healthy dentate gyrus remains unclear.

Keywords: neuropeptide Y; Y1 receptors; Y2 receptors; epilepsy

Neuropeptide Y (NPY) is a 36 amino acid peptide isolated due to its C-terminal amidation common
first isolated from porcine brain extracts (Tatemoto for many neuropeptides. NPY belongs to a larger
et al., 1982). It is so named because it possesses family of pancreatic peptides that in addition
tyrosine residues at both the C- and the N-terminal includes peptide YY (PYY) and pancreatic poly-
ends and three other tyrosine residues within its peptide (PP). In the periphery, NPY is primarily
amino acid sequence. It was discovered and expressed in sympathetic neurons and in the
adrenal gland, and also in neurons of the enteric
Corresponding author. Tel.: +43-512-9003-71210; nervous system including the submucosal ganglion
E-mail: guenther.sperk@i-med.ac.at (G. Sperk) (Lundberg et al., 1983; Sundler et al., 1983). Unlike
Corresponding author.Tel.: 780-492-1933; the other members of the pancreatic PP family,
E-mail: William.colmers@ualberta.ca (W.F. Colmers) NPY is abundant in the central nervous system

DOI: 10.1016/S0079-6123(07)63017-9 285


286

(de Quidt and Emson, 1986; Morris, 1989). It is occasionally on spines of granule cells (Kohler et
preferentially located in subpopulations of GABA- al., 1986; Deller and Leranth, 1990). While most of
ergic interneurons in the telencephalon (Hendry these axons pass through the granule cell layer and
et al., 1984) in noradrenergic neurons originating do not form contacts with somata of granule cells,
from the locus coeruleus (Everitt et al., 1984; a moderate number of NPY-IR synapses are also
Harfstrand et al., 1987; Smialowska, 1988), and in present on dendrites in the inner molecular layer
hypothalamic neurons of the arcuate and other nu- and on the cell body of granule cells. The few
clei (Smith and Parent, 1986; Elmquist et al., 2005). synaptic contacts formed are symmetric, consistent
with the GABAergic nature of these interneurons.
Furthermore, numerous symmetric NPY-IR
Neurons containing NPY in the dentate gyrus and synapses are found on dendrites and somata of
their synaptic contacts neurons in the hilar area, which support the phar-
macological results reported below. Dendrites of
NPY-immunoreactive (IR) neurons, all of which some NPY-IR neurons in the hilar area receive
express GABA, are abundant in the dentate gyrus input from granule cell axons, the mossy fibers
of humans (Chan-Palay et al., 1986; Furtinger axon collaterals. From transection studies, there is
et al., 2001), monkeys (Smith et al., 1985; Smith and also evidence that NPY-IR dendrites are inner-
Parent, 1986; Kohler et al., 1986), rodents (Kohler vated by perforant pathway axon terminals in the
et al., 1986; Haas et al., 1987; Morris, 1989; Deller outer molecular layer (Deller and Leranth, 1990).
and Leranth, 1990) and pigs (Holm et al., 1992). Commissurotomy revealed direct commissural in-
NPY-containing perikarya are present in all layers put to NPY-IR dendrites at the border of the inner
of the fascia dentata. The largest population of and outer molecular layer and in the hilus.
hippocampal NPY-IR neurons is contained in the Although the vast majority of NPY-IR neurons
polymorphic region of the dentate gyrus, the dent- appear to be local circuit neurons, 2% of hilar
ate hilus (Kohler et al., 1986). In their careful ex- NPY-IR neurons seem to project to the contra-
amination of NPY-IR neurons in the rat dentate lateral hippocampus, because they become retro-
gyrus, Deller and Leranth (1990) classified four gradely labeled when horseradish peroxidase
different types of neurons: The majority (>60%, (coupled to wheat germ agglutinin) is injected into
type 2 cells) were identified as medium-sized mul- the contralateral hippocampus (Deller and Lera-
tipolar and fusiform hilar neurons with dendrites nth, 1990). These authors also report that sym-
occasionally reaching the outer molecular layer; metric NPY-IR synapses are found on the cell
the next most common (20%, type 3 cells) were bodies and dendrites of hilar neurons.
pyramidal-shaped cells in the granule cell layer with In primates (humans and monkeys) the distri-
long apical dendrites reaching the outer molecular bution of NPY-IR neurons is similar to the rat
layer. Comprising the remaining 20% were large (Chan-Palay et al., 1986; Kohler et al., 1986). The
multipolar cells located in the deep hilus (type 1); innervation of the outer molecular layer by hilar
and small multipolar NPY-IR cells located in the neurons appears to be, however, more prominent
molecular layer (type 4). About half of the hilar than in the rat. In the monkey, a small portion of
NPY-IR neurons (types 1 and 2) are also IR for NPY-IR hilar interneurons contains parvalbumin
somatostatin and the projection of these neurons (Nitsch and Leranth, 1991). In humans, unlike
has a high density of somatostatin-IR axon termi- naı̈ve rats, faint NPY-IR is also observed in mossy
nals (Kohler et al., 1986; Freund and Buzsaki, fiber terminals (Chan-Palay et al., 1986; Furtinger
1996). et al., 2001).
Light and electron microscopic studies show
that the majority of NPY-IR terminals are located Neuropeptide Y receptors in the fascia dentata
in the outer molecular layer of the rat dentate
gyrus, where they establish symmetric (Gray type Several types of NPY receptors (Y1, Y2, Y4, Y5)
2) synaptic contacts on dendritic shafts and have been identified in the brain (Dumont et al.,
287

1992; Parker and Herzog, 1999). In the hippo- Electrophysiological effects of NPY on dentate
campus of rats, mice, and humans, Y1 and Y2 neurons
receptors are the most abundant. Y1 receptor
mRNA is prominently expressed in granule cells Mossy fiber terminals
and correspond to Y1 receptor binding sites on
granule cell dendrites in the molecular layer, As it does in area CA1 (Colmers et al., 1987, 1988;
where they receive their input from hilar NPY Colmers and Bleakman, 1994), NPY also nega-
interneurons (Dumont et al., 1998; Gackenheimer tively modulates synaptic transmission at the
et al., 2001). Expression of Y1 receptors on other mossy fiber-to-CA3 synapse. Synaptic excitation,
cells of the dentate gyrus (e.g., hilar interneurons) evoked by stimulation of the mossy fiber pathway,
has not unequivocally been demonstrated so far, is potently and reversibly inhibited by bath appli-
although electrophysiological results strongly cation of NPY in a concentration-dependent man-
suggest their presence (below). Interestingly ner (Klapstein and Colmers, 1993; Guo et al.,
NPY seems to augment proliferation of progen- 2002; El Bahh et al., 2005). The mechanism
itor cells located in the subgranular zone (SGZ) underlying this NPY-mediated inhibition of gluta-
by activation of Y1 receptors, a mechanism re- mate release from mossy fibers was determined by
duced in Y1 receptor knock out mice (Howell quantal synaptic analysis. Measurements of the
et al., 2005). This suggests that Y1 receptors frequency and amplitude of spontaneous (sEPSC)
are also present on dentate granule cell (DGC) and miniature (mEPSC) excitatory postsynaptic
progenitors. currents were made in CA3 pyramidal neurons in
The site of Y2 receptor expression seems to be in the whole-cell patch configuration. Application of
part species-dependent (Dumont et al., 1998). In NPY and the Y2-preferring agonist [ahx5 24]NPY
all species investigated so far, Y2 receptor mRNA increased the interval and decreased the amplitude
is highly expressed in pyramidal cells. Whereas in of sEPSCs (Fig. 1) but had no effect on either pa-
most rat strains, notably in Sprague-Dawley rats, rameter when mEPSCs were measured in the pres-
no Y2 message was demonstrated in DGCs, ence of the voltage-dependent Na+ channel
whereas Wistar rats and most mouse strains seem blocker, tetrodotoxin (McQuiston and Colmers,
to express Y2 mRNA there (Parker and Herzog, 1996). This indicates that Y2 receptors on mossy
1999; Wolak et al., 2003; Kishi et al., 2005). Cu- fiber terminals inhibit glutamate release and
riously, there is strong evidence that mossy fiber decrease excitatory input into CA3 pyramidal
terminals in Sprague-Dawley rats express highly neurons. Subsequent experiments in area CA1
functional Y2 receptors (McQuiston and Colmers, determined that NPY inhibits glutamate release
1996), and some evidence exists for Y2 receptor- via a suppression of presynaptic calcium influx
mediated postsynaptic actions in some DGCs (Qian et al., 1997). Controversy developed because
(McQuiston et al., 1996). In humans, prominent it was unclear whether the Y5 receptor also played
expression of Y2 mRNA was demonstrated in a role in this inhibition (Guo et al., 2002). How-
granule cells, and Y2 receptor binding was found ever, in experiments at the mossy fiber-CA3 and
in the terminal areas of mossy fibers in the dentate stratum radiatum-CA1 synapse in hippocampal
hilus and stratum lucidum of humans but not rats. slices, presynaptic inhibitory responses to the
Recent studies using antibodies to Y1 and Y2 re- application either of the Y2-preferring agonist
ceptors support the findings using receptor auto- [ahx5 24]NPY, or the Y5-preferring agonist,
radiography (Stanic et al., 2006). In addition, AlaAib NPY, were both blocked entirely by pre-
Stanic et al. (2006) provided evidence that Y2 re- treatment with the potent and selective Y2 recep-
ceptors may also be located on hilar interneurons, tor antagonist, BIIE0246 (El Bahh et al., 2005). As
indicating that these receptors may also mediate this antagonist has no activity at the Y5 receptor
inhibition of GABA release from interneurons in (Doods et al., 1999), this indicates that the Y2
addition to suppressing glutamate release at mossy receptor is the only one mediating this action
fibers. of NPY, and that at high concentrations, the
288

a.
Control NPY (1 µM) Washout

50 pA

500 msec

b. c.

NPY
Control
300 pA

Control NPY13-36

20 msec

d. e.

[Leu31Pro34] - NPY
PYY
Control
Control

Fig. 1. Y2- and Y1-mediated effects in the dentate gyrus. (a) Spontaneous synaptic currents recorded in CA3 pyramidal neuron.
Sweeps in each column represent consecutive traces from neuron in control (left), in the presence of 1 mM NPY (center) and after
60 min washout (right). Note the reduction in frequency of excitatory postsynaptic currents (downward deflections) with NPY, and
increase upon washout. Subsequent experiments with selective agonists in this paper demonstrate the Y2 nature of this response
(adapted from McQuiston and Colmers, 1996, with permission). (b–e) Ca2+currents in acutely isolated rat dentate granule cells are
inhibited by agonists at Y1 receptors. Shown superimposed are current traces evoked in this neuron in control and in the presence of
the respective agonist. NPY and PYY are pan-agonists, while Leu31, Pro34 NPY is a Y1/Y5-preferring agonist, and NPY13 36 is a Y2/
Y5-preferring agonist. In 30% of neurons tested, a Y2 receptor agonist also inhibited the Ca2+current, while the Y1 agonist was
effective in every neuron tested (adapted from McQuiston et al., 1996, with permission. r 1996, by the Society for Neuroscience).
289

Y5-preferring agonist has some activity at Y2 re- Ca2+channel subtypes. Calcium currents were
ceptors (El Bahh et al., 2005). This conclusion is inhibited by the NPY Y1- and Y5-preferring
further supported by the observation that in hip- agonist, [Leu31Pro34]NPY but were much less
pocampal slices prepared from mutant mice lack- commonly inhibited by the NPY Y2- (and Y5)-
ing functional Y2 receptors, there is no inhibition preferring agonist, NPY13-36 (Fig. 1). These ob-
of either the mossy fiber or stratum radiatum- servations were consistent with the actions of NPY
evoked field excitatory postsynaptic potentials on Ca2+ influx preferentially occurring via Y1
(EPSPs) by NPY or the Y5-preferring agonist receptor activation. Selective blockade of N-type
AlaAib NPY (El Bahh et al., 2005). The Y2 Ca2+ channels occluded all actions of NPY,
receptor is therefore responsible for NPY-medi- consistent with an action of Y1 receptors to selec-
ated inhibition of glutamate release from mossy tively suppress current through this subtype of
fiber terminals, most likely by suppression of volt- voltage-dependent Ca2+channel (VDCC)
age-dependent Ca2+ influx through presynaptic (McQuiston et al., 1996). Functionally, the inhibi-
calcium channels. tion of calcium currents by NPY was postulated to
regulate the release of dynorphin from DGC dend-
rites, which is mediated by L- and N-type VDCCs
Actions of NPY at somata and dendrites of dentate (Simmons et al., 1995), but this has not been
granule cells (DGCs) investigated further. Certainly, the Y1 receptor
does mediate the inhibition of N-type VDCCs in
In the dentate molecular layer, the effect of NPY DGCs. Given the remarkable amounts of NPY
on EPSPs is, if at all detectable, certainly far less in the dentate, particularly in the molecular layer,
pronounced. Stimulation of the perforant path or it is reasonable to presume that there will be some
commissural inputs evokes EPSPs in the molecular significant physiological consequences of Y1 re-
layer that are either unaltered (Klapstein and Col- ceptor activation on the dendrites and soma of
mers, 1993), or show very minor inhibition (Bijak DGCs.
and Smialowska, 1995) with bath application of
NPY. This apparent absence of NPY actions
within the dentate gyrus itself was puzzling, given NPY effects on hilar interneurons
the significant levels of NPY receptor expression
previously reported in this region (Chan-Palay The presence of NPY receptors in the hilus is
et al., 1986; Kohler et al., 1986), and prompted controversial, with some studies showing no
further investigations. immunoreactivity (Dumont et al., 1998; Gack-
Using patch clamp recording and simultaneous enheimer et al., 2001) and others demonstrating
calcium imaging, McQuiston et al. (1996) evoked their existence (Kopp et al., 2002; Paredes et al.,
action potentials in rat DGCs in brain slices. Bath 2003). Electrophysiological recordings of hilar in-
application of NPY did not alter resting Ca2+lev- terneurons in the mouse hippocampus performed
els in the soma or dendrites of the DGCs, but did by Paredes et al. (2003) demonstrated the pres-
significantly and reversibly decrease the depolari- ence of a G-protein-coupled inwardly rectifying
zation-induced Ca2+ influx in the soma and dend- potassium current (GIRK) that was activated by
rites of these same DGCs. Because voltage clamp NPY in 8/17 dentate hilar interneurons studied.
conditions are less than ideal in neurons with their Furthermore, the Y1- and Y5-preferring agonist,
dendritic trees intact, as occurs in brain slices, [Leu31Pro34]NPY mimicked the effects of NPY,
mechanistic studies of this postsynaptic NPY also in about half the neurons tested, consistent
action were undertaken in acutely isolated DGCs, with a Y1 receptor-mediated modulation of
which are electrically more tractable. Somatic GIRK currents. This is similar to a Y1 recep-
Ca2+ currents were isolated pharmacologically tor-mediated action described in reticular and
and NPY and receptor subtype-preferring agonists ventrobasal nucleus neurons of the thalamus (Sun
were applied, as were selective blockers of different et al., 2001a, b). Paredes et al. (2003) also used
290

immunocytochemisty to determine that NPY- NPY effects on synaptosomes


containing interneurons of the dentate hilus
contain Y1 receptors, consistent with their elect- Experiments using synaptosomes to examine NPY
rophysiological findings. These observations sug- function have produced interesting results and in-
gest that a significant subpopulation of hilar terpretations. Synaptosomes, a biochemical prep-
interneurons, similar in percentage to that ex- aration of isolated synaptic terminals, are
pressing NPY itself, are inhibited by NPY. This commonly used for the analysis of synaptic re-
inhibition of inhibitory interneurons is a potential lease. The application of a high extracellular K+
explanation for the proconvulsant actions of Y1 solution stimulates synaptosomal glutamate re-
agonists in the dentate in vivo (Gariboldi et al., lease, which can be quantitated fluorimetrically.
1998). This preparation can also be used to study presy-
naptic calcium influx, using fluorescent calcium
NPY effects on long-term potentiation (LTP) indicators. Preincubation of synaptosomes with
the compound of interest allows testing of possible
LTP is reliably evoked with high frequency stim- effects on transmitter release. The release of gluta-
ulation of the perforant pathway in the dentate mate from synaptosomes prepared from the dent-
gyrus in vivo. Induction of perforant pathway ate gyrus is significantly reduced by preincubation
LTP occurs as a result of DGC dendrites under- of the Y1- and Y5-preferring agonist [Leu31-
going an NMDA receptor-mediated AMPA re- Pro34]NPY and by the Y2- and Y5-preferring ago-
ceptor insertion that is responsible for the nist NPY13-36 (Silva et al., 2001). These effects
transformation of ‘silent synapses’ to active ones occurred both in the presence of normal and very
(Bi and Poo, 1998; Lin et al., 2006; Moga et al., low levels of extracellular Ca2+, indicating that the
2006). Whittaker et al. (1999) injected NPY in- mechanism is Ca2+-independent. Thus, in syna-
tracerebroventricularly (ICV) prior to LTP in- ptosomes prepared from the rat dentate gyrus, Y1
duction in vivo. This resulted in an inhibition of and Y2 receptors appear to mediate suppression of
the induction and maintenance of perforant path glutamate release. With the recent discovery of the
LTP. The previously described Y1 receptor-me- Y5 receptor and Y5 selective agonists, the Y5 re-
diated inhibition of dendritic calcium currents ceptor has also been investigated in the dentate
(McQuiston et al., 1996) must be considered a gyrus. As is the case with Y1 and Y2 agonists, the
potential candidate postsynaptic mechanism for Y5 agonist NPY (19-23)-(Gly(1),Ser(3),Gln(4),-
the observed inhibition of LTP. Alternatively, Y1 Thr(6),Ala(31), Aib(32),Gln(34))-pancreatic PP
receptor activation may also alter downstream also inhibits P/Q-type VDCC-mediated glutamate
effectors necessary for LTP induction, such as release from synaptosomes (Silva et al., 2003).
calcium-calmodulin kinase II, protein kinase C or Interestingly, despite the significant actions of Y1-
mitogen activated/extracellular –signal-regulated and Y5-preferring agonists in this preparation,
kinase (Lin et al., 2006). An alternative interpre- neither the Y1 receptor antagonist BIBP3226 nor
tation is that NPY inhibits perforant pathway the Y5 receptor antagonist L-152,804 are able to
glutamate release on to dendrites of DGCs (Whit- block the inhibition mediated by these agonists
taker et al., 1999). However, there is scant evi- The Y2 antagonist BIIE0246, on the other hand,
dence for any presynaptic actions of NPY on potently blocks the inhibition of glutamate release,
perforant path inputs, as discussed above (Klap- which has lead to the suggestion that Y2 receptors
stein and Colmers, 1993). Therefore, the possibil- are functionally coupled to Y1 and Y5 receptors
ity exists that NPY may inhibit perforant and can override their modulatory effects (Silva et
pathway LTP by reducing postsynaptic Ca2+in- al., 2003). However, an alternate interpretation is
flux through its documented action at N-type suggested by the way the experiments in this paper
VDCCs, or by inhibiting a downstream signal were conducted: the investigators added the ago-
effector. nist and antagonists at the same time, spun the
291

synaptosomes, then reapplied the mixture when NPY effects on neurogenesis


resuspending them. Because the agonist peptides
are much larger and ‘‘stickier’’ molecules than are In the last decade, it has become clear that neural
the antagonists, it is possible that more of the precursor cells in the SGZ of the dentate gyrus
agonists remained behind when the supernatant supply new granule cells on an ongoing basis (see
was removed, and preferentially occupied the re- chapter by Parent, this volume). In the SGZ, neu-
ceptors, giving an impression that the antagonists ronal precursors proliferate and continually mi-
were ineffective. This would not be the case for the grate into the granule cell layer where they mature
Y2 receptor antagonist, BIIE0246, which exhibits and become functional granule cells. The specific
an irreversible form of antagonism upon pro- factors that influence dentate gyrus neurogenesis
longed exposure to the receptor (Dautzenberg & are, therefore, of significant interest because their
Neysari, 2005; El Bahh et al., 2005). It might be impact will modify the memory formation process
possible to test this hypothesis by pretreating the (Aimone et al., 2006). Exercise, growth factors,
synaptosomes with the antagonist alone, then re- environmental enrichment, aging, hormones such
suspending them after centrifugation in the pres- as estrogen and prolactin, glutamatergic neuro-
ence of both antagonist and agonist. transmission, adrenal hormones, LTP, lesions, sei-
Notwithstanding this possibility, it appears that zure activity, and the presence of NPY are all
in synaptosomes prepared from the dentate gyrus, associated with a regulation of dentate neurogen-
Y1, Y2, and Y5 agonists can inhibit glutamate re- esis (cf. Parent, this volume). In NPY knockout
lease. Moreover, the most prominent inhibition mice, it was fortuitously observed that prolifera-
occurs with the Y2 agonist that likely mediates the tion of olfactory precursor cells was impaired, and
inhibitory effect (Silva et al., 2003). These results NPY was seen to promote neurogenesis there, via
differ from those from the electrophysiological a Y1 receptor mechanism (Hansel et al., 2001).
studies which suggest that NPY-mediated inhibi- Howell et al. (2005) examined whether NPY might
tion of glutamate release does not occur in the affect proliferation of neural progenitors in the
dentate gyrus (Klapstein and Colmers, 1993; SGZ by culturing neuroblast precursor cells from
Mcquiston and Colmers, 1996; Mcquiston et al., the early postnatal dentate gyrus in the absence or
1996; El Bahh et al., 2002, 2005). A number of presence of NPY in the culture medium. NPY
potential differences in methods may explain these treatment significantly increased the number of
differences. First, while the electrophysiological neurons and neuroblasts that incorporated Brd-U,
studies represent the evoked or spontaneous ac- consistent with a proliferative action of NPY. The
tivity in defined neurons and synaptic pathways, strongly Y1-preferring agonist, F7,P34 NPY mim-
synaptosome preparations contain all terminals icked the effect of NPY application in vitro, and
that survive isolation, including inputs to inter- the effect of NPY was blocked by application of
neurons, terminals from other pathways, etc. in the Y1 antagonist BIBP3226 in the presence of
addition to the perforant path inputs studied. Fur- NPY. The neuroproliferative action of NPY was
thermore, mossy fiber terminals, which have mul- also abolished by inhibiting the extracellular sig-
tiple release sites (Lawrence and McBain, 2003) nal-regulated kinase (ERK)1/2, a subgroup of mi-
appear to respond to Y2 receptor agonists togen activated kinases (Scharfman et al., 2000).
(McQuiston and Colmers, 1996), and although Therefore, it is likely that a downstream effect of
there is some evidence for the presence of Y5 the Y1 receptor is the activation of ERK1/2. These
receptors in the dentate gyrus (Guo et al., 2002), authors then validated the observations in Y1
this is disputed (El Bahh et al., 2005). However, receptor knockout mice in vivo. Y1 / mice had a
the presence of Y2 receptors on the terminals significantly lower SGZ proliferation rate and a de-
of mossy fibers may be sufficient to explain creased number of neurons expressing doublecor-
the observed inhibition of glutamate release in tin than in wild types. Thus, data from in vitro and
synaptosomes. in vivo models supports a Y1 receptor-mediated
292

neuroproliferative action of NPY in the hilus of expression of Y5 receptors seems to be negligible,


the dentate gyrus. This may also play a role in at least in the mouse hippocampus and a crucial
depression (below). role of Y2 receptors has been shown in the anti-
convulsive action of NPY (El Bahh et al., 2005;
see also below).
Neuropeptide Y and its receptors in disease In hippocampal tissue surgically removed from
patients for treatment of temporal lobe epilepsy,
Epilepsy enhanced expression of NPY mRNA has been
observed in interneurons of the dentate hilus,
In animal models of epilepsy and in human together with a markedly increased total length of
temporal lobe epilepsy, the NPY system under- NPY-positive fibers in the inner and outer dentate
goes considerable changes in the dentate gyrus, molecular layer, and the stratum lacunosum-mole-
both in the expression of the peptide and of its culare (Furtinger et al., 2001). Interestingly, the
receptors (Vezzani et al., 1999). With repeated, pattern of interneuronal axon sprouting over-
mildly convulsive stimuli, such as kindling, lapped in part with terminal areas both of normal
expression of NPY and its Y2 receptors become mossy fibers and the collaterals that sprout into
up-regulated in hilar interneurons (Fig. 2) (Gruber the inner molecular layer in the epileptic brain
et al., 1994; Vezzani et al., 1999), and NPY ex- (mossy fiber sprouting). In dentate gyrus of epi-
pression is up-regulated in DGCs (Marksteiner lepsy patients, Y2 receptors were up-regulated and
et al., 1990; Sperk et al., 1992; Goodman and Y1 receptors down-regulated in mossy fibers and
Sloviter, 1993; Causing et al., 1996). Viral vectors in the molecular layer, respectively (Furtinger
inducing overexpression of NPY have clear anti- et al., 2001). This indicates that NPY, released
convulsive effects when infused locally into the from sprouted interneurons could also reach Y2
hippocampus of rats (Richichi et al., 2004). On the receptors located on mossy fibers. In epileptic rats,
other hand, prolonged seizures (e.g., during status NPY has been shown to have tonic inhibitory ac-
epilepticus) cause degeneration of hilar interneu- tions, mediated via Y2 receptors (Tu et al., 2005).
rons (Sloviter, 1983; Sperk et al., 1992). However, These changes in NPY-related neuronal circuitry
surviving neurons may still up-regulate NPY may be the basis for the anticonvulsive action of
expression. In rat models of temporal lobe epi- NPY demonstrated ex vivo in hippocampal tissue
lepsy, expression of Y2 receptors is up-regulated obtained from patients with temporal lobe epilepsy
simultaneously in granule cells and mossy fibers (Patrylo et al., 1999).
(Fig. 2) (Schwarzer et al., 1998). Over-expression
of NPY and its Y2 receptors seems to have a
protective and anticonvulsive role caused by the Alzheimer’s disease
Y2-mediated presynaptic inhibition of glutamate
release from mossy fibers (Klapstein and Colmers, Chan-Palay et al. (1986) observed a marked
1993; Greber et al., 1994; El Bahh et al., 2005). reduction in hilar NPY neurons in Alzheimer’s
Conversely, Y1 receptors located on granule cell Disease (AD) patients at autopsy. Assuming that
dendrites are down-regulated with seizure activity the degeneration of the septo-hippocampal path-
(Kofler et al., 1997). Because Gariboldi et al. way may be related to the cognitive impairment in
(1998) demonstrated that Y1 receptor agonists in- AD patients, it is interesting that lesions of the
jected into the dentate gyrus have proconvulsant septo-hippocampal pathway result in an initial in-
action in vivo, this change could ameliorate seizure crease in NPY-IR in the dentate hilus (Hortnagl
activity or compensate for the increased excitation et al., 1990; Bayer and Milner, 1993) and a sub-
(see also below). The prominent anticonvulsive sequent loss of hilar NPY neurons (Milner et al.,
action of NPY has been suggested to be mediated 1997). It is possible that the initial increase NPY-
by Y5 receptor stimulation (Woldbye et al., 1997). IR may reflect sustained stimulation of these
This has been debated, however, especially since neurons leading then to their loss through
293

Fig. 2. NPY and its receptors in control and epileptic rats. NPY-IR is present in numerous interneurons of the dentate gyrus that
project to the outer molecular layer and have collaterals within the dentate hilus (see the respective labeling in a, representing a high
magnification image of NPY-IR in the dorsal dentate gyrus). In epileptic rats (b; 30 days after kainic acid-induced seizures; lower
magnification image of the dorsal hippocampus) varying portions of hilar NPY neurons degenerate (note the loss in NPY-IR in the
outer molecular layer). After seizures mossy fibers strongly express NPY-IR unlike in naı̈ve animals. The arrow in b marks sprouted
mossy fiber terminals in the inner molecular layer also containing NPY. (c) Y1 receptors are postsynaptic receptors located mainly on
dendrites of granule cells (here labeled with the Y1/Y5 ligand 125I-Pro34-PYY) and become reduced after kainic acid-induced seizures
(arrow in d; 8 days after application of kainic acid injection). Panels e and f depict brain sections labeled with the Y2/Y5 specific ligand
125
I-PYY(3 36). Note the strong Y2-specific labeling of strata oriens and radiatum in control rats (e). In kainic acid treated rats strong
labeling of mossy fibers can be seen representing presynaptic Y2 receptors (arrow in f; 8 days after application of kainic acid injection).

excitotoxic mechanisms (Chan-Palay et al., 1986; 1994), and Y1 receptor mRNA expression in
Milner et al., 1997). DGCs (Madsen et al., 2000) of naı̈ve rats. In
FSL rats, ECT also increases NPY mRNA and Y1
mRNA expression, and reduces Y1 receptor bind-
Depression ing in the dentate gyrus (Jimenez-Vasquez et al.,
2006). This altered NPY expression was measured
In ‘‘depressed’’ Flinder’s Sensitive Line (FSL) rats, after an ECT-induced reduction in depressive
basal expression of NPY and Y1 mRNA in the behavior, measured with the Porsolt swim test.
dentate gyrus is significantly lower than in control Interestingly, ECT increases seizure threshold in
animals (Jimenez-Vasquez et al., 2006). An up- epileptic rats, which correlates with a reduction in
regulation of Y1 receptor binding in the dentate NPY binding sites in the dentate gyrus (Bolwig
gyrus was also observed, most likely due to an et al., 1999). Fluoxetine, a serotonin-selective re-
increased affinity for and/or decreased internali- uptake inhibitor used in the treatment of depres-
zation of NPY receptors (Jimenez-Vasquez et al., sion, has also been observed to increase NPY
2006). Electroconvulsive therapy (ECT), a com- mRNA in the dentate gyrus of Flinders Sensitive
mon treatment for depressive disorders, raises Line rats (Caberlotto et al., 1999), although this
NPY mRNA levels in the hilus (Mikkelsen et al., was not replicated in a different lab using tricyclic
294

antidepressants in Sprague-Dawley rats (Bellmann MEPSC miniature excitatory postsynaptic


and Sperk, 1993). Thus, a downregulation of Y1 current
receptor mRNA is intimately related to the patho- NPY neuropeptide Y
physiology of depression and is altered by treat- NPY-IR neuropeptide Y immunoreactive
ments that have been efficacious in treating NPY-LI neuropeptide Y-like immunore-
depressive disorders, possibly via an increase in activity
NPY receptor expression. PYY peptide YY
These effects are not only limited to animal PP pancreatic polypeptide
models. Depressed humans have a decreased SEPSC spontaneous excitatory postsy-
NPY-like immunoreactivity (NPY-LI) in their naptic current
cerebrospinal fluid compared to non-depressed SGZ subgranular zone
control subjects (Widerlov et al., 1988; Gjerris et VDCC voltage dependent Ca2+ channel
al., 1992). Consistent with the hypothesis that a
hypofunction of the NPY system is related to de-
References
pression, depressed human subjects exhibit in-
creased NPY-LI after ECT (Mathe, 1999). Lastly, Aimone, J.B., Wiles, J. and Gage, F.H. (2006) Potential role for
it appears that neurogenesis in the dentate gyrus adult neurogenesis in the encoding of time in new memories.
is necessary for the actions of antidepressants Nat. Neurosci., 9(6): 723–727.
(Dranovsky and Hen, 2006). It is therefore tempt- Bayer, L.E. and Milner, T.A. (1993) Transient increases in ne-
uropeptide Y-like immunoreactivity in dentate hilar neurons
ing to speculate that some of the antidepressant
following fimbria/fornix transection. J. Neurosci. Res., 34(4):
actions of the NPY system may be mediated 434–441.
through neuroproliferative actions. Bellmann, R. and Sperk, G. (1993) Effects of antidepressant
drug treatment on levels of NPY or prepro-NPY-mRNA in
the rat brain. Neurochem. Int., 22(2): 183–187.
Summary Bi, G.Q. and Poo, M.M. (1998) Synaptic modifications in cul-
tured hippocampal neurons: dependence on spike timing,
NPY is an extraordinarily abundant peptide in the synaptic strength, and postsynaptic cell type. J. Neurosci.,
dentate gyrus, but the roles that it plays remain 18(24): 10464–10472.
Bijak, M. and Smialowska, M. (1995) Effects of neuropeptide Y
somewhat obscure. It is clear that the dentate on evoked potentials in the CA1 region and the dentate gyrus
gyrus NPY peptide-receptor system is extremely of the rat hippocampal slice. Pol. J. Pharmacol., 47(4):
sensitive to ongoing activity, and is poised to reg- 333–338.
ulate many important aspects of the plasticity of Bolwig, T.G., Woldbye, D.P. and Mikkelsen, J.D. (1999) Elec-
troconvulsive therapy as an anticonvulsant: a possible role of
dentate gyrus circuitry. Certainly, the links be-
neuropeptide Y (NPY). J. ECT, 15(1): 93–101.
tween NPY and excitability, memory formation Caberlotto, L., Jimenez, P., Overstreet, D.H., Hurd, Y.L., Ma-
and depression are important and warrant inten- the, A.A. and Fuxe, K. (1999) Alterations in neuropeptide Y
sive study. levels and Y1 binding sites in the Flinders Sensitive Line rats,
a genetic animal model of depression. Neurosci. Lett., 265(3):
191–194.
Abbreviations Causing, C.G., Makus, K., Ma, Y., Miller, F.D. and Colmers,
W.F. (1996) Selective upregulation of Ta1 a-tubulin and Ne-
AD Alzheimer’s disease uropeptide Y mRNAs after intermittent excitatory stimula-
tion in adult rat hippocampus in vivo. J. Comp. Neurol.,
DGC dentate granule cell 367(1): 132–146.
ECT electroconvulsive therapy Chan-Palay, V., Kohler, C., Haesler, U., Lang, W. and
EPSP excitatory postsynaptic potential Yasargil, G. (1986) Distribution of neurons and axons
GABA gamma-aminobutyric acid immunoreactive with antisera against neuropeptide Y in the
normal human hippocampus. J. Comp. Neurol., 248(3):
GIRK G-protein-coupled inwardly-rec-
360–375.
tifying potassium current Colmers, W.F. and Bleakman, D. (1994) Neuropeptide Y
ICV intracerebroventricular effects on the electrical properties of neurons. Trends Ne-
LTP long-term potentiation urosci., 17(9): 373–379.
295

Colmers, W.F., Lukowiak, K.D. and Pittman, Q.J. (1987) Pre- receptor subtypes in the mouse brain. Peptides, 22(3):
synaptic action of neuropeptide Y in area CA1 of the rat 335–341.
hippocampal slice. J. Physiol., 383: 285–299. Gariboldi, M., Conti, M., Cavaleri, D., Samanin, R. and
Colmers, W.F., Lukowiak, K. and Pittman, Q.J. (1988) Ne- Vezzani, A. (1998) Anticonvulsant properties of BIBP3226, a
uropeptide Y action in the rat hippocampal slice: site and non-peptide selective antagonist at neuropeptide Y Y1 re-
mechanism of presynaptic inhibition. J. Neurosci., 8(10): ceptors. Eur. J. Neurosci., 10(2): 757–759.
3827–3837. Gjerris, A., Widerlov, E., Werdelin, L. and Ekman, R. (1992)
Dautzenberg, F.M. and Neysari, S. (2005) Irreversible binding Cerebrospinal fluid concentrations of neuropeptide Y in de-
kinetics of neuropeptide Y ligands to Y2 but not to Y1 and pressed patients and in controls. J. Psychiatry Neurosci.,
Y5 receptors. Pharmacology, 75(1): 21–29. 17(1): 23–27.
Deller, T. and Leranth, C. (1990) Synaptic connections of ne- Goodman, J.H. and Sloviter, R.S. (1993) Cocaine neurotoxicity
uropeptide Y (NPY) immunoreactive neurons in the hilar and altered neuropeptide Y immunoreactivity in the rat hip-
area of the rat hippocampus. J. Comp. Neurol., 300(3): pocampus; a silver degeneration and immunocytochemical
433–447. study. Brain Res., 616(1–2): 263–272.
Doods, H., Gaida, W., Wieland, H.A., Dollinger, H., Greber, S., Schwarzer, C. and Sperk, G. (1994) Neuropeptide Y
Schnorrenberg, G., Esser, F., Engel, W., Eberlein, W. and inhibits potassium-stimulated glutamate release through Y2
Rudolf, K. (1999) BIIE0246: a selective and high affinity ne- receptors in rat hippocampal slices in vitro. Br. J. Pharma-
uropeptide Y Y(2) receptor antagonist. Eur. J. Pharmacol., col., 113(3): 737–740.
384(2-3): R3–R5. Gruber, B., Greber, S., Rupp, E. and Sperk, G. (1994) Differ-
Dranovsky, A. and Hen, R. (2006) Hippocampal neurogenesis: ential NPY mRNA expression in granule cells and interneu-
regulation by stress and antidepressants. Biol. Psychiatry, rons of the rat dentate gyrus after kainic acid injection.
59(12): 1136–1143. Hippocampus, 4(4): 474–482.
Dumont, Y., Jacques, D., Bouchard, P. and Quirion, R. (1998) Guo, H., Castro, P.A., Palmiter, R.D. and Baraban, S.C. (2002)
Species differences in the expression and distribution of the Y5 receptors mediate neuropeptide Y actions at excitatory
neuropeptide Y Y1, Y2, Y4, and Y5 receptors in rodents, synapses in area CA3 of the mouse hippocampus. J. Ne-
guinea pig, and primates brains. J. Comp. Neurol., 402(3): urophysiol., 87(1): 558–566.
372–384. Haas, H.L., Hermann, A., Greene, R.W. and Chan-Palay, V.
Dumont, Y., Martel, J.C., Fournier, A., St-Pierre, S. and Quir- (1987) Action and location of neuropeptide tyrosine (Y) on
ion, R. (1992) Neuropeptide Y and neuropeptide Y receptor hippocampal neurons of the rat in slice preparations. J.
subtypes in brain and peripheral tissues. Prog. Neurobiol., Comp. Neurol., 257(2): 208–215.
38(2): 125–167. Hansel, D.E., Eipper, B.A. and Ronnett, G.V. (2001) Neuro-
El Bahh, B., Balosso, S., Hamilton, T., Herzog, H., Beck-Sick- peptide Y functions as a neuroproliferative factor. Nature,
inger, A.G., Sperk, G., Gehlert, D.R., Vezzani, A. and 410(6831): 940–944.
Colmers, W.F. (2005) The anti-epileptic actions of neuro- Harfstrand, A., Fuxe, K., Terenius, L. and Kalia, M. (1987)
peptide Y in the hippocampus are mediated by Y and not Y Neuropeptide Y-immunoreactive perikarya and nerve termi-
receptors. Eur. J. Neurosci., 22(6): 1417–1430. nals in the rat medulla oblongata: relationship to cytoarchi-
El Bahh, B., Cao, J.Q., Beck-Sickinger, A.G. and Colmers, W.F. tecture and catecholaminergic cell groups. J. Comp. Neurol.,
(2002) Blockade of neuropeptide Y(2) receptors and suppres- 260(1): 20–35.
sion of NPY’s anti-epileptic actions in the rat hippocampal Hendry, S.H., Jones, E.G., DeFelipe, J., Schmechel, D., Bran-
slice by BIIE0246. Br. J. Pharmacol., 136(4): 502–509. don, C. and Emson, P.C. (1984) Neuropeptide-containing
Elmquist, J.K., Coppari, R., Balthasar, N., Ichinose, M. and neurons of the cerebral cortex are also GABAergic. Proc.
Lowell, B.B. (2005) Identifying hypothalamic pathways con- Natl. Acad. Sci. USA, 81(20): 6526–6530.
trolling food intake, body weight, and glucose homeostasis. J. Holm, I.E., Geneser, F.A. and Zimmer, J. (1992) Somatostatin-
Comp. Neurol., 493(1): 63–71. and neuropeptide Y-like immunoreactivity in the dentate
Everitt, B.J., Hokfelt, T., Terenius, L., Tatemoto, K., Mutt, V. area, hippocampus, and subiculum of the domestic pig. J.
and Goldstein, M. (1984) Differential co-existence of neuro- Comp. Neurol., 322(2): 390–408.
peptide Y (NPY)-like immunoreactivity with catecholamines Hortnagl, H., Sperk, G., Sobal, G. and Maas, D. (1990)
in the central nervous system of the rat. Neuroscience, 11(2): Cholinergic deficit induced by ethylcholine aziridinium
443–462. (AF64A) transiently affects somatostatin and neuropeptide
Freund, T.F. and Buzsaki, G. (1996) Interneurons of the hip- Y levels in rat brain. J. Neurochem., 54(5): 1608–1613.
pocampus. Hippocampus, 6(4): 347–470. Howell, O.W., Doyle, K., Goodman, J.H., Scharfman, H.E.,
Furtinger, S., Pirker, S., Czech, T., Baumgartner, C., Herzog, H., Pringle, A., Beck-Sickinger, A.G. and Gray,
Ransmayr, G. and Sperk, G. (2001) Plasticity of Y1 and W.P. (2005) Neuropeptide Y stimulates neuronal precursor
Y2 receptors and neuropeptide Y fibers in patients with tem- proliferation in the post-natal and adult dentate gyrus. J.
poral lobe epilepsy. J. Neurosci., 21(15): 5804–5812. Neurochem., 93(3): 560–570.
Gackenheimer, S.L., Schober, D.A. and Gehlert, D.R. (2001) Jimenez-Vasquez, P.A., Diaz-Cabiale, Z., Caberlotto, L., Bel-
Characterization of neuropeptide Y Y1-like and Y2-like lido, I., Overstreet, D., Fuxe, K. and Mathe, A.A. (2006)
296

Electroconvulsive stimuli selectively affect behavior and ne- Mikkelsen, J.D., Woldbye, D., Kragh, J., Larsen, P.J. and
uropeptide Y (NPY) and NPY Y(1) receptor gene expres- Bolwig, T.G. (1994) Electroconvulsive shocks increase the
sions in hippocampus and hypothalamus of Flinders expression of neuropeptide Y (NPY) mRNA in the piriform
Sensitive Line rat model of depression. Eur. Neuropsychop- cortex and the dentate gyrus. Brain Res. Mol. Brain Res.,
harmacol., 17(4): 298–308. 23(4): 317–322.
Kishi, T., Aschkenasi, C.J., Choi, B.J., Lopez, M.E., Lee, C.E., Milner, T.A., Wiley, R.G., Kurucz, O.S., Prince, S.R. and
Liu, H., Hollenberg, A.N., Friedman, J.M. and Elmquist, Pierce, J.P. (1997) Selective changes in hippocampal neuro-
J.K. (2005) Neuropeptide Y Y1 receptor mRNA in rodent peptide Y neurons following removal of the cholinergic septal
brain: distribution and colocalization with melanocortin-4 inputs. J. Comp. Neurol., 386(1): 46–59.
receptor. J. Comp. Neurol., 482(3): 217–243. Moga, D.E., Shapiro, M.L. and Morrison, J.H. (2006) Bidi-
Klapstein, G.J. and Colmers, W.F. (1993) On the sites of pre- rectional redistribution of AMPA but not NMDA receptors
synaptic inhibition by neuropeptide Y in rat hippocampus in after perforant path simulation in the adult rat hippocampus
vitro. Hippocampus, 3(1): 103–111. in vivo. Hippocampus, 16(11): 990–1003.
Kofler, N., Kirchmair, E., Schwarzer, C. and Sperk, G. (1997) Morris, B.J. (1989) Neuronal localisation of neuropeptide Y
Altered expression of NPY-Y1 receptors in kainic acid in- gene expression in rat brain. J. Comp. Neurol., 290(3):
duced epilepsy in rats. Neurosci. Lett., 230(2): 129–132. 358–368.
Kohler, C., Eriksson, L., Davies, S. and Chan-Palay, V. (1986) Nitsch, R. and Leranth, C. (1991) Neuropeptide Y (NPY)-
Neuropeptide Y innervation of the hippocampal region in the immunoreactive neurons in the primate fascia dentata; occa-
rat and monkey brain. J. Comp. Neurol., 244(3): 384–400. sional coexistence with calcium-binding proteins: a light and
Kopp, J., Xu, Z.Q., Zhang, X., Pedrazzini, T., Herzog, H., electron microscopic study. J. Comp. Neurol., 309(4):
Kresse, A., Wong, H., Walsh, J.H. and Hokfelt, T. (2002) 430–444.
Expression of the neuropeptide Y Y1 receptor in the CNS of Paredes, M.F., Greenwood, J. and Baraban, S.C. (2003) Ne-
rat and of wild-type and Y1 receptor knock-out mice. Focus uropeptide Y modulates a G protein-coupled inwardly rec-
on immunohistochemical localization. Neuroscience, 111(3): tifying potassium current in the mouse hippocampus.
443–532. Neurosci. Lett., 340(1): 9–12.
Lawrence, J.J. and McBain, C.J. (2003) Interneuron diversity Parker, R.M. and Herzog, H. (1999) Regional distribution of
series: containing the detonation–feedforward inhibition in Y-receptor subtype mRNAs in rat brain. Eur. J. Neurosci.,
the CA3 hippocampus. Trends Neurosci., 26(11): 631–640. 11(4): 1431–1448.
Lin, Y.W., Yang, H.W., Wang, H.J., Gong, C.L., Chiu, T.H. Patrylo, P.R., van den Pol, A.N., Spencer, D.D. and William-
and Min, M.Y. (2006) Spike-timing-dependent plasticity at son, A. (1999) NPY inhibits glutamatergic excitation in the
resting and conditioned lateral perforant path synapses on epileptic human dentate gyrus. J. Neurophysiol., 82(1):
granule cells in the dentate gyrus: different roles of N-methyl- 478–483.
D-aspartate and group I metabotropic glutamate receptors. Qian, J., Colmers, W.F. and Saggau, P. (1997) Inhibition of
Eur. J. Neurosci., 23(9): 2362–2374. synaptic transmission by neuropeptide Y in rat hippocampal
Lundberg, J.M., Terenius, L., Hokfelt, T. and Goldstein, M. area CA1: modulation of presynaptic Ca2+ entry. J. Neuro-
(1983) High levels of neuropeptide Y in peripheral nor- sci., 17(21): 8169–8177.
adrenergic neurons in various mammals including man. Ne- de Quidt, M.E. and Emson, P.C. (1986) Distribution of neuro-
urosci. Lett., 42(2): 167–172. peptide Y-like immunoreactivity in the rat central nervous
Madsen, T.M., Greisen, M.H., Nielsen, S.M., Bolwig, T.G. and system–II. Immunohistochemical analysis. Neuroscience,
Mikkelsen, J.D. (2000) Electroconvulsive stimuli enhance 18(3): 545–618.
both neuropeptide Y receptor Y1 and Y2 messenger RNA Richichi, C., Lin, E.J., Stefanin, D., Colella, D., Ravizza, T.,
expression and levels of binding in the rat hippocampus. Grignaschi, G., Veglianese, P., Sperk, G., During, M.J. and
Neuroscience, 98(1): 33–39. Vezzani, A. (2004) Anticonvulsant and antiepileptogenic
Mathe, A.A. (1999) Neuropeptides and electroconvulsive treat- effects mediated by adeno-associated virus vector neuropep-
ment. J. ECT, 15(1): 60–75. tide Y expression in the rat hippocampus. J. Neurosci.,
Marksteiner, J., Ortler, M., Bellmann, R. and Sperk, G. (1990) 24(12): 3051–3059.
Neuropeptide Y biosynthesis is markedly induced in mossy Scharfman, H.E., Goodman, J.H. and Sollas, A.L. (2000)
fibers during temporal lobe epilepsy of the rat. Neurosci. Granule-like neurons at the hilar/CA3 border after status
Lett., 112(2-3): 143–148. epilepticus and their synchrony with area CA3 pyramidal
McQuiston, A.R. and Colmers, W.F. (1996) Neuropeptide Y2 cells: functional implications of seizure-induced neurogenesis.
receptors inhibit the frequency of spontaneous but not min- J. Neurosci., 20(16): 6144–6158.
iature EPSCs in CA3 pyramidal cells of rat hippocampus. J. Schwarzer, C., Kofler, N. and Sperk, G. (1998) Up-regulation
Neurophysiol., 76(5): 3159–3168. of neuropeptide Y-Y2 in an animal model of temporal lobe
McQuiston, A.R., Petrozzino, J.J., Connor, J.A. and Colmers, epilepsy. Mol. Pharmacol., 53(1): 6–13.
W.F. (1996) Neuropeptide Y1 receptors inhibit N-type cal- Silva, A.P., Carvalho, A.P., Carvalho, C.M. and Malva, J.O.
cium currents and reduce transient calcium increases in rat (2001) Modulation of intracellular calcium changes and
dentate granule cells. J. Neurosci., 16(4): 1422–1429. glutamate release by neuropeptide Y1 and Y2 receptors in the
297

rat hippocampus: differential effects in CA1, CA3 and dent- Sun, Q.Q., Huguenard, J.R. and Prince, D.A. (2001b) Neuro-
ate gyrus. J. Neurochem., 79(2): 286–296. peptide Y receptors differentially modulate G-protein-acti-
Silva, A.P., Carvalho, A.P., Carvalho, C.M. and Malva, J.O. vated inwardly rectifying K+ channels and high-voltage-
(2003) Functional interaction between neuropeptide Y re- activated Ca2+ channels in rat thalamic neurons. J. Physiol.,
ceptors and modulation of calcium channels in the rat hip- 531(1): 67–79.
pocampus. Neuropharmacology, 44(2): 282–292. Sundler, F., Moghimzadeh, E., Hakanson, R., Ekelund, M.
Simmons, M.L., Terman, G.W., Gibbs, S.M. and Chavkin, C. and Emson, P. (1983) Nerve fibers in the gut and pancreas
(1995) L-type calcium channels mediate dynorphin neuro- of the rat displaying neuropeptide-Y immunoreactivity.
peptide release from dendrites but not axons of hippocampal Intrinsic and extrinsic origin. Cell Tissue Res., 230(3):
granule cells. Neuron, 14(6): 1265–1272. 487–493.
Sloviter, R.S. (1983) ‘‘Epileptic’’ brain damage in rats induced Tatemoto, K., Carlquist, M. and Mutt, V. (1982) Neuropeptide
by sustained electrical stimulation of the perforant path. I. Y—a novel brain peptide with structural similarities to pep-
Acute electrophysiological and light microscopic studies. tide YY and pancreatic polypeptide. Nature, 296(5858):
Brain Res. Bull., 10(5): 675–697. 659–660.
Smialowska, M. (1988) Neuropeptide Y immunoreactivity in Tu, B., Timofeeva, O., Jiao, Y. and Nadler, J.V. (2005) Spon-
the locus coeruleus of the rat brain. Neuroscience, 25(1): taneous release of neuropeptide Y tonically inhibits recurrent
123–131. mossy fiber synaptic transmission in epileptic brain. J. Ne-
Smith, Y. and Parent, A. (1986) Neuropeptide Y-immunoreac- urosci., 25(7): 1718–1729.
tive neurons in the striatum of cat and monkey: morpholog- Vezzani, A., Sperk, G. and Colmers, W.F. (1999) Neuropeptide
ical characteristics, intrinsic organization and co-localization Y: emerging evidence for a functional role in seizure mod-
with somatostatin. Brain Res., 372(2): 241–252. ulation. Trends Neurosci., 22(1): 25–30.
Smith, Y., Parent, A., Kerkerian, L. and Pelletier, G. (1985) Widerlov, E., Lindstrom, L.H., Wahlestedt, C. and Ekman, R.
Distribution of neuropeptide Y immunoreactivity in the ba- (1988) Neuropeptide Y and peptide YY as possible cerebro-
sal forebrain and upper brainstem of the squirrel monkey spinal fluid markers for major depression and schizophrenia,
(Saimiri sciureus). J. Comp. Neurol., 236(1): 71–89. respectively. J. Psychiatr. Res., 22(1): 69–79.
Sperk, G., Marksteiner, J., Gruber, B., Bellmann, R., Mahata, Whittaker, E., Vereker, E. and Lynch, M.A. (1999)
M. and Ortler, M. (1992) Functional changes in neuropeptide Neuropeptide Y inhibits glutamate release and long-term
Y- and somatostatin-containing neurons induced by limbic potentiation in rat dentate gyrus. Brain Res., 827(1-2):
seizures in the rat. Neuroscience, 50(4): 831–846. 229–233.
Stanic, D., Brumovsky, P., Fetissov, S., Shuster, S., Herzog, H. Wolak, M.L., DeJoseph, M.R., Cator, A.D., Mokashi, A.S.,
and Hokfelt, T. (2006) Characterization of neuropeptide Y2 Brownfield, M.S. and Urban, J.H. (2003) Comparative dis-
receptor protein expression in the mouse brain. I. Distribu- tribution of neuropeptide Y Y1 and Y5 receptors in the rat
tion in cell bodies and nerve terminals. J. Comp. Neurol., brain by using immunohistochemistry. J. Comp. Neurol.,
499(3): 357–390. 464(3): 285–311.
Sun, Q.Q., Akk, G., Huguenard, J.R. and Prince, D.A. (2001a) Woldbye, D.P., Larsen, P.J., Mikkelsen, J.D., Klemp, K.,
Differential regulation of GABA release and neuronal excit- Madsen, T.M. and Bolwig, T.G. (1997) Powerful inhibition
ability mediated by neuropeptide Y1 and Y2 receptors in rat of kainic acid seizures by neuropeptide Y via Y5-like recep-
thalamic neurons. J. Physiol., 531(1): 81–94. tors. Nat. Med., 3(7): 761–764.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 18

Norepinephrine and the dentate gyrus

Carolyn W. Harley

Department of Psychology, Memorial University of Newfoundland, St. John’s, NL, A1B 3X9, Canada

Abstract: Norepinephrine’s role in the dentate gyrus is assessed based on a review of what is known about
its innervation and receptor patterns and its functional effects at both cellular and behavioral levels. The
data support seven hypotheses: (1) Norepinephrine’s functional actions are primarily mediated by b ad-
renoceptors and include electrophysiological enhancement of responses to excitatory input and glycogen-
olytic metabolic support of excitatory synaptic activity. (2) At the cellular level, locus coeruleus burst
release of norepinephrine transiently inhibits feedforward interneurons and either excites or inhibits sub-
populations of feedback interneurons. Consistent with reduced feedforward inhibition, granule cell firing is
transiently increased. Concomitant EEG effects include transient increases in theta power and decreases in
beta and gamma power. (3) Norepinephrine selectively promotes the processing of medial perforant path
spatial input. This effect is mediated both through short- and long-term potentiation of cell excitability and
through delayed potentiation of synaptic input. A critical level of norepinephrine release is required for
long-term effects to norepinephrine alone. Norepinephrine release switches early phase frequency-induced
long-term potentiation of perforant path input to an enduring late phase form and can reinstate decayed
long-term potentiation. Norepinephrine also promotes frequency-induced potentiation of granule cell out-
put at the mossy fiber to CA3 connection. (4) Local increases in norepinephrine accompany glutamate
release and release of other neurotransmitters providing a mechanism for norepinephrine enhancement
effects independent of locus coeruleus firing. (5) Stimuli, such as novelty and reward and punishment,
which activate locus coeruleus neurons, enhance responses to medial perforant path input and engage late
phase frequency-induced long-term potentiation through b adrenoceptor activation. (6) Behavioral studies
are consistent with the mechanistic evidence for a norepinephrine role in promoting learning and memory
and assisting retrieval. (7) The overall profile suggests lower levels of norepinephrine may facilitate pattern
completion or memory retrieval while higher levels would recruit global remapping and promote long-term
episodic memory.

Keywords: locus coeruleus; LTP; LTD; novelty; glycogen metabolism; theta EEG; gamma EEG; global
remapping; feedback inhibition; feedforward inhibition; alpha adrenoceptors; beta adrenoceptors; medial
perforant path; lateral perforant path; synaptic plasticity

The present chapter examines norepinephrine DG cells, and reviews NE’s role in promoting
(NE) innervation and receptor patterns in the long-term plasticity, based primarily on rodent
dentate gyrus (DG), considers NE’s effects on data.

Corresponding author. Tel.: +1 709 737 7974;


Fax: +1 709 737 4000; E-mail: charley@mun.ca

DOI: 10.1016/S0079-6123(07)63018-0 299


300

NE innervation and granule cell somata, symmetric synapses and


appositions are typical (Milner and Bacon, 1989b).
Blackstad first characterized NE innervation of Biochemically, NE content in the hippocampus
the rat hippocampus, reporting the densest NE is highest in DG (about twice that in CA1) with
innervation in the hilus of the DG, particularly higher levels in ventral (600 ng/g) compared to
the subgranular zone (Blackstad et al., 1967). NE dorsal DG (360 ng/g) (Loy et al., 1980) or similar
fibers were less common in cell body layers, sug- values (500 ng/g) in both regions (Hortnagl et al.,
gesting axodendritic contacts. Within the molec- 1991) depending on the study. Quantification of
ular layer, the dendritic zone of DG granule cells, varicosities in DG (2.4 million/cubic mm) also sug-
NE fibers were denser in the middle molecular gest twice as many varicosities there as in CA1
layer, a target of medial perforant path fibers from (Oleskevich et al., 1989); however when divided by
the medial entorhinal cortex, as compared to the cell number, the ratio of NE varicosities to cells
outer molecular layer, the target of lateral ent- (20–40/1) is smaller in DG than CA1 (180/1). The
orhinal fibers. ratio is lowest in the granule cell layer itself (2–4/
With retrograde tracing, the origin of the DG 1), consistent with the low total number of vari-
NE innervation was identified as multipolar and cosities in this layer. Seventy percent of the input
fusiform cells in the dorsal part of the dorsal from the LC to DG courses through the cingulum,
pontine locus coeruleus (LC) (Haring and Davis, with the fornix and ventral amygdaloid pathways
1985). Electron microscope (EM) images of the each accounting for 15% (Loy et al., 1980).
NE fibers in DG reveal small granular vesicles, Developmentally, the NE input, as character-
indicating NE content in 0.4–1.2 mm diameter ized by dopamine-b-hydroxylase immunocyto-
varicosities. Each varicosity contains 20 vesi- chemistry, is very sparse in the DG at four days
cles, of which 55% can be classified as small postnatal, and is heaviest at this time point in
granular vesicles. Large granular vesicles (6%) CA1. By 10 days postnatal, a band of fibers in
and a few mitochondria are also present (Koda the subgranular zone becomes detectable (Moudy
and Bloom, 1977; Milner and Bacon, 1989b). In- et al., 1993). By 21 days, a mainly adult-like
tervaricosity axons are 0.1–0.15 mm in diameter pattern is developed.
(Milner and Bacon, 1989b). In 6600 mm2 samples,
the hilus has 1 NE bouton (varicosity) for every
400 boutons of other types; while the granule cell
layer has 1/500, and the molecular layer has 1/ Adrenoceptor distribution
4000 NE boutons. Total NE bouton density
differs among regions. An average of 1500 and Both a and b adrenoceptors occur in DG. In an
2000 boutons are located in hilar and molecular early study using several methods to assess ad-
layer samples, respectively, while only 500 bou- renoceptors, Crutcher and Davis (1980) reported b
tons are present in each granule cell layer sample adrenoreceptors at uniform levels in the dentate
(Koda and Bloom, 1977). These estimates of NE (146 fmol/mg) and hippocampal gyri (178 fmol/
innervation density may be compared to estimates mg), while a1 receptors were more concentrated in
of 1/8800–1/14,500 boutons for the neocortex the DG (269 fmol/mg). The uniformity of b ad-
(Lapierre et al., 1973). renoceptor distribution led the authors to suggest
The majority of NE axon contacts in DG end on these receptors might occur separately from LC
small dendrites in the subgranular layer where a terminals, while a1 adrenoreceptors showed better
third make symmetric synapses, a third form correspondence with LC innervation. Both recep-
asymmetric synapses, and a third type make close tors occurred most densely in the synaptosomal/
associations without specializations. Terminals in mitochondrial fractions. The a1 receptor popula-
the molecular layer have a similar typology. Thirty tion was not changed by 6-OHDA lesions (U’Pri-
percent of terminals end at spines and do not have chard et al., 1980) suggesting a postsynaptic
specialized synaptic profiles. On large dendrites localization.
301

b Adrenoceptors occur in spines. Axonal b adrenoceptors occur in


both axons and in axon terminals ending on spines
Milner’s more recent work (Milner et al., 2000) (Milner et al., 2000).
using both EM and light microscope methodol- The b adrenoreceptors on astrocytes are usu-
ogies, suggests NE adrenoceptor distribution is ally next to terminals that make asymmetric
best identified using EM. Her quantified EM synapses on dendrites and which are thought
receptor counts show laminar differences in b to be excitatory inputs. Many of the receptors
adrenoceptors that were not evident in light on astrocytes are closely apposed to the terminals
microscope studies. Approximately 500 b ad- that form synapses on spines (Milner et al.,
renoceptor-immunoreactive profiles per 6000 mm2 2000). The astrocytic adrenoceptor distribution,
sample occur in the middle molecular layer, while in close association with synapses, may support
only 250 profiles are seen in the inner and outer glycogen breakdown at active synaptic sites
molecular layers, and in the subgranular and hilar (see Physiology, below). Astrocytic b adrenocep-
zones; 125 b adrenoceptor-immunoreactive pro- tors also occur around blood vessels.
files/sample occur in the granule cell layer. The NE axons, identified by tryrosine hydroxylase
molecular layer distribution shows interesting immunoreactivity, occur close (1–2 mm) to both
parallels with b adrenoceptor physiological astrocytic and dendritic b adrenoceptors, but
effects (see Physiology, below) and is consistent direct contacts were not observed (Milner et al.,
with Blackstad’s initial description of NE fiber 2000). The b adrenoceptors in DG are well-
innervation. placed to modify granule cell function, and some
Granule cell somata express b adrenoceptor interneurons, either directly through postsynap-
immunoreactivity in the endoplasmic reticulum, tic receptors or indirectly, through effects on
consistent with granule cell production of b glial processes near synapses. The common
adrenoceptors. A few hilar interneurons, includ- postsynaptic location of b adrenoceptors on the
ing those reactive for parvalbumin, which marks dendrites of granule cells near, or with, asym-
basket cells (Kosaka et al., 1987; Ribak et al., metric synaptic input and their predominance in
1990, 1992), are also immunoreactive for b ad- the middle molecular layer, is consistent with a
renoceptors. selective modulation of responses to glut-
In the molecular layers, 40–50% of the b ad- amatergic input from the medial entorhinal cor-
renoceptor reactive profiles are associated with tex to the DG (see Physiology, below). Activation
dendrites, and 50–60% with astrocytes. In the of presynaptic axonal b adrenoceptors and,
granule cell layer, 30% are somatic, dendritic and possibly, astrocytic b adrenoreceptors, may play
astrocytic, respectively. In the subgranular region a role in the increase in glutamate release also
(an 55 mm zone below the granule cells), 65% reported with b-adrenoceptor activation (see
are astrocytic, while 25% are dendritic. Below Physiology, below).
the subgranular zone, in the hilar region, 30% of Molecular and binding studies suggest the
the b adrenoceptor reactive sites are in axons or majority of b adrenoceptors in the DG are b1
axon terminals and over 50% are in astrocytes. adrenoceptors (Minneman et al., 1979; Rainbow
Some b adrenoceptors are present within parval- et al., 1984), but both b1 and b2 adrenoceptors
bumin-immunoreactive terminals. Reactive sites occur in the molecular layer and hilus (Duncan
in axons are less common outside the hilar re- et al., 1991; Booze et al., 1993) and granule
gion, comprising only 2–10% of the b adrenocep- cells produce mRNA for both b1 and b2 ad-
tor-reactive sites in other layers (Milner et al., renoceptors with a stronger signal from the b2
2000). mRNA probe (Nicholas et al., 1993b). The hip-
Dendritic receptors are associated with postsy- pocampus has a high proportion of high-affinity
naptic densities on both large (inner molecular b adrenoceptors, consistent with a strong func-
layer) and small (middle and outer molecular lay- tional role for these receptors (Arango et al.,
ers) dendritic profiles. Immunoreactive sites also 1990).
302

a Adrenoceptors 10% on axons, 50% on dendrites and 40% on glia.


Throughout the molecular layer, a2a receptors on
Three a1 receptor subtypes are recognized. Specific dendrites are primarily on spines at asymmetric
probes for each (a1a, a1b, a1d) reveal that granule synapses. Astroctyic a2a receptors are also near
cells have an intense mRNA signal for the a1d asymmetric synapses. No direct synaptic contacts
subtype (Pieribone et al., 1994; Day et al., 1997), between the postsynaptic a2a receptors and tyro-
but do not contain the a1b subtype mRNA. The sine hyroxylase reactive axons are observed. The
a1a subtype is restricted to polymorph cells of the granule cell layer has the fewest a2a profiles (2/3rd
hilar region (Day et al., 1997). The a1 receptor is less than the molecular layer) and 50% are somatic
described as half as dense in hippocampus as in profiles. In the subgranular zone, there is a paucity
neocortex, despite the fact that the hippocampus of dendritic profiles, with mainly axonal (44%)
has twice the NE innervation density of neocortex and glial (46%) profiles. In the hilar region 60%
(Zilles et al., 1991). Zilles reports that a1 receptors are in axonal elements. The remaining a2a profiles
are more concentrated in the inner 2/3rd of the are in glia. The axonal profiles throughout are
molecular layer, but that overall a1 receptor den- predominantly in non-noradrenergic axons. A
sity is similar in molecular, granule and hilar zones small subpopulation of hilar interneurons, with
and does not correspond with NE innervation characteristics of the somatostatin interneuron
(both a1a and a1b receptors are represented at subtype, contained a2a receptors (Milner et al.,
similar levels in contrast to the mRNA data). 1998).
Zilles’ a1 pattern differs from earlier work using a In situ studies examining mRNA for the three
different a1-ligand in which a1-receptor density a2 receptor subtypes report more of a2c subtype in
closely followed NE fiber density (Jones et al., DG granule cells than a2a (Nicholas et al., 1993a;
1985). Since neither a2a nor b adrenoceptors are in Scheinin et al., 1994) and none of a2b subtype. A
clear synaptic association with NE terminals, a1 mouse study did find evidence of granule cell a2b
adrenoceptors are candidates, by default, for sites receptor expression, with the most signal in more
of synaptic specialization for tyrosine hydroxylase mature granule cells (located at the border with the
fibers. No a1 adrenoceptor subtype distribution inner molecular layer) and none in the subgranular
has been described at the EM level in DG. zone (Wang et al., 2002). Developmental studies
Milner (Milner et al., 1998) characterized the report that the a2a mRNA signal is moderate at
a2a receptor at the light microscope and EM levels 1–3 days postnatal and then becomes weak after
and describes its distribution as complementary to maturity (Winzer-Serhan et al., 1997a), while the
the b adrenoceptors in that the majority of recep- a2c signal intensity is high from 11 to 14 days
tors are presynaptic — in axons and axon termi- postnatal, and remains a moderately-intense signal
nals — rather than postsynaptic. Nonetheless, a2a through adulthood (Winzer-Serhan et al., 1997b).
receptors exist both pre- and postsynaptically, and It would be of interest to know the localization of
while a2a receptors are implicated in the negative a2c receptors.
feedback regulation of NE release, most presy-
naptic profiles are in unmyelinated non-nor-
adrenergic axons. Physiology
Granule cells demonstrate immunoreactivity for
a2a-like receptors in association with endoplasmic Glycogen metabolism
reticulum, suggesting granule cell synthesis of a2a
receptors. The a2a-immunoreactive sites are The astrocytic metabolic enzyme, glycogen phos-
evenly distributed in the middle and outer molec- phorylase, which breaks down glycogen to glu-
ular layers with about 40% on axons, 30% on as- cose, is strongly activated by NE in DG, as
trocytes and another 30% on dendrites or spines. demonstrated pharmacologically using an a2 ad-
There are about 1/3rd fewer immunoreactive a2a renoceptor antagonist (Fara-On et al., 2005), or
receptors in the inner molecular layer with only physiologically, using glutamate activation of the
303

LC (unpublished observations). In 1971, Phelps decay of excitatory postsynaptic currents (EPSCs)


found high levels of glycogen accumulation in as- and increases temporal summation in other hippo-
trocytic processes near synapses in the DG in bar- campal principal cells (Lancaster et al., 2001).
biturate-treated mice (Phelps, 1972). Earlier Gray and Johnston blocked K+ currents and
evidence that NE release depleted glycogen, while showed that b adrenoceptor activation increased
suppression of NE release increased glycogen, led a voltage-dependent Ca++ current, which is ac-
Phelps to hypothesize that NE controlled glycogen tivated when cells are depolarized to at least
breakdown in the DG at sites of synaptic activa- 15 mV, and is therefore likely to be L-type cur-
tion through b adrenoceptors. This hypothesis is rent (Gray and Johnston, 1987). They showed an
consistent with the common association of as- increase in Ca++ channel open times and a
trocytic b adrenoceptors and asymmetric synaptic dependence on cAMP elevation. No changes in
contacts. the N-type Ca++ current were reported. The a2
At the light microscope level, a patchy, modular agonist, clonidine, was without effect. NE also
distribution of glycogen phosphorylase, which co- produced an increase in the amplitude and dura-
exists with glycogen, is evident in the DG. Patches tion of Ca++-dependent action potentials, which
are more numerous during the dark phase of the contribute to back-propagating action potentials
daily cycle, when rodents are aroused, and pre- and spike-timing dependent plasticity (Dan and
dominate (60%) in the middle molecular layer, Poo, 2006). Thus, NE might be expected to en-
with the remaining patches equally divided be- hance this form of plasticity. Intracellular calcium
tween inner and outer layers (Harley and Rusak, is also increased in granule cells by dose-depend-
1993). Glycogen phosphorylase activity is reduced ent activation of a1 adrenoceptors (Kusaka et al.,
during the theta EEG state, possibly in relation to 2004).
a higher overall inhibition in the DG during theta Lacaille and Schwartzkroin used focal applica-
rhythm (Uecker et al., 1997). NE activation of tion of a b adrenoceptor agonist in hippocampal
glycogen phosphorylase and glycogenolysis in slices to evaluate the effects of NE. They replicated
vitro requires b adrenoceptor coupled cAMP ac- the depolarization of the granule cell membrane
tivation (Edwards et al., 1974; Nahorski et al., potential (Lacaille and Schwartzkroin, 1988). They
1975; Quach et al., 1978). NE also increases as- also showed that the depolarization was accom-
trocyte metabolism through a adrenoceptor acti- panied by an increased input resistance, which
vation (Subbarao and Hertz, 1991). they attributed to a blockade of K+ channels that
are open under normal conditions.
Collectively, one would predict that these b ad-
Intracellular recording renoceptor effects would enhance the excitability
and responsivity of granule cells to afferent input.
Granule cells No other adrenoceptor subtype has been impli-
cated in direct modulation of granule cell mem-
Three studies have examined NE effects in vitro. brane characteristics.
Haas and Rose found that NE and the b ad-
renoceptor agonist, isoproterenol, produced a small
depolarization of the membrane potential, an at- Hilar interneurons
tenuated afterhyperpolarization, and a decrease in
accommodation (Haas and Rose, 1987). The at- Misgeld and Bijak sampled hilar interneurons
tenuated afterhyperpolarization was attributed to a in the subgranular zone (Bijak and Misgeld, 1995).
cAMP-mediated block of Ca++-activated K+ cur- Aspiny interneurons, which are GABAergic and
rents. In a few granule cells, a b adrenoceptor in- inhibitory, have strong afterhyperpolarizations
itiated block of the transient outward K+ current and little spike accommodation. EM studies show
(A-type current) was reported. Blocking the slow they receive NE input (Milner and Bacon, 1989a).
Ca++-activated afterhyperpolarization slows the Spiny hilar neurons, which are glutamatergic and
304

excitatory (mossy cells), have strong accommoda- cells, but when paired with sweet milk it evoked
tion and little afterhyperpolarization. NE blocked excitation; LC stimulation enhanced the cell exci-
both aspiny neuron afterhyperpolarizations and tation as well as inhibition (Segal and Bloom,
spiny neuron accommodation through b ad- 1976b). This suggests a circuit-dependent LC mod-
renoceptors. NE, or a b adrenoceptor agonist, ulation of interneuron responses.
increased the firing frequency of the excitatory The inhibition appeared to be a direct effect of
spiny neurons to current injection, but did not NE release, since LC cells transplanted to a den-
alter the response of the aspiny inhibitory inter- ervated hippocampus also inhibited spontaneously
neurons to current injection. Spontaneous firing of active units in the DG. A brief stimulus train to the
both interneuron types increased with NE or b LC transplant inhibited DG cell firing for 30 s
adrenoceptor activation, as did the frequency of (Bjorklund et al., 1979). This inhibition was an-
excitatory postsynaptic potentials (EPSPs) and tagonized by a b adrenoceptor antagonist.
inhibitory postsynaptic potentials (IPSPs). In two pharmacological studies, theta inter-
IPSPs increased in neighboring granule cells neurons (putative inhibitory interneurons) in the
with increased activity in inhibitory interneurons. hilus and near the granule cell layer were excited
With glutamate release blocked, GABA-A recep- by NE and by a2 and b agonists, while granule
tor-mediated IPSPs still increased in granule cells, cells were inhibited by NE and by a1 agonists, but
but NE no longer increased inhibitory interneuron excited by a2 and b agonists (Pang and Rose,
firing. a adrenoceptor agonists suppressed 1987; Rose and Pang, 1989). The excitatory
inhibitory interneuron activity. Thus, NE activa- effects of a2 receptors appeared to be postsynap-
tion of both b adrenoceptors and a adrenoceptors tic, since they occurred in the presence of high
either did not change or else suppressed, inhibitory concentrations of magnesium in the extracellular
interneurons. The enhancement of spontaneous buffer.
IPSPs, without increased inhibitory interneuron A more recent study using glutamatergic acti-
firing, suggests presynaptic b adrenoceptors vation of LC neurons demonstrates both inhibi-
enhance GABA release. Presynaptic b adrenocep- tory and excitatory effects on subgranular and
tors also appear to enhance glutamate release. The hilar interneurons with natural NE release (Brown
outcome of natural NE release on inhibitory et al., 2005). Using NE release from LC activation
interneurons, and network inhibition, will reflect (natural NE release) has the advantage of reveal-
the balance of a and b adrenoceptor activation on ing in situ actions of NE rather than pharmaco-
all elements of the circuit. Extracellular unit logical actions, but non-NE effects may also be
recording (see Extracellular unit recording, below) recruited. The nature of the LC-NE effect de-
shows that the effects of NE naturally released pended on the physiological identity of the inter-
include decreases and increases in inhibitory neurons. Feedforward interneurons, activated by
interneuron firing, depending on the subtype of perforant path input from the entorhinal cortex
the interneuron. and firing with a lower threshold than granule
cells, were consistently inhibited (60 s) by brief
LC activation (see Fig. 1). Feedback interneurons,
Extracellular unit recording recruited only after granule cells discharged, were
either excited or inhibited, suggesting differences
Segal and Bloom were the first to record changes between subpopulations. Consistent with inhibi-
in unit activity in DG in response to electrical tion of feedforward interneurons was the observa-
stimulation of the LC (Segal and Bloom, 1976a). tion of a strongly increased granule cell response
The cells they monitored had relatively high firing after LC activation. These results reveal that nat-
rates and therefore were likely to be interneurons ural release of NE produces selective modulation
(Jung and McNaughton, 1993). LC stimulation of DG circuitry. More work is needed to classify
inhibited this cell type in anesthetized rats. In the feedback interneurons that respond differen-
awake rats, a loud tone also inhibited these DG tially to NE release.
305

Fig. 1. Normalized firing rate of 16 feedforward interneurons in the dentate gyrus following ejection of glutamate in the locus
coeruleus demonstrated the robust effects on this cell type. Glutamate was ejected at 0 min. Adapted with permission from The Society
for Neuroscience, 2005.

EEG recording A pattern of enhanced theta and suppressed


beta/gamma activity is consistent with the hypoth-
In the same study (Brown et al., 2005), LC acti- esis that disengagement from established represen-
vation enhanced DG EEG theta in the 4–8 Hz tations and enhancement of processes that
range while suppressing beta (12–20 Hz) and promote incorporation of new information is an
gamma (20–40 Hz) frequencies (see Fig. 2). This effect of LC activation (Brown et al., 2005). Con-
result appears consistent with concomitant inhibi- sistent with this view, Bouret and Sara have pro-
tion of the feedforward interneuron population, posed ‘network resetting’ as a primary LC
which potentially mediates the more distant influ- function (Bouret and Sara, 2005).
ences implicated in beta oscillations (Kopell et al.,
2000), and excitation of a feedback subpopulation
enhancing theta amplitude as reported for 5-HT Evoked potential recording
modulation (Nitz and McNaughton, 1999).
Gray originally proposed that LC activation in- NE potentiates the perforant path-evoked popu-
creases hippocampal theta power specifically in a lation spike both transiently, and in a sustained
7.5–7.8 Hz range in awake rats (Gray and Ball, manner, depending on the degree of NE exposure
1970). He observed this increase when rats were in the DG. Transient potentiation of the perforant
responding to novelty or a disconfirmation of ex- path-evoked population spike amplitude, which is
pectation. The low driving threshold for this fre- b adrenoceptor-dependent, has been reported in
quency range was lost following NE depletion or numerous in vivo experiments after either ex-
blockade (Gray et al., 1975). We have observed ogenous NE application (Neuman and Harley,
that this theta frequency (7.5–7.8 Hz) increases 1983; Winson and Dahl, 1985; Babstock and
with LC activation in awake rats (unpublished Harley, 1993; Harley et al., 1996; Chaulk and
observations). Harley, 1998) or activation of LC (Dahl and
Segal reported that electrical LC stimulation in- Winson, 1985; Harley and Milway, 1986; Harley et
duces theta burst firing in medial septal neurons al., 1989; Washburn and Moises, 1989; Babstock
(Segal, 1976). Berridge and Foote later showed and Harley, 1992; Frizzell and Harley, 1994;
that tonic cholinergic LC activation produces a Klukowski and Harley, 1994). NE or b adrenocep-
dramatic increase in hippocampal theta in anest- tor agonists in vitro increase the perforant path-
hetized rats (Berridge and Foote, 1991), which de- evoked EPSP slope, as well as the population
pends on b adrenoceptors in medial septum spike, although enhanced E-S coupling is also ob-
(Berridge et al., 1996). served (Lacaille and Harley, 1985). The duration
306

Fig. 2. Frequency-dependent increases and decreases in dentate gyrus EEG in the 4–40 Hz range following ejection of glutamate in the
locus coeruleus. Asterisks indicate significant effects at po0.05. Upward arrow denotes time of ejection. Adapted with permission from
The Society for Neuroscience, 2005.

of in vitro effects are transient when agonist con- NE-induced plasticity


centrations are low or briefly applied (Lacaille and
Harley, 1985; Stanton and Sarvey, 1985c). Spike potentiation in anesthetized rats
The discovery of sustained NE-induced in-
creases in the perforant path-evoked potential or While studies at the cellular and EEG level suggest
NE-induced long-term potentiation (NE-LTP) transient LC-NE modulation of the DG network,
(Neuman and Harley, 1983) was the first direct studies of the perforant path-evoked potential, as
evidence that LC-NE could provide a heterosy- noted, reveal a sustained component of LC-NE
naptic signal to initiate long-lasting increases in modulation of the DG response to entorhinal in-
DG responses to glutamate-mediated information. put. Iontophoresed NE was first reported in 1983
NE-LTP is consistent with Kety’s early hypothe- to produce a long-lasting potentiation of the per-
sis, based on pharmacological and behavioral ev- forant path-evoked DG population spike in anest-
idence, that LC-NE strengthens brain responses to hetized rats that endured for hours (Neuman and
significant stimuli (Harley, 1987), and NE-LTP Harley, 1983). NE-induced LTP of the population
will be examined in detail in the next section. spike has been repeatedly confirmed in vivo using
307

direct NE application (Winson and Dahl, 1985; Sarvey, 1985c), strongly suggesting an associative
Harley et al., 1996; Chaulk and Harley, 1998), LC component to NE-LTP. However in vitro studies
activation by glutamate (Harley and Milway, failed to find evidence that pairing electrical
1986; Harley and Sara, 1992; Klukowski and stimulation of perforant path input with NE bath-
Harley, 1994; Walling and Harley, 2004) and LC application was critical for NE-LTP (Lacaille and
activation by orexin (Walling et al., 2004). Re- Harley, 1985; Dahl and Sarvey, 1990) (but see Spike
peated LC electrical stimulation can also induce potentiation and pairing requirements, below).
LTP of the perforant path-evoked potential Frizzell and Harley found an NMDA receptor-
(Harley et al., 1989). LC electrical stimulation- independent potentiation of EPSP slope and pop-
induced potentiation is controversial, however, ulation spike using LC-NE activation and keta-
with respect to dependence on b adrenoceptor mine anesthesia in vivo, but potentiation was
activation. There is evidence both for (Washburn typically short-lived (Frizzell and Harley, 1994). In
and Moises, 1989) and against (Harley et al., 1989) vitro, Dahl reported that two applications of a low
such dependence. DG evoked response potentiat- b adrenoceptor agonist dose (75 nM), spaced by
ion associated with glutamate- or orexin-induced 30 min, which, individually, elicit only weak tran-
activation of LC neurons is consistently blocked sient potentiation, induce NE-LTP of the perfo-
by b adrenoceptor antagonists, either systemically- rant path population spike only, and NMDA
administered (Harley and Milway, 1986; Harley et receptor activation was not required (Dahl and Li,
al., 1989; Babstock and Harley, 1992; Walling and 1994). This effect of a spaced NE agonist suggests
Harley, 2004; Walling et al., 2004) or locally-de- a mechanism by which LC-NE might contribute to
livered to the DG (Harley and Evans, 1988). The a stronger memory with spaced training.
requirement for DG b adrenoceptor activation is
consistent with the intracellular evidence that b
adrenoceptor activation increases the excitability Pathway selectivity
of granule cells to exogenous input. NE-LTP does
not depend, however, on continued exposure to The in vitro NE-LTP of EPSP slope occurred
NE, as demonstrated by the efficacy of brief ion- only for medial perforant path input, while the
tophoretic applications (Neuman and Harley, same application of NE with an a adrenoceptor
1983; Winson and Dahl, 1985) and by measure- antagonist, or of a b adrenoceptor agonist alone,
ment of hippocampal NE during NE-LTP induced produced a long-term depression (LTD) of the
by exogenous NE application (Harley et al., 1996) lateral perforant path EPSP slope (Dahl and
or LC activation (Walling et al., 2004). (For fur- Sarvey, 1989). It may be relevant for this selective
ther discussion, see LC firing and NE release pat- effect that b adrenoceptors are densest in the ter-
terns, below). minal zone of the medial perforant path (Milner
et al., 2000) and that enhancement of back-prop-
agating calcium spikes by NE would be more
Synaptic and spike potentiation in vitro likely to induce LTP in the adjacent medial per-
forant path target than the more distal lateral
In vitro studies of NE and perforant path stimu- perforant path target (Dan and Poo, 2006).
lation differ from in vivo studies because potent- A possible explanation for the failure to induce
iation of both synaptic and population spike NE-LTP of the field EPSP in vivo could be that
components of the perforant path-evoked poten- stimulation in vivo, particularly the stronger
tial occurred consistently in vitro but not in vivo stimulation that is required to evoke a DG pop-
(Lacaille and Harley, 1985; Stanton and Sarvey, ulation spike, would activate a mixture of medial
1987). In vitro NE-LTP depends on concomitant and lateral perforant path fibers. NE-LTP in
NMDA receptor activation (Burgard et al., 1989; vivo occurs for short-latency population spikes,
Sarvey et al., 1989), as well as on b1 adrenoceptor which reflect medial perforant path activation.
activation and protein synthesis (Stanton and In one attempt to examine the selectivity of the
308

NE effect in vivo, lateral perforant path input was granule cells increases. Although mixed medial
activated by stimulating the lateral olfactory and lateral perforant path input is a possible
tract. The synaptic input produced by lateral explanation of the lack of EPSP potentiation
olfactory stimulation was decreased by pairing observed in vivo, the ubiquity of studies that see
with paragigantocellularis stimulation, a major no change in EPSP slope suggests an increase in
glutamatergic afferent pathway to the LC E-S coupling, or granule cell excitability, is a
(Babstock and Harley, 1993). Depression of the prominent feature of LC-NE modulation. Recent
lateral perforant path input depended on b evidence provides support for the hypothesis
adrenoceptor activation. Since paragiganto- that an increase in intrinsic excitability, through
cellularis stimulation enhances the medial perfo- such mechanisms as reduced K+ conductance in
rant path population spike through a b activated dendrites (Frick et al., 2004), is an im-
adrenoceptor-dependent mechanism (Babstock portant component of associative plasticity func-
and Harley, 1992), the combined results are con- tioning similar to, and independent of, increases in
sistent with enhanced medial perforant path input synaptic drive per se (Daoudal and Debanne,
and depressed lateral perforant path input with 2003).
LC-NE activation in vivo. Evidence for associative plasticity in LC-NE ac-
One functional interpretation of selectivity is tivation effects on DG has been lacking. In vitro
suggested by a new perspective on the entorhinal- studies have not found that NE-LTP requires con-
hippocampal complex, which suggests two forms current pairing of bath-applied NE and perforant
of spatial mapping occur in hippocampus. Global path stimulation (Lacaille and Harley, 1985; Dahl
remapping, reflecting a completely new context, is and Sarvey, 1990). However, we have recently
associated with changes in medial perforant path found that pairing of LC glutamate activation and
‘grid cell’ input; while rate remapping, which perforant path input is critical for NE-LTP in vivo
reflects firing changes in the same place cell map, is (Reid and Harley, 2005). With LC activation
ascribed to alterations in nonspatial sensory 10 min after cessation of perforant path stimula-
input mediated by the lateral perforant path tion and 10 min prior to resumption of perforant
(McNaughton et al., 2006). If strong LC-NE ac- path stimulation, no NE-LTP occurs. The same
tivation potentiates medial, and depresses lateral, LC activation during concurrent perforant path
perforant path inputs, it would favor global re- stimulation produces reliable NE-LTP. Both EPSP
mapping in DG over rate remapping. Global re- slope and spike potentiation occurred in our in
mapping would provide a new context for memory vivo study (Reid and Harley, 2005). The impor-
storage. tance of timing in LC-NE activation for potent-
It should be noted that high frequency-induced iation effects is consistent with evidence that
LTP of either the medial or lateral perforant path conduction velocity in LC-NE axons varies across
requires b adrenoceptor activation (Bramham species to maintain a constant delay-to-signal
et al., 1997). Thus, while b adrenoceptors may arrival in forebrain structures (Aston-Jones et al.,
gate weaker inputs to favor the medial perforant 1985).
path, when signals are strong, plasticity should be Two differences may be noted between the in
promoted by b adrenoceptor activation in both vitro and in vivo pairing experiments. First, there
pathways similarly. was no delay between the offset of perforant path
stimulation and the beginning of NE perfusion in
the in vitro studies. Second, in vitro bath applica-
Spike potentiation and pairing requirements tion of NE is associated with an extended elevation
of cAMP levels in DG, permitting perforant path
In vivo experiments provide evidence that LC-NE interaction with cAMP effects even at late time
activation induces long-lasting increases in the points (Stanton and Sarvey, 1985b). This is un-
medial perforant path-evoked population spike, likely to be the case with LC activation (Siggins
suggesting that excitability in the postsynaptic et al., 1973).
309

Delayed long-term synaptic potentiation in awake measured by microdialysis in the DG is only asso-
rats ciated with NE-LTP in the DG when the NE in-
crease exceeds a critical threshold, estimated to be
In contrast to the evidence that NE induces rapid 400 nM, intrasynaptically (Harley et al., 1996).
potentiation of perforant path synaptic input from
in vitro studies is new in vivo data from awake rats
in which potentiation of perforant path synaptic LC firing and NE release patterns
input is first observed 24 h after LC activation
(Walling and Harley, 2004). In the Walling and LC neurons fire tonically at rates from less than 1
Harley study, LC burst activation initially had no to 5 Hz (Aston-Jones and Bloom, 1981b) as a
effect on synaptic input, but induced the usual function of level of arousal. They also fire phasic-
rapid and enduring potentiation of the perforant ally in bursts in which 2 or 3 spikes occur in rapid
path-evoked population spike. Twenty-four hours succession (10–15 Hz) followed by 200–500 ms
later, both the perforant path field EPSP and pauses (Aston-Jones and Bloom, 1981a). Burst
the population spike were strongly potentiated events are typically associated with responses to
(Fig. 3). The 24 h increase in field EPSP slope pre- environmental stimuli (Aston-Jones and Bloom,
dicted the 24 h increase in population spike am- 1981a). Harley and Sara have shown that gluta-
plitude. This delayed NE-LTP of EPSP slope and mate activation of LC produces a strong burst
population spike amplitude depended on b ad- followed by a pause, lasting minutes, in LC firing
renoceptor activation at the time of LC activation, (Harley and Sara, 1992). Higher levels of NE re-
and on protein synthesis. lease are associated with phasic rather than tonic
Aplysia also demonstrates a 24 h delayed potent- firing patterns (Florin-Lechner et al., 1996).
iation of synaptic input in response to heterosy- Glutamate administration to the LC in vivo in-
naptic activation of a cAMP-dependent cascade creases NE release in hippocampus by 200%,
(Brunelli et al., 1976; Schacher et al., 1988). Wall- measured by microdialysis, in the first 20 min after
ing and Harley suggest mammalian delayed synap- LC activation, subsequently, NE levels return to
tic potentiation implies a special role for DG NE baseline (Walling et al., 2004). Orexin A is also a
modulation in long-term memory. Such an associ- potent activator of LC neurons, and its adminis-
ation is likely to require relatively strong NE tration to the LC led to a similar increase in NE,
release. NE applied intracerebroventricularly and restricted to the first 20 min (Walling et al., 2004).

Fig. 3. Delayed potentiation of the perforant path-evoked field EPSP with immediate and delayed potentiation of the perforant path-
evoked population spike in awake rats following infusion of glutamate in the locus coeruleus at the arrow. Adapted with permission
from The Society for Neuroscience, 2005.
310

Both orexin A and glutamate activation of LC presence of normal magnesium levels and without
produce a b adrenoceptor-dependent increase in membrane depolarization (Pittaluga et al., 2000,
population spike amplitude that lasts for hours 2005; Risso et al., 2004). This provides a mechanism
and is thus triggered, but not maintained by, hip- for local NE release in the absence of direct termi-
pocampal NE release. High frequency electrical nal activation. Finally, cognitive enhancers signifi-
stimulation of the perforant path also increases cantly facilitate NE release in response to
NE release in the DG (Bronzino et al., 2001), and glutamate, even in the presence of the endogenous
may contribute to the level and duration of LTP NMDA receptor antagonist, kynurenate (Pittaluga
induced (Bronzino et al., 1999). et al., 1997). This facilitation is suggested to be an
Voltammetry provides better temporal resolu- important mediator of cognitive enhancement
tion of NE release. Electrical stimulation of the LC (Desai et al., 1995; Pittaluga et al., 2001).
for 2 s at 50 Hz induces a 10 fold increase in ox- Conversely, NE enhances glutamate release in
idation signal in the DG (maximal extracellular DG, an effect which depends on presynaptic b
signal 0.18 mM NE), appearing immediately after adrenoceptors (Lynch and Bliss, 1986). NE can
stimulation and returning to baseline within also enhance GABA release in hippocampus
10–15 s (Yavich et al., 2005). Repeated stimula- through a2 adrenoceptor activation (Pittaluga
tion at 5 min intervals yields stable NE signals, and Raiteri, 1987; Maura et al., 1988). Reversible
while stimulation at 30 s intervals shows decreas- EPSP slope potentiation of perforant path input
ing responses, suggesting a slow rate of vesicle re- to the DG has been related to b adrenoceptor-
filling. Natural stimuli that activate LC (tail pinch mediated phosphorylation of the presynaptic pro-
or vibrissae stimulation) do not produce a visible teins synapsins I and II (Parfitt et al., 1991, 1992).
NE signal without enhancement of NE release us- The positive feedback between NE and gluta-
ing the a2 receptor-antagonist, idazoxan. mate release further suggests local increases in NE
Glutamate activation of LC produces an aver- could occur independently of LC activation. The
age increase of 0.1 mM in hippocampal NE be- release of NE by glutamate and other neurotrans-
ginning 30 s after glutamate infusion, peaking at mitters may explain extended limbic NE elevation
1.5 min and returning to baseline by 7 min in ur- when rats receive shock in a novel context (McIn-
ethane-anesthetized rats (Palamarchouk et al., tyre et al., 2002). Suggestively, memory for the
2000). NE patterns in awake rats with glutamate context correlates positively with NE levels at the
infusion were similar but peaked at 5 min and re- time of acquisition. LTP of perforant path fibers
turned to baseline within 20 min (Palamarchouk also evokes longer lasting increases in NE release
et al., 2002). than LC activation alone (Bronzino et al., 1999,
2001).

NE release modulation by glutamate and vice-versa:


local effects LC-NE modulation of DG in relation to
environmental events
Both NMDA and AMPA receptors are present
presynaptically on NE axon terminals in hippo- Novel objects investigated by a rat using a board
campus, and when activated by glutamate or gluta- with objects contained in holes (holeboard) elicit
mate agonists, induce an increase in NE release LC bursts (Vankov et al., 1995). Perforant path-
(Pittaluga and Raiteri, 1992; Raiteri et al., 1992). evoked population spikes are potentiated for
NMDA-induced NE increases are greatest in DG 1 min after a rat places its nose in a hole in re-
microslices (Andres et al., 1993). Nictotine and sponse to a novel object (nose poke). Spike po-
somatostatin receptors are also present on NE ter- tentiation depends on b adrenoceptor activation
minals in hippocampus and induce NE release (Kitchigina et al., 1997); as does the enhanced ex-
alone (nicotine) or in concert with NMDA receptor ploration of novelty (Sara et al., 1995). Prolonged
stimulation (nicotine and somtatostatin) in the spike enhancement occurs upon first exposure to
311

the novel holeboard environment, possibly be- Restoration of recently-decayed LTP in the DG
cause the novel environment elicits a larger LC occurs following electrical stimulation of the LC
response (Kitchigina et al., 1997). These effects (Ezrokhi et al., 1999). This phenomenon may re-
appear to model the attentional (Berridge and late to other evidence for a role of NE in memory
Waterhouse, 2003), rather than the memory, retrieval (Sara and Devauges, 1989; Devauges and
effects of NE. Sara, 1991; Murchison et al., 2004). NE’s promo-
Novel environment exposure can promote mem- tion of theta EEG and the reported dependence of
ory-like effects when a weak, high-frequency LTP engram spread on b adrenoceptor activation could
stimulus is also presented (Straube et al., 2003). also contribute to retrieval processes (Flexner et
Placement in a novel environment converts early- al., 1985; Givens, 1996). Aston-Jones and Cohen’s
phase LTP to late-phase LTP in the DG. This proposal that phasic LC-NE optimizes decision-
conversion depends on b adrenoceptor activation, driven behavior and decision-driven memory rep-
suggesting LC-NE activation by novelty induces resentations provides further theoretical support
an enhancement of the weak LTP stimulus. NE in for an NE retrieval function (Aston-Jones and
the ventricles mediates a similar conversion of Cohen, 2005).
early to late-LTP in the DG (Almaguer-Melian et Exploratory behavior is increased with DG NE
al., 2005). b adrenoceptor activation is critical for infusions (Flicker and Geyer, 1982) and this effect
late DG LTP both in vitro (Stanton and Sarvey, is blocked by a b adrenoceptor antagonist. A tonic
1985a) and in vivo (Straube and Frey, 2003). Only LC-NE driven modulation of diversive explora-
the use of strong LTP protocols (for example, 15 tion (Usher et al., 1999; Mansour et al., 2003) is
0.25 ms pulses at 200 Hz every 10 s for 200 s) with consistent with lesion evidence suggesting the DG
repeated 15 pulse protocols twice within 5 min can has an important role in encoding spatial novelty
produce LTP in vivo that is independent of b ad- (Lee et al., 2005; Jerman et al., 2006).
renoceptor activation (Straube and Frey, 2003). In Vasopressin in the DG facilitates memory for
addition to facilitation by NE of enduring LTP in passive avoidance and increases NE turnover in
the entorhinal cortex-to-granule cell perforant DG (Kovacs et al., 1979b). Both consolidation and
pathway, there is LTP facilitation in the granule retrieval are enhanced, and these effects require
cell-to-CA3 pathway when subthreshold LTP intact LC-NE innervation (Kovacs et al., 1979a).
stimuli are combined with b adrenoceptor activa- Vasopressin in dorsal hippocampus also facilitates
tion (Hopkins and Johnston, 1984). Thus, weak memory for spatial learning (Paban et al., 2003).
patterns for LTP recruit enduring LTP in both the NE dependence of this effect remains to be as-
input and output components of DG circuitry sessed.
when NE is elevated. Corticotrophin releasing factor infused in the
Reward also promotes the conversion of early DG also enhances passive avoidance memory,
phase perforant path-LTP to late phase perforant while loss of local NE innervation, or blockade of
path-LTP (Seidenbecher et al., 1997), and is known local b adrenoceptors or NMDA receptors, pre-
to activate LC neurons (Sara and Segal, 1991). vents this memory enhancement (Lee et al., 1993).
Punishment has similar effects on activation of LC Lee et al. suggested that b adrenoceptors promote
(Sara and Segal, 1991) and the conversion of early NMDA-mediated synaptic plasticity to support
to late DG LTP. Conversions of early to late LTP passive avoidance memory. A dependence of NE-
by reward and punishment depend b adrenoceptor induced plasticity on NMDA receptors was also
activation (Seidenbecher et al., 1997). These results described in some in vitro studies (Sarvey et al.,
suggest reward- or punishment-related LC activa- 1989).
tion ‘reinforces’ plasticity in the DG. LC-NE may As we come to better understand the behavioral
also contribute to long-lasting circuit changes role of DG, new probes of NE’s function in the
through the promotion of the survival of new neu- DG should emerge. Aston-Jones and Cohen have
rons in DG (Rizk et al., 2006) and/or the promo- proposed, based on data from primates, that pha-
tion of neurogenesis (Kulkarni et al., 2002). sic LC-NE activation facilitates decision-driven
312

responses or memories, while tonic LC-NE activity NE release include reduction of feedforward inhi-
facilitates exploration (Aston-Jones and Cohen, bition and increases in granule cell excitability,
2005). The latter hypothesis is supported by in- such that granule cell responses to entorhinal input
creased exploration following NE infusion in DG increase. NE’s modulation of feedback interneu-
(Flicker and Geyer, 1982), but the relation of pha- rons enhances theta to facilitate throughput and
sic activation to decision-driven behavior and/or associative change in synaptic strength depending
memory is unclear with respect to DG activity. on theta phase (Pavlides et al., 1988; Orr et al.,
Bouret and Sara’s more general description 2001). NE enhances and depresses inputs in a
(Bouret and Sara, 2005) of the function of the pathway-dependent manner and recruits glial met-
LC-NE system as ‘network resetting’ is broadly abolic support for neuronal activation. Strong NE
consistent with NE-associated changes in EEG input engages processes that recruit long-term
and with the promotion of long-term synaptic modifications of excitability and/or delayed in-
plasticity in DG input and output. creases in synaptic strength in the DG pathway.
The promotion of frequency-induced long-term Together these effects support both attentional
potentiation (Dragoi et al., 2003) by NE reviewed and memory roles for DG NE.
above and the selective enhancement of the spatial
map matrix input from the medial entorhinal cor-
tex (Hafting et al., 2005) described earlier (see Acknowledgment
Pathway selectivity) lead to the prediction that, for
the DG, strong activation of NE input would pro- The author wishes to express appreciation for the
mote global remapping of spatial context. Weaker support of Canada’s NSERC Granting Council
activation of NE input and an associated transient via Grant A9791.
reduction in network inhibition (see Extracellular
unit recording, above) might enhance input gener-
alization, or pattern completion to assist retrieval, References
but strong activation would alter the DG map.
Global remapping provides a new framework for Almaguer-Melian, W., Rojas-Reyes, Y., Alvare, A., Rosillo,
the encoding of associative learning, reducing in- J.C., Frey, J.U. and Bergado, J.A. (2005) Long-term potent-
terference among associations and increasing iation in the dentate gyrus in freely moving rats is reinforced
memory capacity. Global remapping is assumed by intraventricular application of norepinephrine, but not
oxotremorine. Neurobiol. Learn. Mem., 83: 72–78.
to underlie episodic memory. The prediction that Andres, M.E., Bustos, G. and Gysling, K. (1993) Regulation of
strong LC-NE activation recruits global remap- [3H]norepinephrine release by N-methyl-D-aspartate recep-
ping in DG is consistent with the hypothesis that tors in minislices from the dentate gyrus and the CA1-CA3
activation of the LC and the related sympathetic area of the rat hippocampus. Biochem. Pharmacol., 46:
1983–1987.
system is part of a response to strongly significant
Arango, V., Ernsberger, P., Reis, D.J. and Mann, J.J. (1990)
events that promotes long-term memories for sig- Demonstration of high- and low-affinity beta-adrenergic re-
nificant events (Cahill et al., 1994). Global remap- ceptors in slide-mounted sections of rat and human brain.
ping with novelty exposure has been shown for Brain Res., 516: 113–121.
CA3 cells (Leutgeb et al., 2006), which like DG Aston-Jones, G. and Bloom, F.E. (1981a) Norepinephrine-con-
(Chawla et al., 2005; Rolls and Kesner, 2006), are taining locus coeruleus neurons in behaving rats exhibit pro-
nounced responses to non-noxious environmental stimuli. J.
implicated in sparse encoding and orthogonal en- Neurosci., 1: 887–900.
vironmental representations. The occurrence of Aston-Jones, G. and Bloom, F.E. (1981b) Activity of nor-
global remapping in DG concomitant with strong epinephrine-containing locus coeruleus neurons in behaving
LC activation remains to be demonstrated, but is rats anticipates fluctuations in the sleep- waking cycle. J.
Neurosci., 1: 876–886.
suggested by the data reviewed here.
Aston-Jones, G. and Cohen, J.D. (2005) An integrative theory
In summary, evoked potential and cellular data of locus coeruleus-norepinephrine function: adaptive gain
support various effects of NE on information and optimal performance. Annu. Rev. Neurosci., 28:
processing in the DG. Transient effects of phasic 403–450.
313

Aston-Jones, G., Foote, S.L. and Segal, M. (1985) Impulse Brunelli, M., Castellucci, V. and Kandel, E.R. (1976) Synaptic
conduction properties of noradrenergic locus coeruleus axons facilitation and behavioral sensitization in Aplysia: possible
projecting to monkey cerebrocortex. Neuroscience, 15: role of serotonin and cyclic AMP. Science, 194: 1178–1181.
765–777. Burgard, E.C., Decker, G. and Sarvey, J.M. (1989) NMDA
Babstock, D.M. and Harley, C.W. (1992) Paragigantocellularis receptor antagonists block norepinephrine-induced long-last-
stimulation induces beta-adrenergic hippocampal potentiat- ing potentiation and long-term potentiation in rat dentate
ion. Brain Res. Bull., 28: 709–714. gyrus. Brain Res., 482: 351–355.
Babstock, D.M. and Harley, C.W. (1993) Lateral olfactory Cahill, L., Prins, B., Weber, M. and McGaugh, J.L. (1994)
tract input to dentate gyrus is depressed by prior nor- Beta-adrenergic activation and memory for emotional events.
adrenergic activation using nucleus paragigantocellularis Nature, 371: 702–704.
stimulation. Brain Res., 629: 149–154. Chaulk, P.C. and Harley, C.W. (1998) Intracerebroventricular
Berridge, C.W., Bolen, S.J., Manley, M.S. and Foote, S.L. norepinephrine potentiation of the perforant path-evoked
(1996) Modulation of forebrain electroencephalographic potential in dentate gyrus of anesthetized and awake rats: a
activity in halothane-anesthetized rat via actions of nor- role for both alpha- and beta-adrenoceptor activation. Brain
adrenergic beta-receptors within the medial septal region. J. Res., 787: 59–70.
Neurosci., 16: 7010–7020. Chawla, M.K., Guzowski, J.F., Ramirez-Amaya, V., Lipa,
Berridge, C.W. and Foote, S.L. (1991) Effects of locus co- P., Hoffman, K.L., Marriott, L.K., Worley, P.F., McNaugh-
eruleus activation on electroencephalographic activity in ton, B.L. and Barnes, C.A. (2005) Sparse, environmentally
neocortex and hippocampus. J. Neurosci., 11: 3135–3145. selective expression of Arc RNA in the upper blade of the
Berridge, C.W. and Waterhouse, B.D. (2003) The locus co- rodent fascia dentata by brief spatial experience. Hippocam-
eruleus-noradrenergic system: modulation of behavioral state pus, 15: 579–586.
and state-dependent cognitive processes. Brain Res. Brain Crutcher, K.A. and Davis, J.N. (1980) Hippocampal alpha- and
Res. Rev., 42: 33–84. beta-adrenergic receptors: comparison of [3H]dihydroalpre-
Bijak, M. and Misgeld, U. (1995) Adrenergic modulation of nolol and [3H]WB 4101 binding with noradrenergic innerva-
hilar neuron activity and granule cell inhibition in the tion in the rat. Brain Res., 182: 107–117.
guinea-pig hippocampal slice. Neuroscience, 67: 541–550. Dahl, D. and Li, J. (1994) Induction of long-lasting potentiat-
Bjorklund, A., Segal, M. and Stenevi, U. (1979) Functional ion by sequenced applications of isoproterenol. Neuroreport,
reinnervation of rat hippocampus by locus coeruleus im- 5: 657–660.
plants. Brain Res., 170: 409–426. Dahl, D. and Sarvey, J.M. (1989) Norepinephrine induces
Blackstad, T.W., Fuxe, K. and Hokfelt, T. (1967) Noradren- pathway-specific long-lasting potentiation and depression in
aline nerve terminals in the hippocampal region of the rat the hippocampal dentate gyrus. Proc. Natl. Acad. Sci.
and the guinea pig. Z. Zellforsch. Mikrosk. Anat., 78: U.S.A., 86: 4776–4780.
463–473. Dahl, D. and Sarvey, J.M. (1990) Beta-adrenergic agonist-in-
Booze, R.M., Crisostomo, E.A. and Davis, J.N. (1993) Beta- duced long-lasting synaptic modifications in hippocampal
adrenergic receptors in the hippocampal and retrohippocam- dentate gyrus require activation of NMDA receptors, but
pal regions of rats and guinea pigs: autoradiographic and not electrical activation of afferents. Brain Res., 526:
immunohistochemical studies. Synapse, 13: 206–214. 347–350.
Bouret, S. and Sara, S.J. (2005) Network reset: a simplified Dahl, D. and Winson, J. (1985) Action of norepinephrine in the
overarching theory of locus coeruleus noradrenaline func- dentate gyrus. I. Stimulation of locus coeruleus. Exp. Brain
tion. Trends Neurosci., 28: 574–582. Res., 59: 491–496.
Bramham, C.R., Bacher-Svendsen, K. and Sarvey, J.M. (1997) Dan, Y. and Poo, M.M. (2006) Spike timing-dependent plas-
LTP in the lateral perforant path is beta-adrenergic receptor- ticity: from synapse to perception. Physiol. Rev., 86:
dependent. Neuroreport, 8: 719–724. 1033–1048.
Bronzino, J.D., Kehoe, P., Hendriks, R., Vita, L., Golas, B., Daoudal, G. and Debanne, D. (2003) Long-term plasticity of
Vivona, C. and Morgane, P.J. (1999) Hippocampal neuro- intrinsic excitability: learning rules and mechanisms. Learn.
chemical and electrophysiological measures from freely Mem., 10: 456–465.
moving rats. Exp. Neurol., 155: 150–155. Day, H.E., Campeau, S., Watson Jr., S.J. and Akil, H. (1997)
Bronzino, J.D., Kehoe, P., Mallinson, K. and Fortin, D.A. Distribution of alpha 1a-, alpha 1b- and alpha 1d-adrenergic
(2001) Increased extracellular release of hippocampal NE is receptor mRNA in the rat brain and spinal cord. J. Chem.
associated with tetanization of the medial perforant pathway Neuroanat., 13: 115–139.
in the freely moving adult male rat. Hippocampus, 11: Desai, M.A., Valli, M.J., Monn, J.A. and Schoepp, D.D. (1995)
423–429. 1-BCP, a memory-enhancing agent, selectively potentiates
Brown, R.A.M., Walling, S.G., Milway, J.S. and Harley, C.W. AMPA-induced [3H]norepinephrine release in rat hippocam-
(2005) Locus ceruleus activation suppresses feedforward pal slices. Neuropharmacology, 34: 141–147.
interneurons and reduces beta-gamma electroencephalogram Devauges, V. and Sara, S.J. (1991) Memory retrieval enhance-
frequencies while it enhances theta frequencies in rat dentate ment by locus coeruleus stimulation: evidence for mediation
gyrus. J. Neurosci., 25: 1985–1991. by beta-receptors. Behav. Brain Res., 43: 93–97.
314

Dragoi, G., Harris, K.D. and Buzsaki, G. (2003) Place repre- Hafting, T., Fyhn, M., Molden, S., Moser, M.B. and Moser,
sentation within hippocampal networks is modified by long- E.I. (2005) Microstructure of a spatial map in the entorhinal
term potentiation. Neuron, 39: 843–853. cortex. Nature, 436: 801–806.
Duncan, G.E., Little, K.Y., Koplas, P.A., Kirkman, J.A., Haring, J.H. and Davis, J.N. (1985) Differential distribution of
Breese, G.R. and Stumpf, W.E. (1991) Beta-adrenergic re- locus coeruleus projections to the hippocampal formation:
ceptor distribution in human and rat hippocampal anatomical and biochemical evidence. Brain Res., 325:
formation: marked species differences. Brain Res., 561: 366–369.
84–92. Harley, C., Milway, J.S. and Lacaille, J.C. (1989) Locus co-
Edwards, C., Nahorski, S.R. and Rogers, K.J. (1974) In vivo eruleus potentiation of dentate gyrus responses: evidence for
changes of cerebral cyclic adenosine 30 ,50 -monophosphate two systems. Brain Res. Bull., 22: 643–650.
induced by biogenic amines: association with phosphorylase Harley, C. and Rusak, B. (1993) Daily variation in active gly-
activation. J. Neurochem., 22: 565–572. cogen phosphorylase patches in the molecular layer of rat
Ezrokhi, V.L., Zosimovskii, V.A., Korshunov, V.A. and Mark- dentate gyrus. Brain Res., 626: 310–317.
evich, V.A. (1999) Restoration of decaying long-term po- Harley, C.W. (1987) A role for norepinephrine in arousal,
tentiation in the hippocampal formation by stimulation of emotion and learning: limbic modulation by norepinephrine
neuromodulatory nuclei in freely moving rats. Neuroscience, and the Kety hypothesis. Prog. Neuropsychopharmacol.
88: 741–753. Biol. Psychiatry, 11: 419–458.
Fara-On, M., Evans, J.H. and Harley, C.W. (2005) Idazoxan Harley, C.W. and Evans, S. (1988) Locus coeruleus-induced
activates rat forebrain glycogen phosphorylase in vivo: a his- enhancement of the perforant path-evoked potential: effects
tochemical study. Brain Res., 1059: 83–92. of intradentate beta blockers. In: Woody, C.D., Alkon, D.L.
Flexner, J.B., Flexner, L.B., Church, A.C., Rainbow, T.C. and and McGaugh, J.L. (Eds.), Cellular Mechanisms of Condi-
Brunswick, D.J. (1985) Blockade of beta 1- but not of beta 2- tioning and Behavioral Plasticity. Plenum Publishing
adrenergic receptors replicates propranolol’s suppression of Corporation, New York, New York, pp. 415–423.
the cerebral spread of an engram in mice. Proc. Natl. Acad. Harley, C.W., Lalies, M.D. and Nutt, D.J. (1996) Estimating
Sci. U.S.A., 82: 7458–7461. the synaptic concentration of norepinephrine in dentate gyrus
Flicker, C. and Geyer, M.A. (1982) Behavior during hippo- which produces beta-receptor mediated long-lasting potent-
campal microinfusions. I. Norepinephrine and diversive ex- iation in vivo using microdialysis and intracerebroventricular
ploration. Brain Res., 257: 79–103. norepinephrine. Brain Res., 710: 293–298.
Florin-Lechner, S.M., Druhan, J.P., Aston-Jones, G. and Harley, C.W. and Milway, J.S. (1986) Glutamate ejection in
Valentino, R.J. (1996) Enhanced norepinephrine release in the locus coeruleus enhances the perforant path-evoked
prefrontal cortex with burst stimulation of the locus co- population spike in the dentate gyrus. Exp. Brain Res., 63:
eruleus. Brain Res., 742: 89–97. 143–150.
Frick, A., Magee, J. and Johnston, D. (2004) LTP is accom- Harley, C.W. and Sara, S.J. (1992) Locus coeruleus bursts in-
panied by an enhanced local excitability of pyramidal neuron duced by glutamate trigger delayed perforant path spike am-
dendrites. Nat. Neurosci., 7: 126–135. plitude potentiation in the dentate gyrus. Exp. Brain Res., 89:
Frizzell, L.M. and Harley, C.W. (1994) The N-methyl-D-as- 581–587.
partate channel blocker ketamine does not attenuate, but Hopkins, W.F. and Johnston, D. (1984) Frequency-dependent
enhances, locus coeruleus-induced potentiation in rat dentate noradrenergic modulation of long-term potentiation in the
gyrus. Brain Res., 663: 173–178. hippocampus. Science, 226: 350–352.
Givens, B. (1996) Stimulus-evoked resetting of the dentate theta Hortnagl, H., Berger, M.L., Sperk, G. and Pifl, C. (1991)
rhythm: relation to working memory. Neuroreport, 8: Regional heterogeneity in the distribution of neurotransmit-
159–163. ter markers in the rat hippocampus. Neuroscience, 45:
Gray, J.A. and Ball, G.G. (1970) Frequency-specific relation 261–272.
between hippocampal theta rhythm, behavior, and am- Jerman, T., Kesner, R.P. and Hunsaker, M.R. (2006) Discon-
obarbital action. Science, 168: 1246–1248. nection analysis of CA3 and DG in mediating encoding but
Gray, J.A., McNaughton, N., James, D.T. and Kelly, P.H. not retrieval in a spatial maze learning task. Learn. Mem., 13:
(1975) Effect of minor tranquillisers on hippocampal theta 458–464.
rhythm mimicked by depletion of forebrain noradrenaline. Jones, L.S., Gauger, L.L. and Davis, J.N. (1985) Anatomy of
Nature, 258: 424–425. brain alpha 1-adrenergic receptors: in vitro autoradiography
Gray, R. and Johnston, D. (1987) Noradrenaline and beta-ad- with [125I]-heat. J. Comp. Neurol., 231: 190–208.
renoceptor agonists increase activity of voltage-dependent Jung, M.W. and McNaughton, B.L. (1993) Spatial selectivity of
calcium channels in hippocampal neurons. Nature, 327: unit activity in the hippocampal granular layer. Hippocam-
620–622. pus, 3: 165–182.
Haas, H.L. and Rose, G.M. (1987) Noradrenaline blocks po- Kitchigina, V., Vankov, A., Harley, C. and Sara, S.J. (1997)
tassium conductance in rat dentate granule cells in vitro. Novelty-elicited, noradrenaline-dependent enhancement of
Neurosci. Lett., 78: 171–174. excitability in the dentate gyrus. Eur. J. Neurosci., 9: 41–47.
315

Klukowski, G. and Harley, C.W. (1994) Locus coeruleus ac- Leutgeb, S., Leutgeb, J.K., Moser, E.I. and Moser, M.B. (2006)
tivation induces perforant path-evoked population spike po- Fast rate coding in hippocampal CA3 cell ensembles.
tentiation in the dentate gyrus of awake rat. Exp. Brain Res., Hippocampus, 16: 765–774.
102: 165–170. Loy, R., Koziell, D.A., Lindsey, J.D. and Moore, R.Y. (1980)
Koda, L.Y. and Bloom, F.E. (1977) A light and electron mi- Noradrenergic innervation of the adult rat hippocampal for-
croscopic study of noradrenergic terminals in the rat dentate mation. J. Comp. Neurol., 189: 699–710.
gyrus. Brain Res., 120: 327–335. Lynch, M.A. and Bliss, T.V. (1986) Noradrenaline modulates
Kopell, N., Ermentrout, G.B., Whittington, M.A. and Traub, the release of [14C]glutamate from dentate but not from
R.D. (2000) Gamma rhythms and beta rhythms have differ- CA1/CA3 slices of rat hippocampus. Neuropharmacology,
ent synchronization properties. Proc. Natl. Acad. Sci. 25: 493–498.
U.S.A., 97: 1867–1872. Mansour, A.A., Babstock, D.M., Penney, J.H., Martin, G.M.,
Kosaka, T., Katsumaru, H., Hama, K., Wu, J.Y. and Heiz- McLean, J.H. and Harley, C.W. (2003) Novel objects in a
mann, C.W. (1987) GABAergic neurons containing the holeboard probe the role of the locus coeruleus in curiosity:
Ca2+-binding protein parvalbumin in the rat hippocampus support for two modes of attention in the rat. Behav.
and dentate gyrus. Brain Res., 419: 119–130. Neurosci., 117: 621–631.
Kovacs, G.L., Bohus, B. and Versteeg, D.H. (1979a) The Maura, G., Pittaluga, A., Ulivi, M. and Raiteri, M. (1988) En-
effects of vasopressin on memory processes: the role of hancement of endogenous GABA release from rat synapto-
noradrenergic neurotransmission. Neuroscience, 4: somal preparations is mediated by alpha 2-adrenoceptors
1529–1537. pharmacologically different from alpha 2-autoreceptors. Eur.
Kovacs, G.L., Bohus, B., Versteeg, D.H., de Kloet, E.R. and J. Pharmacol., 157: 23–29.
de, W.D. (1979b) Effect of oxytocin and vasopressin on McIntyre, C.K., Hatfield, T. and McGaugh, J.L. (2002) Am-
memory consolidation: sites of action and catecholaminergic ygdala norepinephrine levels after training predict inhibitory
correlates after local microinjection into limbic-midbrain avoidance retention performance in rats. Eur. J. Neurosci.,
structures. Brain Res., 175: 303–314. 16: 1223–1226.
Kulkarni, V.A., Jha, S. and Vaidya, V.A. (2002) Depletion of McNaughton, B.L., Battaglia, F.P., Jensen, O., Moser, E.I. and
norepinephrine decreases the proliferation, but does not in- Moser, M.B. (2006) Path integration and the neural basis of
fluence the survival and differentiation, of granule cell pro- the ‘cognitive map’. Nat. Rev. Neurosci., 7: 663–678.
genitors in the adult rat hippocampus. Eur. J. Neurosci., 16: Milner, T.A. and Bacon, C.E. (1989a) GABAergic neurons in
2008–2012. the rat hippocampal formation: ultrastructure and synaptic
Kusaka, K., Morinobu, S., Kawano, K. and Yamawaki, S. relationships with catecholaminergic terminals. J. Neurosci.,
(2004) Effect of neonatal isolation on the noradrenergic 9: 3410–3427.
transduction system in the rat hippocampal slice. Synapse, Milner, T.A. and Bacon, C.E. (1989b) Ultrastructural local-
54: 223–232. ization of tyrosine hydroxylase-like immunoreactivity in the
Lacaille, J.C. and Harley, C.W. (1985) The action of rat hippocampal formation. J. Comp. Neurol., 281:
norepinephrine in the dentate gyrus: beta-mediated 479–495.
facilitation of evoked potentials in vitro. Brain Res., 358: Milner, T.A., Lee, A., Aicher, S.A. and Rosin, D.L. (1998)
210–220. Hippocampal alpha2a-adrenergic receptors are located pre-
Lacaille, J.C. and Schwartzkroin, P.A. (1988) Intracellular dominantly presynaptically but are also found postsynaptic-
responses of rat hippocampal granule cells in vitro to ally and in selective astrocytes. J. Comp. Neurol., 395:
discrete applications of norepinephrine. Neurosci. Lett., 89: 310–327.
176–181. Milner, T.A., Shah, P. and Pierce, J.P. (2000) Beta-adrenergic
Lancaster, B., Hu, H., Ramakers, G.M. and Storm, J.F. (2001) receptors primarily are located on the dendrites of granule
Interaction between synaptic excitation and slow afterhyper- cells and interneurons but also are found on astrocytes and a
polarization current in rat hippocampal pyramidal cells. J. few presynaptic profiles in the rat dentate gyrus. Synapse, 36:
Physiol., 536: 809–823. 178–193.
Lapierre, Y., Beaudet, A., Demianczuk, N. and Descarries, L. Minneman, K.P., Hegstrand, L.R. and Molinoff, P.B. (1979)
(1973) Noradrenergic axon terminals in the cerebral cortex of Simultaneous determination of beta-1 and beta-2-adrenergic
rat. II. Quantitative data revealed by light and electron mi- receptors in tissues containing both receptor subtypes. Mol.
croscope radioautography of the frontal cortex. Brain Res., Pharmacol., 16: 34–46.
63: 175–182. Moudy, A.M., Kunkel, D.D. and Schwartzkroin, P.A.
Lee, E.H., Lee, C.P., Wang, H.I. and Lin, W.R. (1993) (1993) Development of dopamine-beta-hydroxylase-posi-
Hippocampal CRF, NE, and NMDA system interactions in tive fiber innervation of the rat hippocampus. Synapse, 15:
memory processing in the rat. Synapse, 14: 144–153. 307–318.
Lee, I., Hunsaker, M.R. and Kesner, R.P. (2005) The role of Murchison, C.F., Zhang, X.Y., Zhang, W.P., Ouyang, M., Lee,
hippocampal subregions in detecting spatial novelty. Behav. A. and Thomas, S.A. (2004) A distinct role for nor-
Neurosci., 119: 145–153. epinephrine in memory retrieval. Cell, 117: 131–143.
316

Nahorski, S.R., Rogers, K.J. and Edwards, C. (1975) Cerebral Pieribone, V.A., Nicholas, A.P., Dagerlind, A. and Hokfelt, T.
glycogenolysis and stimulation of beta-adrenoreceptors and (1994) Distribution of alpha 1 adrenoceptors in rat brain
histamine H2 receptors. Brain Res., 92: 529–533. revealed by in situ hybridization experiments utilizing sub-
Neuman, R.S. and Harley, C.W. (1983) Long-lasting potent- type-specific probes. J. Neurosci., 14: 4252–4268.
iation of the dentate gyrus population spike by nor- Pittaluga, A., Bonfanti, A. and Raiteri, M. (2000) Somatostatin
epinephrine. Brain Res., 273: 162–165. potentiates NMDA receptor function via activation of
Nicholas, A.P., Pieribone, V. and Hokfelt, T. (1993a) Distri- InsP(3) receptors and PKC leading to removal of the Mg2+
butions of mRNAs for alpha-2 adrenergic receptor subtypes block without depolarization. Br. J. Pharmacol., 130:
in rat brain: an in situ hybridization study. J. Comp. Neurol., 557–566.
328: 575–594. Pittaluga, A., Feligioni, M., Ghersi, C., Gemignani, A. and
Nicholas, A.P., Pieribone, V.A. and Hokfelt, T. (1993b) Cel- Raiteri, M. (2001) Potentiation of NMDA receptor function
lular localization of messenger RNA for beta-1 and beta-2 through somatostatin release: a possible mechanism for the
adrenergic receptors in rat brain: an in situ hybridization cognition-enhancing activity of GABA(B) receptor antago-
study. Neuroscience, 56: 1023–1039. nists. Neuropharmacology, 41: 301–310.
Nitz, D.A. and McNaughton, B.L. (1999) Hippocampal EEG Pittaluga, A., Feligioni, M., Longordo, F., Arvigo, M. and
and unit activity responses to modulation of serotonergic Raiteri, M. (2005) Somatostatin-induced activation and up-
median raphe neurons in the freely behaving rat. Learn. regulation of N-methyl-D-aspartate receptor function: medi-
Mem., 6: 153–167. ation through calmodulin-dependent protein kinase II,
Oleskevich, S., Descarries, L. and Lacaille, J.C. (1989) Quan- phospholipase C, protein kinase C, and tyrosine kinase in
tified distribution of the noradrenaline innervation in the hippocampal noradrenergic nerve endings. J. Pharmacol.
hippocampus of adult rat. J. Neurosci., 9: 3803–3815. Exp. Ther., 313: 242–249.
Orr, G., Rao, G., Houston, F.P., McNaughton, B.L. and Pittaluga, A. and Raiteri, M. (1987) GABAergic nerve termi-
Barnes, C.A. (2001) Hippocampal synaptic plasticity is mod- nals in rat hippocampus possess alpha 2-adrenoceptors reg-
ulated by theta rhythm in the fascia dentata of adult and aged ulating GABA release. Neurosci. Lett., 76: 363–367.
freely behaving rats. Hippocampus, 11: 647–654. Pittaluga, A. and Raiteri, M. (1992) N-methyl-D-aspartic acid
Paban, V., Soumireu-Mourat, B. and escio-Lautier, B. (2003) (NMDA) and non-NMDA receptors regulating hippocampal
Behavioral effects of arginine8-vasopressin in the Hebb-Will- norepinephrine release. I. Location on axon terminals and
iams maze. Behav. Brain Res., 141: 1–9. pharmacological characterization. J. Pharmacol. Exp. Ther.,
Palamarchouk, V.S., Swiergiel, A.H. and Dunn, A.J. (2002) 260: 232–237.
Hippocampal noradrenergic responses to CRF injected into Pittaluga, A., Vaccari, D. and Raiteri, M. (1997) The ‘‘kynuren-
the locus coeruleus of unanesthetized rats. Brain Res., 950: ate test,’’ a biochemical assay for putative cognition enhanc-
31–38. ers. J. Pharmacol. Exp. Ther., 283: 82–90.
Palamarchouk, V.S., Zhang, J., Zhou, G., Swiergiel, A.H. and Quach, T.T., Rose, C. and Schwartz, J.C. (1978) [3H]Glycogen
Dunn, A.J. (2000) Hippocampal norepinephrine-like voltam- hydrolysis in brain slices: responses to neurotransmitters and
metric responses following infusion of corticotropin-releasing modulation of noradrenaline receptors. J. Neurochem., 30:
factor into the locus coeruleus. Brain Res. Bull., 51: 319–326. 1335–1341.
Pang, K. and Rose, G.M. (1987) Differential effects of nor- Rainbow, T.C., Parsons, B. and Wolfe, B.B. (1984) Quantita-
epinephrine on hippocampal complex-spike and theta-neu- tive autoradiography of beta 1- and beta 2-adrenergic recep-
rons. Brain Res., 425: 146–158. tors in rat brain. Proc. Natl. Acad. Sci. U.S.A., 81:
Parfitt, K.D., Doze, V.A., Madison, D.V. and Browning, M.D. 1585–1589.
(1992) Isoproterenol increases the phosphorylation of the Raiteri, M., Garrone, B. and Pittaluga, A. (1992) N-methyl-D-
synapsins and increases synaptic transmission in dentate aspartic acid (NMDA) and non-NMDA receptors regulating
gyrus, but not in area CA1, of the hippocampus. Hippocam- hippocampal norepinephrine release. II. Evidence for func-
pus, 2: 59–64. tional cooperation and for coexistence on the same axon ter-
Parfitt, K.D., Hoffer, B.J. and Browning, M.D. (1991) Nor- minal. J. Pharmacol. Exp. Ther., 260: 238–242.
epinephrine and isoproterenol increase the phosphorylation Reid, A.T. and Harley, C.W. (2005) Unpaired perforant
of synapsin I and synapsin II in dentate slices of young but path and locus ceruleus activations only transiently potenti-
not aged Fisher 344 rats. Proc. Natl. Acad. Sci. U.S.A., 88: ate the evoked dentate gyrus population spike. Online
2361–2365. Abstracts Viewer/Itinerary Planner and CD-ROM. Ref
Pavlides, C., Greenstein, Y.J., Grudman, M. and Winson, J. Type: Abstract.
(1988) Long-term potentiation in the dentate gyrus is induced Ribak, C.E. (1992) Local circuitry of GABAergic basket cells in
preferentially on the positive phase of theta-rhythm. Brain the dentate gyrus. Epilepsy Res. Suppl., 7: 29–47.
Res., 439: 383–387. Ribak, C.E., Nitsch, R. and Seress, L. (1990) Proportion of
Phelps, C.H. (1972) Barbiturate-induced glycogen accumula- parvalbumin-positive basket cells in the GABAergic inner-
tion in brain. An electron microscopic study. Brain Res., 39: vation of pyramidal and granule cells of the rat hippocampal
225–234. formation. J. Comp. Neurol., 300: 449–461.
317

Risso, F., Grilli, M., Parodi, M., Bado, M., Raiteri, M. and potentiation in the dentate gyrus of rat hippocampal slices.
Marchi, M. (2004) Nicotine exerts a permissive role on J. Neurosci., 5: 2169–2176.
NMDA receptor function in hippocampal noradrenergic ter- Stanton, P.K. and Sarvey, J.M. (1985b) The effect of high-
minals. Neuropharmacology, 47: 65–71. frequency electrical stimulation and norepinephrine on cyclic
Rizk, P., Salazar, J., Raisman-Vozari, R., Marien, M., Ruberg, AMP levels in normal versus norepinephrine-depleted rat
M., Colpaert, F. and Debeir, T. (2006) The alpha2-ad- hippocampal slices. Brain Res., 358: 343–348.
renoceptor antagonist dexefaroxan enhances hippocampal Stanton, P.K. and Sarvey, J.M. (1985c) Blockade of nor-
neurogenesis by increasing the survival and differentiation of epinephrine-induced long-lasting potentiation in the hippo-
new granule cells. Neuropsychopharmacology, 31: 1146–1157. campal dentate gyrus by an inhibitor of protein synthesis.
Rolls, E.T. and Kesner, R.P. (2006) A computational theory of Brain Res., 361: 276–283.
hippocampal function, and empirical tests of the theory. Stanton, P.K. and Sarvey, J.M. (1987) Norepinephrine regu-
Prog. Neurobiol., 79: 1–48. lates long-term potentiation of both the population spike and
Rose, G.M. and Pang, K.C. (1989) Differential effect of nor- dendritic EPSP in hippocampal dentate gyrus. Brain Res.
epinephrine upon granule cells and interneurons in the dent- Bull., 18: 115–119.
ate gyrus. Brain Res., 488: 353–356. Straube, T. and Frey, J.U. (2003) Involvement of beta-ad-
Sara, S.J. and Devauges, V. (1989) Idazoxan, an alpha-2 an- renergic receptors in protein synthesis-dependent late
tagonist, facilitates memory retrieval in the rat. Behav. Neu- long-term potentiation (LTP) in the dentate gyrus of
ral Biol., 51: 401–411. freely moving rats: The critical role of the LTP induction
Sara, S.J., Dyon-Laurent, C. and Herve, A. (1995) Novelty strength. Neuroscience, 119: 473–479.
seeking behavior in the rat is dependent upon the integrity of Straube, T., Korz, V., Balschun, D. and Frey, J.U. (2003) Re-
the noradrenergic system. Brain Res. Cogn. Brain Res., 2: quirement of beta-adrenergic receptor activation and protein
181–187. synthesis for LTP-reinforcement by novelty in rat dentate
Sara, S.J. and Segal, M. (1991) Plasticity of sensory responses gyrus. J. Physiol., 552: 953–960.
of locus coeruleus neurons in the behaving rat: implications Subbarao, K.V. and Hertz, L. (1991) Stimulation of energy
for cognition. Prog. Brain Res., 88: 571–585. metabolism by alpha-adrenergic agonists in primary cultures
Sarvey, J.M., Burgard, E.C. and Decker, G. (1989) Long-term of astrocytes. J. Neurosci. Res., 28: 399–405.
potentiation: studies in the hippocampal slice. J. Neurosci. Uecker, A., Barnes, C.A., McNaughton, B.L. and Reiman,
Methods, 28: 109–124. E.M. (1997) Hippocampal glycogen metabolism, EEG, and
Schacher, S., Castellucci, V.F. and Kandel, E.R. (1988) cAMP behavior. Behav. Neurosci., 111: 283–291.
evokes long-term facilitation in Aplysia sensory neurons that U’Prichard, D.C., Reisine, T.D., Mason, S.T., Fibiger, H.C.
requires new protein synthesis. Science, 240: 1667–1669. and Yamamura, H.I. (1980) Modulation of rat brain
Scheinin, M., Lomasney, J.W., Hayden-Hixson, D.M., Scham- alpha- and beta-adrenergic receptor populations by
bra, U.B., Caron, M.G., Lefkowitz, R.J. and Fremeau Jr., lesion of the dorsal noradrenergic bundle. Brain Res., 187:
R.T. (1994) Distribution of alpha 2-adrenergic receptor sub- 143–154.
type gene expression in rat brain. Brain Res. Mol. Brain Res., Usher, M., Cohen, J.D., Servan-Schreiber, D., Rajkowski, J.
21: 133–149. and Aston-Jones, G. (1999) The role of locus coeruleus
Segal, M. (1976) Brain stem afferents to the rat medial septum. in the regulation of cognitive performance. Science, 283:
J. Physiol., 261: 617–631. 549–554.
Segal, M. and Bloom, F.E. (1976a) The action of nor- Vankov, A., Herve-Minvielle, A. and Sara, S.J. (1995) Re-
epinephrine in the rat hippocampus. III. Hippocampal cel- sponse to novelty and its rapid habituation in locus coeruleus
lular responses to locus coeruleus stimulation in the awake neurons of the freely exploring rat. Eur. J. Neurosci., 7:
rat. Brain Res., 107: 499–511. 1180–1187.
Segal, M. and Bloom, F.E. (1976b) The action of nor- Walling, S.G. and Harley, C.W. (2004) Locus ceruleus activa-
epinephrine in the rat hippocampus. IV. The effects of locus tion initiates delayed synaptic potentiation of perforant path
coeruleus stimulation on evoked hippocampal unit activity. input to the dentate gyrus in awake rats: a novel beta-ad-
Brain Res., 107: 513–525. renergic- and protein synthesis-dependent mammalian plas-
Seidenbecher, T., Reymann, K.G. and Balschun, D. (1997) A ticity mechanism. J. Neurosci., 24: 598–604.
post-tetanic time window for the reinforcement of long-term Walling, S.G., Nutt, D.J., Lalies, M.D. and Harley, C.W.
potentiation by appetitive and aversive stimuli. Proc. Natl. (2004) Orexin-A infusion in the locus ceruleus triggers nor-
Acad. Sci. U.S.A., 94: 1494–1499. epinephrine (NE) release and NE-induced long-term potent-
Siggins, G.R., Battenberg, E.F., Hoffer, B.J., Bloom, F.E. and iation in the dentate gyrus. J. Neurosci., 24: 7421–7426.
Steiner, A.L. (1973) Noradrenergic stimulation of cyclic Wang, G.S., Chang, N.C., Wu, S.C. and Chang, A.C. (2002)
adenosine monophosphate in rat Purkinje neurons: an Regulated expression of alpha2B adrenoceptor during devel-
immunocytochemical study. Science, 179: 585–588. opment. Dev. Dyn., 225: 142–152.
Stanton, P.K. and Sarvey, J.M. (1985a) Depletion of nor- Washburn, M. and Moises, H.C. (1989) Electrophysiological
epinephrine, but not serotonin, reduces long-term correlates of presynaptic alpha 2-receptor-mediated
318

inhibition of norepinephrine release at locus coeruleus rat brain development. II. Alpha 2C messenger RNA expression
synapses in dentate gyrus. J. Neurosci., 9: 2131–2140. and [3H]rauwolscine binding. Neuroscience, 76: 261–272.
Winson, J. and Dahl, D. (1985) Action of norepinephrine in the Yavich, L., Jakala, P. and Tanila, H. (2005) Noradrenaline
dentate gyrus. II. Iontophoretic studies. Exp. Brain Res., 59: overflow in mouse dentate gyrus following locus coeruleus
497–506. and natural stimulation: real-time monitoring by in vivo
Winzer-Serhan, U.H., Raymon, H.K., Broide, R.S., Chen, Y. voltammetry. J. Neurochem., 95: 641–650.
and Leslie, F.M. (1997a) Expression of alpha 2 adrenocep- Zilles, K., Gross, G., Schleicher, A., Schildgen, S., Bauer, A.,
tors during rat brain development. I. Alpha 2A messenger Bahro, M., Schwendemann, G., Zech, K. and Kolassa, N.
RNA expression. Neuroscience, 76: 241–260. (1991) Regional and laminar distributions of alpha 1-ad-
Winzer-Serhan, U.H., Raymon, H.K., Broide, R.S., Chen, Y. and renoceptors and their subtypes in human and rat hippocam-
Leslie, F.M. (1997b) Expression of alpha 2 adrenoceptors during pus. Neuroscience, 40: 307–320.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 19

Endocannabinoids in the dentate gyrus

Charles J. Frazier1,2,

1
Department of Pharmacodynamics, University of Florida, College of Pharmacy, Gainesville, FL 32610, USA
2
Department of Neuroscience, University of Florida, College of Medicine, Gainesville, FL 32610, USA

Abstract: Recent years have produced rapid and enormous growth in our understanding of end-
ocannabinoid-mediated signaling in the CNS. While much of the recent progress has focused on other areas
of the brain, a significant body of evidence has developed that indicates the presence of a robust system for
endocannabinoid-mediated signaling in the dentate gyrus. This chapter will provide an overview of our
current understanding of that system based on available anatomical and physiological data.

Keywords: dentate gyrus; cannabinoids; CB1; retrograde transmission; short term plasticity; presynaptic
modulation.

Introduction depolarization-induced suppression of inhibition


(DSI) (Ohno-Shosaku et al., 2001; Wilson and Ni-
Both chronic and acute use of cannabis (mari- coll, 2001). As originally described by Pitler and
juana) produces a wide array of physiological, Alger in hippocampal pyramidal cells (1992), and
cognitive, and analgesic effects in humans that by Marty in cerebellar Purkinje cells (Llano et al.,
have been widely recognized for centuries. The 1991), DSI is a form of short term synaptic plas-
active ingredient in marijuana, delta-9-tetra- ticity whereby depolarization of a single neuron
hydrocannabinol (D9-THC), was first identified in results in transient inhibition of GABAergic affer-
1964 (Gaoni and Mechoulam, 1964); the first pri- ents to that same neuron. A principal role for ECs
mary cannabinoid receptor in the CNS (CB1) was in DSI is indicated by the fact DSI is blocked by
identified and cloned in 1990 (Howlett et al., 1990; CB1 antagonists, occluded by CB1 receptor ago-
Matsuda et al., 1990), and the principal endog- nists, and absent in CB1 / animals (Kreitzer and
enous ligands, arachidonoylethanolamide (AEA) Regehr, 2001a; Ohno-Shosaku et al., 2001; Wilson
and 2-arachidonoylglycerol (2-AG), were identi- and Nicoll, 2001; Wilson et al., 2001; Varma et al.,
fied soon thereafter (Devane et al., 1992; Sugiura 2001; Yoshida et al., 2002). The retrograde nature
et al., 1995). of the EC-mediated signaling was originally (and
Efforts to understand the neurophysiological most simply) indicated by failure of depolariza-
consequences of CB1 receptor activation received tion-induced EC release (or bath application syn-
an enormous boost in 2001 with the discovery that thetic agonists) to inhibit responses to local
endogenous cannabinoids (ECs) act as retrograde application of exogenous GABA, despite effec-
signaling molecules in a phenomenon known as tively reducing evoked responses. This and other
physiological data is strongly supported by an ar-
Corresponding author. Tel.: +1 352 392 3447; ray of immunohistochemical studies that indicate
Fax: +1 352 392 9187; E-mail: cjfraz@ufl.edu prominent expression of CB1 receptors both

DOI: 10.1016/S0079-6123(07)63019-2 319


320

presynaptically and on GABAergic terminals (for as a surprise that fascinating metaplastic roles for
general review see Freund et al., 2003, and below). ECs are also now coming to light. For example,
Although it has only been a short time since these Castillo and colleagues have recently and elegantly
landmark studies, EC-dependent DSI governed by demonstrated that EC-mediated LTD can lower the
fundamentally similar mechanisms has now been threshold for traditional NMDA-dependent LTP in
identified in numerous other brain areas including spatially confined parts of the CA1 pyramidal cell
the amygdala, substantia nigra, basal ganglia, neo- dendritic tree (Chevaleyre and Castillo, 2004).
cortex, brainstem, and recently, in both granule These types of findings are mentioned here in
cells and mossy cells of the dentate gyrus (Trettel brief simply to underscore our growing apprecia-
and Levine, 2003; Yanovsky et al., 2003; Bodor et tion for the enormous breadth of EC-dependent
al., 2005; Isokawa and Alger, 2005; Mukhtarov et signaling in CNS physiology, and to highlight the
al., 2005; Zhu and Lovinger, 2005; Engler et al., extremely rapid nature of recent progress. In my
2006; Hofmann et al., 2006). Cumulatively, these view, this trend shows no sign of slowing. In fact,
findings highlight what is now the obviously broad in this chapter, I will attempt to make the case that
significance of EC-mediated retrograde signaling in the dentate gyrus is an area of significant interest,
modulating inhibitory transmission in the CNS, but and strong potential, for this field; albeit one that
nevertheless, represent only a small percentage of has received comparatively little attention to date.
recent advances. For example, it is now clear that In making that argument I will highlight a signifi-
cannabinoid receptors are expressed on many glut- cant body of existing evidence clearly indicating
amatergic terminals as well. Accordingly, depolari- the presence of a robust system for EC-dependent
zation-induced EC-mediated inhibition of signaling in the dentate gyrus, I will review the
excitation (DSE) has now been described in the growing list of studies that reveal specific neuro-
cerebellum (Kreitzer and Regehr, 2001b), ventral physiological consequences of EC-dependent
tegmental area (Melis et al., 2004a, b), hypo- signaling at identified synapses in the dentate,
thalamus (Di et al., 2005), amygdala (Domenici et and I will cover recent work suggesting several
al., 2006), and in the hippocampus (Ohno-Shosaku novel roles for EC signaling (from neuroprotection
et al., 2002b), where the role of ECs in modulating to neurogenesis) in which the dentate may be an
Schaffer collateral inputs to CA1 pyramidal cells area of particular interest. Potential future direc-
remains a particularly active area of investigation tions for some of this work will also be discussed.
(Hajos and Freund, 2002; Hoffman et al., 2005; For a broader overview of specific roles for ECs in
Katona et al., 2006; Takahashi and Castillo, 2006). CNS physiology the interested reader is referred to
Further, efforts to understand the role of postsy- any number of recent and outstanding reviews
naptic metabotropic glutamate receptors in EC re- (Schlicker and Kathmann, 2001; Alger, 2002; Wil-
lease have proven instrumental in identifying EC- son and Nicoll, 2002; Freund et al., 2003; Cheval-
mediated long-term depression (LTD) in a number eyre et al., 2006; Marsicano and Lutz, 2006).
of areas. At present, both heterosynaptic and ho-
mosynaptic forms of EC-dependent LTD have been
described, affecting both GABAergic and glut- A brief overview of the endocannabinoid system
amatergic synapses respectively (Marsicano et al.,
2002; Robbe et al., 2002; Chevaleyre and Castillo, There are currently two known cannabinoid re-
2003; Hoffman et al., 2003). While presynaptic ex- ceptors: the CB1 receptor, cloned in 1990 (Mats-
pression still seems to be the norm in this form of uda et al., 1990), and the CB2 receptor, identified 3
EC-dependent synaptic plasticity, there has even years later (Munro et al., 1993). Both receptors are
been a recent example of EC-mediated, mGluR classic metabotropic receptors coupled to Gi/o. Of
dependent LTD in the cerebellum that is expressed these, the CB1 receptor appears to be intimately
postsynaptically (Safo and Regehr, 2005). Given involved in mediating most central effects of ECs.
the extensive nature of EC-mediated effects on It is among the most prominently expressed me-
synaptic transmission, it should perhaps not come tabotropic receptors in the mammalian CNS, with
321

overall expression levels approaching that of iono- a highly calcium-dependent enzyme, N-acyltransf-
tropic receptors for glutamate and GABA. Partic- erase, which is capable of producing N-arachido-
ularly high levels of CB1 expression are found in noyl-PE from phosphatidylethanolamine (Di Marzo
hippocampus (including the dentate gyrus), cortex, et al., 1994; Cadas et al., 1996; Piomelli, 2003).
cerebellum, and basal ganglia (Herkenham et al., While it is clear that increases in intracellular cal-
1991a, c; Tsou et al., 1998a). This distribution is cium can also increase production of 2-AG through
consistent with well-recognized effects of can- a PLC and DAG dependent pathway (Stella et al.,
nabinoids on movement and memory. The CB2 1997), the overall story here is somewhat more
receptor by contrast, has been reported in the CNS complicated. For example, in some systems calcium
only in striatum and cerebellum (Skaper et al., dependent production of 2-AG is likely to be sig-
1996; Van Sickle et al., 2005), but is far more nificantly enhanced if the calcium influx occurs co-
abundant in the immune system where it is incident with synaptic activation of certain
thought to be involved in regulation of both im- metabotropic receptors (notably metabotropic
mune function and inflammatory responses (for glutamate receptors and/or muscarinic acetylcholine
review see Berdyshev, 2000; De et al., 2000). Since receptors) capable of activating a calcium-dependent
the CB2 receptor is believed to be absent in the isoform of phospholipase C (Hashimotodani et al.,
dentate, it will not be reviewed further here. 2005; Maejima et al., 2005). On the other hand,
There are likely several endogenous ligands for there also appear to be metabolic routes through
the CB1 receptor, of which two have been partic- which metabotropic receptors can promote largely
ularly well characterized to date: arachidonoyl- calcium-independent production of 2-AG (Maejima
ethanolamide (anandamide, abbreviated AEA) et al., 2001; Varma et al., 2001; Ohno-Shosaku et al.,
and 2-AG. Both of these ligands are fatty acid 2002a, 2005; Fukudome et al., 2004).
derivatives, although they differ significantly in Once AEA has reached the extracellular space,
terms of their synthetic pathway. AEA is thought uptake is mediated by a fast, selective, saturable,
to be synthesized by phospholipase D induced hy- temperature-dependent, and thus apparently car-
drolysis of a lipid precursor, N-arachidonoyl rier-mediated process (Beltramo et al., 1997; Hillard
phosphatidylethanolamine (N-arachidonoyl-PE) et al., 1997; Piomelli et al., 1999). Subsequently,
(Di Marzo et al., 1994, 1999; Cadas et al., 1997). AEA is rapidly broken down to arachidonic acid
One of the primary immediate precursors for and ethanolamine by fatty acid amide hydrolase
synthesis of 2-AG, by contrast, appears to be di- (FAAH), an enzyme that has an expression pattern
acylglycerol, which is produced from phosphatidy- in the CNS that closely corresponds to that of the
linositol by phospholipase C, and subsequently CB1 receptor (Hillard et al., 1995; Cravatt et al.,
converted to 2-AG by diacylglycerol lipase 1996; Egertova et al., 1998; Piomelli et al., 1999;
(DAGL) (Stella et al., 1997; Piomelli, 2003). Deutsch et al., 2002). The corresponding story is
Because both AEA and 2-AG are membrane per- again more complicated for 2-AG. While it is clear
meant, conventional vesicular release mechanisms that the anandamide transporter can also facilitate
are implausible. Thus, regulation of EC-mediated uptake of 2-AG, and that FAAH can also break it
signaling is likely to depend principally on regula- down, it is too early to rule out the possibility of an
tion of the synthetic pathways described above, on additional transport mechanism for 2-AG, and it is
uptake from the extracellular space, and on subse- clear that at least one other major route for break-
quent breakdown. These processes also differ some- down, in this case mediated by monoglyceride
what between AEA and 2-AG. Synthesis of AEA is lipase (MGL), is available (Dinh et al., 2002).
increased significantly by rapid increases in intra- Activation of metabotropic CB1 receptors lo-
cellular calcium concentration (that reach low mi- cated primarily on presynaptic terminals may be
cromolar levels) such as occur with membrane coupled to adenylyl cyclase, or directly to either
depolarization. The calcium dependence of AEA Ca2+ or K+ conductances, and is typically (but
production is tied to the fact that availability of N- perhaps not exclusively) observed to inhibit cal-
arachidonoyl-PE depends heavily on the activity of cium dependent release of neurotransmitter. The
322

general consensus is that such activation of en- receptors and a strong tendency to partition into
dogenous CB1 receptors typically occurs through biological membranes (Harris et al., 1978; Roth and
retrograde transmission of endocannabinoids syn- Williams, 1979; Thomas et al., 1992). However, in
thesized postsynaptically, and further that 2-AG 1991 two landmark studies capitalized on the fa-
and/or AEA are likely to be the retrograde mes- vorable properties of a synthetic CB1 agonist,
sengers in most if not all forms of EC-mediated CP55,940, to provide one of the first complete sur-
retrograde signaling identified to date. It is inter- veys of putative cannabinoid receptor binding sites
esting to note that 2-AG is present in brain at in the CNS (Herkenham et al., 1991a, b). These
concentrations approximately 200-fold greater studies demonstrated that the molecular layer of the
than AEA, although this may simply reflect a dentate gyrus was among the most intensely labeled
more central role in lipid metabolism and not nec- areas in the rat brain. Comparatively low to mod-
essarily a similarly dominant role in retrograde erate levels of 3H-CP55,940 binding were also re-
signaling (Sugiura et al., 1995; Stella et al., 1997). ported in the hilus, with only sparse labeling over
For interesting and recent discussion of the rela- the granule cell layer. Interestingly, these studies
tive roles of 2-AG and AEA in retrograde signa- also reported differential binding of 3H-CP55,940
ling see Chevaleyre et al. (2006). along the septo-temporal axis of the hippocampus,
While this represents a brief and general summary with denser binding on the dorsal (septal) end.
of the EC system as currently understood in the These findings, which were reported from rat brain,
brain (including the dentate) it is important to re- were largely consistent with later studies using the
iterate that this field is advancing rapidly. Significant same radioligand in human (Glass et al., 1997).
advances in our conceptions of this system are cer- While the original studies with 3H-CP55,940
tainly possible, and perhaps even probable. One represented a significant advance in their day, it is
area where our knowledge is currently developing at noteworthy that they were completed at about the
a rapid pace concerns CB1 receptor-mediated mod- same time that the CB1 receptor was isolated and
ulation of glutamatergic transmission (see below). cloned (Matsuda et al., 1990). That feat enabled
Further, it is likely that additional sites of EC action two additional approaches to the study of CB1
will be identified, which may include currently un- receptor expression and distribution. The first to
cloned CB-like receptors and/or conventional non- be employed was in situ hybridization for CB1
CB receptors not previously associated with EC ac- mRNA (Mailleux and Vanderhaeghen, 1992;
tion. Thus it seems likely that our understanding of Mailleux et al., 1992; Matsuda et al., 1993; Marsi-
the possible modulatory effects of ECs on synaptic cano and Lutz, 1999). In general, these studies
transmission in the CNS will only broaden with noted high levels of CB1 mRNA in a subpopula-
time. A more extensive discussion of the molecular tion of cells in the subgranular region of the hilus,
aspects of EC signaling is available in a number of where 3H-CP55,940 was low to moderate, with
other reviews (Di Marzo and Deutsch, 1998; Lutz, comparatively less concentrated expression of CB1
2002; Piomelli, 2003; Sugiura et al., 2006). mRNA in the molecular layer, where 3H-CP55,940
binding had been intense. This differential distri-
bution between CB1 mRNA expression and 3H-
Markers of the EC system in the dentate gyrus CP55,940 binding was the basis for one of the first
arguments that functional CB1 expression may be
CB1 receptor expression in the dentate gyrus: initial relegated largely to processes distal from the soma.
findings indicate prominent role in GABAergic The second major advance enabled by cloning
neurons of the CB1 receptor was the immunohistochemical
characterization of receptor distribution in the
Early efforts to map the distribution of putative CB CNS. This technique offered several significant
receptors in brain were hindered by the fact that advantages over earlier efforts. First, because the
they relied largely on radiolabeled derivatives of D9- initial antibodies employed were targeted to spe-
THC that had generally poor affinity for the cific amino acid sequences on the N-terminal of
323

the cloned CB1 receptor, they offered significantly (Katona et al., 1999; Tsou et al., 1999). Further
greater specificity than earlier radioligand binding examination of these results at the electron micro-
studies. Second, the higher resolution of immuno- scopic level indicated cellular expression of CB1
histochemical techniques allowed the first detailed was largely or exclusively confined to cytoplasmic
examination of CB1 expression at the subcellular organelles, while surface expression of CB1 was
level. Third, the eventual application of elegant very high on CCK-positive axon terminals that
dual-labeling immunohistochemical techniques al- were often clustered around principal cell somas
lowed simultaneous analysis of CB1 expression and proximal dendrites (Katona et al., 1999). Al-
and cell phenotype. As such, our current under- though these particular observations were origi-
standing of CB1 expression and distribution in the nally made primarily in hippocampal cells in CA1
CNS depends heavily on these techniques, and re- and CA3, it now seems reasonable to predict that
sults relevant to the dentate will be reviewed in surface expression of CB1 on GABAergic neurons
some detail below. in the dentate may also be restricted to nerve ter-
The first immunohistochemical studies of CB1 minals. This hypothesis is consistent with the
distribution in the CNS clearly demonstrated in- differential distribution of 3H-CP55,940 binding
tense immunoreactivity in the molecular layer of the and in situ hybridization described earlier, and is
dentate gyrus (Pettit et al., 1998; Tsou et al., 1998a; also strongly supported by a study from the
Katona et al., 1999). Consistent with early binding Freund laboratory the following year that specifi-
studies, the intensity of this signal was reported cally identified CB1-positive axon terminals in the
to be highest in the inner third of the molecular hilus (Acsady et al., 2000). Although this study did
layer, and on the septal end of the hippocampus not directly perform dual-labeling experiments for
(Fig. 1AA, Tsou et al., 1998a). However, based on both CB1 and CCK, it did make the striking ob-
the greater resolution of these techniques, it now servation that both CB1-positive and (in separate
became apparent that this area specifically con- experiments) CCK-positive terminals displayed
tained many CB1-immunoreactive nerve fibers, but very similar and yet highly unusual target selec-
very few labeled cell bodies. By contrast, intensely tivity. Specifically they found these terminals made
immunoreactive neuronal cell bodies were identified abundant synaptic contacts on the soma and prox-
in the hilus, just below the granule cell layer. These imal dendrites of the GluR2/3 and calcitonin gene-
cells were reported to represent as much as 62% of related peptide (CGRP)-positive hilar mossy cells,
all CB1-immunoreactive cell bodies in the dentate, but largely avoided both parvalbumin and sub-
and had characteristic pyramidal morphology, with stance P (and thus also CCK)-immunoreactive in-
a single primary dendrite that often extended with- terneurons in the hilus.
out branching through the granule cell layer toward Cumulatively, the data described so far clearly
the molecular layer (Fig. 1AB C, Tsou et al., 1998a; indicates prominent expression of CB1 receptors in
Katona et al., 1999). The remaining CB1-positive the dentate, particularly on presynaptic terminals of
cell bodies were located largely in the hilus and ex- CCK-positive interneurons, many of which have
hibited a variety of morphologies. cell bodies located in the subgranular zone. The
The subsequent application of dual-labeling specific distribution of CB1-positive terminals fur-
immunohistochemical techniques represented ther suggests a likely role for CB1 activation in
a major advance in this area, as it allowed the regulating GABAergic transmission to both gran-
determination that the vast majority of CB1- ule cells and hilar mossy cells. Nevertheless the ap-
immunoreactive cell bodies were also positive for parent absence of CCK and CB1positive terminals
CCK. Specifically, between 80 and 95% of all making synaptic contact with other basket cells
CCK-positive cell bodies in the dentate were re- clearly distinguishes hilar circuitry from that in
ported to also be CB1-positive, with a comparably other hippocampal subfields, and may hold clues to
high percentage of CB1-positive cells being CCK- understanding the unique physiology of this area.
positive. In stark contrast, virtually none of the While the vast majority of the work described
parvalbumin-positive basket cells expressed CB1 above was completed in either Sprague-Dawley or
324

Fig. 1. CB1 receptor expression in the dentate gyrus. (AA) An early study demonstrates intense immunoreactivity for CB1 receptors in
the dentate gyrus, especially in the inner molecular layer (mo). (AB C) Although dentate granule cells appear to be CB1 negative, there
are many intensely stained neurons found at the base of the granule cell layer, with characteristic pyramidal morphology. Other
intensely stained neurons were found in the hilus (not shown). Note that labeled cell bodies were generally absent in the molecular
layer. Scale bars: 200 mm, 50 mm, 20 mm in AA–C respectively. Adapted with permission from Tsou et al. (1998a). (B) A recent
immunohistochemical analysis of CB1 receptor expression in wild type (WT), CaMK-CB1 / , GABA-CB1 / , and complete CB1 /
mice provided clear evidence of CB1 expression on glutamatergic terminals (as well as GABAergic terminals) in the inner molecular
layer of the dentate gyrus. Areas indicated by black boxes in the top row are shown in higher magnification in the bottom row. Scale
bar in H is 100 mm. Adapted with permission from Monory et al. (2006).

Wistar rats, the basic features of CB1 expression in to emphasize that the data as summarized above do
the dentate described so far appear to be largely not represent the full picture with respect to likely
conserved in gray mouse lemur, primate, and hu- sites of EC action in the dentate. Our current un-
man brain (Ong and Mackie, 1999; Katona et al., derstanding of this topic with respect to glut-
2000; Harkany et al., 2005). Further, although amatergic systems is reviewed below.
much of this work has relied on the N-terminal
antibodies to the CB1 receptor that are identical or
similar to the one originally developed by Tsou et CB1 receptor expression in the dentate gyrus:
al. (1998a), key features of this expression pattern emphasis on glutamatergic neurons
have also been observed using antibodies that target
intracellular sites on the C-terminal end of the CB1 Compared to the results summarized above, the
receptor (Egertova and Elphick, 2000; Hajos et al., question of CB1 receptor expression by glut-
2000). Nevertheless, at this juncture it is important amatergic neurons has so far been a contentious
325

one. This is particularly true in areas CA1 and was notably greater than had been observed pre-
CA3 where apparent expression of low levels of viously, particularly in the inner third of the mo-
CB1 mRNA in pyramidal neurons as indicated by lecular layer. Further, upon examination of the
in situ hybridization generally has not been easy to results at the electron microscopic level it became
reconcile with the apparent lack of CB1 immuno- clear that, in contrast to results with earlier anti-
reactivity on glutamatergic cell bodies or axon bodies, as many as 80% of asymmetrical (i.e. ex-
terminals as indicated by immunohistochemical citatory) synapses in the inner molecular layer, and
techniques. This discrepancy has been made all the 30–50% of asymmetrical synapses in other layers
more acute by divergent reports from a number of were unambiguously positive for CB1. No differ-
outstanding laboratories regarding the physiolog- ences in this pattern were observed between
ical effects of ECs on glutamatergic transmission C57BL/6 and CD1 mice, and the authors noted
in the Schaffer collateral pathway (Hajos and that preliminary data suggests similar results will
Freund, 2002; Hoffman et al., 2005; Katona et al., be obtained using this antibody in both rat and
2006; Takahashi and Castillo, 2006). However, in human hippocampus.
general, the current details of this debate are out- A second extremely recent study has reproduced
side the scope of this review. The dentate gyrus has and extended these findings using yet another
to a large extent been spared from a similar debate unique antibody, in this case a goat anti-CB1 an-
in part because, unlike in area CA1 and CA3, in tibody targeting the C-terminal 77 amino acids of
situ hybridization tends to agree with both the protein (Monory et al., 2006). While providing
immunohistochemical analysis and early radiolig- other significant insights to be reviewed later in
and binding studies in suggesting that dentate this chapter, this study confirms the presence of
granule cells are CB1-negative. Nevertheless, it CB1-positive glutamatergic terminals in the inner
would be inaccurate to say there have been no di- third of the molecular layer by elegantly demon-
vergent results in this area at all. For example, a strating clear immunohistochemical labeling for
unique N-terminal antibody targeting amino acids CB1 that is resistant to selective CB1 knockouts
83–91 of the CB1 receptor employed by Pettit et al. targeted specifically at GABAergic neurons, but
(1998) does label dentate granule cells, while the reduced by selective CB1 knockouts targeting
Tsou et al. (1998a) antibody that fails to label glutamatergic cells (Fig. 1B, for more on this tech-
dentate granule cells in rats and humans has been nology, also see Marsicano et al., 2003). The ob-
reported to do so in non-human primates. While servation that apparently glutamate-specific CB1
the precise reasons for these discrepancies in the labeling was concentrated in the inner third of the
dentate gyrus have still not been completely re- molecular layer, combined with the observation
solved, very recent work using novel C-terminal that dentate granule cells continued to lack
antibodies to CB1 has simultaneously reinforced immunoreactivity for CB1 in all models, suggested
the conclusion that dentate granule cells are CB1- the hypothesis that glutamate-specific CB1 labe-
negative, while elegantly demonstrating a poten- ling could be related to presynaptic expression on
tially major role for EC signaling via the other the axon terminals of hilar mossy cells. Consistent
glutamatergic cell type in the dentate, the hilar with that hypothesis, these authors further use
mossy cells. dual-labeling in situ hybridization techniques to
The first of two significant reports in this area show strong co-localization of mRNA for CB1
was from the laboratory of Tamas Freund and for the vesicular glutamate transporter type 1
(Katona et al., 2006). This study used a guinea (VGlutT1) both in the soma of hilar mossy cells
pig anti-CB1 antibody originally characterized by and in presynaptic terminals forming asymmetric
Fukudome et al. (2004) that targets amino acids synaptic contacts in the inner molecular layer.
443–473 on the C-terminal end of the protein. The These findings do not conflict with earlier work
pattern of CB1 immunoreactivity observed with using other antibodies that indicated robust ex-
this antibody had similar distribution to that pre- pression of CB1 on CCK-positive axon terminals
viously reported, however the intensity of labeling in this area. Instead, they indicate additional
326

expression of CB1 on glutamatergic terminals These findings suggest the possibility of significant
likely belonging to hilar mossy cells that was not cross talk between several transmitter systems in
previously detected. This is a striking conclusion regulating the activity of CCK-positive interneurons
by any measure, and I suspect it will have a sig- in the dentate.
nificant impact on future work in this area. Nev-
ertheless, the precise reasons that CB1 expression
by glutamatergic terminals in the molecular layer Other markers of the EC system in the dentate
was not detected earlier and continues to be some- gyrus
what unclear. One likely possibility is that the N-
terminal antibodies generally employed in earlier Although a few additional approaches have been
work simply lack the sensitivity to detect lower employed, most efforts to demonstrate the pres-
levels of CB1 expression present on glutamatergic ence of the EC system in the hippocampus and
terminals. Another compatible possibility that has dentate gyrus without relying on detection of CB1
been raised, but to my knowledge not yet formally have focused on the two primary enzymes for
reported, is that there exists a ‘CB1 receptor-in- degradation of AEA and 2-AG reviewed earlier:
teracting protein’ that is exclusively expressed by FAAH and MGL, respectively.
excitatory cells and is capable of interfering with At this point, immunohistochemical studies tar-
sites targeted by some of the earlier C-terminal geting various areas on the C-terminal end of
antibodies (Niehaus et al., 2004). FAAH have indicated that expression is predom-
inantly postsynaptic and largely confined to hippo-
campal principal cells and dentate granule cells
CB1 receptor expression in the dentate gyrus: co- (Tsou et al., 1998b; Romero et al., 2002; Egertova
expression with other receptor subtypes et al., 2003; Gulyas et al., 2004). This distribution is
consistent with that reported for mRNA for FAAH
While a concerted effort has been made to charac- (Thomas et al., 1997), and indicates that FAAH
terize CB1 expression with respect to glutamatergic expression is largely restricted to neurons known to
vs. GABAergic neurons and presynaptic vs. somatic receive CB1-positive afferent inputs. The diffuse
expression as described above, some otherwise iso- labeling for FAAH outside the cell layers noted by
lated reports have specifically examined co-expres- Egertova et al. (2003) may be an artifact of tissue
sion of CB1 with other receptor subtypes. preparation, or as noted by the authors, could in-
Noteworthy examples in the dentate include a re- dicate an ability of principal cells to excrete FAAH
port by Cristino et al. (2006) who used dual-labeling into the extracellular space. Interestingly, Gulyas et
immunohistochemical techniques to suggest co-ex- al. (2004) has also noted that intracellular expres-
pression of both CB1 and the vanilloid receptor sion of FAAH was most often localized to the sur-
TRPV1 in many non-pyramidal bipolar neurons of face of organelles associated with calcium storage.
the dentate, particularly in the molecular layer. Fur- In contrast to the predominantly postsynaptic
ther, Morales and Backman (2002) used double localization of FAAH, MGL appears to be ex-
labeling in situ hybridization techniques to demon- pressed primarily presynaptically in both the hip-
strate a strong co-localization of mRNA for both pocampus and dentate gyrus. Specifically,
CB1 receptors and 5-HT3A receptors in the sub- immunoreactivity for MGL has been identified
granular interneurons of the dentate hilus. This re- on the mossy fiber axon terminals of dentate gran-
sult was generally consistent with Hermann et al. ule cells, as well as on the terminals of CA3 py-
(2002) who also demonstrated co-localization of ramidal cells and some interneurons (Gulyas et al.,
CB1 with 5-HT1B, and 5-HT3 receptors, and further 2004). In contrast, absence of MGL immunoreac-
noted the extent of co-expression was higher in tivity in certain areas of CA3 and the molecular
neurons with high levels of CB1 mRNA. This study layer noted in the same study suggested that CA1
also was noteworthy in identifying co-expression of and perforant path inputs are likely to lack this
CB1 with D2 receptors throughout the dentate. enzyme.
327

Another noteworthy technique in this area has findings to date are consistent with predictions
been to examine the expression of DAGL, a bio- based on the existing anatomical information re-
synthetic enzyme central to the production of 2- viewed above.
AG. Two very recent studies that were published
almost simultaneously have both demonstrated
that DAGL, as detected by immunohistochemical DSI in the dentate gyrus
analysis using several different antibodies, is highly
expressed by principal neurons in the hippocam- Anatomical studies reviewed earlier strongly sug-
pus and granule cells (and likely mossy cells) of the gest that CB1-positive GABAergic terminals of
dentate gyrus, and has further indicated that such CCK-positive basket cells make numerous synap-
expression is almost completely restricted to the tic contacts with both granule cells and hilar mossy
head and neck of dendritic spines (Katona et al., cells. These findings suggest such inputs may be
2006; Yoshida et al., 2006). Katona et al. (2006) modulated by CB1 agonists. Further, the presence
also showed specifically that presynaptic terminals of FAAH and/or DAGL in both granule cells and
opposite DAGL positive spines are CB1-positive, mossy cells suggests they may represent sources of
indicating that DAGL is perfectly positioned to ECs. Detailed physiological studies have now
promote retrograde transmission, presumably me- strongly reinforced these conclusions.
diated by 2-AG at numerous sites throughout the The first evidence that GABAergic inputs to
hippocampus and dentate gyrus. dentate granule cells were sensitive to exogenous
cannabinoids was published in 2000 when Hajos et
al. used whole-cell recording techniques to dem-
Physiological role for ECs in the dentate onstrate that evoked inhibitory postsynaptic cur-
rents (eIPSCs) recorded from dentate granule cells
Early efforts to examine the neurophysiological were reduced to 3676% of baseline by bath ap-
effects of cannabinoids began well before isolation plication of WIN55,212-2 in wild type but not in
and cloning of the CB1 receptor. Both in vivo and CB1 / mice (Hajos et al., 2000). Several years
in vitro approaches were employed; experimental later, Nakatsuka et al. (2003) similarly reported a
measures ranged from auditory and sensory WIN55,212-2-mediated, AM-251-sensitive reduc-
evoked potentials to extracellular field potentials tion in both frequency and amplitude of sponta-
and unit recordings. Using such techniques, a neous IPSCs (sIPSCs) recorded from dentate
number of studies specifically demonstrated effects granule cells in the human dentate gyrus. These
of cannabinoids on identified cells or circuits in the findings indicated that GABAergic inputs to dent-
dentate gyrus (see Kujtan et al., 1983; Campbell et ate granule cells are indeed inhibited by CB1 re-
al., 1986a, b; Wilkison and Pontzer, 1987; Hamp- ceptor activation, however they did not directly
son et al., 1989). However, just as cloning of the address the question of whether granule cells were
CB1 receptor and development of powerful capable of activating these receptors by signaling
immunohistochemical techniques dramatically ex- via ECs.
panded our understanding of CB1 expression in That question was answered in the affirmative
the brain, major advances in electrophysiological with the first complete characterization of DSI in
techniques have now provided the technology nec- dentate granule cells in 2005 (Isokawa and Alger,
essary to dissect the effects of both endogenous 2005). This study specifically demonstrated that
and exogenous cannabinoids at the synaptic level. direct depolarization of a single dentate granule
Admittedly, applications of such technology to the cell reduces the amplitude of both sIPSCs and
study of cannabinoid-dependent systems in the IPSCs evoked by stimulation of the molecular
dentate gyrus are only recently beginning to ap- layer (Fig. 2A). Endocannabinoids were impli-
pear in the literature. Nevertheless, the emphasis cated in this process because both CP55,942 and
below will be on such recent findings, with a par- WIN55,212-2 mimicked and occluded DSI and
ticular effort to note when and where physiological because DSI was blocked by the CB1 antagonist
328

Fig. 2. DSI in dentate granule cells and hilar mossy cells. (AA) Dentate gyrus of the hippocampal slice (scale, 50 mm). (AB) Differential
interference contrast image of dentate granule cells (scale, 10 mm). (AC) Single dentate granule cell filled, and visualized with flu-
orescence microscopy during whole cell recording (scale, 10 mm). (AD) DSI of eIPSCs (left, top) and sIPSCs (right, top) is observed in
dentate granule cells. In both cases, magnitude of DSI is reduced by bath application of thapsigargin (TG, 2 mM, bottom). (AE)
Thapsigargin also reduced the calcium transients produced by DSI-inducing depolarizing voltage steps. Adapted with permission from
Isokawa and Alger (2005). (B) Depolarization induced release of endogenous cannabinoids from hilar mossy cells preferentially
inhibits calcium-dependent exocytosis. DSI in of eIPSCs is apparent in an individual mossy cell. DSI of miniature IPSCs is absent in
the same cell. Following bath application of KCl and CaCl2, DSI of mIPSCs becomes apparent. Summary data are presented in BD,
where numbers on the bars are n values. Adapted with permission from Hofmann et al. (2006). (See Color Plate 19.2 in color plate
section.)

SR141716A. These authors further demonstrated calcium stores in the mechanism of DSI, as deple-
that the magnitude of DSI correlated with the ex- tion of those stores reduced both [Ca2+]i and the
tent of depolarization-induced increases in [Ca2+]i magnitude of DSI.
and that DSI could be blocked by chelating Interestingly, an earlier study had demonstrated
postsynaptic calcium with internal BAPTA. Fi- that although DSI of sIPSCs was absent in dentate
nally, this study indicated a particular role for granule cells in control conditions, it could be de-
calcium release from ryanodine sensitive internal tected 1 week or more following induction of
329

febrile seizures (Chen et al., 2003). Induction of within the dentate gyrus. Nevertheless, I think it
febrile seizures also enhanced EC-mediated signa- is fair to say that despite a number of studies
ling in area CA1, apparently due to increases in clearly indicating antiepileptic (Consroe et al.,
CB1 expression that were restricted to CCK-pos- 1975; Consroe and Wolkin, 1977; Wallace et al.,
itive GABAergic terminals. While the finding that 2001, 2002, 2003; Blair et al., 2006), and neuro-
febrile seizures upregulate EC-mediated signaling protective (Nagayama et al., 1999; Braida et al.,
throughout the hippocampus and dentate gyrus 2000; Van der Stelt et al., 2001a, b, 2002;
remains quite intriguing, the apparent discrepancy Khaspekov et al., 2004) effects of cannabinoids,
with respect to the ability to detect DSI in dentate there is a notable lack of physiological studies that
granule cells in control conditions remains unex- directly test the hypothesis that ECs can modulate
plained. excitatory transmission at these synapses. For ex-
Also consistent with predictions based on ana- ample, to my knowledge, there are currently no
tomical data is a recent report from my laboratory published reports directly testing effects of ECs on
which provides the first published characterization mossy fiber transmission. Preliminary data from
of DSI as observed in hilar mossy cells (Hofmann my laboratory is consistent with the weight of ex-
et al., 2006). Specifically, we reported DSI of isting anatomical evidence in suggesting that iso-
eIPSCs in hilar mossy cells that depends on both lated mossy fiber inputs to CA3 pyramidal cells are
postsynaptic calcium influx and presynaptic CB1 indeed CB1-negative; however it must be noted
receptors. The magnitude of DSI in hilar mossy that neither our preliminary work nor existing an-
cells was directly dependent on depolarization du- atomical data can at present explicitly rule out the
ration, and enhanced by bath application of car- possibility of additional non-CB1 receptors for
bachol (CCh). Further, the presynaptic mechanism endocannabinoids at these terminals.
of inhibition was found to be largely or exclusively There are somewhat more available data with
selective for calcium-dependent release events (Fig. respect to perforant path inputs to dentate granule
2B). We also used dual whole cell recording tech- cells. Although these inputs were initially reported
niques to demonstrate that there are tight spatial to be insensitive to D9-THC using in vivo extra-
constraints on diffusion of ECs released from hilar cellular recording techniques (Wilkison and Po-
mossy cells. Finally, we were intrigued to observe ntzer, 1987), two later reports using different
two different forms of expression of DSI of sIPSCs agonists and activation strategies have been con-
that depended on whether the sIPSCs had high sistent with the hypothesis that these terminals
spectral power at theta frequencies. Cumulatively, may express presynaptic CB1 receptors. One such
these results suggested a prominent role for EC- study explicitly demonstrated that bath applica-
mediated signaling between hilar mossy cells and tion of WIN55,212-2 produced a rightward shift in
GABAergic afferents in normal hilar function, and the input/output curve, and a reduction in paired
raise interesting questions about the effects of both pulse facilitation, of perforant path synaptic po-
ECs and CCh on coordinated activity in GAB- tentials recorded from the outer third of the mo-
Aergic networks. An earlier report from another lecular layer (Kirby et al., 1995). Further, another
laboratory indicating DSI of sIPSCs could be de- study examining field potential responses pro-
tected in hilar mossy cells appeared in abstract duced by stimulation of the medial perforant path
form (Howard et al., 2003). strikingly found that reductions in fEPSPs pro-
duced by an acetylcholinesterase inhibitor
(physostigmine) were blocked by the CB1 recep-
Evidence that ECs modulate glutamatergic tor antagonist AM-251, but not by methoctra-
transmission in the dentate gyrus mine, an antagonist presumably selective for
presynaptic (m2) muscarinic acetylcholine recep-
The axon terminals of mossy fibers, perforant path tors. These results suggested that cholinergic fa-
inputs, mossy cells, and back-projecting CA3 py- cilitation of EC release can lead to tonic EC-
ramidal cells all form glutamatergic synapses mediated inhibition of perforant path inputs to
330

dentate granule cells, and as such may be consist- the subgranular region of the adult rat dentate
ent with anatomical studies which indicate low gyrus are immunoreactive for CB1. These striking
levels of CB1 expression in the outer two thirds of findings, coupled with work completed in culture,
the molecular layer. have clearly suggested the hypothesis that EC-me-
There is also very limited available data that diated signaling might modulate neurogenesis in
directly addresses whether ECs play a role in mod- vivo. Consistent with that hypothesis, Jiang et al.
ulating release of glutamate from mossy cell axon (2005) also demonstrated that chronic administra-
terminals. Likely sensitivity to inhibition by CB1 tion of a synthetic cannabinoid (HU210) increased
activation has been suggested, of course, by recent neurogenesis in the adult dentate gyrus, coincident
and striking immunohistochemical findings re- with anxiolytic and antidepressant like effects. At
viewed earlier, but it is also implicated by the present, two other studies have made a compatible
demonstration that WIN55,212-2 sensitive eIPSCs observation that BrdU-labeling of newborn cells in
can be produced in dentate granule cells by stim- the adult dentate gyrus is dramatically reduced in
ulation in the inner molecular layer, and perhaps CB1 knockout animals compared to wild type
most notably by the observation that CB1 receptor mice (Jin et al., 2004; Kim et al., 2006). However,
knockouts that are selective for glutamatergic neu- the full story here is likely to be quite complicated
rons reduce the threshold for kainic acid-induced because apparently paradoxical increases in ne-
seizures (Marsicano et al., 2003; Monory et al., urogenesis produced by administration of CB1
2006). Nevertheless, it is likely that much remains antagonists, including SR141716A and AM-251,
to be learned from a detailed electrophysiological prior to BrdU-labeling have been noted by several
examination of EC-mediated signaling at this laboratories (Rueda et al., 2002; Jin et al., 2004).
synapse. Finally, there is also a complete lack of Further, in at least one case, an AEA-mediated
information in the literature about the EC sensi- reduction in neurogenesis in the dentate gyrus has
tivity of back projecting associational/commis- been reported (Rueda et al., 2002). One possible
sural inputs to hilar neurons or dentate granule explanation is that apparent inhibitory effects of
cells. As the functional consequences of CB1 re- cannabinoids on neurogenesis, when they occur,
ceptor activation on Schaffer collateral inputs to may be CB1 receptor-independent, and perhaps
CA1 pyramidal cells is currently a matter of much mediated by direct activation of VR1. This is con-
debate and investigation (see ‘Introduction’) it will sistent with the observation by Jin et al. (2004) that
not be reviewed further here. SR141716A-mediated increases in neurogenesis
were preserved in CB1 knockouts, but absent in
VR1 knockouts, and also compatible with a much
A role for endocannabinoids in neurogenesis? earlier report indicating that both AEA and
SR141716A may interact with VR1 receptors
In contrast to original conclusions, groundbreak- (Zygmunt et al., 1999). In contrast, the most re-
ing work over the last 10–15 years has clearly cent mechanism suggested for CB1-dependent in-
demonstrated that neurogenesis does occur in se- creases in neurogenesis implicate down regulation
lective regions of the adult mammalian CNS, most of nitric oxide synthase (Di Marzo et al., 2002;
notably in the olfactory bulb and dentate gyrus Kim et al., 2006).
(for recent review, see Lledo et al., 2006). Indeed,
the subgranular zone of the dentate gyrus, previ-
ously recognized for intense expression of CB1 re- Non-CB receptor targets of ECs relevant to the
ceptors in a subset of inhibitory interneurons, has dentate
now also been recognized as a primary neuropro-
liferative zone. Further, Jiang et al. (2005) has used A final topic worthy of particular consideration here
immunohistochemical techniques coupled with in- is the possible abundance of non-CB1 targets
jections of 5-bromo-2-deoxyuridine (BrdU, a base for EC-mediated signaling in the dentate gyrus. Al-
analog of thymidine) to show that dividing cells in though 2-AG has been postulated to be the
331

retrograde messenger in many forms of EC-medi- functional assays (Christopoulos and Wilson,
ated signaling, a role for AEA is far from ruled out. 2001; Lanzafame et al., 2004).
In fact, a role for AEA is implied by the presence of
FAAH in hippocampal principal cells and dentate
granule cells, and has occasionally been implicated Conclusions and future directions
in depolarization-induced as opposed to synaptical-
ly driven release of ECs (for review, see Chevaleyre The data reviewed here indicate the presence of a
et al., 2006). The extent to which EC-mediated robust system for EC-mediated signaling in the
signaling involves AEA is a particularly important dentate gyrus that is clearly composed of both
question in large part because AEA and its met- synthetic and degradative enzymes for end-
abolites, far more than 2-AG, have been associated ocannabinoids as well as cannabinoid receptors.
with action at non-CB receptor sites (for review, see A significant body of work reviewed here has
Di Marzo et al., 2002). largely coalesced to indicate that there is promi-
One such site that may have an important role nent expression of CB1 receptors on the axon ter-
to play in the dentate gyrus is the VR1 receptor. Of minals of CCK-positive basket cells in the dentate,
course, VR1 receptor expression is typically asso- and that these CB1-positive terminals likely form
ciated with sensory fibers, and thus much of the numerous synaptic contacts with both granule
data implicating an interaction of AEA and VR1 is cells and hilar mossy cells. Consistent with that
based on apparent AEA-mediated increases in view, recent work has demonstrated robust mod-
VR1-dependent release of CGRP. However, AEA ulation of GABAergic synapses to both granule
and its analogs have also been shown to directly cells and mossy cells that does indeed depend on
activate recombinant VR1 receptors in several retrograde activation of CB1 receptors by ECs
types of assays and expression systems (Bisogno et (Isokawa and Alger, 2005; Hofmann et al., 2006).
al., 2001; Di Marzo et al., 2001; Ralevic et al., Until very recently, available data suggested that
2001). Further, recent immunohistochemical data we would be unlikely to observe major effects of
has indicated a prominent expression of VR1 in ECs at other types of synapses in the dentate.
the brain, with particularly high levels in several However, dramatic improvements in immuno-
areas, including the dentate gyrus (Toth et al., histochemical techniques have now clearly chal-
2005). Collectively these findings suggest an inter- lenged that conclusion by indicating robust
action between AEA and VR1, which may be rel- expression of CB1 receptors on excitatory termi-
evant in understanding EC-mediated signaling in nals in the inner third of the molecular layer that
the dentate. This argument is further supported, are likely to arise from hilar mossy cells, and that
albeit indirectly, by reports in other areas of the appear to be key players in determining threshold
CNS where VR1 activation by AEA and/or caps- to kainic acid-induced seizures (Marsicano et al.,
aicin has been shown to directly modulate synaptic 2003; Katona et al., 2006; Monory et al., 2006).
transmission (Marinelli et al., 2002, 2003). Other These striking findings have strongly reinforced
potential targets for AEA that could be significant the notion of the dentate gyrus as a key player in
in the dentate include a7-containing nicotinic ace- EC-mediated signaling, and may highlight it as an
tylcholine receptors and 5-HT3 receptors. Inter- attractive system for future research on EC-medi-
estingly, in both these cases, action of AEA has ated modulation of glutamatergic transmission.
been shown to have inhibitory effects on recom- That being said, it seems clear that we still have
binant receptors expressed in Xenopus oocytes (Oz much to learn from a continued characterization
et al., 2002, 2003), while additional lines of evi- of the EC system at the electron microscopic level,
dence implicating cannabinoids in serotonergic particularly when coupled with detailed physio-
function have been previously reviewed (Morales, logical studies at the synaptic level. In fact, a
2006). Further, a potential non-competitive inter- number of issues beyond transmitter phenotype
action between AEA and muscarinic acetylcholine are likely to be of particular importance. These
receptors has been suggested in both binding and include developing a better understanding of
332

physiological stimuli that trigger EC-mediated Bodor, A.L., Katona, I., Nyiri, G., Mackie, K., Ledent, C.,
signaling and more accurate measures of the scope Hajos, N. and Freund, T.F. (2005) Endocannabinoid signa-
ling in rat somatosensory cortex: laminar differences and in-
of the effects. This will likely involve further ex-
volvement of specific interneuron types. J. Neurosci., 25:
amination of the role of postsynaptic metabotrop- 6845–6856.
ic receptors in the release process and careful Braida, D., Pozzi, M. and Sala, M. (2000) CP 55,940 protects
evaluation of potential heterosynaptic and meta- against ischemia-induced electroencephalographic flattening
plastic signaling mechanisms. At a larger level, it and hyperlocomotion in Mongolian gerbils. Neurosci. Lett.,
will also be important to understand broader 296: 69–72.
Cadas, H., di, T.E. and Piomelli, D. (1997) Occurrence and
effects of ECs on network activity, where both in biosynthesis of endogenous cannabinoid precursor, N-ara-
vitro and in vivo studies focused on the dentate chidonoyl phosphatidylethanolamine, in rat brain. J. Neuro-
gyrus will likely be central to elucidating the ne- sci., 17: 1226–1242.
uroprotective and antiepileptic effects of can- Cadas, H., Gaillet, S., Beltramo, M., Venance, L. and Piomelli,
D. (1996) Biosynthesis of an endogenous cannabinoid pre-
nabinoids. Finally, the potential role of ECs in
cursor in neurons and its control by calcium and cAMP. J.
neurogenesis is an extremely important topic that Neurosci., 16: 3934–3942.
clearly has broad implications for neurobiology. Campbell, K.A., Foster, T.C., Hampson, R.E. and Deadwyler,
Although the dentate gyrus has arguably been S.A. (1986a) Delta 9-tetrahydrocannabinol differentially
partially overlooked in the rapid pace of the last 5 affects sensory-evoked potentials in the rat dentate gyrus. J.
years, it now seems clear that those motivated to Pharmacol. Exp. Ther., 239: 936–940.
Campbell, K.A., Foster, T.C., Hampson, R.E. and Deadwyler,
understand EC-mediated signaling in the CNS S.A. (1986b) Effects of delta 9-tetrahydrocannabinol on sen-
have cause to look toward the dentate, while con- sory-evoked discharges of granule cells in the dentate gyrus
versely, those who seek a more thorough under- of behaving rats. J. Pharmacol. Exp. Ther., 239: 941–945.
standing of the dentate may find significant value Chen, K., Ratzliff, A., Hilgenberg, L., Gulyas, A., Freund,
in further studies of cannabinoids. T.F., Smith, M., Dinh, T.P., Piomelli, D., Mackie, K. and
Soltesz, I. (2003) Long-term plasticity of endocannabinoid
signaling induced by developmental febrile seizures. Neuron,
39: 599–611.
References Chevaleyre, V. and Castillo, P.E. (2003) Heterosynaptic LTD of
hippocampal GABAergic synapses: a novel role of end-
Acsady, L., Katona, I., Martinez-Guijarro, F.J., Buzsaki, G. ocannabinoids in regulating excitability. Neuron, 38:
and Freund, T.F. (2000) Unusual target selectivity of peri- 461–472.
somatic inhibitory cells in the hilar region of the rat hippo- Chevaleyre, V. and Castillo, P.E. (2004) Endocannabinoid-medi-
campus. J. Neurosci., 20: 6907–6919. ated metaplasticity in the hippocampus. Neuron, 43: 871–881.
Alger, B.E. (2002) Retrograde signaling in the regulation of Chevaleyre, V., Takahashi, K.A. and Castillo, P.E. (2006)
synaptic transmission: focus on endocannabinoids. Prog. Endocannabinoid-mediated synaptic plasticity in the CNS.
Neurobiol., 68: 247–286. Annu. Rev. Neurosci., 29: 37–76.
Beltramo, M., Stella, N., Calignano, A., Lin, S.Y., Ma- Christopoulos, A. and Wilson, K. (2001) Interaction of anand-
kriyannis, A. and Piomelli, D. (1997) Functional role of amide with the M(1) and M(4) muscarinic acetylcholine re-
high-affinity anandamide transport, as revealed by selective ceptors. Brain Res., 915: 70–78.
inhibition. Science, 277: 1094–1097. Consroe, P. and Wolkin, A. (1977) Cannabidiol: antiepileptic
Berdyshev, E.V. (2000) Cannabinoid receptors and the regula- drug comparisons and interactions in experimentally induced
tion of immune response. Chem. Phys. Lipids, 108: 169–190. seizures in rats. J. Pharmacol. Exp. Ther., 201: 26–32.
Bisogno, T., Hanus, L., De, P.L., Tchilibon, S., Ponde, D.E., Consroe, P.F., Wood, G.C. and Buchsbaum, H. (1975) Antic-
Brandi, I., Moriello, A.S., Davis, J.B., Mechoulam, R. and onvulsant nature of marihuana smoking. JAMA, 234:
Di Marzo, V. (2001) Molecular targets for cannabidiol and 306–307.
its synthetic analogues: effect on vanilloid VR1 receptors and Cravatt, B.F., Giang, D.K., Mayfield, S.P., Boger, D.L.,
on the cellular uptake and enzymatic hydrolysis of anand- Lerner, R.A. and Gilula, N.B. (1996) Molecular character-
amide. Br. J. Pharmacol., 134: 845–852. ization of an enzyme that degrades neuromodulatory fatty-
Blair, R.E., Deshpande, L.S., Sombati, S., Falenski, K.W., acid amides. Nature, 384: 83–87.
Martin, B.R. and Delorenzo, R.J. (2006) Activation of the Cristino, L., De, P.L., Pryce, G., Baker, D., Guglielmotti, V.
cannabinoid type-1 receptor mediates the anticonvulsant and Di Marzo, V. (2006) Immunohistochemical localization
properties of cannabinoids in the hippocampal neuronal cul- of cannabinoid type 1 and vanilloid transient receptor po-
ture models of acquired epilepsy and status epilepticus. J. tential vanilloid type 1 receptors in the mouse brain. Neuro-
Pharmacol. Exp. Ther., 317: 1072–1078. science, 139: 1405–1415.
333

De, P.L., Melck, D., Bisogno, T. and Di Marzo, V. (2000) Egertova, M., Giang, D.K., Cravatt, B.F. and Elphick, M.R.
Endocannabinoids and fatty acid amides in cancer, inflam- (1998) A new perspective on cannabinoid signalling: com-
mation and related disorders. Chem. Phys. Lipids, 108: plementary localization of fatty acid amide hydrolase and the
191–209. CB1 receptor in rat brain. Proc. Biol. Sci., 265: 2081–2085.
Deutsch, D.G., Ueda, N. and Yamamoto, S. (2002) The fatty Engler, B., Freiman, I., Urbanski, M. and Szabo, B. (2006)
acid amide hydrolase (FAAH). Prostaglandins Leukot. Es- Effects of Exogenous and Endogenous Cannabinoids on
sent. Fatty Acids, 66: 201–210. GABAergic Neurotransmission between the Caudate-Put-
Devane, W.A., Hanus, L., Breuer, A., Pertwee, R.G., Steven- amen and the Globus Pallidus in the Mouse. J. Pharmacol.
son, L.A., Griffin, G., Gibson, D., Mandelbaum, A., Etinger, Exp. Ther., 316: 608–617.
A. and Mechoulam, R. (1992) Isolation and structure of a Freund, T.F., Katona, I. and Piomelli, D. (2003) Role of en-
brain constituent that binds to the cannabinoid receptor. dogenous cannabinoids in synaptic signaling. Physiol. Rev.,
Science, 258: 1946–1949. 83: 1017–1066.
Di Marzo, V., Bisogno, T., De, P.L., Brandi, I., Jefferson, R.G., Fukudome, Y., Ohno-Shosaku, T., Matsui, M., Omori, Y.,
Winckler, R.L., Davis, J.B., Dasse, O., Mahadevan, A., Ra- Fukaya, M., Tsubokawa, H., Taketo, M.M., Watanabe, M.,
zdan, R.K. and Martin, B.R. (2001) Highly selective CB(1) Manabe, T. and Kano, M. (2004) Two distinct classes of
cannabinoid receptor ligands and novel CB(1)/VR(1) van- muscarinic action on hippocampal inhibitory synapses: M2-
illoid receptor ‘‘hybrid’’ ligands. Biochem. Biophys. Res. Co- mediated direct suppression and M1/M3-mediated indirect
mmun., 281: 444–451. suppression through endocannabinoid signalling. Eur. J. Ne-
Di Marzo, V., Bisogno, T., De, P.L., Melck, D. and Martin, urosci., 19: 2682–2692.
B.R. (1999) Cannabimimetic fatty acid derivatives: the Gaoni, Y. and Mechoulam, R. (1964) Isolation, structure and
anandamide family and other endocannabinoids. Curr. partial synthesis of an active constituent of hashish. J. Am.
Med. Chem., 6: 721–744. Chem. Soc., 86: 1646–1647.
Di Marzo, V., De, P.L., Fezza, F., Ligresti, A. and Bisogno, T. Glass, M., Dragunow, M. and Faull, R.L. (1997) Cannabinoid
(2002) Anandamide receptors. Prostaglandins Leukot. Es- receptors in the human brain: a detailed anatomical and
sent. Fatty Acids, 66: 377–391. quantitative autoradiographic study in the fetal, neonatal
Di Marzo, V. and Deutsch, D.G. (1998) Biochemistry of the and adult human brain. Neuroscience, 77: 299–318.
endogenous ligands of cannabinoid receptors. Neurobiol. Gulyas, A.I., Cravatt, B.F., Bracey, M.H., Dinh, T.P., Piomelli,
Dis., 5: 386–404. D., Boscia, F. and Freund, T.F. (2004) Segregation of two
Di Marzo, V., Fontana, A., Cadas, H., Schinelli, S., Cimino, endocannabinoid-hydrolyzing enzymes into pre- and postsy-
G., Schwartz, J.C. and Piomelli, D. (1994) Formation and naptic compartments in the rat hippocampus, cerebellum and
inactivation of endogenous cannabinoid anandamide in cen- amygdala. Eur. J. Neurosci., 20: 441–458.
tral neurons. Nature, 372: 686–691. Hajos, N. and Freund, T.F. (2002) Pharmacological separation
Di, S., Boudaba, C., Popescu, I.R., Weng, F.J., Harris, C., of cannabinoid sensitive receptors on hippocampal excitatory
Marcheselli, V.L., Bazan, N.G. and Tasker, J.G. (2005) Ac- and inhibitory fibers. Neuropharmacology, 43: 503–510.
tivity-dependent release and actions of endocannabinoids in Hajos, N., Katona, I., Naiem, S.S., Mackie, K., Ledent, C.,
the rat hypothalamic supraoptic nucleus. J. Physiol., 569: Mody, I. and Freund, T.F. (2000) Cannabinoids inhibit hip-
751–760. pocampal GABAergic transmission and network oscillations.
Dinh, T.P., Carpenter, D., Leslie, F.M., Freund, T.F., Katona, Eur. J. Neurosci., 12: 3239–3249.
I., Sensi, S.L., Kathuria, S. and Piomelli, D. (2002) Brain Hampson, R.E., Foster, T.C. and Deadwyler, S.A. (1989)
monoglyceride lipase participating in endocannabinoid inac- Effects of delta-9-tetrahydrocannabinol on sensory evoked
tivation. Proc. Natl. Acad. Sci. U.S.A., 99: 10819–10824. hippocampal activity in the rat: principal components anal-
Domenici, M.R., Azad, S.C., Marsicano, G., Schierloh, A., ysis and sequential dependency. J. Pharmacol. Exp. Ther.,
Wotjak, C.T., Dodt, H.U., Zieglgansberger, W., Lutz, B. and 251: 870–877.
Rammes, G. (2006) Cannabinoid receptor type 1 located on Harkany, T., Dobszay, M.B., Cayetanot, F., Hartig, W., Siege-
presynaptic terminals of principal neurons in the forebrain mund, T., Aujard, F. and Mackie, K. (2005) Redistribution
controls glutamatergic synaptic transmission. J. Neurosci., of CB1 cannabinoid receptors during evolution of cholinergic
26: 5794–5799. basal forebrain territories and their cortical projection areas:
Egertova, M., Cravatt, B.F. and Elphick, M.R. (2003) Com- a comparison between the gray mouse lemur (Microcebus
parative analysis of fatty acid amide hydrolase and cb(1) murinus, primates) and rat. Neuroscience, 135: 595–609.
cannabinoid receptor expression in the mouse brain: evidence Harris, L.S., Carchman, R.A. and Martin, B.R. (1978) Evi-
of a widespread role for fatty acid amide hydrolase in reg- dence for the existence of specific cannabinoid binding sites.
ulation of endocannabinoid signaling. Neuroscience, 119: Life Sci., 22: 1131–1137.
481–496. Hashimotodani, Y., Ohno-Shosaku, T., Tsubokawa, H., Ogata,
Egertova, M. and Elphick, M.R. (2000) Localisation of can- H., Emoto, K., Maejima, T., Araishi, K., Shin, H.S. and
nabinoid receptors in the rat brain using antibodies to the Kano, M. (2005) Phospholipase Cbeta serves as a coincidence
intracellular C-terminal tail of CB. J. Comp. Neurol., 422: detector through its Ca2+ dependency for triggering retro-
159–171. grade endocannabinoid signal. Neuron, 45: 257–268.
334

Herkenham, M., Lynn, A.B., de Costa, B.R. and Richfield, (2000) GABAergic interneurons are the targets of can-
E.K. (1991a) Neuronal localization of cannabinoid receptors nabinoid actions in the human hippocampus. Neuroscience,
in the basal ganglia of the rat. Brain Res., 547: 267–274. 100: 797–804.
Herkenham, M., Lynn, A.B., Johnson, M.R., Melvin, L.S., de Katona, I., Sperlagh, B., Sik, A., Kafalvi, A., Vizi, E.S., Mac-
Costa, B.R. and Rice, K.C. (1991b) Characterization and kie, K. and Freund, T.F. (1999) Presynaptically located CB1
localization of cannabinoid receptors in rat brain: a quanti- cannabinoid receptors regulate GABA release from axon
tative in vitro autoradiographic study. J. Neurosci., 11: terminals of specific hippocampal interneurons. J. Neurosci.,
563–583. 19: 4544–4558.
Herkenham, M., Lynn, A.B., Johnson, M.R., Melvin, L.S., de Katona, I., Urban, G.M., Wallace, M., Ledent, C., Jung, K.M.,
Costa, B.R. and Rice, K.C. (1991c) Characterization and Piomelli, D., Mackie, K. and Freund, T.F. (2006) Molecular
localization of cannabinoid receptors in rat brain: a quanti- composition of the endocannabinoid system at glutamatergic
tative in vitro autoradiographic study. J. Neurosci., 11: synapses. J. Neurosci., 26: 5628–5637.
563–583. Khaspekov, L.G., Brenz Verca, M.S., Frumkina, L.E., Her-
Hermann, H., Marsicano, G. and Lutz, B. (2002) Coexpression mann, H., Marsicano, G. and Lutz, B. (2004) Involvement of
of the cannabinoid receptor type 1 with dopamine and se- brain-derived neurotrophic factor in cannabinoid receptor-
rotonin receptors in distinct neuronal subpopulations of the dependent protection against excitotoxicity. Eur. J. Neuro-
adult mouse forebrain. Neuroscience, 109: 451–460. sci., 19: 1691–1698.
Hillard, C.J., Edgemond, W.S., Jarrahian, A. and Campbell, Kim, S.H., Won, S.J., Mao, X.O., Ledent, C., Jin, K. and
W.B. (1997) Accumulation of N-arachidonoylethanolamine Greenberg, D.A. (2006) Role for neuronal nitric oxide synt-
(anandamide) into cerebellar granule cells occurs via facili- hase in cannabinoid-induced neurogenesis. J. Pharmacol.
tated diffusion. J. Neurochem., 69: 631–638. Exp. Ther., 319: 150–154.
Hillard, C.J., Wilkison, D.M., Edgemond, W.S. and Campbell, Kirby, M.T., Hampson, R.E. and Deadwyler, S.A. (1995) Can-
W.B. (1995) Characterization of the kinetics and distribution nabinoids selectively decrease paired-pulse facilitation of per-
of N-arachidonylethanolamine (anandamide) hydrolysis by forant path synaptic potentials in the dentate gyrus in vitro.
rat brain. Biochim. Biophys. Acta, 1257: 249–256. Brain Res., 688: 114–120.
Hoffman, A.F., Macgill, A.M., Smith, D., Oz, M. and Lupica, Kreitzer, A.C. and Regehr, W.G. (2001a) Cerebellar depolari-
C.R. (2005) Species and strain differences in the expression of zation-induced suppression of inhibition is mediated by en-
a novel glutamate-modulating cannabinoid receptor in the dogenous cannabinoids. J. Neurosci., 21: RC174.
rodent hippocampus. Eur. J. Neurosci., 22: 2387–2391. Kreitzer, A.C. and Regehr, W.G. (2001b) Retrograde inhibition
Hoffman, A.F., Oz, M., Caulder, T. and Lupica, C.R. (2003) of presynaptic calcium influx by endogenous cannabinoids
Functional tolerance and blockade of long-term depression at excitatory synapses onto Purkinje cells. Neuron, 29:
at synapses in the nucleus accumbens after chronic can- 717–727.
nabinoid exposure. J. Neurosci., 23: 4815–4820. Kujtan, P.W., Carlen, P.L. and Kapur, B.M. (1983) Delta 9-
Hofmann, M.E., Nahir, B. and Frazier, C.J. (2006) End- tetrahydrocannabinol and cannabidiol: dose-dependent
ocannabinoid mediated depolarization-induced suppression effects on evoked potentials in the hippocampal slice. Can.
of inhibition in hilar mossy cells of the rat dentate gyrus. J. J. Physiol. Pharmacol., 61: 420–426.
Neurophysiol., 96: 2501–2512. Lanzafame, A.A., Guida, E. and Christopoulos, A. (2004)
Howard, A.L., Ratzliff, A.D.H. and Soltesz, I. (2003) Explor- Effects of anandamide on the binding and signaling proper-
ing Endocannabinoid signaling in the dentate gyrus. Soc. ties of M1 muscarinic acetylcholine receptors. Biochem.
Neurosci. Abstr., 808.11. Pharmacol., 68: 2207–2219.
Howlett, A.C., Bidaut-Russell, M., Devane, W.A., Melvin, Llano, I., Leresche, N. and Marty, A. (1991) Calcium entry
L.S., Johnson, M.R. and Herkenham, M. (1990) The can- increases the sensitivity of cerebellar Purkinje cells to applied
nabinoid receptor: biochemical, anatomical and behavioral GABA and decreases inhibitory synaptic currents. Neuron,
characterization. Trends Neurosci., 13: 420–423. 6: 565–574.
Isokawa, M. and Alger, B.E. (2005) Retrograde end- Lledo, P.M., Alonso, M. and Grubb, M.S. (2006) Adult ne-
ocannabinoid regulation of GABAergic inhibition in the rat urogenesis and functional plasticity in neuronal circuits. Nat.
dentate gyrus granule cell. J. Physiol., 567: 1001–1010. Rev. Neurosci., 7: 179–193.
Jiang, W., Zhang, Y., Xiao, L., Van, C.J., Ji, S.P., Bai, G. and Lutz, B. (2002) Molecular biology of cannabinoid receptors.
Zhang, X. (2005) Cannabinoids promote embryonic and Prostaglandins Leukot. Essent. Fatty Acids, 66: 123–142.
adult hippocampus neurogenesis and produce anxiolytic- and Maejima, T., Hashimoto, K., Yoshida, T., Aiba, A. and Kano,
antidepressant-like effects. J. Clin. Invest., 115: 3104–3116. M. (2001) Presynaptic inhibition caused by retrograde signal
Jin, K., Xie, L., Kim, S.H., Parmentier-Batteur, S., Sun, Y., from metabotropic glutamate to cannabinoid receptors. Neu-
Mao, X.O., Childs, J. and Greenberg, D.A. (2004) Defective ron, 31: 463–475.
adult neurogenesis in CB1 cannabinoid receptor knockout Maejima, T., Oka, S., Hashimotodani, Y., Ohno-Shosaku, T.,
mice. Mol. Pharmacol., 66: 204–208. Aiba, A., Wu, D., Waku, K., Sugiura, T. and Kano, M.
Katona, I., Sperlagh, B., Magloczky, Z., Santha, E., Kofalvi, (2005) Synaptically driven endocannabinoid release requires
A., Czirjak, S., Mackie, K., Vizi, E.S. and Freund, T.F. Ca2+-assisted metabotropic glutamate receptor subtype 1 to
335

phospholipase C {beta}4 signaling cascade in the cerebellum. Zieglgansberger, W., Wotjak, C.T., Mackie, K., Elphick,
J. Neurosci., 25: 6826–6835. M.R., Marsicano, G. and Lutz, B. (2006) The end-
Mailleux, P., Parmentier, M. and Vanderhaeghen, J.J. (1992) ocannabinoid system controls key epileptogenic circuits in
Distribution of cannabinoid receptor messenger RNA in the the hippocampus. Neuron, 51: 455–466.
human brain: an in situ hybridization histochemistry with Morales, M. (2006) Cannabinoids and the central serotonergic
oligonucleotides. Neurosci. Lett., 143: 200–204. system. In: Onaivi E.S., Sugiura T. and Di Marzo V. (Eds.),
Mailleux, P. and Vanderhaeghen, J.J. (1992) Distribution of Endocannabinoids. CRC Press, Boca Raton, FL, pp.
neuronal cannabinoid receptor in the adult rat brain: a com- 249–260.
parative receptor binding radioautography and in situ hy- Morales, M. and Backman, C. (2002) Coexistence of serotonin
bridization histochemistry. Neuroscience, 48: 655–668. 3 (5-HT3) and CB1 cannabinoid receptors in interneurons of
Marinelli, S., Di Marzo, V., Berretta, N., Matias, I., Maccarr- hippocampus and dentate gyrus. Hippocampus, 12: 756–764.
one, M., Bernardi, G. and Mercuri, N.B. (2003) Presynaptic Mukhtarov, M., Ragozzino, D. and Bregestovski, P. (2005)
facilitation of glutamatergic synapses to dopaminergic neu- Dual Ca2+ modulation of glycinergic synaptic currents in
rons of the rat substantia nigra by endogenous stimulation of rodent hypoglossal motoneurones. J. Physiol., 569: 817–831.
vanilloid receptors. J. Neurosci., 23: 3136–3144. Munro, S., Thomas, K.L. and bu-Shaar, M. (1993) Molecular
Marinelli, S., Vaughan, C.W., Christie, M.J. and Connor, M. characterization of a peripheral receptor for cannabinoids.
(2002) Capsaicin activation of glutamatergic synaptic trans- Nature, 365: 61–65.
mission in the rat locus coeruleus in vitro. J. Physiol., 543: Nagayama, T., Sinor, A.D., Simon, R.P., Chen, J., Graham,
531–540. S.H., Jin, K. and Greenberg, D.A. (1999) Cannabinoids and
Marsicano, G., Goodenough, S., Monory, K., Hermann, H., neuroprotection in global and focal cerebral ischemia and in
Eder, M., Cannich, A., Azad, S.C., Cascio, M.G., Gutierrez, neuronal cultures. J. Neurosci., 19: 2987–2995.
S.O., Van der Stelt, M., Lopez-Rodriguez, M.L., Casanova, Nakatsuka, T., Chen, H.X., Roper, S.N. and Gu, J.G. (2003)
E., Schutz, G., Zieglgansberger, W., Di Marzo, V., Behl, C. Cannabinoid receptor-1 activation suppresses inhibitory
and Lutz, B. (2003) CB1 cannabinoid receptors and on-de- synaptic activity in human dentate gyrus. Neuropharmacol-
mand defense against excitotoxicity. Science, 302: 84–88. ogy, 45: 116–121.
Marsicano, G. and Lutz, B. (1999) Expression of the can- Niehaus, J.L., Wallis, K.T., Elphick, M.R. and Lewis, D.L.,
nabinoid receptor CB1 in distinct neuronal subpopulations in (2004) CRIP1B, a novel CB1 cannabinoid receptor interact-
the adult mouse forebrain. Eur. J. Neurosci., 11: 4213–4225. ing protein, interacts with CB1 at the distal C-terminal tail.
Marsicano, G. and Lutz, B. (2006) Neuromodulatory functions Soc. Neurosci. Abstr., 623.19.
of the endocannabinoid system. J. Endocrinol. Invest., 29: Ohno-Shosaku, T., Hashimotodani, Y., Maejima, T. and Kano,
27–46. M. (2005) Calcium signaling and synaptic modulation: reg-
Marsicano, G., Wotjak, C.T., Azad, S.C., Bisogno, T., Ram- ulation of endocannabinoid-mediated synaptic modulation
mes, G., Cascio, M.G., Hermann, H., Tang, J., Hofmann, C., by calcium. Cell Calcium, 38: 369–374.
Zieglgansberger, W., Di Marzo, V. and Lutz, B. (2002) The Ohno-Shosaku, T., Maejima, T. and Kano, M. (2001) Endog-
endogenous cannabinoid system controls extinction of avers- enous cannabinoids mediate retrograde signals from depo-
ive memories. Nature, 418: 530–534. larized postsynaptic neurons to presynaptic terminals.
Matsuda, L.A., Bonner, T.I. and Lolait, S.J. (1993) Localiza- Neuron, 29: 729–738.
tion of cannabinoid receptor mRNA in rat brain. J. Comp. Ohno-Shosaku, T., Shosaku, J., Tsubokawa, H. and Kano, M.
Neurol., 327: 535–550. (2002a) Cooperative endocannabinoid production by neu-
Matsuda, L.A., Lolait, S.J., Brownstein, M.J., Young, A.C. and ronal depolarization and group I metabotropic glutamate
Bonner, T.I. (1990) Structure of a cannabinoid receptor and receptor activation. Eur. J. Neurosci., 15: 953–961.
functional expression of the cloned cDNA. Nature, 346: Ohno-Shosaku, T., Tsubokawa, H., Mizushima, I., Yoneda,
561–564. N., Zimmer, A. and Kano, M. (2002b) Presynaptic can-
Melis, M., Perra, S., Muntoni, A.L., Pillolla, G., Lutz, B., nabinoid sensitivity is a major determinant of depolarization-
Marsicano, G., Di Marzo, V., Gessa, G.L. and Pistis, M. induced retrograde suppression at hippocampal synapses. J.
(2004a) Prefrontal cortex stimulation induces 2-arachido- Neurosci., 22: 3864–3872.
noyl-glycerol-mediated suppression of excitation in dopa- Ong, W.Y. and Mackie, K. (1999) A light and electron micro-
mine neurons. J. Neurosci., 24: 10707–10715. scopic study of the CB1 cannabinoid receptor in primate
Melis, M., Pistis, M., Perra, S., Muntoni, A.L., Pillolla, G. and brain. Neuroscience, 92: 1177–1191.
Gessa, G.L. (2004b) Endocannabinoids mediate presynaptic Oz, M., Ravindran, A., az-Ruiz, O., Zhang, L. and Morales, M.
inhibition of glutamatergic transmission in rat ventral teg- (2003) The endogenous cannabinoid anandamide inhibits al-
mental area dopamine neurons through activation of CB1 pha7 nicotinic acetylcholine receptor-mediated responses in
receptors. J. Neurosci., 24: 53–62. Xenopus oocytes. J. Pharmacol. Exp. Ther., 306: 1003–1010.
Monory, K., Massa, F., Egertova, M., Eder, M., Blaudzun, H., Oz, M., Zhang, L. and Morales, M. (2002) Endogenous can-
Westenbroek, R., Kelsch, W., Jacob, W., Marsch, R., Ekker, nabinoid, anandamide, acts as a noncompetitive inhibitor on
M., Long, J., Rubenstein, J.L., Goebbels, S., Nave, K.A., 5-HT3 receptor-mediated responses in Xenopus oocytes.
During, M., Klugmann, M., Wolfel, B., Dodt, H.U., Synapse, 46: 150–156.
336

Pettit, D.A., Harrison, M.P., Olson, J.M., Spencer, R.F. and Takahashi, K.A. and Castillo, P.E. (2006) The CB1 can-
Cabral, G.A. (1998) Immunohistochemical localization of the nabinoid receptor mediates glutamatergic synaptic suppres-
neural cannabinoid receptor in rat brain. J. Neurosci. Res., sion in the hippocampus. Neuroscience, 139: 795–802.
51: 391–402. Thomas, B.F., Wei, X. and Martin, B.R. (1992) Characteriza-
Piomelli, D. (2003) The molecular logic of endocannabinoid tion and autoradiographic localization of the cannabinoid
signalling. Nat. Rev. Neurosci., 4: 873–884. binding site in rat brain using [3H]11-OH-delta 9-THC-
Piomelli, D., Beltramo, M., Glasnapp, S., Lin, S.Y., Gout- DMH. J. Pharmacol. Exp. Ther., 263: 1383–1390.
opoulos, A., Xie, X.Q. and Makriyannis, A. (1999) Structural Thomas, E.A., Cravatt, B.F., Danielson, P.E., Gilula, N.B. and
determinants for recognition and translocation by the anand- Sutcliffe, J.G. (1997) Fatty acid amide hydrolase, the degra-
amide transporter. Proc. Natl. Acad. Sci. U.S.A., 96: dative enzyme for anandamide and oleamide, has selective
5802–5807. distribution in neurons within the rat central nervous system.
Pitler, T.A. and Alger, B.E. (1992) Postsynaptic spike firing J. Neurosci. Res., 50: 1047–1052.
reduces synaptic GABAA responses in hippocampal pyram- Toth, A., Boczan, J., Kedei, N., Lizanecz, E., Bagi, Z., Papp,
idal cells. J. Neurosci., 12: 4122–4132. Z., Edes, I., Csiba, L. and Blumberg, P.M. (2005) Expression
Ralevic, V., Kendall, D.A., Jerman, J.C., Middlemiss, D.N. and and distribution of vanilloid receptor 1 (TRPV1) in the adult
Smart, D. (2001) Cannabinoid activation of recombinant and rat brain. Brain Res. Mol. Brain Res., 135: 162–168.
endogenous vanilloid receptors. Eur. J. Pharmacol., 424: Trettel, J. and Levine, E.S. (2003) Endocannabinoids mediate
211–219. rapid retrograde signaling at interneuron right-arrow pyram-
Robbe, D., Kopf, M., Remaury, A., Bockaert, J. and Manzoni, idal neuron synapses of the neocortex. J. Neurophysiol., 89:
O.J. (2002) Endogenous cannabinoids mediate long-term 2334–2338.
synaptic depression in the nucleus accumbens. Proc. Natl. Tsou, K., Brown, S., Sanudo-Pena, M.C., Mackie, K. and
Acad. Sci. U.S.A., 99: 8384–8388. Walker, J.M. (1998a) Immunohistochemical distribution of
Romero, J., Hillard, C.J., Calero, M. and Rabano, A. (2002) cannabinoid CB1 receptors in the rat central nervous system.
Fatty acid amide hydrolase localization in the human central Neuroscience, 83: 393–411.
nervous system: an immunohistochemical study. Brain Res. Tsou, K., Mackie, K., Sanudo-Pena, M.C. and Walker, J.M.
Mol. Brain Res., 100: 85–93. (1999) Cannabinoid CB1 receptors are localized primarily on
Roth, S.H. and Williams, P.J. (1979) The non-specific mem- cholecystokinin-containing GABAergic interneurons in the
brane binding properties of delta9-tetrahydrocannabinol and rat hippocampal formation. Neuroscience, 93: 969–975.
the effects of various solubilizers. J. Pharm. Pharmacol., 31: Tsou, K., Nogueron, M.I., Muthian, S., Sanudo-Pena, M.C.,
224–230. Hillard, C.J., Deutsch, D.G. and Walker, J.M. (1998b) Fatty
Rueda, D., Navarro, B., Martinez-Serrano, A., Guzman, M. acid amide hydrolase is located preferentially in large neu-
and Galve-Roperh, I. (2002) The endocannabinoid anand- rons in the rat central nervous system as revealed by
amide inhibits neuronal progenitor cell differentiation immunohistochemistry. Neurosci. Lett., 254: 137–140.
through attenuation of the Rap1/B-Raf/ERK pathway. J. Van der Stelt, M., Veldhuis, W.B., Bar, P.R., Veldink, G.A.,
Biol. Chem., 277: 46645–46650. Vliegenthart, J.F. and Nicolay, K. (2001a) Neuroprotection
Safo, P.K. and Regehr, W.G. (2005) Endocannabinoids control by delta9-tetrahydrocannabinol, the main active compound
the induction of cerebellar LTD. Neuron, 48: 647–659. in marijuana, against ouabain-induced in vivo excitotoxicity.
Schlicker, E. and Kathmann, M. (2001) Modulation of trans- J. Neurosci., 21: 6475–6479.
mitter release via presynaptic cannabinoid receptors. Trends Van der Stelt, M., Veldhuis, W.B., van Haaften, G.W., Fezza,
Pharmacol. Sci., 22: 565–572. F., Bisogno, T., Bar, P.R., Veldink, G.A., Vliegenthart, J.F.,
Skaper, S.D., Buriani, A., Dal, T.R., Petrelli, L., Romanello, S., Di, M.V. and Nicolay, K. (2001b) Exogenous anandamide
Facci, L. and Leon, A. (1996) The ALIAmide palm- protects rat brain against acute neuronal injury in vivo. J.
itoylethanolamide and cannabinoids, but not anandamide, Neurosci., 21: 8765–8771.
are protective in a delayed postglutamate paradigm of ex- Van der Stelt, M., Veldhuis, W.B., Maccarrone, M., Bar, P.R.,
citotoxic death in cerebellar granule neurons. Proc. Natl. Nicolay, K., Veldink, G.A., Di, M.V. and Vliegenthart, J.F.
Acad. Sci. U.S.A., 93: 3984–3989. (2002) Acute neuronal injury, excitotoxicity, and the end-
Stella, N., Schweitzer, P. and Piomelli, D. (1997) A second en- ocannabinoid system. Mol. Neurobiol., 26: 317–346.
dogenous cannabinoid that modulates long-term potentiat- Van Sickle, M.D., Duncan, M., Kingsley, P.J., Mouihate, A.,
ion. Nature, 388: 773–778. Urbani, P., Mackie, K., Stella, N., Makriyannis, A., Piomelli,
Sugiura, T., Kondo, S., Sukagawa, A., Nakane, S., Shinoda, A., D., Davison, J.S., Marnett, L.J., Di Marzo, V., Pittman, Q.J.,
Itoh, K., Yamashita, A. and Waku, K. (1995) 2-Arachidonoyl- Patel, K.D. and Sharkey, K.A. (2005) Identification and
glycerol: a possible endogenous cannabinoid receptor ligand in functional characterization of brainstem cannabinoid CB2
brain. Biochem. Biophys. Res. Commun., 215: 89–97. receptors. Science, 310: 329–332.
Sugiura, T., Oka, S., Ikeda, S. and Waku, K. (2006) Occur- Varma, N., Carlson, G.C., Ledent, C. and Alger, B.E. (2001)
rence, biosynthesis, and metabolism of endocannabinoids. In: Metabotropic glutamate receptors drive the end-
Onaivi E.S., Sugiura T. and Di Marzo V. (Eds.), End- ocannabinoid system in hippocampus. J. Neurosci., 21:
ocannabinoids. CRC Press, Boca Raton, FL, pp. 177–214. RC188.
337

Wallace, M.J., Blair, R.E., Falenski, K.W., Martin, B.R. and Yanovsky, Y., Mades, S. and Misgeld, U. (2003) Retrograde
Delorenzo, R.J. (2003) The endogenous cannabinoid system signaling changes the venue of postsynaptic inhibition in rat
regulates seizure frequency and duration in a model of tem- substantia nigra. Neuroscience, 122: 317–328.
poral lobe epilepsy. J. Pharmacol. Exp. Ther., 307: 129–137. Yoshida, T., Fukaya, M., Uchigashima, M., Miura, E.,
Wallace, M.J., Martin, B.R. and Delorenzo, R.J. (2002) Evi- Kamiya, H., Kano, M. and Watanabe, M. (2006) Localiza-
dence for a physiological role of endocannabinoids in the tion of diacylglycerol lipase-{alpha} around postsynaptic
modulation of seizure threshold and severity. Eur. J. Pharma- spine suggests close proximity between production site of an
col., 452: 295–301. endocannabinoid, 2-arachidonoyl-glycerol, and presynaptic
Wallace, M.J., Wiley, J.L., Martin, B.R. and Delorenzo, R.J. cannabinoid CB1 receptor. J. Neurosci., 26: 4740–4751.
(2001) Assessment of the role of CB1 receptors in cannabinoid Yoshida, T., Hashimoto, K., Zimmer, A., Maejima, T., Araishi,
anticonvulsant effects. Eur. J. Pharmacol., 428: 51–57. K. and Kano, M. (2002) The cannabinoid CB1 receptor me-
Wilkison, D.M. and Pontzer, N.J. (1987) The actions of THC diates retrograde signals for depolarization-induced suppres-
on the intact hippocampus: a comparison of dentate and sion of inhibition in cerebellar Purkinje cells. J. Neurosci., 22:
CA1 responses. Brain Res. Bull., 19: 63–67. 1690–1697.
Wilson, R.I., Kunos, G. and Nicoll, R.A. (2001) Presynaptic Zhu, P.J. and Lovinger, D.M. (2005) Retrograde end-
specificity of endocannabinoid signaling in the hippocampus. ocannabinoid signaling in a postsynaptic neuron/synaptic
Neuron, 31: 453–462. bouton preparation from basolateral amygdala. J. Neurosci.,
Wilson, R.I. and Nicoll, R.A. (2001) Endogenous cannabinoids 25: 6199–6207.
mediate retrograde signalling at hippocampal synapses. Na- Zygmunt, P.M., Petersson, J., Andersson, D.A., Chuang, H.,
ture, 410: 588–592. Sorgard, M., Di Marzo, V., Julius, D. and Hogestatt, E.D.
Wilson, R.I. and Nicoll, R.A. (2002) Endocannabinoid signa- (1999) Vanilloid receptors on sensory nerves mediate the
ling in the brain. Science, 296: 678–682. vasodilator action of anandamide. Nature, 400: 452–457.
Plate 19.2. DSI in dentate granule cells and hilar mossy cells. (AA) Dentate gyrus of the hippocampal slice (scale, 50 mm). (AB)
Differential interference contrast image of dentate granule cells (scale, 10 mm). (AC) Single dentate granule cell filled, and visualized
with fluorescence microscopy during whole cell recording (scale, 10 mm). (AD) DSI of eIPSCs (left, top) and sIPSCs (right, top) is
observed in dentate granule cells. In both cases, magnitude of DSI is reduced by bath application of thapsigargin (TG, 2 mM, bottom).
(AE) Thapsigargin also reduced the calcium transients produced by DSI-inducing depolarizing voltage steps. Adapted with permission
from Isokawa and Alger (2005). (B) Depolarization induced release of endogenous cannabinoids from hilar mossy cells preferentially
inhibits calcium-dependent exocytosis. DSI in of eIPSCs is apparent in an individual mossy cell. DSI of miniature IPSCs is absent in
the same cell. Following bath application of KCl and CaCl2, DSI of mIPSCs becomes apparent. Summary data are presented in BD,
where numbers on the bars are n values. Adapted with permission from Hofmann et al. (2006). (For B/W version, see page 328 in the
volume.)
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 20

Pro-inflammatory cytokines and their effects


in the dentate gyrus

Mark Pickering and John J. O’Connor

UCD School of Biomolecular and Biomedical Science, Conway Institute of Biomolecular and Biomedical Research,
University College Dublin, Belfield, Dublin 4, Ireland

Abstract: The older notion of a central nervous system existing in essential isolation from the immune
system has changed dramatically in recent years as the body of evidence relating to the interactions between
these two systems has grown. Here we address the role of a particular subset of immune modulatory
molecules, the pro-inflammatory cytokines, in regulating neuronal function and viability in the dentate
gyrus of the hippocampus. These inflammatory mediators are known to be elevated in many neuropath-
ological conditions, such as Alzheimer’s disease, Parkinson’s disease and ischaemic injury that follows
stroke. Pro-inflammatory cytokines, such as tumour necrosis factor-a (TNF-a), interleukin 1-b (IL-1b) and
interleukin 18 (IL-18), have been shown to regulate neurotoxicity; although, due to the complexity of the
cytokine action in neurons and glia, the effect may be either facilitatory or protective, depending on the
circumstances. As well as their role in neurotoxicity and neuroprotection, the pro-inflammatory cytokines
have also been shown to be potent regulators of synaptic function. In particular, TNF-a, IL-1b and IL-18
have all been shown to inhibit long-term potentiation, a form of neuronal plasticity widely believed to
underlie learning and memory, both in the early p38 mitogen activated protein kinase-dependant phase and
the later protein synthesis-dependant phase. In this article we address the mechanisms underlying these
cytokine effects in the dentate gyrus of the hippocampus.

Keywords: TNF-a; IL-1b; IL-18; long-term potentiation; dentate gyrus; mGluR5

Introduction Chun, 2001). Chief amongst these are the cyto-


kines — multifunctional proteins that play crucial
The older notion of two ‘‘super-systems’’, nervous roles in cellular communication and activation.
and immune, existing in relative isolation from Cytokines have been classified as pro-inflamma-
each other due largely to the blood-brain barrier, tory or anti-inflammatory depending on the bal-
has given way in recent years to a new consensus, ance of their effects on the immune system
as it became apparent that many immune mole- (Mosmann et al., 1986). Cytokines may have an
cules may be used by the nervous system in inter- indirect modulatory effect on the nervous system
cellular communication (Boulanger et al., 2001; via their effects on the hypothalamic-pituitary-
adrenal axis (Besedovsky et al., 1991). However,
here we will address the direct effect roles of the
Corresponding author. Tel.: +353 1 716 6765; pro-inflammatory cytokines, in particular tumour
Fax: +353 1 716 7417; E-mail: john.oconnor@ucd.ie necrosis factor-a (TNF-a), interleukin-1b (IL-1b)

DOI: 10.1016/S0079-6123(07)63020-9 339


340

and interleukin-18 (IL-18) on the CNS in general, brain by astrocytes, microglial and some neurons
and the dentate gyrus in particular. (Lieberman et al., 1989; Morganti-Kossman et al.,
1997; Chung and Benveniste, 1990). Under various
pathological conditions, such as trauma, ischemia
Distribution of pro-inflammatory cytokines
and inflammatory diseases, the expression and
in the CNS
release of TNF-a is rapidly increased, in some
cases as early as 1 h after the brain insult and well
Direct action of the pro-inflammatory cytokines
before neuronal death (Liu et al., 1994; Wang et al.,
TNF-a and IL-1b on the CNS has been known for
1994; Allan and Rothwell, 2001)
some time (Plata-Salaman et al., 1988). Elevated
Two different TNFR (p55 and p75) have been
CNS expression of various pro-inflammatory cyto-
identified (Beutler and Van Huffel, 1994a, b;
kines have been noted in many neuropathological
Wajant and Scheurich, 2001) and shown to medi-
situations, both chronic, such as Alzheimer’s disease
ate differential cellular responses using distinct
(Cacquevel et al., 2004) and multiple sclerosis
pathways (Kinouchi et al., 1991; Tartaglia et al.,
(Merrill, 1992), and acute, such as ischemia and
1991). For instance, the signal transduction path-
stroke (Liu et al., 1994; Klein et al., 2000; Yu and
way used by the p55 TNFR results in the activa-
Lau, 2000), and infection (Waage et al., 1989).
tion of the transcription factor nuclear factor
Additionally, pro-inflammatory cytokines are con-
kappaB (NF-kB) (Kolesnick and Golde, 1994;
stitutively expressed in unperturbed cells of the
Goodman and Mattson, 1996; Mattson et al.,
CNS (Benveniste and Benos, 1995; Yu and Lau,
1997a, b) (see Fig. 1). TNF-a activates the NF-kB
2000).
family of transcription factors, which are ubiqui-
While pro-inflammatory cytokine binding sites
tously expressed and are pivotal in controlling
are found throughout the hippocampus, in addi-
diverse cellular processes, including immune re-
tion to other brain regions, there is some evidence
sponses, cell proliferation and differentiation
that some of the functions of these cytokines are
(Israel, 2000; Silverman and Maniatis, 2001;
distinct in different hippocampal regions. For ex-
Ghosh and Karin, 2002; Li and Verma, 2002).
ample, it has been demonstrated that local injec-
Lack of the p50 subunit of NF-kB increases the
tion of IL-1 receptor antagonist (IL-1ra) into the
vulnerability of hippocampal neurons to excito-
dentate gyrus and CA3 regions of the hippocam-
toxicity (Yu et al., 1999). Increasingly, it has
pus leads to a reduction in the febrile response of
become evident that NF-kB also plays important
rats to peripheral injection of lipopolysaccharide
roles in the CNS (O’Neill and Kaltschmidt, 1997)
(Cartmell et al., 1999), whereas IL-1ra injection
and recently Sheridan et al. (2006) have shown
has no effect in the CA1 region of the hippocam-
that TNF-a can selectively upregulate NF-kB
pus. Here, we will address some of the many roles
expression in hippocampal sub-neuronal popula-
played by TNF-a and the pro-inflammatory IL1-
tions. For example in the dentate gyrus and CA3
type cytokines in the dentate gyrus, highlighting
regions there were clear subpopulations with
differences between the dentate gyrus and the
respect to activation of NF-kB. In the dentate
pyramidal cell regions of the hippocampus where
gyrus there is a clear distinction between the outer
appropriate.
4/5th of the stratum granulosum, the cells of which
expressed high nuclear NF-kB activity and the
Action of TNF-a in the dentate gyrus inner 1–2 cell layers that expressed much lower lev-
els of the active form of this transcription factor.
The pro-inflammatory cytokine TNF-a is a One of the most active areas of investigation
17-kDa peptide and forms multimers which are ac- regarding TNF-a is in relation to its role in
tive in binding TNF receptors (TNFR) that are neurotoxicity. Although TNF-a was originally
constitutively expressed on both neurons and glia in named for its degeneration-inducing action in some
the central nervous system (Benveniste and Benos, types of tumor cells, our current understanding of
1995). TNF-a can be synthesized and released in the TNF-a is such that it cannot be simply considered
341

Fig. 1. Putative TNF-a signalling pathway in the dentate gyrus. Downstream signalling by TNF-a involves recruitment by the TNF-a
receptor of the death domain containing adaptor protein, TNF receptor-associated death domain (TRADD) and the TNF receptor-
associated factor 2 (TRAF-2) which then recruits the receptor interacting protein (RIP) eventually leading to activation and trans-
location of NF-kB to the nucleus.

as an inducer or facilitator of cell death. On the one neuronal cell death. For example TNF-a treat-
hand, TNF-a has been shown to induce cell death ment can protect against focal cerebral
in septo-hippocampal cultures. Zhao et al. (2001) ischemia (Nawashiro et al., 1997). Also, while it
detected alpha-spectrin fragments in these septo- has been shown that TNF-a mediates damage to
hippocampal cultures treated with TNF-a and myelin and oligodendrocytes (Selmaj and Raine,
found elevated 120 kDa fragments, indicative of 1998), Garcia et al. (1992) found that it was not
caspase-3 activity, but not 145-kDa fragments, in- toxic to rat-cultured CNS neurons, including those
dicative of calpain activity. Unlike calpain, which is from the spinal cord, ventral mesencephalon, cer-
associated with both necrotic and apoptotic cell ebellum, septum and striatum, in vitro. TNF-a
death, caspase-3 is exclusively characteristic of under in vitro conditions may protect neurons
apoptosis-like cell death (Armstrong et al., 1996; against metabolic, excitotoxic or oxidative insults
Wang et al., 1996; Nath et al., 1998). by promoting maintenance of intracellular Ca2+
On the other hand, a number of investigators homeostasis, suppression of reactive oxygen spe-
have presented evidence that, under different con- cies (Cheng et al., 1994), and by activation of
ditions, TNF-a may play a protective role against transcription factor NF-kB (Barger et al., 1995).
342

Mice genetically deficient in TNF-receptor p55 (for review, see Haydon, 2001; Volterra and
(R1) or both R1 and p75 (R2), also show exacer- Steinhauser, 2004); neuronal–glial interactions
bated neuronal damage compared to wild type are seen as important to many neural processes
controls following middle cerebral artery occlusion and phenomena. Of particular interest in the con-
(Bruce et al., 1996; Gary et al., 1998). text of this review is the relationship between
Some of the reasons for the manifest complexity neurons, glial cells and TNF-a in glutamate ex-
of the relationship between TNF-a and neuronal citotoxicity.
cell death may be due to differences between the
role of TNF-a in the early, acute post-injury
phase, where it appears to be largely detrimental, Role of TNF-a in excitotoxicity
and late post-injury phase, where it may be pro-
tective because it activates cellular repair mecha- Excitotoxicity in general is linked to excessive
nisms. It must also be remembered that there is glutamate activation of receptors, particularly the
some evidence that the different TNF receptors N-methyl-D-aspartate (NMDA) receptor. Cell
may have differing roles in TNF-a toxicity. death resulting from excessive levels of glutamate
It has been known for some time that the p55 and overstimulation of glutamate receptors is
TNF-R1 plays a major role in the anti-bacterial known to be caused by impaired uptake of gluta-
response and sensitivity to septic shock (Pfeffer mate by glial cells (Choi, 1988). In vivo, it has been
et al., 1993), but the role of TNF-R2 was unclear shown that mice lacking expression of the excita-
until the membrane form of TNF receptor was tory amino-acid transporter, EAAT2/GLT-1,
recognized as the physiological activator of this develop epilepsy and increased susceptibility to
TNF receptor (Grell et al., 1995). The more recent acute injury as a result of excessive extracellular
development of TNF-R1 and TNF-R2 knockout glutamate (Tanaka et al., 1997). The expression of
mice has allowed opportunities to tease apart the this transporter has been shown recently to be
roles of the separate receptor pathways. While both positively and negatively regulated by NF-kB
looking at cell death in retinal ischemia-reperfu- (Sitcheran et al., 2005). This study showed that the
sion experiments in these knockout mice, Fontaine increased binding of NF-kB to the EAAT2 pro-
et al. (2002) demonstrated opposing actions of the moter in H4 astroglioma cells was regulated by
two TNF receptors, with TNF-R1 aggravating epidermal growth factor (EGF), but decreased
neuronal damage and TNF-R2 promoting neuro- expression was caused by TNF-a inducing the
protection via an Akt/PKB signal pathway. The classical IkappaB degradation pathway to trigger
work of Fontaine et al. (2002) is in agreement with NF-kB nuclear translocation and DNA binding to
Kassiotis and Kollias (2001), who also showed a repress EAAT2 expression. In this situation, the
dual role for TNF-a in experimental autoimmune presence of elevated TNF-a concentrations leads
encephalomyelitis (EAE), a mouse model for to elevated extracellular glutamate concentration,
multiple sclerosis. Indeed, this differential role thereby increasing the risk of glutamate excitotox-
for the two TNF receptors may be of key impor- icity. Indeed, Zou and Crews (2005) also showed
tance in understanding TNF-a systems. By ‘‘fine that in rat organotypic hippocampal slice cultures,
tuning’’ the balance between the two receptors, the which possess a cytoarchitecture comparable to
exact role of TNF-a in cell damage can be con- that in vivo, TNF-a increased glutamate neuro-
trolled from species to species and tissue to tissue. toxicity. They also demonstrated that the effect was
This makes the TNF-a system very evolutionary mediated by NMDA and not AMPA receptors.
versatile and advantageous. In addition to this, TNF-a and glutamate
Another aspect of the complexity may relate to have also been implicated together recently in
the role of TNF-a in the interactions between the b-amyloid-induced microglia-related cell death.
neurons and glia. Increasingly, glial cells are no Abundant activated microglia are prominent in
longer seen as passive supporters of neurons, but the brains of patients with Alzheimer’s disease
as active participants in information processing (Griffin et al., 1989), and are associated with beta
343

amyloid plaques (Griffin et al., 1995; Frautschy glutamate release induced by activation of the
et al., 1998). It has been proposed that inefficient chemokine receptor CXCR4 is accompanied by
phagocytosis of peptide by microglia could lead to release of TNF-a, and that the TNF-a release is
hyperactivation of cells and release of inflamma- dramatically enhanced by microglia. In light of
tory mediators and neurotoxic factors, thereby this evidence of glial–glial and glial–neuronal inter-
contributing to neurodegenerative processes actions, it is possible to see how a cascading
(Akiyama et al., 2000). It is widely believed that neuronal–glial interaction could occur, leading to
the microglia play a direct role in the neuronal cell death. Activation of TNF receptors on as-
death in Alzheimer’s disease. Floden et al. (2005) trocytes leads to increased extracellular glutamate,
showed that applying media from b-amyloid- which may lead to neurotoxicity in itself, while also
stimulated microglial cultures to neurons led to activating mGlu2 on microglia, which releases more
neuronal cell death that was dependant on the TNF-a, in addition to other pro-inflammatory
synergistic co-activation of TNF-a and NMDA cytokines. Also, it has been known for some time
receptors; the NMDA receptor antagonists meman- that the neurotoxic effects of TNF-a may
tine and 2-amino-5-phosphopetanoic acid as well as be mediated by activation of glutamate AMPA
soluble TNF-R1 applied to the neurons protected receptor subtypes (Gelbard et al., 1993).
them from cell death. Interestingly, blockade of
either the TNF-R1 or NMDA receptors alone was
insufficient to induce neuronal cell death. TNF-a and synaptic transmission
There is, however, evidence that other glutamate
receptors are involved in the relationship between The subject of TNF-a in glial–neuronal interac-
neuronal cell death, glial cells and TNF-a. Taylor tions also emerges when looking at synaptic trans-
et al. (2005) examined the effect of metabotropic mission. Beattie et al. (2002) showed that glial
glutamate (mGlu) receptor stimulation on TNF-a TNF-a causes an increase in surface expression of
release. They found that stimulating rat primary- neuronal AMPA receptors, which would increase
cultured microglial mGlus for 24 h induced micro- synaptic efficacy, and that the removal of endog-
glial activation. This in turn induced caspase-3 enous TNF-a induced a decrease in AMPA recep-
activation in cerebellar granule neurons in culture, tor expression at the cell surface. Further evidence
both those treated with microglial-conditioned for this relationship between TNF-a and AMPA
media as well as in neuronal/microglial co-cul- receptor density came from Stellwagen and Male-
tures. This microglial neurotoxicity was mediated nka (2006), who found the same elevated AMPA
by TNF-a released by the microglia via neuronal receptor density in the presence of elevated
TNF-R1 and caspase-3 activation. Importantly, it TNF-a. This TNF-a-induced AMPA receptor ex-
was the specific group II mGluR agonist 2S,20 R,30 - ocytosis has recently been shown to be mediated
2-(20 ,30 -dicarboxy-cyclopropyl)glycine, which led by activation of TNF-R1 receptors through a PI3
to the release of TNF-a and consequent toxicity; kinase-dependent pathway. (Stellwagen et al.,
N-acetyl-L-aspartyl-L-glutamate, a specific mGlu3 2005). The newly expressed AMPA receptors were
agonist, did not induce microglial activation or shown to have lower stoichiometric amounts
neurotoxicity. TNF-a was only neurotoxic in the of GluR2, making the receptors permeable to
presence of microglia or microglial-conditioned Ca2+ ions. Additionally, this study also showed a
medium; possibly this was due to the presence of surprising concurrent endocytosis of inhibitory
microglial-derived Fas ligand. However, TNF-a GABA receptors.
neurotoxicity was prevented when the neurons
were exposed to conditioned medium from micro-
glia stimulated by the specific group III agonist L- TNF-a and synaptic plasticity
2-amino-4-phosphono-butyric acid.
Glial–glial interactions also influence the effects Changes in neuronal excitability brought about by
of TNF-a. Bezzi et al. (2001) showed that astrocyte TNF-a have important implications for synaptic
344

plasticity (Carroll et al., 2001). Indeed, TNF-a is kinase. Inhibition of p38 MAP kinase with SB
known to act as a regulator of synaptic plasticity 203580 blocked the impairment of the early phase
in the dentate gyrus, in addition to playing a role of LTP by TNF-a, but the late phase was un-
in apoptotic events. As previously mentioned, affected, suggesting that the inhibition of LTP by
elevated levels of TNF-a have been observed in TNF-a is biphasic in character. If the early stage
several neuropathological states that are associ- inhibition of LTP is mediated by a p38 MAP kin-
ated with learning and memory deficits, such as ase pathway, it may be that the late phase inhibi-
Alzheimer’s disease, leading to the search for a tion of LTP by TNF-a is regulated by altered
possible role in plasticity. To this end, much work protein synthesis; not an unlikely possibility given
has been done concerning the role of TNF-a in the the aforementioned regulation of NF-kB by TNF-
hippocampus. TNF-a has been shown to regulate a. As well as a role for the p38 MAPK, (Cumiskey
the development of the hippocampus; TNF-a et al., 2004; Cumiskey and O’ Connor, 2006) also
alpha knockout mice demonstrate decreased provided evidence for a role of metabotropic gluta-
arbourization of the apical dendrites of the CA1 mate receptors in the TNF-a-mediated inhibition of
and CA3 regions while at the same time causing LTP in the dentate gyrus (see Fig. 2).
accelerated development in the dentate gyrus In addition to the effect of TNF-a on LTP, it
(Golan et al., 2004), probably via activation of has been shown that TNF receptor knockout mice
the TNF-R2 receptor, which, as mentioned earlier, demonstrate an impairment of LTD in the CA1
does not lead to caspase-3 activation, but is known region of the hippocampus (Albensi and Mattson,
to transduce the trophic effect of TNF-a by lead- 2000). The effect is mimicked by NF-kB decoy
ing to the activation of several transcription fac- DNA, which implicates the TNF-a/NF-kB signal-
tors (Yang et al., 2002). ling pathway in LTD. Additionally, Wang et al.
Two forms of synaptic plasticity in the hippo- (2006) showed that metabotropic glutamate recep-
campus, long-term potentiation (LTP) and long- tor mediated LTD in the dentate, induced with
term depression (LTD) involve glutamate receptor (RS)-3,5-dihydroxyphenylglycine (DHPG) could
activation and increased intracellular Ca2+ levels, not be induced in TNF-R1 knockout mice.
with LTP induction dependant on the activation Interestingly, the findings relating to the effect of
of Ca2+/calmodulin kinase II, PKC and PKA, and TNF-a on synaptic plasticity seem to have some
LTD on activation of serine/threonine phospha- behavioural correlates in vivo. TNF-a knockout
tases 1, 2A and 2B (Mayford et al., 1995; Coussens mice showed increased performance in spatial mem-
and Teyler, 1996; Silva et al., 1997). LTP is a long- ory and learning as measured in the Morris water
lasting increase in synaptic efficacy, which is maze task when compared to wild type animals
thought to be an important underlying mecha- (Golan et al., 2004). Conversely, Aloe et al. (1999)
nism of learning and memory formation (Bliss and demonstrated a significant impairment in spatial
Collindridge, 1993). learning in two lines of mice that over-expressed
TNF-a has in fact been shown to inhibit LTP in human recombinant TNF-a, also using the same
the CA1 region, as well as the dentate gyrus of water maze task. However, these results must be
the rat hippocampus, when pathophysiological interpreted in the context of the study showing de-
levels of TNF-a are used (Tancredi et al., 1992; velopmental differences mentioned previously; the
Cunningham et al., 1996; Butler et al., 2004; ‘‘substrate’’ of learning and memory is not neces-
Cumiskey et al., 2004, 2007). There are, however, sarily the same in the knockout and wild type mice.
differences in the mechanisms by which TNF-a Recent work has also shown that TNF-a plays a
inhibits early phase (i.e., o120 min post-tetanic crucial role in the development of synaptic scaling;
stimulation) and late phase (>120 min) LTP. But- a form of homeostatic synaptic plasticity that
ler et al. (2002) investigated these differences in the scales synaptic strengths to compensate for pro-
dentate gyrus region of rat hippocampal slices. longed changes in activity. This form of plasticity
They found, using immunohistochemical tech- is essential to stabilize activity across networks
niques, that TNF-a activated the p38 MAP (Turrigiano and Nelson, 2004), as it regulates
345

Fig. 2. Schematic diagram of the role of the mGluR5 and NMDA receptor in mediating the TNF-a effects on LTP. TNF-a inhibits the
early phase of LTP by activation of TNF-R1 (p55) and is dependent on p38 activation. Activation of the G-protein-linked mGluRs
leads to inositol-1,4,5-trisphosphate (IP3) receptor-mediated Ca2+ release via phospholipase C (PLC). TNF-a may also alter Ca2+
concentrations. Combined activation of TNF-R and mGluR5 may lead to the impairment of LTP. TNF-a has negative effects on the
NMDA receptor EPSPs.

synaptic ‘‘gain’’, thus preventing excessive quiet- Action of IL-1b in the dentate gyrus
ening or activity across the network. Stellwagen
and Malenka (2006) showed, using TNF receptor IL-1b receptors
knockout mice, that signalling from the constitu-
tive expression of TNF-a was not necessary for the As described for TNF-a, IL-1 receptors have also
induction of LTP or LTD, but was necessary for been shown to be present in many brain regions,
the development of synaptic scaling in the longer with high levels in the hippocampus and hypo-
term. Additionally, they showed that the source of thalamus (Ban et al., 1991). There are a number of
this TNF-a was not neuronal but was glial. This receptors to which IL-1 can bind (O’Neill, 1997).
finding is not surprising when one considers that The principal mediator of IL-1 signalling in the
the glia are in a unique position to sense the overall CNS is believed to be the type I IL-1 receptor
level of local network activity (presumably using (IL-1R1), which initiates intracellular events upon
the index of glutamate overflow from synapses) IL-1b binding (O’Neill, 1996), although the type II
and adjust the activity accordingly. IL-1 receptor (IL-1R2) is also present, which
While Stellwagen and Malenka (2006) did not although having a similar affinity for IL-1b
directly implicate TNF-a in synaptic scaling in the to the type I receptor, has no ability to initiate
dentate gyrus, the hippocampal slice component signal transduction events, and so may serve as a
of their study instead focusing on the CA1 region ‘‘decoy’’ receptor, attenuating IL-1b-induced sig-
in intact hippocampal slices, the evidence from nalling. Other receptors are the IL-1 receptor
dissociated hippocampal cultures (presumably accessory protein (IL-RAcp) that partakes in
containing a cohort of granule neurons) would IL-1R1 signalling and the IL-1 receptor-related
seem to suggest that TNF-a may also play a role in protein (IL-1Rcp), which is the receptor for IL-18,
the development of synaptic scaling in the dentate a member of the IL-1 family (Okamura et al.,
gyrus. 1998), and will be discussed in detail later.
346

The presence of IL-1 receptors in brain was MAP kinases. The three MAP kinase cascades: the
shown in 1990 by Takao et al. using 125I-labelled p42/44 MAP kinases (also known as extracellular
IL-1a (both IL-1a and IL-1b can activate IL-1 re- regulated kinases — ERKs); the p38 MAP kinases
ceptors) binding to crude membrane preparations (also known as reactivating kinase — RK) and jun
of mouse hippocampus. Their autoradiographic kinases (JNK; also known as p54 or stress-
localization studies revealed very low densities of activated protein kinases — SAPKs) have all been
[1251]-IL-1a binding sites throughout the brain, implicated in IL-1 signalling.
with highest densities present in the molecular and
granular layers of the dentate gyrus of the hippo-
campus. This finding was later confirmed by fur- IL-1b and synaptic plasticity
ther binding studies (Haour et al., 1990; Ban et al.,
1991) and then by affinity cross-linking of IL-1a IL-1b has been found to exert a number of effects
(Parnet et al., 1994). Further characterization of in the CNS, and has been implicated in a number
the distribution of IL-1 receptors in brain utilized of processes such as centrally-mediated fever, slow
antisense cRNA against the type I IL-1 receptor wave sleep patterns, appetite suppression and
in mouse brain (Cunningham et al., 1992), and neurodegeneration (for reviews, see Rothwell and
RT-PCR (Parnet et al., 1994). These studies Hopkins, 1995; Rothwell, 1999). Oitzl et al. (1993)
demonstrate the presence of both type I and type showed that IL-1b can disrupt the acquisition of
II IL-1 receptors in mouse hippocampus, with spatial learning in the Morris water maze via
particularly high expression in the cell-dense gran- a mechanism which seemed to be independent of
ule cell layer of the dentate gyrus. IL-1b’s pyrogenic properties. More recently, it was
Much like TNF-R2 signalling, IL-1 signalling shown that IL-1R knockout mice exhibited deficits
leads to a change in gene expression — up to 90 in a number of learning paradigms, as well as in
genes have been shown to be affected by IL-1b LTP, suggesting that there is a physiological role for
(O’Neill and Green, 1998). These changes include IL-1b in learning and memory, and possibly the in-
upregulation of mRNA transcription for a number duction of synaptic plasticity (Avital et al., 2003);
of cytokines, growth factors, adhesion molecules although, as with the TNFR knockout studies,
and acute-phase proteins. Again, similar to TNF- the evidence is confounded by possible effects of
a, IL-1b regulates gene transcription is by activa- cytokine receptor deficits on development. The first
tion of the transcription factor NF-kB. IL-1 binds reported action of IL-1b on synaptic plasticity was
to IL-1R1, which is complexed with the IL-1 Racp. inhibition of NMDA receptor-independent LTP
This in turn leads to recruitment of the IL-1 in the mouse mossy-fibre-CA3 pathway in vitro
receptor-associated kinase (IRAK; O’Neill and (Katuski et al., 1990). Concentrations as low as
Green, 1998). IRAK then associates with TRAF6 50 pg/ml were found to produce a significant
(tumour necrosis factor (TNF)-receptor-associated inhibition of LTP, and this inhibitory effect
factor), causing activation of IcB kinase (IKK). was blocked by Lys-D-Pro-Thr, a tripeptide ana-
IKK activation leads to phosphorylation of IKB, logue of IL-1b. No effects on baseline transmission
which marks it for ubiquitination and subsequent were observed. In the Schaeffer-collateral-CA1
degradation (O’Neill, 1997; O’Neill and Green, pathway, Bellinger et al. (1993) reported an inhi-
1998). The p50-p65 heterodimer, now free of 1cB, bition by IL-1b of the potentiation of both the
can translocate to the nucleus, bind to its nuclear field EPSP and the population spike induced by
KB site and activate gene transcription, leading to high-frequency stimulation. IL-1b treatment was
expression of proteins such as IL-6, cycloxygenase- also reported to cause a transient decrease in pop-
2, inducible-nitric oxide synthase, IL-1b, TNF-a ulation spike amplitude and a persistent decrease
and interferons (O’Neill and Kaltschmidt, 1997; in EPSP slope. Bellinger et al. (1993) reported
see Fig. 3). preliminary results that suggested that treatment
One family of protein kinases that have been of slices with the IL-1ra led to the induction of a
very strongly implicated in IL-1 signalling are the larger LTP than control conditions. A later study
347

Fig. 3. Putative IL-1b signalling pathway in the dentate gyrus. In IL-1 signalling, the signal-mediating IL-1 receptor type I (IL-1R1)
forms a heterodimer with a second molecule the IL-1 receptor accessory protein (IL-1RAcP). Binding of IL-1b to this receptor complex
leads to activation of the transcription factor NFB through different signalling molecules. DNA-binding motifs for NFB are found in
the promotors and enhancers of many genes that are known to be activated upon inflammation.

of the MPP-dentate granule cell pathway showed Schneider et al. (1998) and Coogan et al. (1999) have
an inhibition of the induction of LTP by treatment shown that the maintenance phase of LTP can be
with IL-1b, and this effect was antagonized by co- blocked by administration of IL-1ra. This suggests
application of IL-1ra, and again, as in the CA3 a requirement for IL-1b in the sustained expression
region, IL-1b was found to have no effect on of LTP, i.e., the endogenous IL-1b production by
baseline transmission in the dentate gyrus tetanic stimulation is required for sustained expres-
(Cunningham et al., 1996). This IL-1b-induced sion of LTP, while, somewhat paradoxically, the
inhibition of LTP was accompanied by a decrease presence of elevated IL-1b can inhibit the initial
in Ca2+ influx into slices compared with tetanized induction of LTP. It seems that it is the sequence
slices, suggesting that IL-1b inhibits LTP in the and timing of IL-1b changes that is important in
dentate gyrus by inhibiting the Ca2+ influx neces- regulating its effect on LTP.
sary for LTP induction. In the dentate gyrus, the IL-1b-induced inhibition
Schneider et al. (1998) showed that in the CA1 of LTP in slices was found to be concomi-
region, expression of IL-1b transcripts was signifi- tant with a decrease in Ca2+ influx following tetanic
cantly upregulated 1 h following tetanization both in stimulation, either pre- or postsynaptically, as
vitro and in vivo, which may explain why judged from synaptosomal preparations (Cunning-
Bellinger et al. (1993) found that treatment of slices ham et al., 1996). Further investigations into the
solely with IL-1ra produces a larger potentiation actions of IL-1b in the dentate gyrus showed that
of EPSPs than in untreated slices. Along with the treatment of slices with IL-1b caused a large de-
increased gene expression of IL-1b during LTP, pression of the pharmacologically isolated NMDA
348

receptor EPSP (NMDA-EPSP; Coogan and O’Con- both LTP induction and NMDA-EPSPs in the
nor, 1997). These effects were thought to be specific dentate gyrus (Coogan et al., 1999). SB203580 did
for IL-1R1, because the depression was fully antag- cause a small but significant increase in baseline
onized by IL-1ra. This might explain the decrease in transmission. It is possible that there may be a high
Ca2+ influx, possibly through the NMDA receptor constitutive expression of p38 MAP kinase activity
channel, and so may serve to inhibit LTP induction in the hippocampus, which may exert an inhibitory
by preventing sufficient Ca2+ influx via the NMDA effect on synaptic transmission, and its inhibition
receptor to trigger the intracellular processes that may lead to the observed increase in synaptic trans-
underlie the expression of LTP (Bliss and Collind- mission. When the effects of a lower dose of
ridge, 1993). The fact that IL-1b seems to have an PD98059 (10 mM, a dose that does not inhibit
effect on NMDA receptor-mediated EPSPs suggests LTP) were examined on the IL-1b inhibition of
a postsynaptic locus of action for IL-1b in the dent- LTP, it was found that it did not antagonize both
ate gyrus. The inhibitory effects of IL-1b have not the inhibition of LTP or the inhibition of the iso-
been restricted to tetanically induced LTP since it lated NMDA receptor-mediated EPSP (Coogan et
has been shown that IL-1b can inhibit a tetraethyl- al., 1999). In addition to p38 MAPK, a role for c-
ammonium-induced synaptic potentiation in the rat jun n-terminal kinase in the inhibition of LTP by
dentate gyrus (Coogan and O’Connor, 1999). IL-1b and long-term depression has also been dem-
IL-1b has also been shown to inhibit the release of onstrated (Curran et al., 2003). A role for MAP
glutamate from synaptosomes prepared from hippo- kinases in LTD has also been shown in the
campi of young rats, via a mechanism that seemed to CA3–CA1 synapse (Bolshakov et al., 2000).
be coupled to phospholipase A2(Murray et al., 1997).
However, IL-1b was not observed to have any effect
on glutamate release from synaptosomes and hippo- IL-1b and Ca2+ channels
campi from aged animals (Murray et al., 1997;
Lynch, 1998b) The inhibition of LTP by IL-1b in While, as previously mentioned, IL-1 receptor acti-
vivo has been proposed to be mediated by a decrease vation leads to altered gene expression, since most
in glutamate release and also via an induced increase of the inhibitory effects of IL-1b in the hippocampus
in the formation of reactive oxygen radicals, leading have been reported following acute treatment of
to an increase in lipid peroxidation and a decrease in slices, it seems unlikely that gene transcription plays
arachidonic acid levels (Murray and Lynch, 1998). a role in the observed effects. Therefore it seems
As arachidonic acid has been shown to act as a more likely that IL-1b mediates its effects by acti-
retrograde (postsynaptic-to-presynaptic) messenger vation of cytoplasmic cascades with cytoplasmic/
during LTP induction (Medina and Izquierdo, membrane bound substrates. The mechanism by
1995), this has been suggested as a mechanism of which IL-1b exerts its effects on hippocampal LTP
IL-1b-induced inhibition of LTP in the dentate gyrus. remains controversial, although a number of possi-
As discussed above, the mitogen-activated protein bilities have been raised. In dissociated guinea-pig
(MAP) kinases have been implicated in the induc- CA1 pyramidal neurones it has been shown that IL-
tion of LTP. PD98059, a specific inhibitor of MAP 1b rapidly induces a depression of inward, voltage-
kinase kinase (MEK) and thus, p42/44 MAP kinase dependent Ca2+ currents, (both peak and late cur-
activation (Pang et al., 1995), has been shown to rent) in a dose-dependent fashion from 0.197 to
block induction of LTP in the area CA1 as well as 3.1 ng/ml. This could be reversed by pretreatment
the dentate gyrus of the hippocampus (English and with IL-1ra (Plata-Salaman and Ffrench-Mullen,
Sweatt, 1996; Coogan et al., 1999). When these 1992), indicating a specific effect on IL-1R1.
effects were further investigated to try to elucidate Further investigation of this effect showed it to be
the intracellular pathway through which IL-1b inhibited by a non-hydrolysable GTP analogue and
exerts its action on LTP, it was found that pre- pertussis toxin, and also by protein kinase C inhib-
incubation of slices with the p38 MAP kinase in- itors H-7 and staurosporine (Plata-Salaman and
hibitor SB203580 inhibited the effects of IL-1b on Ffrench-Mullen, 1994), suggesting that the effect on
349

Ca2+ channels is mediated by G-proteins and pro- et al., 1999). Again, like TNF-a and IL-1b, acute
tein kinase C, although H-7 and staurosporine can application of IL-18 has been shown to inhibit the
also inhibit other kinases, such as protein kinase A induction of LTP in the dentate gyrus, without
and Ca2+-calmodulin kinase (Hikada et al., 1991). affecting baseline neurotransmission (Curran and
IL-1ra was found to completely block (excess of 25) O’Connor, 2001, 2003; Cumiskey et al., 2007).
the PKC-dependent IL-1b induced decrease in Ca2+ Another similarity between IL-1b and IL-18 is the
currents. It is therefore likely that IL-1b interacts observation that IL-18 also activates p38 MAP
with the IL-1R1 membrane site. It must be noted, kinase (Thomassen et al., 1998), which, as men-
however, that low concentrations of IL-1b (3.5 ng/ tioned above, is implicated in the IL-1b inhibition
ml) have been shown to decrease intracellular Ca2+ of LTP. Recent work has also shown interesting
concentrations while higher levels (100 ng/ml) mark- similarities between the mechanism of IL-18 inhi-
edly increase it (Campbell and Lynch, 1998). bition of LTP and the mechanism by which TNF-a
inhibits LTP. Cumiskey et al. (2007) showed that,
like TNF-a, IL-18 inhibition of LTP involves
mGluRs. The study suggested that IL-18 inhibits
IL-18 LTP in the dentate gyrus by an LTD-like mech-
anism. In these experiments, the application of the
IL-18 is a member of the IL-1 pro-inflammatory group II mGluR antagonist MTPG prevented the
cytokine family, closely related to IL-1b (Bazan inhibition of LTP by IL-18. Group II and III
et al., 1996), and, like other cytokines, can be mGluRs are negatively coupled to adenylyl cyclase
detected in many brain regions, including the hip- (Tanabe et al., 1992), and their activation can induce
pocampus (Culhane et al., 1998). It is expressed synaptic depression (Manahan-Vaughan and Rey-
mainly in glia (both astrocytes and microglia) and mann, 1997), thus suggesting that IL-18 might be
is secreted in response to LPS stimulation (Conti acting through an LTD-like mechanism.

Fig. 4. Schematic diagram of the role of the mGluRs and IL-18R in LTP in the dentate gyrus. IL-18 inhibits the early phase of LTP by
activation of IL-18R. Activation of the G-protein-linked mGluR5 leads to inositol-1,4,5-trisphosphate (IP3) receptor-mediated Ca2+
release via phospholipase C (PLC). Activation of the group II mGluRs leads to inhibition of cAMP. p38 has also been implicated in
IL-18 mediated inhibition of LTP.
350

An interesting aspect of the involvement of References


mGluRs in the effect of IL-18 on synaptic trans-
mission and plasticity comes from an apparently Akiyama, H., Barger, S., Barnum, S., Bradt, B., Bauer, J., Cole,
G.M., Cooper, N.R., Eikelenboom, P., Emmerling, M., Fie-
contradictory study. Kanno et al. (2004) found that
bich, B.L., Finch, C.E., Frautschy, S., Griffin, W.S., Hampel,
IL-18 facilitates basal synaptic transmission as a H., Hull, M., Landreth, G., Lue, L., Mrak, R., MacKenzie,
result of stimulating presynaptic glutamate release I.R., McGeer, P.L., O’Banion, M.K., Pachter, J., Pasinetti,
and enhancing AMPA receptor responses in the G., Plata-Salaman, C., Rogers, J., Rydel, R., Shen, Y., Streit,
CA1, but had no significant effect on LTP in that W., Strohmeyer, R., Tooyoma, I., Van Muiswinkel, F.L.,
region. The key difference would appear to be the Veerhuis, R., Walker, D., Webster, S., Wegrzyniak, B.,
Wenk, G. and Wyss-Coray, T. (2000) Inflammation and
region of the hippocampus under investigation. The Alzheimer’s disease. Neurobiol. Aging, 21: 383–421.
expression of mGluRs is known to be very different Albensi, B.C. and Mattson, M.P. (2000) Evidence for the
in the CA1 and dentate regions (Fotuhi et al., involvement of TNF-a and NF-kappaB in hippocampal
1994), with high levels of mGluR2, mGluR5 and synaptic plasticity. Synapse, 35: 151–159.
mGluR7 and low levels of mGluR1, mGluR3 and Allan, S.M. and Rothwell, N.J. (2001) Cytokines and acute
neurodegeneration. Nat. Rev. Neurosci., 2: 734–744.
mGluR4 in the dentate, while the CA1 exhibits low Aloe, L., Properzi, F., Probert, L., Akassoglou, K., Kassiotis,
mGluR5 and high mGluR1 distribution. This G., Micera, A. and Fiore, L. (1999) Learning abilities, NGF
would seem to suggest that the effect of IL-18 on and BDNF brain levels in two lines of TNFalpha transgenic
synaptic transmission and plasticity is dependant in mice, one characterized by neurological disorders, the other
phenotypically normal. Brain Res., 840: 125–137.
some way in the balance between mGluR1 and
Armstrong, R.C., Aja, T., Xiang, J., Gaur, S., Krebs, J.F.,
mGluR5 signalling. The exact nature of this rela- Hoang, K., Bai, X., Korsmeyer, S.J., Karanewsky, D.S.,
tionship will remain unknown until further studies Fritz, L.C. and Tomaselli, K.J. (1996) Fas-induced activation
can shed light on the signalling paths from IL-18 to of the cell death-related protease CPP32 is inhibited by Bcl-2
the mGluRs, which would also provide further in- and by ICE family protease inhibitors. J. Biol. Chem., 271:
sight into the complex nature of IL-18 signalling in 16850–16855.
Avital, A., Goshen, I., Kamsler, A., Segal, M., Iverfeldt, K.,
the hippocampus (Fig. 4). Richter-Levin, G. and Yirmiya, R. (2003) Impaired inter-
leukin-1 signaling is associated with deficits in hippocampal
memory processes and neural plasticity. Hippocampus, 13:
Concluding remarks 826–834.
Ban, E., Milon, G., Prunhom, N., Fillion, G. and Haour, F.
(1991) Receptors for interleukin-1 (a and b) in mouse brain:
The overriding theme among studies of pro-inflam- mapping and neuronal localization in hippocampus. Neuro-
matory cytokines in the dentate gyrus, and other science, 43: 21–30.
brain regions, is that of complexity. A primary ex- Barger, S.W., Horster, D., Furukawa, K., Goodman, Y.,
ample is the potential for opposing actions. In ad- Krieglstein, J. and Mattson, M.P. (1995) Tumor necrosis
dition, the action of these cytokines on neurons is factors alpha and beta protect neurons against amyloid beta-
peptide toxicity: evidence for involvement of a kappa
often subtle and multi-faceted, with very different B-binding factor and attenuation of peroxide and Ca2+ ac-
short- and long-term consequences, making their cumulation. Proc. Natl. Acad. Sci. U.S.A., 92: 9328–9332.
elucidation all the more challenging. A more com- Bazan, J.F., Timans, J.C. and Kastelein, R.A. (1996) A newly
prehensive understanding cannot be achieved until defined interleukin-1b. Nature, 379: 591.
the interactions between the cytokine receptors and Beattie, E.C., Stellwagen, D., Morishita, W., Bresnahan, J.C.,
Ha, B.K., Von Zastrow, M., Beattie, M.S. and Malenka,
other neuronal mechanisms are better understood. R.C. (2002) Control of synaptic strength by glial TNFalpha.
Science, 295: 2282–2285.
Bellinger, F.P., Madamba, S. and Siggins, G.R. (1993)
Acknowledgements Interleukin-1b inhibits synaptic strength and long-term
potentiation in the rat CAI hippocampus. Brain Res., 628:
We would like to thank the Health Education 227–234.
Benveniste, E.N. and Benos, D.J. (1995) TNF-alpha- and IFN-
Authority, Ireland, PRTLI Cycle 3 for funding
gamma-mediated signal transduction pathways: effects on
and Dr Derval Cumiskey for assistance on figure glial cell gene expression and function. FASEB J., 9:
production. 1577–1584.
351

Besedovsky, H.O., del Rey, A., Klusman, I., Furukawa, H., Chun, J. (2001) Selected comparison of immune and nervous
Monge Arditi, G. and Kabiersch, A. (1991) Cytokines as system development. Adv. Immunol., 77: 297–322.
modulators of the hypothalamus-pituitary-adrenal axis. Chung, I.Y. and Benveniste, E.N. (1990) Tumor necrosis
J. Steroid Biochem. Mol. Biol., 40: 613–618. factor-alpha production by astrocytes. Induction by lip-
Beutler, B. and Van Huffel, C. (1994a) An evolutionary and opolysaccharide, IFN-gamma, and IL-1 beta. J. Immunol.,
functional approach to the TNF receptor/ligand family. Ann. 144: 2999–3007.
N.Y. Acad. Sci., 730: 118–133. Conti, B., Park, L.C., Calingasan, N.Y., Kim, Y., Kim, H.,
Beutler, B. and Van Huffel, C. (1994b) Unraveling function in Bae, Y., Gibson, G.E. and Joh, T.H. (1999) Cultures of
the TNF ligand and receptor families. Science, 264: 667–668. astrocytes and microglia express interleukin 18. Brain Res.
Bezzi, P., Domercq, M., Brambilla, L., Galli, R., Schols, D., De Mol. Brain Res., 67: 46–52.
Clercq, E., Vescovi, A., Bagetta, G., Kollias, G., Meldolesi, J. Coogan, A. and O’Connor, J.J. (1997) Inhibition of NMDA
and Volterra, A. (2001) CXCR4-activated astrocyte gluta- receptor-mediated synaptic transmission in the rat dentate
mate release via TNF: amplification by microglia triggers gyrus in vitro by IL-1b. Neuroreport, 8(9): 2107–2110.
neurotoxicity. Nat. Neurosci., 4: 702–707. Coogan, A. and O’Connor, J.J. (1999) Interleukin-1b inhibits
Bliss, T.V.P. and Collindridge, G.L. (1993) A synaptic model of a tetraethylammonium-induced synaptic potentiation in the
memory: long-term potentiation in the hippocampus. Nature, rat dentate gyrus in vitro. Eur. J. Pharmacol., 374(2):
361: 31–38. 197–206.
Bolshakov, V.Y., Carboni, L., Cobb, M.H., Siegelbaum, S.A. Coogan, A., O’Neill, L.A.J. and O’Connor, J.J. (1999) The p38
and Belardetti, F. (2000) Dual MAP kinase pathways medi- MAP kinase inhibitor SB203580 antagonises the inhibitory
ate opposing forms of long-term plasticity at CA3–CA1 effect of interleukin-1b on long-term potentiation in the rat
synapses. Nat. Neurosci., 3: 1107–1112. dentate gyrus in vitro. Neuroscience, 93(1): 57–69.
Boulanger, L.M., Huh, G.S. and Shatz, C.J. (2001) Neuronal Coussens, C.M. and Teyler, T.J. (1996) Protein kinase and
plasticity and cellular immunity: shared molecular mecha- phosphatase activity regulate the form of synaptic plasticity
nisms. Curr. Opin. Neurobiol., 11: 568–578. expressed. Synapse, 24: 97–103.
Bruce, A.J., Boling, W., Kindy, M.S., Peschon, J., Kraemer, Culhane, A.C., Hall, M.D., Rothwell, N.J. and Luheshi, G.N.
P.J., Carpenter, M.K., Holtsberg, F.W. and Mattson, M.P. (1998) Cloning of rat brain interleukin-18 cDNA. Mol.
(1996) Altered neuronal and microglial responses to excito- Psychiatry, 3: 362–366.
toxic and ischemic brain injury in mice lacking TNF recep- Cumiskey, D., Butler, M.P., Moynagh, P.N. and O’Connor, JJ.
tors. Nat. Med., 2: 788–794. (2004) The inhibitory effect of tumour necrosis factor-a on
Butler, M., O’Connor, J.J. and Moynagh, P. (2004) Dissection long-term potentiation is attenuated by type 1 metabotropic
of TNF-a inhibition of LTP reveals a p38 MAPK-dependent glutamate receptor blockade. J. Physiol., 560P: C23.
mechanism which maps to early but not late-phase LTP. Cumiskey, D. and O’ Connor, J.J. (2006) A role for mGluRs
Neuroscience, 124: 319–326. and Ca2+ in the inhibitory effect of tumour necrosis factor-a
Butler, M.P., Moynagh, P.N. and O’Connor, J.J. (2002) on long-term potentiation. Proc. Br. Pharmacol. Soc., 3(4):
Methods of detection of the transcription factor NF-kappa 055P.
B in rat hippocampal slices. J. Neurosci. Methods, 119: Cumiskey, D., Pickering, M. and O’Connor, J.J. (2007)
185–190. Interleukin-18 mediated inhibition of LTP in the rat dentate
Cacquevel, M., Lebeurrier, N., Cheenne, S. and Vivien, D. gyrus is attenuated in the presence of mGluR antagonists.
(2004) Cytokines in neuroinflammation and Alzheimer’s dis- Neurosci. Lett., 412: 206–210.
ease. Curr. Drug Targets, 5(6): 529–534. Cunningham, A.J., Murray, C.A., O’Neill, L.A., Lynch, M.A.
Campbell, V. and Lynch, M.A. (1998) Biphasic modulation and O’Connor, J.J. (1996) Interleukin-1 beta (IL-1 beta)
of [Ca2] by interleukin-1b in rat cortical synatosomes. and tumour necrosis factor (TNF) inhibit long-term po-
Involvement of a PTX-sensitive G-protein and p42 MAP tentiation in the rat dentate gyrus in vitro. Neurosci. Lett.,
kinase. Neuroreport, 9: 9–12. 203: 17–20.
Carroll, R.C., Beattie, E.C., von Zastrow, M. and Malenka, Cunningham, E.T., Wada, E., Carter, D.B., Tracey, D.E.,
R.C. (2001) Role of AMPA receptor endocytosis in synaptic Battery, J.F. and De Souza, E.B. (1992) In situ histochemical
plasticity. Nat. Rev. Neurosci., 2(5): 315–324. localization of type I interleukin-1 messenger RNA in the
Cartmell, T., Luheshi, G.N. and Rothwell, N.J. (1999) Brain central nervous system, pituitary and adrenal gland of the
sites of action of endogenous interleukin-1 in the febrile mouse. J. Neurosci., 12: 1101–1114.
response to localized inflammation in the rat. J. Physiol., Curran, B. and O’Connor, J.J. (2001) The novel pro-inflam-
518(2): 585–594. matory cytokine interleukin-18 (IL-18) inhibits long-term
Cheng, B., Christakos, S. and Mattson, M.P. (1994) Tumor potentiation in the rat hippocampus in vitro. Neuroscience,
necrosis factors protect neurons against excitotoxic/ 108(1): 83–90.
metabolic insults and promote maintenance of Ca2+ home- Curran, B. and O’Connor, J.J. (2003) The inhibition of long-
ostasis. Neuron, 12: 139–153. term potentiation by pro-inflammatory cytokines is attenu-
Choi, D.W. (1988) Glutamate neurotoxicity and diseases of the ated in the presence of nicotine. Neurosci. Lett., 344:
nervous system. Neuron, 1: 623–634. 103–106.
352

Curran, B.P., Murray, H.J. and O’Connor, J.J. (2003) A role Griffin, W.S.T., Stanley, L.C., Ling, C., MacLeod, V., Perrot,
for c-jun n-terminal kinase in the inhibition of long-term L.J., White III, C.L., et al. (1989) Brain interleukin-1 and S100
potentiation by interleukin-1beta and long-term depression immunoreactivity are elevated in Alzheimer disease and in
in the rat dentate gyrus in vitro. Neuroscience, 118: Down syndrome. Proc. Natl. Acad. Sci. U.S.A., 86: 7611–7615.
347–357. Haour, F.G., Ban, E.M., Milon, G.M., Baran, D. and Fillion,
English, J.D. and Sweatt, J.D. (1996) Activation of p42 G.M. (1990) Brain interleukin-1 receptors: characterization
MAPkinase in hippocampal LTP. J. Biol. Chem., 271: and modulation after lipopolysaccharide injection. Prog.
24329–24332. Neuroendrochronol., 3: 196–204.
Floden, A.M., Li, S. and Combs, C.K. (2005) Beta-amyloid- Haydon, P.G. (2001) Glia: listening and talking to the synapse.
stimulated microglia induce neuron death via synergistic Nat. Rev. Neurosci., 2: 185–193.
stimulation of tumor necrosis factor alpha and NMDA Hikada, H., Watanabe, M. and Kobayashi, R. (1991) Proper-
receptors. J. Neurosci., 25: 2566–2575. ties and use of H-series compounds as protein kinase inhib-
Fontaine, V., Mohand-Said, S., Hanoteau, N., Fuchs, C., itors. Metab. Enzymol., 201: 328–336.
Pfizenmaier, K. and Eisel, U. (2002) Neurodegenerative and Israel, A. (2000) The IKK complex: an integrator of all signals
neuroprotective effects of tumor necrosis factor (TNF) in that activate NF-kappaB? Trends Cell Biol., 10: 129–133.
retinal ischemia: opposite roles of TNF receptor 1 and TNF Kanno, T., Nagata, T., Yamamoto, S., Okamura, H. and
receptor 2. J. Neurosci., 22: 1–7. Nishizaki, T. (2004) Interleukin-265 18 stimulates synaptic-
Fotuhi, M., Standaert, D.G., Testa, C.M., Penney Jr., J.B. ally released glutamate and enhances postsynaptic AMPA
and Young, A.B. (1994) Differential expression of metabo- receptor responses in the CA1 region of mouse hippocampal
tropic glutamate receptors in the hippocampusand ent- slices. Brain Res., 1012: 190–193.
orhinal cortex of the rat. Brain Res. Mol. Brain Res., Kassiotis, G. and Kollias, G. (2001) Uncoupling the proin-
21(3–4): 283–292. flammatory from the immunosuppressive properties of
Frautschy, SA., Yng, F., Irrizarry, M., Hyman, B., Saido, T.C., tumor necrosis factor (TNF) at the p55 TNF receptor level:
Hsiao, K. and Cole, G.M. (1998) Microglial response to implications for pathogenesis and therapy of autoimmune
amyloid plaques in APPsw transgenic mice. Am. J. Pathol., demyelination. J. Exp. Med., 193: 427–434.
152: 307–317. Katuski, H., Nakai, S., Hirai, Y., Akaji, K.I., Kiso, Y. and
Garcia, J.E., Nonner, D., Ross, D. and Barrett, J.N. (1992) Satoh, M. (1990) Interleukin-l, inhibits long-term potentiat-
Neurotoxic components in normal serum. Exp. Neurol., 118: ion in the CA3 region of mouse hippocampal slices. Eur. J.
309–316. Pharmacol., 181: 323–326.
Gary, D.S., Bruce-Keller, A.J., Kindy, M.S. and Mattson, M.P. Kinouchi, K., Brown, G., Pasternak, G. and Donner, D.B.
(1998) Ischemic and excitotoxic brain injury is enhanced in (1991) Identification and characterization of receptors for
mice lacking the p55 tumor necrosis factor receptor. J. Cereb. tumor necrosis factor-alpha in the brain. Biochem. Biophys.
Blood Flow Metab., 18: 1283–1287. Res. Commun., 181: 1532–1538.
Gelbard, H.A., Dzenko, K.A., DiLoreto, D., del Cerro, C., Klein, B.D., White, H.S. and Callahan, K.S. (2000) Cytokine
del Cerro, M. and Epstein, L.G. (1993) Neurotoxic effects of and intracellular signaling regulation of tissue factor expres-
tumor necrosis factor alpha in primary human neuronal sion in astrocytes. Neurochem. Int., 36: 441–449.
cultures are mediated by activation of the glutamate AMPA Kolesnick, R. and Golde, D.W. (1994) The sphingomyelin
receptor subtype: implications for AIDS neuropathogenesis. pathway in tumor necrosis factor and interleukin-1 signaling.
Dev. Neurosci., 15: 417–422. Cell, 77: 325–328.
Ghosh, S. and Karin, M. (2002) Missing pieces in the Li, Q. and Verma, I.M. (2002) NF-kappaB regulation in the
NF-kappaB puzzle. Cell, 109(Suppl.): S81–S96. immune system. Nat. Rev. Immunol., 2: 725–734.
Golan, H., Levav1, T., Mendelsohn, A. and Huleihel, M. (2004) Lieberman, A.P., Pitha, P.M., Shin, H.S. and Shin, M.L.
Involvement of tumor necrosis factor alpha in hippocampal (1989) Production of tumor necrosis factor and other
development and function. Cereb. Cortex, 14: 97–105. cytokines by astrocytes stimulated with lipopolysaccharide
Goodman, Y. and Mattson, M.P. (1996) Ceramide protects or a neurotropic virus. Proc. Natl. Acad. Sci. U.S.A., 86:
hippocampal neurons against excitotoxic and oxidative 6348–6352.
insults, and amyloid beta-peptide toxicity. J. Neurochem., Liu, T., Clark, R.K., McDonnell, P.C., Young, P.R., White,
66: 869–872. R.F., Barone, F.C. and Feuerstein, G.Z. (1994) Tumor
Grell, M., Wajant, H., Löhden, M., Maxeiner, B., Georgopou- necrosis factor-alpha expression in ischemic neurons. Stroke,
los, S., Kollias, G., Lesslauer, W., Pfizenmaier, K. and 25: 1481–1488.
Scheurich, P. (1995) The transmembrane form of tumor Lynch, M.A. (1998) Age-related impairment in long-term
necrosis factor (TNF) is the prime activating ligand of the 80 potentiation in hippocampus: a role for the cytokine, inter-
kDa TNF receptor. Cell, 83: 793–802. leukin-1 beta? Prog. Neurobiol., 56: 571–589.
Griffin, W.S.T., Sheng, J.G., Roberts, G.W. and Mrak, R.E. Manahan-Vaughan, D. and Reymann, K.G. (1997) Group 1
(1995) Interleukin-1 expression in different plaque types metabotropic glutamate receptors contribute to slow-onset
in Alzheimer’s disease, significance in plaque evolution. potentiation in the rat CA1 region in vivo. Neuropharma-
J. Neuropathol. Exp. Neurol., 54. cology, 36: 1533–1538.
353

Mattson, M.P., Barger, S.W., Furukawa, K., Bruce, A.J., O’Neill, L.A.J. and Green, C. (1998) Signal transduction path-
Wyss-Coray, T., Mark, R.J. and Mucke, L. (1997a) Cellular ways activated by the IL-1 receptor family: ancient signalling
signaling roles of TGF beta, TNF alpha and beta APP in machinery in mammals, insects and plants. J. Leukoc. Biol.,
brain injury responses and Alzheimer’s disease. Brain Res. 63: 650–657.
Brain Res. Rev., 23: 47–61. Oitzl, M.S., van Oers, H., Schobitz, B. and de Kloet, E.R.
Mattson, M.P., Goodman, Y., Luo, H., Fu, W. and Furukawa, (1993) Interleukin-1 beta, but not interleukin-6, impairs spa-
K. (1997b) Activation of NF-kappaB protects hippocampal tial navigation learning. Brain Res., 613: 160–163.
neurons against oxidative stress-induced apoptosis: evidence Okamura, H., Tsutsui, H., Kashiwamura, S., Yoshimoto, T.
for induction of manganese superoxide dismutase and sup- and Nakanishi, K. (1998) Interleukin-18: a novel cytokine
pression of peroxynitrite production and protein tyrosine that augments both innate and acquired immunity. Adv.
nitration. J. Neurosci. Res., 49: 681–697. Immunol., 70: 281–312.
Mayford, M., Wang, J., Kandel, E.R. and O’Dell, T.J. (1995) Pang, L., Sawada, T., Deckers, S.J. and Saltiel, A.R. (1995)
CaMKII regulates the frequency response function of hip- Inhibition of MAP kinase kinase blocks the differentiation of
pocampal synapses for the production of both LTD and PC-12 cells induced by nerve growth factor. J. Biol. Chem.,
LTP. Cell, 81: 891–904. 270: 13585–13588.
Medina, J.H. and Izquierdo, I. (1995) Retrograde messengers, Parnet, P., Amindari, S., Wu, C., Brunke-Reese, D., Goujon,
long-term potentiation and memory. Behav. Brain Res., 58: E., Weyhenmeyer, J.A., Dantzer, R. and Kelley, K.W. (1994)
1–8. Expression of type I and type II interleukin-I receptors in
Merrill, J.E. (1992) Pro-inflammatory and antiinflammatory mouse brain. Mol. Brain Res., 27: 63–70.
cytokines in multiple sclerosis and central nervous system Pfeffer, K., Matsuyama, T., Kunidg, T.M., Wakeham, A.,
acquired immunodeficiency syndrome. J. Immunother., 12: Kishihara, K., Shahinian, A., Wiegmann, K., Ohashi, P.S.,
167–170. Krönke, M. and Mak, T.W. (1993) Mice deficient for the 55 kd
Morganti-Kossman, M.C., Lenzlinger, P.M., Hans, V., Stahel, tumor necrosis factor receptor are resistant to endotoxic shock,
P., Csuka, E., Ammann, E., Stocker, R., Trentz, O. and yet succumb to L. monocytogenes infection. Cell, 73: 457–467.
Kossmann, T. (1997) Production of cytokines following brain Plata-Salaman, C.R. and Ffrench-Mullen, J.M.H. (1992)
injury: beneficial and deleterious for the damaged tissue. Interleukin-lb depresses calcium currents in CA 1 hippocam-
Mol. Psychiatry, 2: 133–136. pal neurons at pathophysiological concentrations. Brain Res.
Mosmann, T.R., Cherwinski, H., Bond, M.W., Giedlin, M.A. Bull., 29: 221–223.
and Coffman, R.L. (1986) Two types of murine helper T Plata-Salaman, C.R. and Ffrench-Mullen, J.M.H. (1994) Inter-
cell clone. I. Definition according to profiles of lymphokine leukin-lb inhibits Ca2, channel currents in hippocampal neu-
activities and secreted proteins. J. Immunol., 136: rons through protein kinase C. Eur. J. Pharmacol., 266: 1–10.
2348–2357. Plata-Salaman, C.R., Oomura, Y. and Kai, Y. (1988) Tumor
Murray, C.A. and Lynch, M.A. (1998) Evidence that increased necrosis factor and interleukin-1 beta: suppression of food
hippocampal expression of the cytokine interleukin-1b is a intake by direct action in the central nervous system. Brain
common trigger for age- and stress-induced impairments in Res., 448: 106–114.
long-term potentiation. J. Neurosci., 18: 2974–2981. Rothwell, N. (1999) Cytokines — killers in the brain. J.
Murray, C.A., McGahon, B., McBennet, S. and Lynch, M.A. Physiol., 514: 3–17.
(1997) Interleukin-l/1 inhibits glutamate release in hippo- Rothwell, N.J. and Hopkins, S.J. (1995) Cytokines and the
campus of young, but not aged, rats. Neurobiol. Ageing, 18: nervous system II: actions and mechanisms of action. Trends
343–348. Neurosci., 18: 130–136.
Nath, R., Robert, A., McGinnis, K.M. and Wang, K.K.W. Schneider, H., Prossi, F., Balschun, D., Wagner, A., Del Rey,
(1998) Evidence for activation of caspase-3-like protease in A. and Besedovsky, H.O. (1998) A neuromodulatory role of
excitotoxins and hypoxia/hypoglycemia-injured cerebrocor- interleukin-1b in the hippocampus. Proc. Natl. Acad. Sci.
tical neurons. J. Neurochem., 71: 186–195. U.S.A., 95: 7778–7783.
Nawashiro, H., Tasaki, K., Ruetzler, C.A. and Hallenbeck, Selmaj, K.W. and Raine, C.S. (1998) Tumor necrosis factor
J.M. (1997) TNF-alpha pretreatment induces protective mediates myelin and oligodendrocyte damage in vitro. Ann.
effects against focal cerebral ischemia in mice. J. Cereb. Neurol., 23: 339–346.
Blood Flow Metab., 17: 483–490. Sheridan, G., Pickering, M., Moynagh, P.N., O’Connor, J.J. and
O’Neill, L.A. and Kaltschmidt, C. (1997) NF-kappa B: a crucial Murphy, K.J. (2006) Tumour necrosis factor-a selectively up-
transcription factor for glial and neuronal cell function. regulates nuclear factor kb expression in hippocampal sub-
Trends Neurosci., 20: 252–258. neuronal populations. Proc. Br. Pharmacol. Soc., 3(4): 113P.
O’Neill, L.A.J. (1996) Pharmacological targets in the immune Silva, A.J., Smith, A.M. and Giese, K.P. (1997) Gene targeting
response. Interleukin-l receptors and signal transduction. and the biology of learning and memory. Annu. Rev. Genet.,
Biochem. Soc. Trans., 24: 207–211. 31: 527–546.
O’Neill, L.A.J. (1997) Towards an understanding of the signal Silverman, N. and Maniatis, T. (2001) NF-kappaB signaling
transduction pathway for interleukin-1. Biochim. Biophys. pathways in mammalian and insect innate immunity. Genes
Acta, 1266: 31–44. Dev., 15: 2321–2342.
354

Sitcheran, R., Gupta, P., Fisher, P.B. and Baldwin, A.S. (2005) Waage, A., Halstensen, A., Shalaby, R., Brandtzaeg, P.,
Positive and negative regulation of EAAT2 by NF-kappaB: a Kierulf, P. and Espevik, T. (1989) Local production of
role for N-myc in TNFalpha-controlled repression. EMBO tumor necrosis factor alpha, interleukin 1, and interleukin 6
J., 24(3): 510–520. in meningococcal meningitis. Relation to the inflammatory
Stellwagen, D., Beattie, E.C., Seo, J.Y. and Malenka, R.C. response. J. Exp. Med., 170: 1859–1867.
(2005) Differential regulation of AMPA receptor and Wajant, H. and Scheurich, P. (2001) Tumor necrosis factor
GABA receptor trafficking by tumor necrosis factor-alpha. receptor associated factor (TRAF) 2 and its role in TNF
J. Neurosci., 25: 3219–3228. signaling. Int. J. Biochem. Cell Biol., 33: 19–32.
Stellwagen, D. and Malenka, R.C. (2006) Synaptic scaling Wang, K.K.W., Nath, R., Raser, K.J. and
mediated by glial TNF-alpha. Nature, 440: 1054–1059. Hajimohammadreza, I. (1996) Maitotoxin induces calpain
Tanabe, Y., Masu, M., Ishii, T., Shigemoto, R. and Nakanishi, activation in SH-SY5Y neuroblastoma cells and cerebrocor-
S. (1992) A family of metabotropic glutamate receptors. tical cultures. Arch. Biochem. Biophys., 331: 208–214.
Neuron, 8: 169–179. Wang, Q., Chang, L., Rowan, M.J. and Anwyl, R. (2006)
Tanaka, K., Watase, K., Manabe, T., Yamada, K., Watanabe, Developmental dependence, the role of the kinases p38
M., Takahashi, K., Iwama, H., Nishikawa, T., Ichihara, N., MAPK and PKC, and the involvement of tumor necrosis
Kikuchi, T., Okuyama, S., Kawashima, N., Hori, S., Taki- factor-R1 in the induction of mGlu-5 LTD in the dentate
moto, M. and Wada, K. (1997) Epilepsy and exacerbation gyrus. Neuroscience, 144: 110–118.
of brain injury in mice lacking the glutamate transporter Wang, X., Yue, T.L., Barone, F.C., White, R.F., Gagnon, R.C.
GLT-1. Science, 276: 1699–1702. and Feuerstein, G.Z. (1994) Concomitant cortical expression
Tancredi, V., D’Arcangelo, G., Grassi, F., Tarroni, P., of TNF-alpha and IL-1 beta mRNAs follows early response
Palmieri, G., Santoni, A. and Eusebi, F. (1992) Tumor ne- gene expression in transient focal ischemia. Mol. Chem.
crosis factor alters synaptic transmission in rat hippocampal Neuropathol., 23: 103–114.
slices. Neurosci. Lett., 146: 176–178. Yang, L., Lindholm, K., Konishi, Y., Li, R. and Shen, Y.
Tartaglia, L.A., Weber, R.F., Figari, I.S., Reynolds, C., (2002) Target depletion of distinct tumor necrosis factor
Palladino Jr., M.A. and Goeddel, D.V. (1991) The two receptor subtype reveals hippocampal neuron death and
different receptors for tumor necrosis factor mediate distinct survival through different signal transduction pathways.
cellular responses. Proc. Natl. Acad. Sci. U.S.A., 88: J. Neurosci., 22: 3025–3032.
9292–9296. Yu, A.C. and Lau, L.T. (2000) Expression of interleukin-1
Taylor, D.L., Jones, F., Kubota, E.S. and Pocock, J.M. (2005) alpha, tumor necrosis factor alpha and interleukin-6 genes in
Stimulation of microglial metabotropic glutamate receptor astrocytes under ischemic injury. Neurochem. Int., 36:
mGlu2 triggers tumor necrosis factor alpha-induced neuro- 369–377.
toxicity in concert with microglial-derived Fas ligand. J. Yu, Z., Zhou, D., Bruce-Keller, A.J., Kindy, M.S. and
Neurosci., 25(11): 2952–2964. Mattson, M.P. (1999) Lack of the p50 subunit of nuclear
Thomassen, E., Bird, T.A., Renshaw, B.R., Kennedy, M.K. factor-kappaB increases the vulnerability of hippocampal
and Sims, J.E. (1998) Binding of interleukin-18 to the neurons to excitotoxic injury. J. Neurosci., 19: 8856–8865.
interleukin-1 receptor homologous receptor IL-1Rrp1 Zhao, X., Bausano, B., Pike, B.R., Newcomb-Fernandez, J.K.,
leads to activation of signaling pathways similar to those Wang, K.K., Shohami, E., Ringger, N.C., DeFord, S.M.,
used by interleukin-1. J. Interferon Cytokine Res., 18: Anderson, D.K. and Hayes, R.L. (2001) TNF-alpha stimu-
1077–1088. lates caspase-3 activation and apoptotic cell death in
Turrigiano, G.G. and Nelson, S.B. (2004) Homeostatic plastic- primary septo-hippocampal cultures. J. Neurosci. Res., 64:
ity in the developing nervous system. Nat. Rev. Neurosci., 5: 121–131.
97–107. Zou, J.Y. and Crews, F.T. (2005) TNF alpha potentiates gluta-
Volterra, A. and Steinhauser, C. (2004) Glial modulation of mate neurotoxicity by inhibiting glutamate uptake in organ-
synaptic transmission in the hippocampus. Glia, 47(3): otypic brain slice cultures: neuroprotection by NF kappa B
249–257. inhibition. Brain Res., 1034(1–2): 11–24.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 21

Role of corticosteroid hormones in the dentate gyrus

Marian Joëls

Swammerdam Institute of Life Sciences, Center for NeuroScience (SILS-CNS), University of Amsterdam,
Kruislaan 320, 1098 SM Amsterdam, The Netherlands

Abstract: Dentate granule cells are enriched with receptors for the stress hormone corticosterone, i.e., the
high-affinity mineralocorticoid receptor (MR), which is already extensively occupied with low levels of the
hormone, and the glucocorticoid receptor (GR), which is particularly activated after stress. More than any
other cell type in the brain studied so far, dentate granule cells require hormone levels to be within the
physiological range. In the absence of corticosteroids, proliferation and apoptotic cell death are dramat-
ically enhanced. Dendritic morphology and synaptic transmission are compromised. Conversely, prolonged
exposure of animals to a high level of corticosterone suppresses neurogenesis and presumably makes
dentate granule cells more vulnerable to delayed cell death. These corticosteroid effects on dentate cell and
network function are translated into behavioral consequences, in health and disease.

Keywords: chronic stress; adrenalectomy; mineralocorticoid receptor; glucocorticoid receptor;


neurogenesis; apoptosis; electrophysiology

Systems activated by stress latter causes an array of behavioral changes in-


cluding increased vigilance, alertness, focused
When organisms are exposed to a stressful situa- attention, and selection of appropriate strategies,
tion, sensory information about the situation is which together result in behavioral adaptation.
processed in the brain, and output regarding the Moreover, the hormones help to store relevant
event is forwarded to the hypothalamus (de Kloet information in brain, for future use.
et al., 1998, 2005; see Fig. 1). From there two The initial phases of behavioral adaptation pri-
hormonal systems are activated, aimed to face the marily involve fast-acting catecholamines, i.e.,
challenge and eventually normalize the disturbed mainly noradrenaline, neuropeptides such as
balance: the sympatho-adrenomedullar system, corticotrophin releasing hormone and vasopressin
which leads to the release of adrenaline from the (de Kloet et al., 2005), and possibly neurosteroids
adrenal medulla; and the hypothalamo-pituitary- (Stell et al., 2003). Moreover, it has become clear
adrenal system, which results in elevated levels of in recent years that corticosterone may also play a
the adrenocortical hormone (corticosterone in ro- role in these early phases, through non-genomic
dents, cortisol in humans). Adrenaline and corti- pathways (Di et al., 2003; Karst et al., 2005). The
costerone are not only active in peripheral organs main role of corticosteroids, though, is to change
but can also strongly affect brain function. The brain function through slow and long-lasting
actions. These long-lasting actions are mediated
Corresponding author. Tel.: +31-20-5257626; by corticosteroid hormone receptors, of which
Fax: +31-20-5257709; E-mail: joels@science.uva.nl two main types have been recognized in the brain

DOI: 10.1016/S0079-6123(07)63021-0 355


356

Fig. 1. Exposure of a rat to stress may activate many brain regions (depending on the type of stressor), including the amygdala (Amy),
hippocampus (Hipp) and prefrontal cortex (PFC). These areas project to the hypothalamus (HYP). Stimulation of cells in the
hypothalamus leads to the activation of the fast-acting sympatho-adrenomedullar system (lower right) and the slower-acting hypo-
thalamo-pituitary-adrenal axis (lower left). Both systems not only affect the function of peripheral organs but also feed back to the
brain, via adrenaline and corticosterone respectively. Adrenaline can, via intermediate steps involving the nucleus tractus solitarius,
give rise to central release of noradrenaline (NA) from the locus coeruleus (LC), which then exerts widespread influence on other areas
such as the amygdala, prefrontal cortex and hippocampus. Corticosterone is distributed throughout the brain but acts only at those
sites where receptors are enriched. Two receptor types are known in the brain, i.e., the mineralocorticoid (MR) and the glucocorticoid
receptor (GR). Principal cells in the hippocampal CA1 area and the dentate gyrus (DG; see inset) are among the few cells in the brain
that express MR (gray dots) as well as GR (black dots). CA3 neurons primarily express MR. In most other parts of the brain, GRs
prevail. ANS ¼ autonomic nervous system; ACTH ¼ adrenocorticotropin hormone; CRH ¼ corticotropin releasing hormone.
Adapted with permission from Elsevier (Joëls et al., 2006).

(Reul and de Kloet, 1985). The first type, i.e., the corticosterone is converted into a biologically
mineralocorticoid receptor (MR), is identical inactive 11-keto isoform, by the enzyme 11-b-
to the protein found in peripheral organs like the hydroxysteroid dehydrogenase type 2 (Seckl and
kidney where it is involved in mineral balance. Walker, 2003). In most brain regions this enzyme
This receptor displays a very high affinity for is not present during adulthood, so corticosterone
the mineralocorticoid aldosterone as well as for (which is present in a 100-fold excess over aldos-
corticosterone (Kd0.5 nM). In kidney tissue, al- terone) under normal conditions is the most likely
dosterone preferentially binds to the MR, since occupant of MRs in brain. MRs show a restricted
357

distribution within the brain, with high abundance hormones, or when animals have been exposed to
in some limbic areas, like the hippocampal subre- excess corticosteroid levels (e.g., due to chronic
gions CA1, CA3 and dentate gyrus, the lateral stress). Each of these conditions will be addressed
septum and central amygdala, and in motor nuclei in the following sections.
of the brain stem (de Kloet et al., 1998). The sec-
ond receptor type in brain is the glucocorticoid
receptor (GR), which has an approximately 10- Dentate function in the absence of corticosteroid
fold lower affinity for corticosterone than the MR. hormones
GRs have a relatively low affinity for aldosterone,
but effectively bind synthetic steroids like dexa- Cell turnover and morphology
methasone. This receptor is much more ubiquitous
than the MR and can be found in nearly all The dentate gyrus is quite unique in that it requires
brain regions — both in neurons and glial cells — corticosteroid receptor activation to prevent apop-
although some areas show a high enrichment, tosis and to preserve neuronal integrity. Nearly
such as the paraventricular nucleus of the hypo- two decades ago it was reported for the first time
thalamus, the hippocampal CA1 region, and the that complete removal of corticosterone by adre-
dentate gyrus. Recent studies have shown that nalectomy (ADX) can lead to substantial loss of
both the MR and GR gene give rise to several dentate granule cells (Sloviter et al., 1989; Gould
splice variants, and the proteins are known to have et al., 1990), due to a process showing all the hall-
multiple isoforms due to alternative translation marks of apoptosis (Sloviter et al., 1993a). The
initiation sites as well as abundant posttransla- degree of apoptosis varies greatly among individ-
tional modifications caused by acetylation, phos- ual animals, but rarely involves more than 25% of
phorylation, ubiquitination and sumoylation the neurons in the first week after ADX. Interest-
(Pascual-Le Tallec and Lombes, 2005; Zhou and ingly, about one out of five animals do not show
Cidlowski, 2005). The functional relevance of any apoptosis upon ADX (Sloviter et al., 1993b;
these variants in brain, though, is presently not Stienstra et al., 1998). It is still not fully under-
understood. stood why these animals escape from apoptosis,
As is evident from the above overview, some but most likely these animals still have some re-
cells in brain express both MR and GR. This is for sidual corticosterone producing cells (despite the
instance the case for granule cells in the dentate fact that circulating corticosterone levels are below
gyrus. The degree of activation of the two receptor the detection limit). Approximately 3 days are re-
types, however, depends on the available levels of quired for cells to fully undergo the apoptosis pro-
the hormone. When the organism is at rest and gram upon removal of corticosterone (Hu et al.,
circadian release of corticosteroids is at its nadir, 1997). Substitution with a low dose of corticoster-
MRs are already activated to a substantial degree one, which preferentially occupies the MR, or with
by corticosterone, but occupation of GRs by corti- aldosterone can fully prevent apoptosis (Woolley
costerone is below 20% (de Kloet et al., 1998). et al., 1991; Stienstra et al., 1998), pointing to a
After exposure to stress or at the circadian crucial role of the MR in preventing apoptosis.
peak, GRs will become increasingly occupied by Why some cells undergo apoptosis but neigh-
corticosterone. Under physiological conditions, boring cells that are seemingly comparable stay
corticosteroid receptor occupancy in dentate healthy is not easy to understand. Presumably,
granule cells will therefore range from a situation relatively ‘young’ granule cells are more resistant
of predominant MR activation to a situation to apoptosis (Cameron and Gould, 1996). We
where both receptor types are fully occupied. tested the hypothesis that surviving cells differ in
Electrophysiological studies over the past decades their gene expression profile triggered by ADX,
have investigated both conditions. In addition, resulting in resistance against apoptosis (Nair
studies have examined what happens to dentate et al., 2004; Fig. 2). Multiple genes expressed in
cells in the complete absence of corticosteroid surviving individual granule cells were assessed by
358

Fig. 2. (A) Slices from animals that were adrenalectomized (ADX) 1, 2 or 3 days previously, or sham-operated rats, were stained in
vitro with Hoechst 33258, a stain for nuclear chromatin. Apoptotic granule cells with condensed nuclear chromatin are brightly colored
(arrows) and could not be recorded with whole cell patch clamp electrodes. Neighboring cells, which were not apoptotic, were recorded
(cross), and RNA was subsequently isolated from the cell. This procedure allows the collection of RNA from cells that have not yet
entered the apoptotic pathways. (B) After RNA collection, multiple gene expression in surviving individual granule cells was assessed
by linear antisense RNA amplification and hybridization to slot blots containing various neuronal cDNAs. (C) Compared to cells from
sham-operated control rats (average of all data from the 3 days, normalized to 10%, black squares), the ratio between Bcl-2 and Bax
RNA expression was significantly enhanced in 2- and 3-day ADX cells (open squares), which at that time had still resisted entering the
apoptotic pathway. A similar enhanced expression ratio was seen 3 days after ADX for calbindin relative to calcium-calmodulin kinase
II (CamKII). (D) To confirm that Bcl-2 and calbindin conferred resistance to entering the apoptotic route, these two genes were
overexpressed unilaterally in the dentate gyrus shortly before ADX; a control virus was injected into the contralateral hemisphere. A
significant reduction in the number of apoptotic cells 3 days after ADX was seen in the ipsilateral versus contralateral (open bars)
dentate gyrus of animals receiving the Bcl-2-expressing virus (gray bars), but not the calbindin-expressing virus (black bar). Adapted
with permission from Blackwell (Nair et al., 2004).

linear antisense RNA amplification and hybridi- relative MR, alpha1A voltage-gated calcium chan-
zation to slot blots containing various neuronal nel, Bcl-2 and NMDA R2C mRNA expression; it
cDNAs. Hierarchical clustering and principal is important to realize that overall dentate expres-
component analysis was performed on two phys- sion of these products may show changes in the
iological variables and 14 mRNA ratios from cells opposite direction (e.g., Cardenas et al., 2002),
collected 1, 2 or 3 days after ADX or sham op- most likely caused by the cells destined to die.
eration. The results indicated that surviving 3-day Some 1- and 2-day ADX cells clustered with these
ADX granule cells display lower membrane 3-day survivors; therefore, one or more compo-
capacitance, lower relative N-methyl-D-aspartate nents of their mRNA expression profile may rep-
(NMDA) R1 mRNA expression and higher resent predictive markers for apoptosis resistance.
359

The functional relevance of two candidate genes activation of the adrenal cortex by adreno-
was tested by in vivo local over-expression in the corticotropin is absent) granule cell density was
same model system; of these, Bcl-2 conferred par- decreased, due to diminished cell proliferation
tial protection when induced shortly before ADX. (Ostwald et al., 2006). Mice with a genetic disrup-
Collectively, the data indicate that cells with a low tion of the MR also exhibited decreased granule
membrane capacitance and high input resistance cell density and a significant reduction in neuro-
— i.e., presumably ‘younger’ dentate granule cells genesis (Gass et al., 2000).
(Liu et al., 1996) — can trigger a gene expression
profile upon ADX which prevents initiation of the
apoptotic program, whereas ‘older’ cells may no Physiology
longer be able to do this.
Probably not all degradation in the dentate fol- If removal of corticosterone leads to loss of gran-
lowing ADX is due to apoptosis. Interestingly, it ule cells within 3 days, this would be expected to
was shown that loss of markers of mature neurons impair functional network properties in the dent-
can occur, rather than cell death, and this can be ate, particularly so if the loss involves mature
reversed by a serotonin-1A receptor agonist neurons that are fully integrated into the network.
(Huang et al., 1997). This could reflect that the The fact that proliferation is enhanced shortly
latter restored antigenicity; more likely, though, it after ADX is less relevant at early time points,
could indicate that ADX leads to neurite retrac- because at that time the newborn cells are not
tion, which is restored by a serotonin-1A receptor yet incorporated into the network. Later on, of
agonist. In agreement with the latter, 3-D recon- course, this burst in neurogenesis may have func-
struction of surviving granule cells after ADX tional consequences too.
revealed that the dendritic length of higher order Initial field potential recordings indeed corrob-
segments is significantly reduced in animals with orated the notion that apoptosis impairs synaptic
undetectable corticosterone levels (Wossink et al., transmission in the dentate gyrus. Thus, stimula-
2001). Yet, in ADX animals with residual circu- tion of the perforant path in vitro leads to a field
lating corticosterone no such reduced dendritic excitatory postsynaptic potential (fEPSP) in the
length was observed, indicating that minute molecular layer of the dentate gyrus. Three days
amounts of corticosterone suffice to maintain den- after ADX, both the slope and the amplitude of
dritic integrity of granule cells. Growth factors, this fEPSP were reduced by approximately 30%
which are sensitive to the presence of corticosteroids (Stienstra et al., 1998). No reduction in fEPSP was
in the dentate gyrus, may play a role in maintaining observed in those animals that did not display
the dendritic structure (Chao and McEwen, 1994). apoptosis. Substitution of animals with a low dose
ADX not only promotes apoptosis and neuro- of corticosterone, sufficient to activate the brain
degeneration but also neurogenesis (Cameron and MRs and prevent apoptosis, fully normalized the
Gould, 1994). Already within a day of corticos- fEPSPS properties.
terone removal the degree of cell proliferation in Yet, three observations indicate that the reduced
the dentate gyrus was enhanced (Cameron et al., field response is not a consequence of apoptotic
1995). This process may depend on intact ent- cell loss. First, the reduced fEPSP slope and am-
orhinal input and NMDA receptor-dependent plitude were already present 1 day after ADX, i.e.,
transmission (Cameron et al., 1995). Data support several days before the first apoptotic cells can be
that facilitation of neurogenesis after ADX not discerned (Stienstra et al., 1998). Second, admin-
only requires MR activation but also GR activa- istration of corticosterone in vitro to slices from
tion (Wong and Herbert, 2005). The importance 3-day ADX rats fully restored the amplitude and
of corticosterone and the MR for neurogenesis slope of the fEPSP, despite the presence of apo-
and dentate granule cell number also follows ptotic cells in the slices (Stienstra and Joëls, 2000).
from observations in mutant mice. Thus, in a Finally, intracellular recording of dentate granule
pro-opiomelanocortin null mutant mouse (where cells — which due to the method is confined to
360

non-apoptotic cells with an intact plasma mem- probably limited. Rather, in parallel with (and not
brane — also revealed a 40% reduction of the as a consequence of) the change in cell turnover,
EPSP amplitude (Joëls et al., 2001; Fig. 3A). This an extensive synaptic reorganization seems to
was observed both for the NMDA- and AMPA occur which largely alters synaptic properties of
receptor-mediated components of the synaptic the surviving cells.
response. The latter suggests that postsynaptic In addition to these altered responses to low-
glutamate receptors are unlikely to be directly frequency stimulation of perforant path afferents,
changed after ADX, but rather that dendritic it has also been found that long-term potentiation
segments carrying both types of receptors, or pre- (LTP) is reduced after ADX (Smriga et al., 1996;
synaptic afferents, are lost. Krugers et al., 2007), although one report found
Most passive (resting membrane potential, input no change (Pavlides et al., 1994). Again, the
resistance) and active membrane properties (action reduced LTP occurs independent of altered cell
potential height and width) were on average not turnover, as LTP (but not cell turnover) can be
significantly changed 3 days after ADX (Joëls fully normalized by adding corticosterone to slices
et al., 2001). As these properties depend on the age from 3 day ADX rats (Krugers et al., 2007).
of the cells (Liu et al., 1996), these observations Ionic conductances of dentate granule cells
also indicate that the functional consequences of have also been studied after ADX. More specifi-
apoptotic cell death of mature dentate cells are cally, the influx of calcium (Ca) through high

Fig. 3. (A) Typical excitatory postsynaptic potentials (EPSPs) induced in dentate granule cells by stimlation of the perforant path. The
EPSP amplitude and area under the signal (voltage  time) was reduced after ADX (A2) compared to sham operated controls (A1) or
ADX rats treated in vivo with a low dose of corticosterone sufficient to occupy the MR but not GR (A3). When the AMPA and
NMDA receptor-mediated components of the EPSPs were pharmacologically isolated, it became apparent that both were reduced by
40–50% 3 days after ADX (A4). Adapted with permission from the American Physiological Society (Joëls et al., 2001). (B) The
distribution of the dentate Ca-current amplitudes in sham-operated control animals can be fitted well with two Gaussian curves (dark
gray line), suggesting that under control conditions most cells have relatively small Ca-current amplitudes, but a subpopulation of
dentate cells displays a large Ca-current amplitude. At short intervals after ADX (1–2 days, light gray line), both populations can still
be discerned. The curve is shifted to the right, reflecting an ADX-dependent increase in Ca-current amplitude. Some days later (stippled
line; 3–7 days after ADX, i.e., at a time point when some of the cells have died by apoptosis), the peak of the amplitude distribution is
more or less at the same position, suggesting that in most neurons the Ca-current amplitude is not further enhanced. Importantly,
though, the curve became unimodal: the subpopulation with large Ca-current amplitudes were no longer present, suggesting that those
cells that had a large Ca-current amplitude to start with died by apoptosis. Adapted with permission from Blackwell (Karst and Joëls,
2001).
361

voltage-activated channels was investigated (Karst exposure to a single stressor. It was found that
and Joëls, 2001). Shortly after ADX (i.e., 1–2 days proliferation is indeed suppressed shortly after
later) Ca-current amplitude and density were in- stress exposure (Heine et al., 2004b). Apoptosis is
creased, whereas ADX rats treated with a low dose increased, not only after stress (Heine et al.,
of corticosterone (thus preferentially occupying 2004b), but also after a single injection with a
the MR) exhibited small Ca currents. Interestingly, high dose of dexamethasone (Hassan et al., 1996),
reduced Ca influx was observed at later times after involving shifts in the expression of pro- and anti-
ADX. Further investigation of the amplitude dis- apoptotic genes (Almeida et al., 2000). The effects
tribution (Fig. 3B) revealed that while the current of acute stress, though, are largely normalized
amplitude distribution in general was shifted to within 24 h (Heine et al., 2004b), indicating that
the right after ADX, a particular subpopulation of the impact of a single stressor is probably limited.
cells with large Ca-current amplitudes had com-
pletely disappeared. It was reasoned that ADX
enhances Ca influx of all dentate granule cells, Physiology
similar to what had been observed earlier in CA1
neurons (Karst et al., 1994). In cells with a large In cells which carry both MRs and GRs, such as
Ca influx — i.e., presumably the relatively ‘old’ pyramidal neurons in the CA1 region or granule
cells (Thibault and Landfield, 1996) — the ADX- cells in the dentate gryus, physiological fluctua-
induced increase in Ca influx may be just sufficient tions in corticosterone levels will result in a recep-
to make them vulnerable to cell death. Combined tor occupation ranging from predominant MR
with a reduced potency to trigger genes that would occupation when the organism is at rest and at the
confer resistance to apoptotic cell death, this circadian nadir, to concomitant MR and GR
enhanced Ca influx may cause the disappearance occupation after stress or at the circadian peak. In
of these ‘older’ cells 3–7 days after ADX. CA1 neurons, these two conditions are associated
In summary, low levels of corticosterone suffi- with distinctly different physiological responses
cient to activate MRs in dentate granule cells seem (Joëls, 1997). Thus, compared to the situation
to be necessary to preserve neuronal integrity in where corticosteroid receptors are unoccupied,
this area. In the absence of corticosteroids and conditions of predominant MR occupation are
MR activation, apoptosis and neurogenesis are associated with restricted Ca influx in the soma
increased. Even prior to apoptosis, synaptic reor- and primary dendrites, limited cell-firing frequency
ganization seems to occur, resulting in reduced accommodation, stable glutamate/GABAergic
synaptic efficacy and plasticity. At this early stage transmission, small hyperpolarizing responses to
too, Ca influx is enhanced. Relatively ‘young’ gran- serotonin (5-HT) and efficient LTP. All of these
ule cells may withstand these adverse conditions, as properties help to promote ongoing transfer of in-
they probably have a small Ca influx to start with, formation through this area. Concomitant activa-
and seem to be able to increase the expression of tion of GRs leads in a slow gene-mediated fashion
survival genes such as Bcl-2. Yet, ‘older’ cells are to enhanced Ca influx, prominent cell-firing
particularly vulnerable and may be destined to frequency accommodation, large hyperpolarizing
degenerate and/or enter the apoptotic pathway. responses to serotonin (5-HT), reduction of exci-
tatory responses to noradrenaline, and reduction
of LTP; there seems to be an activity (energy)-
Physiological variations in corticosteroid levels dependent attenuation of glutamate and gamma
amino butyric acid (GABA) mediated responses,
Cell turnover although postsynaptic responses via AMPA
receptors are enhanced within a time-window of
In contrast to what is seen after ADX (or chronic 2–4 h after GR activation (Karst and Joëls, 2005).
stress, see below), neurogenesis and apoptosis in All in all, GR activation generally may result in
the dentate gyrus are only mildly affected by normalization of activity that was temporary
362

raised by transmitters that reach the CA1 region at interpret. Initial experiments reported that selec-
the time of the stress exposure. As is also evident tive activation of MRs in ADX rats prolongs LTP
from the above discussion, steroid-dependency (Pavlides et al., 1994), whereas very high doses of
for many cell properties in the CA1 area follows corticosterone reduce LTP in vivo (Pavlides et al.,
a U-shaped curve (Joëls, 2006). For instance, Ca 1993) and show a strong tendency toward a de-
currents are large in the absence of corticosterone, crease in vitro (Alfarez et al., 2003). Later studies
small with predominant MR activation, and showed that forced and uncontrollable swim stress
large again when both receptor types are activated prolongs LTP (Korz and Frey, 2003; Kavushansky
(Fig. 4). et al., 2006), via a rapid MR-dependent mecha-
Although the consequences of physiological nism (Korz and Frey, 2003) involving mitogen-
fluctuations in corticosteroid levels are far less in- activated protein kinases (Ahmed et al., 2006) and
vestigated in the dentate gyrus than in the CA1 depending on intact input from the basolateral
region, a U-shaped dose dependency may not exist amygdala (Korz and Frey, 2005). This suggests
for all properties in the dentate (Joëls, 2006; see that LTP in the dentate can be reduced by corti-
Fig. 4). In the dentate, as opposed to the CA1 costerone (as in the CA1 area), but that stress
region (Kerr et al., 1992, 1994), no significant in the intact brain causes a different effect, due to
difference in Ca-current amplitude was observed in modulatory synaptic inputs and the effect of
slices with predominant MR versus slices with stress-released factors other than corticosteroid
concomitant MR and GR activation (Van Gemert hormones (e.g., peptides or neurosteroids).
and Joëls, 2006). It should be noted, though, that In conclusion, activation of MRs in dentate
these data were obtained in animals that had been granule cells is very important for the physiology
handled daily for 3 weeks, which in itself may of the area. In general it serves the same role as in
affect cellular function; still, preliminary data the CA1 area, i.e., it promotes the transfer of sig-
in naı̈ve animals confirm this observation (Van nals through the dentate gyrus. To what extent
Gemert et al., unpublished observation). Similarly, additional GR activation plays a role in the dent-
AMPA receptor-mediated synaptic responses were ate is at this moment less clear. The available data
comparable in slices (from handled rats) with pre- seems to indicate that the strong effects of GR
dominant MR activation and slices where both activation (on top of MR activation) as seen in the
receptor types were activated (Karst and Joëls, CA1 area do not take place in the dentate gyrus.
2003). This was also seen for the slope of the fE-
PSP (Stienstra and Joëls, 2000). Hyperpolarizing
responses to 5-HT do show some differences be- Dentate function after chronic stress
tween conditions of low and high corticosterone
concentrations (Karten et al., 2001), but these are Cell turnover
not as distinct as in the CA1 area. A gene expres-
sion survey of single dentate cells from handled Cell proliferation and apoptosis have been exten-
rats — involving, e.g., transcripts for the subunits sively studied in the dentate in relation to chronic
of glutamate and GABAA receptors and of Ca stress, a condition that is known to be a risk factor
channels — also revealed very little differences due in the etiology of psychiatric diseases like major
to activation of GRs (Qin et al., 2004). Appar- depression (McEwen, 2003; de Kloet et al., 2005).
ently, a rise in corticosterone level does often not This research has recently been boosted by the
lead to efficient activation of GR or GR-depend- observation that anti-depressants promote neuro-
ent pathways in the dentate gyrus. This could be genesis and that, in fact, this proliferative action is
due to local differences in GR variants, in post- necessary for these compounds to exert their
translational modification of the GR or in the behavioral effects in an animal model (Santarelli
presence of proteins that interact with GR. et al., 2003; see further Chapter 38 in this book).
The steroid sensitivity of LTP in the dentate In general, chronic stress — be it from repeated
gyrus seems to be somewhat more complex to restraining, from a psychosocial source, or from a
363

100
90
80
70
60
50
response (% of maximum)

40
30 CA1
CA1
20

100
90
80
70

60
50
40
30 DG
DG
20

MR
MR

GR
GR

stress
Fig. 4. Dose-response relationships for cellular effects of corticosterone in the CA1 hippocampal area and the dentate gyrus (DG). All
responses are expressed as a percentage of the maximal response. The concentration of corticosterone is a rough estimate of the local
concentration, based on the solutions perfused on in vitro preparations or derived from the plasma concentration when fluctuations in
hormone levels were measured in vivo. In the CA1 area, both the amplitude of depolarization-induced Ca currents (open squares) and
the hyperpolarization caused by serotonin-1A receptor activation (filled circles) display a clear U-shaped dose dependency. The
descending limb is linked to activation of MRs, while the ascending limb is associated with gradual GR activation, as occurs after
stress; relative occupation of the two receptors with increasing corticosterone levels is depicted at the bottom. DG granule neurons
show a clear effect on the field potential (filled squares) and single cell response (closed triangles) caused by activation of glutamatergic
AMPA receptors; this effect is linked to MR activation. Although these cells also abundantly express GRs, high doses of corti-
costerone do not lead to additional changes, except when animals had been exposed to 3 weeks of unpredictable stress prior to
recording (open triangles). Adapted with permission from Elsevier (Joëls, 2006).
364

combination of physical and psychological stress- in the dentate gyrus (Marmigere et al., 2003), re-
ors — has been reported to reduce cell prolifera- peated restraint stress reduces BDNF mRNA lev-
tion in the dentate gyrus, at least in males (Czeh els (Smith et al., 1995). In contrast, NT-3 mRNA
et al., 2001; Pham et al., 2003; Westenbroek et al., levels were increased. Stress did not affect the ex-
2004; Heine et al., 2004b). This seems to be largely pression of neurotrophin-4, or tyrosine receptor
due to the elevation in corticosterone level asso- kinases (trkB or C). More recently, though, it was
ciated with chronic stress, as repeated exogenous found that chronic stress increases the levels of the
administration of corticosterone (which enhances catalytic but not of the truncated form of trkB
corticosteroid levels but suppresses circulating lev- (Nibuya et al., 1999). As many newborn cells are
els of CRH and ACTH) also decreased the number found in close association with the vasculature,
of adult newborn cells, due to a reduction in cell mediators in this system were also investigated.
survival (Wong and Herbert, 2004). Importantly, A significant reduction in the level of vascular
administration of a GR antagonist can fully nor- endothelial growth factor and its receptor Flk-1
malize the reduction in cell proliferation and/or was observed after 3 weeks of unpredictable stress
survival after repeated exogenous corticosterone (Heine et al., 2005). These levels were restored
administration (Mayer et al., 2006). It is presently when the period of stress was followed by 3 weeks
still unclear, however, whether exogenous corti- of rest. In addition to these growth factors, several
costerone and chronic stress indeed have the exact cell adhesion factors have been studied. Effects of
same effects on all aspects of neurogenesis, i.e., chronic over-exposure to corticosterone varied,
proliferation, migration, differentiation and sur- however, because it was found to enhance the
vival of newborn cells. Whether or not chronic expression of neural cell adhesion molecule in the
stress results in a reduction of the dentate volume dentate gyrus (Pham et al., 2003), decrease it
may depend on the duration or severity of the (Venero et al., 2002; Nacher et al., 2004) or leave it
stress protocol (Pham et al., 2003; Heine et al., unaffected (Van Gemert et al., 2006). Transcripts
2004b). for the adhesion molecule L1 were increased
Surprisingly, apoptosis was reduced after (Venero et al., 2002). Clearly, the crucial factors
chronic stress when analyzed over the entire dent- mediating the effect of chronic stress on progenitor
ate gyrus and hilus (Heine et al., 2004b). This may cell function still need to be identified.
signify that chronic stress slows down the cell
cycle, causing fewer neurons to be born, but also
fewer to die. This is corroborated by the observa- Physiology
tion that chronic stress increases the level of
p27Kip, an inhibitor of the cell cycle (Heine While temporary elevations in corticosterone level,
et al., 2004a). Both the effect on neurogenesis and leading to brief activation of GRs, seem to be rel-
on apoptosis observed after 3 weeks of unpredict- atively ineffective in the dentate gyrus of handled
able stress or restraint stress were largely reversed or naı̈ve rats (as opposed to the CA1 area), such
when animals were allowed to recover from stress elevations do become very effective when occur-
for 3 weeks (Pham et al., 2003; Heine et al., 2004b). ring against a background of chronic stress. This
As progenitor cells do not express GRs (Garcia was found to be true for GR-evoked changes in
et al., 2004), corticosteroid-dependent effects on gene expression as well as functional endpoints.
their proliferation requires other cells in the neigh- For instance, high doses of corticosterone in
borhood which do contain GRs and can produce vitro do not change the AMPA receptor-mediated
factors that indirectly affect the function of pro- responses in handled animals, but a large increase
genitor cells. Several potential intermediate factors in the amplitude of excitatory postsynaptic cur-
have been investigated, but the emerging picture rents was observed with the same dose of corti-
is somewhat confusing. It was found that, while costerone when slices were prepared from
acute stressors lead to a transient increase in brain- chronically-stressed rats (Karst and Joëls, 2003).
derived neurotrophic factor (BDNF) mRNA levels A significant enhancement was also seen with
365

respect to the amplitude of the sustained Ca cur- hormones that go beyond the borders of a re-
rent, when comparing corticosterone-treated cells stricted, physiological range. When corticosteroid
from control with those of chronically-stressed hormone concentrations drop below a critical
rats (Van Gemert and Joëls, 2006). This effect level, or when they are repeatedly elevated, neu-
of chronic stress was also seen for the relative ronal integrity in the dentate gyrus is at stake. In
expression of the subunit, which forms the pore the CA1 area of the hippocampus, conditions that
of the L-type (sustained) Ca channel (Qin et al., lead to predominant activation of the MR have
2004). been proposed to facilitate ongoing activity as well
It has been found, both in vivo (Pavlides et al., as promote neuronal viability. In the dentate
2002) and in vitro (Alfarez et al., 2003), that in- gyrus, however, activation of MRs is an absolute
duction of LTP is very much impaired in the dent- necessity to preserve neurites and to restrain ex-
ate gyrus of chronically stressed rats. Additional cessive cell turnover. On the other end of the spec-
administration of a high dose of corticosterone is trum, repeated stress seems to slow the cycle of
ineffective (Alfarez et al., 2003), as LTP cannot be proliferating cells and put granule cells at risk,
reduced more than it is already in tissue exposed to by increasing their response to glutamate and
low concentrations of the hormone. enhancing voltage-dependent Ca influx. Clearly,
Taken together, it is evident that repeated much of the unique sensitivity of the dentate gyrus
exposure of the dentate gyrus to high levels of co- to fluctuating corticosteroid levels is caused by the
rticosterone reduces the number of newborn cells, presence of progenitor cells in this area, which do
through mechanisms that are presently only partly not carry corticosteroid receptors themselves but
understood. Over a period of weeks this may not nevertheless are very much influenced in their ac-
lead to a large reduction in the overall number of tivity through neighboring cells that do express
cells or volume of the adult dentate gyrus. Periods MR and GR.
of stress relief may even partially normalize the In CA1 pyramidal cells, which like dentate
changes in cell turnover. However, prolonged granule cells are among the few types of neurons
ongoing stress to which the organism cannot that co-express MR and GR, physiological shifts
adapt may substantially change the composition in corticosteroid level (and hence in receptor oc-
of dentate gyrus granule cells. This could result in cupation ratios) lead to distinct changes in func-
cognitive disturbances, and contribute to the pre- tional properties. As discussed above, a U-shaped
cipitation of clinical symptoms. Interestingly, with dose dependency has been observed for cortico-
respect to physiological properties, dentate granule steroid actions in this region. By contrast, dentate
cells seem to respond stronger to activation of granule cells seem to react in a more linear fashion,
GRs (and thus presumably to acute stressors) with a relatively limited response to temporary
when the organism has experienced some weeks of activation of GRs. For instance, (1) absence of
unpredictable stress. The ensuing increased excit- corticosterone leads to increased neurogenesis and
ability in combination with enhanced Ca influx (2) exposure to an acute stressor leads to a tem-
could make cells more vulnerable to excitotoxic porary reduction in neurogenesis; (3) repetitive
damage. stress exposure results in a more prominent and
sustained suppression of neurogenesis (Fig. 5).
Another example pertains to glutamate receptor-
Functional relevance of corticosteroid effects in the mediated transmission: (1) in the absence of corti-
dentate gyrus costerone, responses to glutamate are small; (2)
activation of MRs increases the responses; (3)
Corticosteroid dose dependency in the dentate gyrus while additional GR activation does not seem to
further enhance the response in control rats, such
While many brain regions are affected by stress- activation does lead to a marked enhancement of
related hormones, the dentate gyrus seems to be AMPA receptor-mediated responses in animals
exquisitely sensitive to concentrations of such with a history of chronic stress exposure.
366

A C handled

6000 * Acute stress

Mol. layer 24 hrs after

# BrdUpositive cells
5000 acute stress

4000
SGZ
h
hilus 3000

2000

GCL 1000

B D
9 * SHAM 3500
* handled
# H3 Thymidine-labeled cells

8
3000 24 hrs after

# BrdUpositive cells
7 ADX
chronic
2500 stress
6

5 2000

4
1500
3
1000
2
1 500

0 0

Fig. 5. (A) Immunohistochemical image of the dentate gyrus, showing proliferating cells mostly in the subgranular zone (SGZ) and far
less in the granular cell layer (GCL), molecular layer and hilus. Proliferating cells were marked with Ki-67. (B) One week after ADX or
sham operation, animals received an injection of 3H thymidine to label dividing cells. Three weeks later, dentate cells with a neuronal
phenotype were counted and expressed in number per 106 mm2. The number of newborn neurons was significantly enhanced after
ADX. Adapted with permission from Elsevier (Cameron and Gould, 1994). (C) Injection of BrdU just before exposure to acute stress
revealed a significant reduction in the number of newborn cells in the dentate gyrus (excluding the hilus). When BrdU was injected one
day after acute stress exposure, the difference in the number of newborn cells no longer was evident. Apparently, acute stress only
results in a transient suppression of proliferation in the dentate gyrus. Adapted from Heine et al., 2004b. (D). If BrdU was injected one
day after chronic stress exposure, the number of newborn cells was significantly reduced, indicating that chronic stress induces a much
longer-lasting suppression of cell proliferation. Adapted from Heine et al., 2004b and unpublished observations.

Importantly, the altered glutamate responses do not have consequences for behavioral functions in
depend on the changes in neurogenesis, but never- which these neurons participate. Indeed, numerous
theless seem to have the same steroid dependency. studies have described how stress and/or selective
Why two cell types that apparently share the same manipulation of MR or GR occupancy alter hip-
steroid receptor expression pattern — i.e., CA1 pocampal-dependent behavior. With respect to
pyramidal neurons and dentate granule cells — dis- tests requiring spatial memory, it was found that
play such a different steroid dose dependency is not MR occupation is important for reaction to novel
understood, but most likely local factors like the information as well as determination of behavioral
intracellular protein composition and network prop- strategy (Oitzl and de Kloet, 1992; Oitzl et al.,
erties play an important role. 1994). Additional GR activation within the learn-
ing context is required for consolidation of spatial
information (Oitzl and de Kloet, 1992; Oitzl et al.,
Functional relevance in health and disease 2001). Importantly, GR activation that is not
related to the learning context may impair rather
The effects of corticosteroid hormones on physi- than improve memory consolidation (De Kloet
ological properties of hippocampal neurons will et al., 1999). It has been proposed that convergence
367

in time and space (i.e., brain area) is necessary for More distinct functional changes may occur in
the facilitating effects of stress on memory to take the dentate gyrus when the corticosteroid levels are
place (Joëls et al., 2006). It should also be realized at extreme levels for a considerable period of time.
that due to stress, many transmitter and hormone As discussed above, this will occur when cortico-
systems are activated, so that not only corticoster- steroid levels are extremely low, such as after ADX
one levels but also those of, e.g., CRH and nor- in experimental animals. In humans though,
adrenaline will be elevated. This is of great relevance extreme adrenal insufficiency will rarely pass un-
since dentate gyrus function, and in particular LTP, noticed for any length of time, so the consequences
is known to be enhanced both by CRH (Wang et al., for brain function will be usually curtailed by re-
2000) and noradrenaline (e.g., Bramham et al., placement therapy. By contrast, prolonged periods
1997). Next to these short-lasting stressors, many of exposure to uncontrollable and/or unpredicta-
studies have addressed the effects of chronic stress ble stressors are not uncommon. Animal studies
on memory function. It was found that chronic indicate that renewed exposure to stress of an
stress exposure impairs spatial memory in males organism that already has a history of repeated
(Luine et al., 1994; Wright and Conrad, 2005); yet, stress experiences induces particularly strong
memory improvement was observed in females changes in gene expression and cell physiology of
(Luine, 2002; McLaughlin et al., 2005). It is un- dentate cells, such as enhanced response to exci-
clear, however, to what extent the spatial orientation tatory input and increased voltage-dependent Ca
studies reflect functions of the dentate gyrus. influx. The consequences of these changes in phys-
In this respect, it may be also of interest to con- iology will be particularly prominent when they
sider effects of stress on anxiety-related behavior, coincide with extensive depolarization of the dent-
which was shown to involve amygdala nuclei as well ate area. This could occur when the dentate is
as hippocampal subregions (LeDoux, 2000; McG- challenged by demanding behavioral events but
augh, 2004). An extensive line of research has re- also in association with pathological conditions
vealed that inhibitory avoidance memory is such as ischemic insults (Krugers et al., 2000). This
facilitated by stress (McGaugh and Roozendaal, fits with the general theory that (prolonged) expo-
2002). Contextual fear conditioning is also pro- sure to glucocorticoids may exacerbate damage
moted by additional stress exposure (Cordero inflicted to hippocampal cells by pathological con-
et al., 2003; Donley et al., 2005). The facilitation ditions (Sapolsky, 1996; McEwen, 2004).
of avoidance behavior critically depends on GR ac- A better understanding of the functional conse-
tivation in the basolateral amygdala and requires quence of corticosteroid actions in the dentate,
interactions between noradrenaline and corticoster- however, awaits studies that more specifically ad-
one to occur within a restricted time-frame (McG- dress this issue. It will be very informative to study
augh and Roozendaal, 2002). Although intra- the influences of variations in corticosteroid level
hippocampal injections of a GR agonist could also using behavioral tests that more precisely reflect
improve avoidance memory, these effects only dentate gyrus function as well as pathological con-
occurred when the basolateral amygdala was intact ditions that critically depend on the contribution
(Roozendaal et al., 1999). A clear interaction be- of this area.
tween basolateral amygdala and the hippocampus,
particularly the dentate gyrus, was also demon- References
strated with electrophysiology. Thus, perforant
path-induced LTP in the dentate was shown to be Ahmed, T., Frey, J.U. and Korz, V. (2006) Long-term effects of
facilitated by amygdala stimulation applied within brief acute stress on cellular signaling and hippocampal LTP.
1 h before tetanic stimulation of the perforant path J. Neurosci., 26: 3951–3958.
Akirav, I. and Richter-Levin, G. (2002) Mechanisms of am-
(Akirav and Richter-Levin, 2002). With longer time
ygdala modulation of hippocampal plasticity. J. Neurosci.,
delays, amygdala stimulation reduced perforant 22: 9912–9921.
path LTP. It was shown that both effects involve Alfarez, D.N., Joëls, M. and Krugers, H.J. (2003) Chronic
noradrenaline as well as corticosterone. unpredictable stress impairs long-term potentiation in rat
368

hippocampal CA1 area and dentate gyrus in vitro. Eur. J. Garcia, A., Steiner, B., Kronenberg, G., Bick-Sander, A. and
Neurosci., 17: 1928–1934. Kempermann, G. (2004) Age-dependent expression of
Almeida, O.F., Conde, G.L., Crochemore, C., Demeneix, B.A., glucocorticoid — and mineralocorticoid receptors on neural
Fischer, D., Hassan, A.H., Meyer, M., Holsboer, F. and precursor cell populations in the adult murine hippocampus.
Michaelidis, T.M. (2000) Subtle shifts in the ratio between Aging Cell, 3: 363–371.
pro- and antiapoptotic molecules after activation of cortico- Gass, P., Kretz, O., Wolfer, D.P., Berger, S., Tronche, F.,
steroid receptors decide neuronal fate. FASEB J., 14: Reichardt, H.M., Kellendonk, C., Lipp, H.P., Schmid, W.
779–790. and Schütz, G. (2000) Genetic disruption of mineral-
Bramham, C.R., Bacher-Svendsen, K. and Sarvey, J.M. (1997) ocorticoid receptor leads to impaired neurogenesis and
LTP in the lateral perforant path is beta-adrenergic receptor- granule cell degeneration in the hippocampus of adult mice.
dependent. Neuroreport, 8: 719–724. EMBO Rep., 1: 447–451.
Cameron, H.A. and Gould, E. (1994) Adult neurogenesis is Gould, E., Woolley, C.S. and McEwen, B.S. (1990) Short-term
regulated by adrenal steroids in the dentate gyrus. Neurosci- glucocorticoid manipulations affect neuronal morphology
ence, 61: 203–209. and survival in the adult dentate gyrus. Neuroscience, 37:
Cameron, H.A. and Gould, E. (1996) Distinct populations 367–375.
of cells in the adult dentate gyrus undergo mitosis or Hassan, A.H., von Rosenstiel, P., Patchev, V.K., Holsboer, F.
apoptosis in response to adrenalectomy. J. Comp. Neurol., and Almeida, O.F. (1996) Exacerbation of apoptosis in
369: 56–63. the dentate gyrus of the aged rat by dexamethasone and
Cameron, H.A., McEwen, B.S. and Gould, E. (1995) Regula- the protective role of corticosterone. Exp. Neurol., 140:
tion of adult neurogenesis by excitatory input and NMDA 43–52.
receptor activation in the dentate gyrus. J. Neurosci., 15: Heine, V.M., Maslam, S., Joëls, M. and Lucassen, P.J. (2004a)
4687–4692. Increased P27KIP1 protein expression in the dentate gyrus
Cardenas, S.P., Parra, C., Bravo, J., Morales, P., Lara, H.E., of chronically stressed rats indicates G1 arrest involvement.
Herrera-Marschitz, M. and Fiedler, J.L. (2002) Corticoster- Neuroscience, 129: 593–601.
one differentially regulates bax, bcl-2 and bcl-x mRNA levels Heine, V.M., Maslam, S., Zareno, J., Joëls, M. and Lucassen,
in the rat hippocampus. Neurosci. Lett., 331: 9–12. P.J. (2004b) Suppressed proliferation and apoptotic changes
Chao, H.M. and McEwen, B.S. (1994) Glucocorticoids and the in the rat dentate gyrus after acute and chronic stress are
expression of mRNAs for neurotrophins, their receptors and reversible. Eur. J. Neurosci., 19: 131–144.
GAP-43 in the rat hippocampus. Brain Res. Mol. Brain Res., Heine, V.M., Zareno, J., Maslam, S., Joëls, M. and Lucassen,
26: 271–276. P.J. (2005) Chronic stress in the adult dentate gyrus re-
Cordero, M.I., Venero, C., Kruyt, N.D. and Sandi, C. (2003) duces cell proliferation near the vasculature and VEGF
Prior exposure to a single stress session facilitates subsequent and Flk-1 protein expression. Eur. J. Neurosci., 21:
contextual fear conditioning in rats. Evidence for a role of 1304–1314.
corticosterone. Horm. Behav., 44: 338–345. Hu, Z., Yuri, K., Ozawa, H., Lu, H. and Kawata, M. (1997)
Czeh, B., Michaelis, T., Watanabe, T., Frahm, J., de Biurrun, The in vivo time course for elimination of adrenalectomy-
G., van Kampen, M., Bartolomucci, A. and Fuchs, E. (2001) induced apoptotic profiles from the granule cell layer of the
Stress-induced changes in cerebral metabolites, hippocampal rat hippocampus. J. Neurosci., 17: 3981–3989.
volume, and cell proliferation are prevented by antidepres- Huang, J., Strafaci, J.A. and Azmitia, E.C. (1997) 5-HT1A
sant treatment with tianeptine. Proc. Natl. Acad. Sci. U.S.A., receptor agonist reverses adrenalectomy-induced loss of
98: 12796–12801. granule neuronal morphology in the rat dentate gyrus.
De Kloet, E.R., Joëls, M. and Holsboer, F. (2005) Stress and Neurochem. Res., 22: 1329–1337.
the brain: from adaptation to disease. Nat. Rev. Neurosci., 6: Joëls, M. (1997) Steroid hormones and excitability in the mam-
463–475. malian brain. Front. Neuroendocrinol., 18: 2–48.
De Kloet, E.R., Oitzl, M.S. and Joëls, M. (1999) Stress and Joëls, M. (2006) Corticosteroid effects in the brain: U-shape it.
cognition: are corticosteroids good or bad guys? Trends Trends Pharmacol. Sci., 27: 244–250.
Neurosci., 22: 422–426. Joëls, M., Pu, Z., Wiegert, O., Oitzl, M.S. and Krugers, H.J.
De Kloet, E.R., Vreugdenhil, E., Oitzl, M.S. and Joëls, M. (2006) Learning under stress: how does it work? Trends
(1998) Brain corticosteroid receptor balance in health and Cogn. Sci., 10: 152–158.
disease. Endocr. Rev., 19: 269–301. Joëls, M., Stienstra, C. and Karten, Y. (2001) Effect of
Di, S., Malcher-Lopes, R., Halmos, K.C. and Tasker, J.G. adrenalectomy on membrane properties and synaptic
(2003) Nongenomic glucocorticoid inhibition via end- potentials in rat dentate granule cells. J. Neurophysiol., 85:
ocannabinoid release in the hypothalamus: a fast feedback 699–707.
mechanism. J. Neurosci., 23: 4850–4857. Karst, H., Berger, S., Turiault, M., Tronche, F., Schütz, G. and
Donley, M.P., Schulkin, J. and Rosen, J.B. (2005) Glucocorti- Joëls, M. (2005) Mineralocorticoid receptors are indispensa-
coid receptor antagonism in the basolateral amygdala and ble for nongenomic modulation of hippocampal glutamate
ventral hippocampus interferes with long-term memory of transmission by corticosterone. Proc. Natl. Acad. Sci.
contextual fear. Behav. Brain Res., 164: 197–205. U.S.A., 102: 19204–19207.
369

Karst, H. and Joëls, M. (2001) Calcium currents in rat dentate glucocorticoid receptor antagonist mifepristone normalises
granule cells are altered after adrenalectomy. Eur. J. Neuro- the corticosterone-induced reduction of adult hippocampal
sci., 14: 503–512. neurogenesis. J. Neuroendocrinol., 18: 629–631.
Karst, H. and Joëls, M. (2003) Effect of chronic stress on McEwen, B.S. (2003) Mood disorders and allostatic load. Biol.
synaptic currents in rat hippocampal dentate gyrus neurons. Psychiatry, 54: 200–207.
J. Neurophysiol., 89: 625–633. McEwen, B.S. (2004) Protection and damage from acute and
Karst, H. and Joëls, M. (2005) Corticosterone slowly chronic stress: allostasis and allostatic overload and relevance
enhances miniature excitatory postsynaptic current ampli- to the pathophysiology of psychiatric disorders. Ann. N.Y.
tude in mice CA1 hippocampal cells. J. Neurophysiol., 94: Acad. Sci., 1032: 1–7.
3479–3486. McGaugh, J.L. (2004) The amygdala modulates the consolida-
Karst, H., Wadman, W.J. and Joëls, M. (1994) Corticosteroid tion of memories of emotionally arousing experiences. Annu.
receptor-dependent modulation of calcium currents in rat Rev. Neurosci., 27: 1–28.
hippocampal CA1 neurons. Brain Res., 649: 234–242. McGaugh, J.L. and Roozendaal, B. (2002) Role of adrenal
Karten, Y.J., Stienstra, C.M. and Joëls, M. (2001) Corticoster- stress hormones in forming lasting memories in the brain.
oid effects on serotonin responses in granule cells of the rat Curr. Opin. Neurobiol., 12: 205–210.
dentate gyrus. J. Neuroendocrinol., 13: 233–238. McLaughlin, K.J., Baran, S.E., Wright, R.L. and Conrad,
Kavushansky, A., Vouimba, R.M., Cohen, H. and Richter- C.D. (2005) Chronic stress enhances spatial memory in
Levin, G. (2006) Activity and plasticity in the CA1, the ovariectomized female rats despite CA3 dendritic retrac-
dentate gyrus, and the amygdala following controllable vs. tion: possible involvement of CA1 neurons. Neuroscience,
uncontrollable water stress. Hippocampus, 16: 35–42. 135: 1045–1054.
Kerr, D.S., Campbell, L.W., Thibault, O. and Landfield, P.W. Nacher, J., Gomez-Climent, M.A. and McEwen, B. (2004)
(1992) Hippocampal glucocorticoid receptor activation Chronic non-invasive glucocorticoid administration
enhances voltage-dependent Ca2+ conductances: relevance decreases polysialylated neural cell adhesion molecule
to brain aging. Proc. Natl. Acad. Sci. U.S.A., 89: 8527–8531. expression in the adult rat dentate gyrus. Neurosci. Lett.,
Korz, V. and Frey, J.U. (2003) Stress-related modulation of 370: 40–44.
hippocampal long-term potentiation in rats: involvement of Nair, S.M., Karst, H., Dumas, T., Phillips, R., Sapolsky, R.M.,
adrenal steroid receptors. J. Neurosci., 23: 7281–7287. Rumpff-van Essen, L., Maslam, S., Lucassen, P.J. and Joëls,
Korz, V. and Frey, J.U. (2005) Bidirectional modulation of M. (2004) Gene expression profiles associated with survival
hippocampal long-term potentiation under stress and no- of individual rat dentate cells after endogenous corticosteroid
stress conditions in basolateral amygdala-lesioned and intact deprivation. Eur. J. Neurosci., 20: 3233–3243.
rats. J. Neurosci., 25: 7393–7400. Nibuya, M., Takahashi, M., Russell, D.S. and Duman, R.S.
Krugers, H.J., van der Linden, S., van Olst, E., Alfarez, D.N., (1999) Repeated stress increases catalytic TrkB mRNA in rat
Maslam, S., Lucassen, P.J. and Joëls, M. (2007) Dissociation hippocampus. Neurosci. Lett., 67: 81–84.
between apoptosis, neurogenesis and synaptic potentiation in Oitzl, M.S., Fluttert, M. and de Kloet, E.R. (1994) The effect of
the dentate gyrus of adrenalectomized rats. Synapse, 61: corticosterone on reactivity to spatial novelty is mediated by
221–230. central mineralocorticosteroid receptors. Eur. J. Neurosci., 6:
Krugers, H.J., Maslam, S., Korf, J. and Joëls, M. (2000) The 1072–1079.
corticosterone synthesis inhibitor metyrapone prevents Oitzl, M.S. and de Kloet, E.R. (1992) Selective corticosteroid
hypoxia/ischemia-induced loss of synaptic function in the antagonists modulate specific aspects of spatial orientation
rat hippocampus. Stroke, 31: 1162–1172. learning. Behav. Neurosci., 106: 62–71.
LeDoux, J.E. (2000) Emotion circuits in the brain. Annu. Rev. Oitzl, M.S., Reichardt, H.M., Joëls, M. and de Kloet, E.R.
Neurosci., 23: 155–184. (2001) Point mutation in the mouse glucocorticoid receptor
Liu, Y.B., Lio, P.A., Pasternak, J.F. and Trommer, B.L. (1996) preventing DNA binding impairs spatial memory. Proc. Natl.
Developmental changes in membrane properties and postsy- Acad. Sci. U.S.A., 98: 12790–12795.
naptic currents of granule cells in rat dentate gyrus. J. Ostwald, D., Karpac, J. and Hochgeschwender, U. (2006)
Neurophysiol., 76: 1074–1088. Effects on hippocampus of lifelong absence of glucocorti-
Luine, V. (2002) Sex differences in chronic stress effects on coids in the pro-opiomelanocortin null mutant mouse
memory in rats. Stress, 5: 205–216. reveal complex relationship between glucocorticoids and
Luine, V., Villegas, M., Martinez, C. and McEwen, B.S. (1994) hippocampal structure and function. J. Mol. Neurosci., 28:
Repeated stress causes reversible impairments of spatial 291–302.
memory performance. Brain Res., 639: 167–170. Pascual-Le Tallec, L. and Lombes, M. (2005) The mineral-
Marmigere, F., Givalois, L., Rage, F., Arancibia, S. and Tapia- ocorticoid receptor: a journey exploring its diversity and
Arancibia, L. (2003) Rapid induction of BDNF expression in specificity of action. Mol. Endocrinol., 19: 2211–2221.
the hippocampus during immobilization stress challenge in Pavlides, C., Kimura, A., Magarinos, A.M. and McEwen,
adult rats. Hippocampus, 13: 646–655. B.S. (1994) Type I adrenal steroid receptors prolong hip-
Mayer, J.L., Klumpers, L., Maslam, S., de Kloet, E.R., Joëls, pocampal long-term potentiation. Neuroreport, 5:
M. and Lucassen, P.J. (2006) Brief treatment with the 2673–2677.
370

Pavlides, C., Nivon, L.G. and McEwen, B.S. (2002) Effects subunit-containing GABAA receptors. Proc. Natl. Acad.
of chronic stress on hippocampal long-term potentiation. Sci. U.S.A., 100: 14439–14444.
Hippocampus, 12: 245–257. Stienstra, C.M., van der Graaf, F., Bosma, A., Karten, Y.J.,
Pavlides, C., Watanabe, Y. and McEwen, B.S. (1993) Effects Hesen, W. and Joëls, M. (1998) Synaptic transmission in the
of glucocorticoids on hippocampal long-term potentiation. rat dentate gyrus after adrenalectomy. Neuroscience, 85:
Hippocampus, 3: 183–192. 1061–1071.
Pham, K., Nachler, J., Hof, P.R. and McEwen, B.S. (2003) Stienstra, C.M. and Joëls, M. (2000) Effect of corticosteroid
Repeated restraint stress suppresses neurogenesis and induces treatment in vitro on adrenalectomy-induced impairment
biphasic PSA-NCAM expression in the adult rat dentate of synaptic transmission in the rat dentate gyrus. J.
gyrus. Eur. J. Neurosci., 17: 879–886. Neuroendocrinol., 12: 199–205.
Qin, Y., Karst, H. and Joëls, M. (2004) Chronic unpredictable Thibault, O. and Landfield, P.W. (1996) Increase in single
stress alters gene expression in rat single dentate granule cells. L-type calcium channels in hippocampal neurons during
J. Neurochem., 89: 364–374. aging. Science, 272: 1017–1020.
Reul, J.M. and de Kloet, E.R. (1985) Two receptor systems for Van Gemert, N.G. and Joëls, M. (2006) Effect of chronic stress
corticosterone in rat brain: microdistribution and differential and RU38486 treatment on voltage-dependent calcium cur-
occupation. Endocrinology, 117: 2505–2511. rents in rat hippocampal dentate gyrus. J. Neuroendocrinol.,
Roozendaal, B., Nguyen, B.T., Power, A.E. and McGaugh, 18: 732–741.
J.L. (1999) Basolateral amygdala noradrenergic influence Van Gemert, N.G., van Riel, E., Meijer, O.C., Fehr, S.,
enables enhancement of memory consolidation induced by Schachner, M. and Joëls, M. (2006) No effect of prolonged
hippocampal glucocorticoid receptor activation. Proc. Natl. corticosterone over-exposure on NCAM, SGK1, and RGS4
Acad. Sci. U.S.A., 96: 11642–11647. mRNA expression in rat hippocampus. Brain Res., 1093:
Santarelli, L., Saxe, M., Gross, C., Surget, A., Battaglia, F., 161–166.
Dulawa, S., Weisstaub, N., Lee, J., Duman, R., Arancio, O., Venero, C., Tilling, T., Hermans-Borgmeyer, I., Schmidt, R.,
Belzung, C. and Hen, R. (2003) Requirement of hippocampal Schachner, M. and Sandi, C. (2002) Chronic stress induces
neurogenesis for the behavioral effects of antidepressants. opposite changes in the mRNA expression of the cell
Science, 301: 805–809. adhesion molecules NCAM and L1. Neuroscience, 115:
Sapolsky, R.M. (1996) Stress, glucocorticoids, and damage to the 1211–1219.
nervous system: the current state of confusion. Stress, 1: 1–19. Wang, H.L., Tsai, L.Y. and Lee, E.H. (2000) Corticotropin-
Seckl, J.R. and Walker, B.R. (2003) Minireview: 11beta- releasing factor produces a protein synthesis — dependent
hydroxysteroid dehydrogenase type 1 — a tissue-specific long-lasting potentiation in dentate gyrus neurons. J.
amplifier of glucocorticoid action. Endocrinology, 142: Neurophysiol., 83: 343–349.
1371–1376. Westenbroek, C., den Boer, J.A., Veenhuis, M. and ter Horst,
Sloviter, R.S., Dean, E. and Neubort, S. (1993a) Electron mi- G.J. (2004) Chronic stress and social housing differentially
croscopic analysis of adrenalectomy-induced hippocampal affect neurogenesis in male and female rats. Brain Res. Bull.,
granule cell degeneration in the rat: apoptosis in the adult 64: 303–308.
central nervous system. J. Comp. Neurol., 330: 337–351. Wong, E.Y. and Herbert, J. (2004) The corticoid environment:
Sloviter, R.S., Sollas, A.L., Dean, E. and Neurbort, S. (1993b) a determining factor for neural progenitors’ survival in the
Adrenalectomy-induced granule cell degeneration in the rat adult hippocampus. Eur. J. Neurosci., 20: 2491–2498.
hippocampal dentate gyrus: characterization of an in vivo Wong, E.Y. and Herbert, J. (2005) Roles of mineralocorticoid
model of controlled neuronal death. J. Comp. Neurol., 330: and glucocorticoid receptors in the regulation of progenitor
324–336. proliferation in the adult hippocampus. Eur. J. Neurosci., 22:
Sloviter, R.S., Valiquette, G., Abrams, G.M., Ronk, E.C., 785–792.
Sollas, A.L., Paul, L.A. and Neubort, S. (1989) Selective loss Woolley, C.S., Gould, E., Sakai, R.R., Spencer, R.L. and
of hippocampal granule cells in the mature rat brain after McEwen, B.S. (1991) Effects of aldosterone or RU28362
adrenalectomy. Science, 243: 535–538. treatment on adrenalectomy-induced cell death in the dentate
Smith, M.A., Makino, S., Kvetnansky, R. and Post, R.M. gyrus of the adult rat. Brain Res., 554: 312–315.
(1995) Stress and glucocorticoids affect the expression of Wossink, J., Karst, H., Mayboroda, O. and Joëls, M. (2001)
brain-derived neurotrophic factor and neurotrophin-3 Morphological and functional properties of rat dentate gran-
mRNAs in the hippocampus. J. Neurosci., 15: 1768–1777. ule cells after adrenalectomy. Neuroscience, 108: 263–272.
Smriga, M., Saito, H. and Nishiyama, N. (1996) Hippocampal Wright, R.L. and Conrad, C.D. (2005) Chronic stress leaves
long- and short-term potentiation is modulated by adrenal- novelty-seeking behavior intact while impairing spatial rec-
ectomy and corticosterone. Neuroendocrinology, 64: 35–41. ognition memory in the Y-maze. Stress, 8: 151–154.
Stell, B.M., Brickley, S.G., Tang, C.Y., Farrant, M. and Mody, Zhou, J. and Cidlowski, J.A. (2005) The human glucocorticoid
I. (2003) Neuroactive steroids reduce neuronal excitability by receptor: one gene, multiple proteins and diverse responses.
selectively enhancing tonic inhibition mediated by delta Steroids, 70: 407–417.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 22

Neurotrophins in the dentate gyrus

Devin K. Binder

Department of Neurological Surgery, University of California, Irvine, CA 92868, USA

Abstract: Since the discovery of nerve growth factor (NGF) in the 1950s and brain-derived neurotrophic
factor (BDNF) in the 1980s, a great deal of evidence has mounted for the roles of neurotrophins (NGF;
BDNF; neurotrophin-3, NT-3; and neurotrophin-4/5, NT-4/5) in development, physiology, and pathology.
BDNF in particular has important roles in neural development and cell survival, as well as appearing
essential to molecular mechanisms of synaptic plasticity and larger scale structural rearrangements of axons
and dendrites. Basic activity-related changes in the central nervous system (CNS) are thought to depend on
BDNF modulation of synaptic transmission. Pathologic levels of BDNF-dependent synaptic plasticity may
contribute to conditions such as epilepsy and chronic pain sensitization, whereas application of the trophic
properties of BDNF may lead to novel therapeutic options in neurodegenerative diseases and perhaps even
in neuropsychiatric disorders. In this chapter, I review neurotrophin structure, signal transduction mech-
anisms, localization and regulation within the nervous system, and various potential roles in disease.
Modulation of neurotrophin action holds significant potential for novel therapies for a variety of
neurological and psychiatric disorders.

Keywords: brain-derived neurotrophic factor; neurotrophin-3; neurotrophin-4/5; nerve growth factor;


epilepsy

Introduction to neurotrophins antiparallel b-strands and cysteine residues in


a cystine knot motif. Initially produced as prone-
Neurotrophin structure urotrophins, prohormone convertases such as furin
cleave the proneurotrophins (molecular weight,
Each neurotrophin (including nerve growth factor, MW 30 kDa) to the mature neurotrophin (MW
NGF; brain-derived neurotrophic factor, BDNF; 14 kDa) (Chao and Bothwell, 2002). Proneuro-
neurotrophin-3, NT-3; neurotrophin-4/5, NT-4/5; trophins have altered binding characteristics and
see below) consists of a noncovalently linked ho- distinct biologic activity in comparison with mature
modimer and contains (1) a signal peptide following neurotrophins (Lee et al., 2001b). Mature ne-
the initiation codon; (2) a pro-region containing an urotrophins are noncovalently linked homodimers
N-linked glycosylation site and a proteolytic cleav- with MW approximately 28 kDa. Dimerization
age site for furin-like pro-protein convertases, fol- appears essential for neurotrophin (NT) receptor
lowed by the mature sequence; (3) a distinctive activation. BDNF shares approximately 50% amino
three-dimensional structure containing two pairs of acid identity with NGF, NT-3, and NT-4/5.
The BDNF gene (in humans mapped to chro-
Corresponding author. Tel.: +1 (714) 456-6966; mosome 11p) has four 50 exons (exons I–IV) that
Fax: +1 (714) 456-8212; E-mail: dbinder@uci.edu are associated with distinct promoters, and one 30

DOI: 10.1016/S0079-6123(07)63022-2 371


372

exon (exon V) that encodes the mature BDNF and function (Eide et al., 1996; Haapasalo et al.,
protein (Metsis et al., 1993; Timmusk et al., 2001, 2002). Expression of truncated trk receptors
1993b). Eight distinct mRNAs are transcribed, on astrocytes is upregulated following injury (Frisen
with transcripts containing exons I–III expressed et al., 1993) and may modulate neuronal vulnera-
predominantly in brain and exon IV found in lung bility (Saarelainen et al., 2000a) and sequestration
and heart (Timmusk et al., 1993b). of BDNF in astrocytes (Biffo et al., 1995; Roback
et al., 1995; Alderson et al., 2000). Recent studies
have shown that BDNF activates glial calcium
Neurotrophin signal transduction signaling by truncated trk receptors (Climent et al.,
2000; Rose et al., 2003).
Each NT binds one or more of the tropomyosin- In addition, all of the NTs bind to the p75
related kinase (trk) receptors, members of the receptor, designated p75NTR. p75NTR, related to
family of receptor tyrosine kinases (RTKs) proteins of the tumor necrosis factor (TNFR) su-
(Patapoutian and Reichardt, 2001). Trk proteins perfamily, has a glycosylated extracellular region
are transmembrane RTKs homologous to other involved in ligand binding, a transmembrane re-
RTKs such as the epidermal growth factor (EGF) gion, and a short cytoplasmic sequence lacking
receptor and insulin receptor family. Ligand- intrinsic catalytic activity (Chao and Hempstead,
induced receptor dimerization results in kinase ac- 1995; Dechant and Barde, 2002). NT binding to
tivation; subsequent receptor autophosphorylation p75NTR is linked to several intracellular signal
on multiple tyrosine residues creates specific binding transduction pathways, including nuclear factor-
sites for intracellular target proteins, which bind to kB (NF-kB), Jun kinase, and sphingomyelin
the activated receptor via SH2 domains (Barbacid, hydrolysis (Dechant and Barde, 2002). p75NTR
1994; Patapoutian and Reichardt, 2001). These in- signaling mediates biologic actions distinct from
clude PLCg1 (phospholipase C), p85 (the noncata- those of the trk receptors, notably the initiation
lytic subunit of PI-3 kinase), and Shc (SH2- of programed cell death (apoptosis) (Casaccia-
containing sequence); activation of these target pro- Bonnefil et al., 1996; Frade et al., 1996; Roux
teins can then lead to a variety of intracellular et al., 1999; Dechant and Barde, 2002). It has also
signaling cascades such as the Ras-MAP (mitogen- been suggested that p75 may serve to determine
activated protein) kinase cascade and phosphorylat- NT binding specificity (Esposito et al., 2001; Lee
ion of cyclic AMP-response element binding protein et al., 2001a; Zaccaro et al., 2001).
(CREB) (Patapoutian and Reichardt, 2001; Segal,
2003).
Binding specificity is conferred via the juxtamem- Nerve growth factor (NGF)
brane Ig-like domain of the extracellular portion of
the receptor in the following pattern (Urfer et al., NGF was discovered in the early 1950s by Rita
1995). TrkA binds NGF (with low-affinity binding Levi-Montalcini and Viktor Hamburger due to its
by NT-3 in some systems); trkB binds BDNF and trophic (survival and growth-promoting) effects on
NT-4/5 with lower-affinity binding by NT-3; and sensory and sympathetic neurons (Levi-Montalcini
trkC binds NT-3 (Barbacid, 1994). Trk receptors and Hamburger, 1951). In addition, NGF supports
exist in both a full-length (trkB.FL) form as well as the survival and neurotransmitter synthesis of
truncated (trkB.T1, trkB.T2) forms lacking the kin- cholinergic neurons in the central nervous system
ase domain (Eide et al., 1996; Fryer et al., 1997). (CNS). In the brain, it is synthesized primarily in
Although most functions attributed to BDNF are cholinergic target tissues such as the cortex, hippo-
associated with full-length trkB, several roles have campal pyramidal layer, and striatum (Gall and
been suggested for truncated receptors, including Isackson, 1989; Rylett and Williams, 1994). The
growth and development (Fryer et al., 1997; trkA NT receptor is expressed primarily on the
Yacoubian and Lo, 2000; Luikart et al., 2003) and axons of NGF-dependent cholinergic neurons
negative modulation of trkB receptor expression (Sobreviela et al., 1994).
373

Table 1. Seizure regulation of NGF expression

Reference Methods Results

NGF mRNA
Gall and Isackson (1989) Hilar electrolytic lesion Increases in DG (4 h) and cortex (17 h)
Gall and Lauterborn (1992) Hilar electrolytic lesion Increase in DG, max at 6 h (10  ) and 24 h (6  )
(biphasic)
Ernfors et al. (1991) Rapid kindling (ventral Increase max at 1 h (DG), 4 h (PC)
hippocampal stimulation)
Bengzon et al. (1993) Traditional kindling (ventral Increase — did not do time course but studied
hippocampal stimulation relationship between development of kindling
CA1–CA2) and neurotrophin induction — found similar
NGF induction (approximately 2  at 2 h time
point) regardless of kindling stage
Schmidt-Kastner et al. (1996) Pilocarpine Increase in DG max at 3–6 h
Mudo et al. (1996) Pilocarpine Increase in DG max at 3 h (approximately 2.5  )
Sato et al. (1996) Traditional amygdala kindling Increase in CA1, CA3, perirhinal cortex 1 h after
stage 5 kindled seizure
Morimoto et al. (1998) Traditional amygdala kindling Increase in DG, max at 2 h (2  )
NGF protein
Bengzon et al. (1992) Rapid hippocampal kindling After 40 stimulations, NGF protein levels
(ventral hippocampal (measured by 2-site ELISA) increased to 150% in
stimulation) NGF ELISA DG at 7 days, 260% in PC at 12 h, and 170% in
parietal cortex at 24 h

NGF gene regulation Basal levels of trkA mRNA in the hippocampus


are very low (Cellerino, 1996; Mudo et al., 1996),
NGF expression levels are regulated by activity. This and do not appear to increase following kindling
has been most clearly demonstrated following the (Bengzon et al., 1993; Merlio et al., 1993) or pi-
intense activity associated with seizures (Table 1). locarpine status epilepticus (Mudo et al., 1996)
Hilar electrolytic lesion-induced (Gall and Isackson, (Table 2).
1989; Gall and Lauterborn, 1992; Lauterborn et al.,
1994) or kindled seizures (Ernfors et al., 1991;
Bengzon et al., 1993; Sato et al., 1996; Morimoto Brain-derived neurotrophic factor (BDNF)
et al., 1998) induce a rapid and transient expression
of NGF mRNA in dentate gyrus granule cells as In 1982, BDNF, the second member of the ‘‘NT’’
well as piriform cortex. Similarly, pilocarpine- family of neurotrophic factors, was shown to pro-
induced status epilepticus increases NGF mRNA mote survival of a subpopulation of dorsal root
expression in dentate gyrus, maximum at approxi- ganglion neurons, and subsequently purified from
mately 3 h (Mudo et al., 1996; Schmidt-Kastner pig brain (Barde et al., 1982). The amino acid
et al., 1996). sequence of BDNF was found to have a strong
Basal levels of NGF protein in the hippocampus homology with NGF. Since then, other members
are very low (Narisawa-Saito and Nawa, 1996). of the NT family such as NT-3 (Maisonpierre
Studies with NGF ELISAs indicate that NGF et al., 1990b) and NT-4/5 (Hallbook et al., 1991; Ip
protein levels do increase following kindling, es- et al., 1992) have been described, each with a dis-
pecially in dentate gyrus, piriform cortex, and pa- tinct profile of trophic effects on subpopulations of
rietal cortex (Bengzon et al., 1992). In contrast to neurons in the peripheral and CNS.
the mRNA increases, NGF protein levels remain BDNF mRNA has a widespread distribution in
elevated for at least 7 days (Bengzon et al., 1992). the CNS (Merlio et al., 1993; Conner et al., 1997),
374

Table 2. Seizure regulation of trk receptor expression

Reference Methods Results

TrkA mRNA
Cellerino (1996) Basal hippocampal Basal not detectable by 35S; can only
detect with 33P
Mudo et al. (1996) Pilocarpine Basal not detectable and pilocarpine
status did not increase levels
Bengzon et al. (1993) Hippocampal kindling No change
Merlio et al. (1993) Rapid hippocampal kindling No change
TrkB mRNA
Bengzon et al. (1993) Rapid hippocampal kindling Increased in DG 2 h (2  ) after focal or
(ventral hippocampal stimulation) generalized seizures
Merlio et al. (1993) Rapid hippocampal kindling Increased threefold in DG at 30 min after
(ventral hippocampal stimulation) 40 stimulations
Elmer et al. (1996a, b) Rapid hippocampal kindling Increased in DG max at 2 h
(ventral hippocampal stimulation)
Humpel et al. (1993) PTZ Increased twofold in DG at 3 h
Nibuya et al. (1995) ECS Increased fivefold (DG) at 2 h
Schmidt-Kastner et al. (1996) Pilocarpine Increase in DG, CA1–3 at 3–6 h
Mudo et al. (1996) Pilocarpine Increase in DG, amygdala, PC at 3 h
Hughes et al. (1998) Hippocampal afterdischarge Increase in DG at 4 h
TrkC mRNA
Bengzon et al. (1993) Rapid hippocampal kindling Increased in DG 2 h (1.5  ) after focal or
(ventral hippocampal stimulation) generalized seizures
Merlio et al. (1993) Rapid hippocampal kindling No change after 40 rapid K stimulations
(ventral hippocampal stimulation)
Mudo et al. (1995) ICV KA or ICV bicuculline Increase max at 3 h (bicuculline) or 12 h
(KA) confined to DG

including limbic forebrain, neocortex, and is abun- Localization, transport and release
dant in all principal neurons of the hippocampus
(Ernfors et al., 1990; Hofer et al., 1990; Wetmore Unlike the classical target-derived trophic factor
et al., 1990). Like BDNF mRNA, constitutive model in which NTs — such as NGF — are
BDNF protein expression is widespread (Conner retrogradely transported, there is now abundant
et al., 1997; Yan et al., 1997b), localized on neu- evidence that BDNF is also anterogradely trans-
ronal cell bodies, axons, and dendrites. The mossy- ported in brain. First, BDNF protein is localized
fiber axons of hippocampal dentate granule cells to nerve terminals (Conner et al., 1997), and path-
display especially intense BDNF immunoreactivity way transection or axonal transport inhibition
(Conner et al., 1997). The principal receptor for abrogates this terminal expression (Altar et al.,
BDNF, trkB, is a receptor tyrosine kinase, which 1997; Conner et al., 1997; Altar and DiStefano,
is found in both catalytic and truncated forms in 1998). Second, higher resolution studies have
the adult forebrain (Fryer et al., 1996; Drake et al., shown that BDNF is associated with dense-core
1999). TrkB mRNA and protein are found in hip- vesicles (Fawcett et al., 1997; Altar and DiStefano,
pocampus (Merlio et al., 1992; Altar et al., 1994; 1998), which are the primary site for neuropeptide
Yan et al., 1997a). Truncated trkB is also found in storage and release from nerve terminals. Third,
the ependymal cells lining the ventricular cavities, further functional studies have supported the
effectively limiting diffusion of intraventricularly anterograde transport hypothesis (Fawcett et al.,
administered BDNF (Yan et al., 1994; Anderson 1998, 2000). Fourth, pro-BDNF is shuttled from
et al., 1995). the trans-Golgi network into secretory granules,
375

where it is cleaved by prohormone convertase 1 cycle, which correlate with its effects on neural
(PC1) (Farhadi et al., 2000). excitability (Scharfman et al., 2003).
In addition, emerging evidence suggests that Distinct BDNF 50 exons are differentially reg-
both BDNF and trk receptors may undergo reg- ulated by stimuli such as neural activity. For
ulated intracellular transport. For example, sei- example, exons I–III, but not exon IV, increase
zures lead to redistribution of BDNF mRNA from after kainic acid-induced seizures (Timmusk et al.,
hippocampal CA3 cell bodies to their apical dend- 1993b) or other stimuli that increase activity
rites (Bregola et al., 2000; Simonato et al., 2002). (Lauterborn et al., 1996; Tao et al., 2002). Pro-
Trk signaling is now thought to include retrograde tein synthesis is required for the effects of activity
transport of intact NT-trk complexes to the neu- on exon I and II, but not III and IV, raising the
ronal cell body (Miller and Kaplan, 2001; Ginty possibility that the latter act as immediate early
and Segal, 2002). genes (Lauterborn et al., 1996; Castrén et al.,
Recent evidence indicates that NTs are released 1998). The transcription factor CaRF (calcium
acutely following neuronal depolarization (Gries- response factor) activates transcription of exon III
beck et al., 1999; Mowla et al., 1999; Hartmann et under the control of a calcium response element,
al., 2001; Egan et al., 2003; Goggi et al., 2003; CaRE1 (Tao et al., 2002). CREB, which can be
Brigadski et al., 2005). In fact, direct activity- stimulated by diverse stimuli ranging from activity
dependent pre- to postsynaptic transneuronal to chronic antidepressant treatment (Nibuya et al.,
transfer of BDNF has been demonstrated using 1995, 1996; Shieh et al., 1998; Tao et al., 1998;
fluorescently labeled BDNF (Kohara et al., 2001). Shieh and Ghosh, 1999), also modulates exon III
The released form of BDNF is thought to be pro- transcription. Recent evidence also indicates that
BDNF (Mowla et al., 2001), raising the possibility neural activity triggers calcium-dependent phos-
of postsecretory proteolytic processing by mem- phorylation and release of MeCP2 (methyl-CpG
brane-associated or extracellular proteases in the binding protein 2) from BDNF promoter III to
modulation of BDNF action (Lee et al., 2001b). derepress transcription (Chen et al., 2003).
Pathologic states are also associated with altera-
tion in BDNF gene expression. For example,
BDNF gene regulation seizures dramatically upregulate BDNF mRNA
(Table 3). A wide variety of seizure paradigms
A multitude of stimuli have been described that (kindling; kainic acid; pilocarpine; pentylenetetrazol,
alter BDNF gene expression in both physiologic PTZ; electroconvulsive shock, ECS) rapidly and
and pathologic states (Lindholm et al., 1994). dramatically increase expression of BDNF mRNA
Physiologic stimuli are known to increase BDNF in dentate gyrus as well as in other areas of the hip-
mRNA content. For example, light stimulation pocampus and cortex (Ernfors et al., 1991; Isackson
increases BDNF mRNA in visual cortex (Castrén et al., 1991; Gall and Lauterborn, 1992; Dugich-
et al., 1992), osmotic stimulation increases BDNF Djordjevic et al., 1992a, b; Bengzon et al., 1993;
mRNA in the hypothalamus (Castrén et al., 1995; Humpel et al., 1993; Nibuya et al., 1995; Mudo
Dias et al., 2003), and whisker stimulation in- et al., 1996; Sato et al., 1996; Schmidt-Kastner et al.,
creases BDNF mRNA expression in somatosen- 1996). This is associated with a transient upregula-
sory barrel cortex (Rocamora et al., 1996). tion of BDNF protein (Nawa et al., 1995; Elmer
Electrical stimuli that induce long-term potentiat- et al., 1996b; Hughes et al., 1998).
ion (LTP) in the hippocampus, a cellular model of TrkB mRNA and protein in the dentate gyrus
learning and memory, increase BDNF and NGF are also upregulated following various seizure
expression (Patterson et al., 1992; Castrén et al., protocols (Bengzon et al., 1993; Merlio et al.,
1993; Bramham et al., 1996). Even physical exer- 1993; Elmer et al., 1996a) (Table 2). TrkB mRNA
cise has been shown to increase NGF and BDNF expression is increased in dentate granule cells
expression in hippocampus (Neeper et al., 1995). 2–6 h after rapid electrical kindling, hippocampal
Interestingly, BDNF levels vary across the estrous after discharge, PTZ kindling, ECS, or pilocarpine
376

Table 3. Seizure regulation of BDNF expression

Reference Methods Results

BDNF mRNA
Isackson et al. (1991) Hilar electrolytic lesion Increase, onset o1.5 h, max at 6 h (12  )
Gall and Lauterborn (1992) Perforant path stimulation (1 AD) Increase, onset 20 min, max at 4 h (12  )
Ernfors et al. (1991) Rapid kindling (ventral Increase max at 30 min (DG/PC), 1 h (CA1)
hippocampal stimulation)
Dugich-Djordjevic et al. (1992b) KA Increase in all hippocampus (P21, P40), no
change despite seizures (P8)
Dugich-Djordjevic et al. (1992a) KA Increase max at 30 min (10  in DG, 2–6  in
CA1, CA3, CA4); later in cortex
Bengzon et al. (1993) Traditional kindling (ventral Increase — did not do time course but studied
hippocampal stimulation CA1–2) relationship between development of kindling
and neurotrophin induction — found similar
BDNF induction (approximately 9  at 2 h time
point) regardless of kindling stage
Humpel et al. (1993) PTZ kindling (30 mg/kg i.p. Increase max at 3 h (DG, PC, amygdala) after
followed by convulsive dose acute convulsive PTZ
(50 mg/kg))
Nibuya et al. (1995) ECS Increase 30-fold 2 h after ECS (DG), fivefold
(PC)
Schmidt-Kastner et al. (1996) Pilocarpine Increase max at 3–6 h (DG, other hippocampal,
neocortex, PC, striatum, thalamus)
Mudo et al. (1996) Pilocarpine Increase max at 3–6 h (DG, amygdala, PC)
Sato et al. (1996) Traditional amygdala kindling 4–5-fold increase in DG 1 h after stage 5 kindled
seizure
BDNF protein
Nawa et al. (1995) Hilar electrolytic lesion Highest levels of basal BDNF in hippocampus
(followed by hypothalamus, neocortex,
cerebellum, thalamus and striatum)
BDNF ELISA Fourfold induction of BDNF protein levels in
hippocampus, maximum at 24 h after HL, down
by 1 week
Elmer et al. (1996a, b) Rapid kindling (ventral Basal levels of BDNF highest in
hippocampal stimulation) DGCA3>CA1>PC
BDNF ELISA After 1 AD: increase to 150% in DG at 6 h, 200%
in CA3 at 12 h, 50% in PC at 6 h
After 40 ADs: increase to 200% in DG at 6 and
24 h, 150% in CA3 at 6 h, 300% in PC at 2 and
6h
After 7 ADs: increase to 200% in CA3 at 24 h
Hughes et al. (1998) Hippocampal after discharge Increase in DG at 4 h

status epilepticus (Bengzon et al., 1993; Humpel et Role(s) of BDNF during development
al., 1993; Nibuya et al., 1995; Elmer et al., 1996a;
Mudo et al., 1996; Schmidt-Kastner et al., 1996; In vitro and in vivo studies have demonstrated
Hughes et al., 1998). Subcellular studies have dem- that BDNF has survival- and growth-promoting
onstrated targeting of BDNF and trkB mRNAs actions on a variety of CNS neurons, including
to dendrites in CA3 neurons following kindled dorsal root ganglion cells, dopaminergic and
seizures (Simonato et al., 2002). cholinergic neurons, retinal ganglion cells, and
377

hippocampal and cortical neurons (Johnson et al., BDNF. Overall, BDNF appears to strengthen
1986; Alderson et al., 1990; Hyman et al., 1991; excitatory (glutamatergic) synapses and weaken in-
Knusel et al., 1991; Acheson et al., 1995; Patel and hibitory (GABAergic) synapses. Schuman and
McNamara, 1995; Huang and Reichardt, 2001). colleagues demonstrated that exposure of adult rat
Certain peripheral sensory neurons, especially hippocampal slices to BDNF led to a long-lasting
those in vestibular and nodose-petrosal ganglia, potentiation of synaptic strength at Schaffer collat-
depend on the presence of BDNF because BDNF eral-CA1 synapses (Kang and Schuman, 1995). Sub-
homozygous knockout (BDNF/) mice lack sequent studies have supported a role of BDNF in
these neurons (Huang and Reichardt, 2001). Un- LTP (Korte et al., 1995; Korte et al., 1996; Patterson
like NGF, sympathetic neurons are not affected, et al., 1996; Kang et al., 1997; Xu et al., 2000). For
nor are motor neurons. BDNF/ mice fail to example, incubation of hippocampal or visual cor-
thrive, demonstrate lack of proper coordination of tical slices with trkB inhibitors inhibits LTP (Figurov
movement and balance, and ultimately die by 3 et al., 1996), and hippocampal slices from BDNF/
weeks of age. However, heterozygous BDNF mice exhibit impaired LTP induction (Korte et al.,
knockout (BDNF+/) mice are viable, and ex- 1995), which is restored by reintroduction of BDNF
hibit a variety of phenotypes, including obesity (Korte et al., 1996; Patterson et al., 1996).
(Lyons et al., 1999; Kernie et al., 2000), decreased Whether BDNF-induced synaptic potentiation
seizure susceptibility (Kokaia et al., 1995), and occurs primarily by a presynaptic action (e.g.
impaired spatial learning (Linnarsson et al., 1997). through enhancement of glutamate release)
Interestingly, conditional postnatal BDNF gene or postsynaptically (e.g. via phosphorylation of
deletion (Rios et al., 2001) and reduction in trkB neurotransmitter receptors) is intensely debated
expression (Xu et al., 2003) also cause obesity. (Schinder and Poo, 2000). A number of studies
Physiologic regulation of BDNF gene expres- have provided evidence for a presynaptic locus
sion may be very important in the development (Xu et al., 2000; Tyler et al., 2002; see also Kafitz
of the brain. For example, BDNF contributes to et al., 1999), yet evidence for postsynaptic actions
activity-dependent development of the visual cor- has also been obtained (Black, 1999; Thakker-
tex. Provision of excess BDNF (Cabelli et al., Varia et al., 2001; reviewed in Poo, 2001). Both
1995) or blockade of BDNF signaling (Cabelli pre- and postsynaptic trkB receptors in the hip-
et al., 1997) leads to abnormal patterning of ocular pocampus may be important (Drake et al., 1999).
dominance columns during a critical period of A role for BDNF in GABAergic synapses was
visual cortex development. This suggests a role for first raised by studies showing that BDNF influ-
BDNF in axonal pathfinding during development. ences GABAergic neuronal phenotype (Marty et al.,
BDNF also has powerful effects on dendritic mor- 1996; McLean Bolton et al., 2000). Subsequently,
phology (McAllister et al., 1997; Murphy et al., BDNF was shown to decrease inhibitory (GAB-
1998; Horch and Katz, 2002; Tolwani et al., 2002). Aergic) synaptic transmission (Tanaka et al., 1997;
Frerking et al., 1998; Wardle and Poo, 2003). Re-
cent evidence shows that BDNF can modulate the
Effects on synaptic transmission function of GABAA receptors via modulation
of phosphorylation state (Jovanovic et al., 2004).
BDNF has an enormous range of physiologic ac- Interestingly, BDNF may also regulate the efficacy
tions at both developing and mature synapses, over- of GABAergic synapses by direct downregulation of
all enhancing synaptic transmission by both pre- and the neuronal K+-Cl cotransporter, which would
postsynaptic mechanisms. The first studies of BDNF impair neuronal Cl extrusion and weaken GAB-
effects on synaptic transmission showed that BDNF Aergic inhibition (Rivera et al., 2002). Similarly,
increased the frequency of miniature excitatory a recent paper found differential effects of BDNF
postsynaptic currents (EPSCs) at Xenopus neuro- on GABA-mediated currents in excitatory and
muscular synapses (Lohof et al., 1993). Since then, inhibitory neuron subpopulations, selectively de-
numerous studies have examined the actions of creasing the efficacy of inhibitory neurotransmission
378

by downregulation of Cl transport (Wardle and humans and animals, appears to be an important site
Poo, 2003). of BDNF action. Rapid and selective induction
of BDNF expression in the hippocampus during
contextual learning has been demonstrated (Hall
Effect on neurogenesis
et al., 2000), and function-blocking antibodies to
BDNF (Alonso et al., 2002), BDNF knockout
An important feature of the dentate gyrus is the
(Linnarsson et al., 1997), knockout of forebrain
lifelong production of new granule cells from pro-
trkB signaling (Minichiello et al., 1999), or overex-
genitor cells located in the subgranular zone
pression of truncated trkB (Saarelainen et al., 2000b)
(Gould and McEwen, 1993; Scharfman, 2004).
in mice impairs spatial learning. Another study dem-
BDNF has also been found to enhance neurogen-
onstrated upregulation of BDNF in monkey parietal
esis. Intraventricular infusion of BDNF or adeno-
cortex associated with tool-use learning (Ishibashi
viral-induced BDNF activity increases the number
et al., 2002). In humans, a valine to methionine
of neurons in the adult olfactory bulb, striatum,
polymorphism at the 50 pro-region of the human
septum, and thalamus (Zigova et al., 1998;
BDNF protein was found to be associated with
Benraiss et al., 2001; Pencea et al., 2001), which
poorer episodic memory; in vitro, neurons trans-
can be potentiated by concurrent inhibition of glial
fected with met-BDNF-green fluorescence protein
differentiation of subependymal progenitor cells
(GFP) exhibited reduced depolarization-induced
(Chmielnicki et al., 2004). Intrahippocampal infu-
BDNF secretion (Egan et al., 2003).
sion of BDNF into adult rats leads to increased
dentate gyrus neurogenesis, accompanied by in-
creased numbers of ectopic granule cells (Scharf- Neurotrophin-3 (NT-3)
man et al., 2005). BDNF+/ mice show decreased
numbers of BrdU-labeled cells in the dentate gyrus NT-3, first described in 1990 (Maisonpierre et al.,
(Lee et al., 2002). Studies of cultured progenitor 1990b), is similar to BDNF in several ways. Like
cells have elucidated some of the signaling mech- BDNF, NT-3 mRNA and protein are widely dis-
anisms, which appear to involve trkB activation, tributed in the adult CNS (Maisonpierre et al.,
followed by activation of the MAP kinase and PI3- 1990a, b, Zhou and Rush, 1994; Katoh-Semba
kinase pathways (Barnabe-Heider and Miller, et al., 1996). While the preferred receptor for NT-3
2003) and downstream modification of basic he- is trkC, NT-3 can also bind to trkA and trkB
lix-loop-helix transcription factors (Ito et al., (Barbacid, 1994; Ryden and Ibanez, 1996; Huang
2003). Although some studies have concluded that and Reichardt, 2003). Like BDNF, NT-3 is in-
the primary effect of BDNF is on proliferation volved in synaptic transmission and neuronal
(Katoh-Semba et al., 2002), other experiments excitability (Thoenen, 1995). Addition of NT-3
suggest an important effect on survival (Lee et al., to hippocampal slices enhances synaptic strength
2002). The effects of BDNF may depend on a at Schaffer collateral-CA1 synapses (Kang and
previous history of ischemic damage (Larsson Schuman, 1995). NT-3 enhances paired-pulse fa-
et al., 2002; Gustafsson et al., 2003). cilitation in the perforant path-dentate gyrus path-
way (Kokaia et al., 1998; Asztely et al., 2000). Like
BDNF, NT-3 reduces GABAergic inhibition (Kim
Effects on learning and memory
et al., 1994). Also like BDNF, NT-3 enhances the
survival and differentiation of neural progenitor
Learning and memory depend on persistent selective
cells (Barnabe-Heider and Miller, 2003).
modification of synapses between CNS neurons.
Since BDNF appears to be involved in activity-
dependent synaptic plasticity, there is great interest NT-3 gene regulation
in its role in learning and memory (Yamada and
Nabeshima, 2003). The hippocampus, which is re- However, whereas NGF and BDNF levels increase
quired for many forms of long-term memory in after seizures, NT-3 levels are reduced in dentate
379

Table 4. Seizure regulation of NT-3 and NT-4 expression

Reference Methods Results

NT-3 mRNA
Gall and Lauterborn (1992) Hilar electrolytic lesion Decrease, onset 12 h max 12 h (20%) in DG
Bengzon et al. (1993) Traditional kindling (ventral Decrease — did not do time course but studied
hippocampal stimulation CA1-CA2) relationship between development of kindling
and neurotrophin induction — found similar NT-
3 decrease (50%) regardless of kindling stage
Schmidt-Kastner and Olson (1995) Pilocarpine Decrease max at 3 h
Mudo et al. (1996) Pilocarpine Decrease max at 12–24 h (40–50%)
Kim et al. (1998) Amygdala kindling Decrease in DG
NT-4 mRNA
Timmusk et al. (1993a, b) RNAse protection analysis from Could not detect NT-4 by in situ but did find very
different tissues low levels in adult brain using RNAse protection
No increase in NT-4 mRNA following either
systemic KA or hippocampal stimulation
Mudo et al. (1996) Pilocarpine Basal not detectable and pilocarpine status did
not increase levels

gyrus granule neurons (Gall and Lauterborn, 5 protects hippocampal and cortical neurons
1992; Bengzon et al., 1993; Schmidt-Kastner and against energy deprivation-induced injury (Cheng
Olson, 1995; Mudo et al., 1996; Kim et al., 1998) et al., 1994) and adrenalectomy-induced apoptosis
(Table 4). This suggests that the potential role of of hippocampal granule cells (Qiao et al., 1996).
NT-3 in seizure progression is different. Whether Application of NT-4/5 enhances excitatory synap-
trkC is elevated appears to depend on the model tic transmission in cultured hippocampal neurons
used; rapid kindling induces no change (Merlio et (Lessmann et al., 1994).
al., 1993) or a transient increase (Bengzon et al.,
1993) in trkC mRNA levels, and ICV KA or ICV
bicuculline transiently increase trkC mRNA in Roles of neurotrophins in epilepsy models
dentate granule cells (Mudo et al., 1995).
The discovery that limbic seizures increase NGF
mRNA levels (Gall and Isackson, 1989) led to the
idea that seizure-induced expression of neurotrophic
Neurotrophin-4/5 factors may contribute to the lasting structural and
functional changes underlying epileptogenesis (Gall
The fourth member of the NT family, NT-4/5, was et al., 1991, 1997; Jankowsky and Patterson, 2001).
discovered after NGF, BDNF, and NT-3 (Hall-
book et al., 1991; Ip et al., 1992). Levels of NT-4/5
in the brain are very low at baseline (Timmusk NGF and epilepsy
et al., 1993a; Katoh-Semba et al., 2003) and are
not increased by seizures (Timmusk et al., 1993a; Is there a functional role for NGF gene upregula-
Mudo et al., 1996) (Table 4). NT-4/5/ mice, un- tion in epileptogenesis? Indeed, intraventricular
like BDNF/ mice, are normal and long-lived administration of NGF antibodies retards amygdala
with no obvious neurological deficits (Conover kindling (Funabashi et al., 1988) and blocks kin-
et al., 1995; Liu et al., 1995). The only loss of dling-induced mossy-fiber sprouting (Van der Zee
neurons in NT-4/5/ mice appears to be a reduc- et al., 1995). Similarly, an NGF inhibitory peptide
tion in the number of sensory neurons in the no- inhibits amygdala kindling and mossy-fiber sprout-
dose-petrosal and geniculate ganglia (Conover ing (Rashid et al., 1995). Conversely, intraventricu-
et al., 1995; Liu et al., 1995). Provision of NT-4/ lar NGF infusion was found to facilitate amygdala
380

and hippocampal kindling and increase mossy-fiber BDNF in transgenic mice leads to spontaneous
sprouting (Adams et al., 1997). seizures (Croll et al., 1999), and intrahippocampal
Whether NGF exerts its effects on kindling and infusion of BDNF is sufficient to induce seizure
kindling-induced morphological changes via trkA activity in vivo (Scharfman et al., 2002).
or p75NTR has been investigated. Inhibition of A separate group of experiments has demon-
Ras, a downstream effector of trkA, inhibits kin- strated that chronic BDNF infusion can inhibit
dling and kindling-associated mossy-fiber sprout- kindling (Larmet et al., 1995; Osehobo et al., 1996;
ing (Li et al., 2003). Peptide inhibitors of NGF Reibel et al., 2000b). These inhibitory effects ap-
binding to trkA but not to p75NTR can inhibit pear to be due to trkB receptor downregulation
kindling, whereas both trkA and p75NTR inhibi- following chronic BDNF administration, and
tion can inhibit mossy-fiber sprouting (Li et al., hence are still consistent with the ‘‘pro-
2005). epileptogenic BDNF’’ hypothesis. This interpreta-
What is the locus of effects of NGF on hippo- tion is supported by the observation that chronic
campal kindling? As described above, there is little exposure to BDNF in vitro leads to downregula-
evidence for trkA expression in hippocampus tion of trkB mRNA and protein (Knusel et al.,
(Sobreviela et al., 1994; Cellerino, 1996). Simi- 1997). Similarly, continuous in vivo intrahippo-
larly, there is little expression of p75 in hippocam- campal BDNF infusion results in downregulation
pus at baseline (Pioro and Cuello, 1990; Sobreviela of trkB protein by as much as 80% (Frank et al.,
et al., 1994). It is likely that NGF-dependent 1996). Thus, whereas chronic BDNF infusion in-
effects are due to modulation of the cholinergic hibits kindling progression, acute microinjections
system, as both trkA and p75NTR (Hofer et al., of BDNF enhance epileptogenesis in the absence
1990) receptors are most strongly expressed in the of effect on trkB expression (Xu et al., 2004). Fur-
basal forebrain cholinergic neurons which project thermore, chronic infusions of BDNF may upreg-
to hippocampus (Sobreviela et al., 1994). Consist- ulate the inhibitory neuropeptide Y (NPY) (Reibel
ent with this hypothesis is that cholinergic agonists et al., 2000a).
and antagonists produce effects on kindling and Whether BDNF has a significant effect on sei-
sprouting parallel to those of NGF (Adams et al., zure-associated mossy-fiber sprouting is not clear.
2002). While mossy-fiber sprouting has been reported in
BDNF+/ mice and following BDNF infusion
(Kokaia et al., 1995; Scharfman et al., 2002), there
BDNF and epilepsy is no effect on mossy-fiber sprouting in BDNF-
overexpressing mice or following chronic infusion
Abundant in vitro and in vivo evidence implicates or bolus injection of BDNF in other studies (Qiao
BDNF in the cascade of electrophysiologic and et al., 2001; Xu et al., 2004). However, BDNF
behavioral changes underlying the epileptic state overexpression does increase dendritic length and
(Binder et al., 2001). BDNF mRNA and protein complexity in the hippocampus (Tolwani et al.,
are markedly upregulated in the hippocampus by 2002). The relative role of BDNF on effect synap-
seizure activity in animal models (Ernfors et al., tic changes vs. larger scale morphological changes
1991; Isackson et al., 1991; Lindvall et al., 1994; during epileptogenesis remains to be clarified.
Nibuya et al., 1995). Infusion of trkB receptor The anatomic locus of action of NTs during
body (a chimera of human IgG-Fc domain and the epileptogenesis has been clarified with the study of
extracellular domain of the trkB receptor) (Binder trk receptor activation (see below).
et al., 1999b) or use of BDNF+/ (Kokaia et al.,
1995) or truncated trkB-overexpressing (Lahteinen
et al., 2002) mice inhibits epileptogenesis in animal NT-3 and epilepsy
models. Conversely, direct application of BDNF
induces hyperexcitability in vitro (Scharfman, In comparison with BDNF, what is the evidence
1997; Scharfman et al., 1999), overexpression of for a role of NT-3 in epileptogenesis? In NT-3+/
381

mice, which have 30% reduction in basal NT-3 correspond to the previously reported threshold for
mRNA levels, amygdala kindling was markedly increase in BDNF gene expression. These observa-
retarded (Elmer et al., 1997). However, compen- tions are examined in greater detail in the next few
satory changes in BDNF and trkB mRNA levels in sections.
these mice made these data difficult to interpret
(Elmer et al., 1997). Chronic intraventricular in-
Anatomy of seizure-induced phospho-trk
fusion of NT-3 retards the development of behavi-
immunoreactivity
oral seizures (Xu et al., 2002), probably in part via
downregulation of trk phosphorylation (Xu et al.,
Following seizure activity, phospho-trk immuno-
2002).
reactivity is selectively increased in dentate hilus
What about the effects of NT-3 on kindling-
and CA3 stratum lucidum of hippocampus (Bin-
induced mossy-fiber sprouting in the dentate
der et al., 1999a). This distribution precisely coin-
gyrus? Chronic infusion of NT-3 inhibits kin-
cides with the ‘‘mossy fiber’’ pathway of dentate
dling-associated mossy-fiber sprouting (Xu et al.,
granule cell axon terminals. In addition, this an-
2002). However, this effect is unclear as infusion of
atomic pattern coincides with the distribution of
NT-3 in the absence of kindling actually enhances
both basal and seizure-induced BDNF protein.
sprouting of mossy fibers in the inner molecular
Basal BDNF protein is also localized in hilus and
layer of the dentate gyrus and CA3 stratum oriens
CA3 stratum lucidum (Conner et al., 1997), and
(Xu et al., 2002).
seizures increase levels of BDNF protein in dentate
gyrus and CA3 (Elmer et al., 1998) and BDNF
NT-4 and epilepsy immunoreactivity in hilus and CA3 stratum lucid-
um (Smith et al., 1997; Yan et al., 1997b; Rudge et
Unlike NGF, BDNF, and NT-3 levels, levels of al., 1998; Vezzani et al., 1999). This precise an-
NT-4/5 do not appear to be regulated by seizure atomic colocalization of increased phospho-trk
activity (Timmusk et al., 1993a; Mudo et al., immunoreactivity and increases in BDNF protein
1996). The amygdala kindling phenotype of NT-4/ suggests that the phospho-trk immunoreactivity is
5/ mice was studied (He et al., 2006). No aspect caused by seizure-induced increases in BDNF.
of the development or persistence of amygdala BDNF, but not NGF, is known to increase levels
kindling was different between NT-4/5/ and of NPY (Croll et al., 1994), and kindling and
wild-type mice (He et al., 2006). kainate-induced seizures increase NPY immuno-
reactivity in hilus and CA3 stratum lucidum
(Marksteiner et al., 1990; Tønder et al., 1994),
Trk receptor activation following seizure activity
further implicating seizure-induced BDNF acting
in the mossy-fiber pathway. While NGF mRNA
The ability to monitor trk receptor activation fol-
content is upregulated by seizures, the anatomic
lowing seizures using phospho-specific trk antibodies
distribution of increased NGF protein is not
enabled identification of the anatomy, time course,
known. Thus, these anatomic considerations are
and threshold characteristics of trk receptor activa-
most consistent with a role for BDNF.
tion in the hippocampus following seizure activity
(Binder et al., 1999a). Kainate-induced status epi-
lepticus or hippocampal electrographic seizures in- Time course of seizure-induced phospho-trk
crease phospho-trk immunoreactivity selectively in immunoreactivity
the hippocampus, primarily confined to the dentate
hilus and CA3 stratum lucidum. This seizure- The time course of known BDNF upregulation
induced phospho-trk immunoreactivity is marked following seizures coincides temporally with in-
but transient, maximal at 24–48 h but back to base- creased phospho-trk immunoreactivity. Using hip-
line by 1 week. The seizure duration threshold for pocampal microdissection and a two-site ELISA
increase in phospho-trk immunoreactivity appears to for BDNF, Elmer et al. showed that after seven
382

ventral hippocampal electrographic seizures, the mRNA content may not only be necessary for
maximum increase in BDNF protein occurs at 12 h any increase in phospho-trk immunoreactivity but
in dentate gyrus and 24 h in CA3 (Elmer et al., also sufficient for maximal increase in phospho-trk
1998). Similarly, maximum increases in BDNF immunoreactivity following seizures.
protein following hilus lesion-induced (Nawa et
al., 1995) or kainate-induced (Rudge et al., 1998)
seizures occur at approximately 24 h in hippocam-
pus. Importantly, BDNF protein levels in both of Evidence that the trk receptor activated by seizures
these studies returned to baseline after 1 week, is trkB
similar to phospho-trk immunoreactivity. In con-
trast, Bengzon et al. found maximal NGF protein Indirect evidence suggests that BDNF-induced
content (measured by two-site immunoassay) 7 trkB activation is responsible for the increased
days after a similar rapid kindling protocol (Ben- phospho-trk immunoreactivity following seizures.
gzon et al., 1992) and did not see NGF protein First, the mRNA content of NGF and BDNF is
increases at earlier time points. Similarly, Lowen- increased following seizures (Ernfors et al., 1991;
stein et al. found maximal NGF-like neurotrophic Isackson et al., 1991; Lindvall et al., 1994; Nibuya
activity of hippocampal extracts from animals 1 et al., 1995) whereas dentate granule NT-3 mRNA
week after KA treatment (Lowenstein et al., 1993). content is decreased (Gall et al., 1991; Gall and
Thus, the time-course data favor a role for BDNF Lauterborn, 1992; Bengzon et al., 1993; Schmidt-
rather than NGF in seizure-induced phospho-trk Kastner and Olson, 1995; Mudo et al., 1996). Sec-
immunoreactivity. ond, protein levels of NGF and BDNF increase
after seizure activity (Bengzon et al., 1992; Elmer
et al., 1998). Third, the time-course data described
above implicate BDNF rather than NGF. Fourth,
Seizure duration threshold for increased phospho- mRNA levels of the other NT known to activate
trk immunoreactivity trkB, NT-4, are very low in adult brain (Timmusk
et al., 1993a) and do not increase after seizures
The seizure duration threshold for increase in (Mudo et al., 1996). Fifth, unlike trkB and trkC,
phospho-trk immunoreactivity further supports a levels of expression of trkA in hippocampus are
role for BDNF. Consistently, increased phospho- barely detectable (Barbacid, 1994; Cellerino,
trk immunoreactivity was observed only in hippo- 1996), suggesting that trkA is unlikely to mediate
campal kindled animals with ESD70 s (Binder seizure-induced increases in phospho-trk immuno-
et al., unpublished data). In a similar ventral hip- reactivity.
pocampal stimulation protocol, Bengzon et al. ob- In order to more directly analyze the role of the
served increases in BDNF mRNA content in trkB receptor in seizure-induced trk receptor acti-
dentate granule cells in an all-or-none manner vation, He et al. studied trk receptor phosphor-
above an electrographic seizure duration of ap- ylation in a mouse mutant with a single point
proximately 70 s (Bengzon et al., 1993). Like the mutation at the shc site (Y490 in humans, Y515 in
increases in mRNA content, increases in phospho- mice) of the trkB receptor (He et al., 2002). Ho-
trk immunoreactivity appeared to be ‘‘all- mozygous trkBshc/shc (Y515F) mice were generated
or-none’’ as no differences were noted in intensity by Minichiello et al. and interestingly display loss
of immunoreactivity between kainate-treated and of NT-4-dependent neurons but have no major
7 hippocampal ES-treated animals despite marked effects on BDNF responses (Minichiello et al.,
differences in seizure duration (hours for kainate 1999). He et al. found that following amygdala
vs. seconds for 7 hippocampal ESs) (Binder et al., kindling stimulation, phospho-trk immunoreactiv-
unpublished data). This strong similarity between ity is increased in wild-type mice in a similar pat-
thresholds as well as all-or-none characteristics tern (hilus and CA3 stratum lucidum) to that seen
suggests that such prior increases in BDNF in the rat experiments (described above). The
383

trkBshc/shc homozygous mice displayed absence of after seizure (dentate hilus and CA3 stratum lucidum
seizure-induced phospho-trk immunoreactivity, of hippocampus) corresponds to the mossy-fiber
and the heterozygotes displayed intermediate pathway of dentate granule cell axon terminals
immunoreactivity (He et al., 2002). These experi- (Binder et al., 1999a). This suggests that the cellular
ments suggest that the trk receptor activated site of phospho-trk immunoreactivity is either on
during kindling stimulation is indeed trkB. mossy-fiber axons and/or targets. Localization on
Interestingly, the Y515F point mutation had mossy-fiber axons represents a parsimonious expla-
no effect on kindling development in the same nation for both hilar and CA3 stratum lucidum
study (He et al., 2002). This is remarkably con- immunoreactivity. In contrast, localization on targets
sistent with the lack of effect of this mutation on requires immunoreactivity on both targets in hilus
synaptic LTP (Korte et al., 2000). More recently, (hilar interneurons) and in CA3 stratum lucidum
this group has generated a distinct mouse with a (pyramidal cell dendrites and/or stratum lucidum in-
point mutation at the PLC site. Unlike trkBshc/shc terneurons).
mice, trkBPLC/PLC mice exhibit impaired LTP Anatomic consideration of trkB-like immuno-
(Minichiello et al., 2002). This direct comparison reactivity may lend insight into the likely cellular
of distinct trkB tyrosine mutants implicates the site of phospho-trk immunoreactivity. In some
PLC signaling pathway as opposed to the MAPK published experiments, an affinity-purified anti-
pathway in trkB activation-induced synaptic body directed against an extracellular trkB peptide
plasticity. sequence was used, which does not distinguish be-
Similarly, other studies have shown that spe- tween full-length and truncated (Barbacid, 1994)
cific stimuli may cause tyrosine-specific phos- trkB receptors. The earlier studies (using light mi-
phorylation of the trkB receptor (i.e. at other croscopy) demonstrated that trkB-like immunore-
tyrosines but not at the shc site). For example, activity is preferentially distributed on cell bodies
Saarelainen et al., in studying the role of en- and dendrites of both cortical and hippocampal
dogenous BDNF and trkB signaling in the neurons (Fryer et al., 1996; Yan et al., 1997a).
mechanism of action of antidepressant drugs, Pyramidal neurons in hippocampus in particular
found that acute and chronic antidepressant demonstrate marked trkB immunoreactivity on
treatment caused trkB receptor phosphorylat- cell bodies and dendrites in comparison with axons
ion and activation, but the pY674/5 site was (Fryer et al., 1996; Yan et al., 1997a). These stud-
selectively phosphorylated compared to the ies utilized an antibody raised against the extra-
pY490 (shc) site (Saarelainen et al., 2003). The cellular portion of trkB (trkB2336) common to
further development of phosphorylation state- both full-length and truncated forms. A more
specific antibodies to distinct tyrosines (pY674/ recent and comprehensive study of cellular and
5, pY785) may prove to be of use in dissecting subcellular localization of trkB immunoreactivity
tyrosine site-specific trkB signal transduction in was carried out by Drake et al. (1999). These
vivo in a variety of paradigms. Furthermore, investigators used acytoplasmic-domain antibody
these results can be compared with antibodies (trkB-in) to selectively label the full-length form of
that recognize activated intracellular signaling trkB and carried out both light and electron
pathways (e.g. phosphoCREB) (Finkbeiner et microscopic analysis. Their conclusion was that
al., 1997). full-length trkB immunoreactivity exists in glut-
amatergic granule and pyramidal cells and was
most intense in axons, axon terminals, and den-
Cellular site of seizure-induced phospho-trk dritic spines and to a lesser extent in somata and
immunoreactivity dendritic shafts. Occasionally, interneurons were
also labeled. Thus, phospho-trkB immunoreactiv-
What is the likely cellular site of seizure-induced ity could represent pre- and/or postsynaptic acti-
phospho-trk immunoreactivity? The light micro- vation of trkB receptors in the mossy-fiber
scopic distribution of phospho-trk immunoreactivity pathway.
384

Potential models for induction of phospho-trk dendritic BDNF mRNA targeting may underlie
immunoreactivity by seizure activity another potential cellular mechanism for BDNF
translation, release, and trk receptor activation
Throughout the brain, BDNF immunoreactivity (Simonato et al., 2002). Determining the ultrastruc-
appears to be preferentially localized in cell bodies tural distribution of phospho-trk immunoreactivity
and axons compared to dendrites (Conner et al., would be necessary to distinguish these possibilities.
1997). In addition, unlike the classical target-de- Since the other primary target of mossy-fiber
rived trophic factor model in which NTs are ret- axons in CA3 is dendrites of stratum lucidum
rogradely transported, abundant recent evidence interneurons (Spruston et al., 1997), it is possible
suggests that CNS BDNF appears to be ante- that phospho-trk immunoreactivity in stratum
rogradely transported (Von Bartheld et al., 1996a; lucidum could reflect activation of trk receptors on
Zhou and Rush, 1996; Altar et al., 1997; Conner et interneurons as well as CA3 pyramidal cell dend-
al., 1997; Fawcett et al., 1998; Tonra et al., 1998). rites. Indeed, quantitative analysis of mossy-fiber
This evidence, together with the anatomic distri- targets in CA3 suggests that the number of synaptic
bution of BDNF immunoreactivity in hippocam- contacts onto GABAergic interneurons vastly out-
pus in a mossy fiber-like pattern, suggests that numbers those onto CA3 dendrites (Acsady et al.,
BDNF protein in hilus and CA3 stratum lucidum 1998). Indeed, any interneuron with dendrites tra-
was synthesized in granule cell bodies and ante- versing stratum lucidum could be a target of mossy-
rogradely transported to mossy-fiber terminals. fiber axons. However, it is unclear whether func-
Furthermore, following seizures there may be tional trkB receptors exist on stratum lucidum in-
increased anterograde transport of BDNF. First, terneurons, as in situ hybridization studies show
using hippocampal microdissections of dentate trkB mRNA localization predominantly in granule
gyrus (which contained hilus) and CA3 (which and pyramidal cells of hippocampus (Bengzon et al.,
contained stratum lucidum), Elmer et al. showed 1993) and only occasional interneurons were found
that maximal BDNF protein levels after seizures to be trkB-immunoreactive in the EM study (Drake
were at 12 h in dentate gyrus but 24 h in CA3 et al., 1999).
(Elmer et al., 1998). This suggests anterograde Furthermore, recent evidence indicates that
transport of seizure-induced BDNF protein. More activated trk receptors may be endocytosed and
recent evidence regarding the time course of retrogradely transported while still tyrosine phos-
BDNF immunoreactivity following seizures dem- phorylated (Grimes et al., 1996; Von Bartheld et al.,
onstrates that there is increased BDNF immuno- 1996b; Bhattacharyya et al., 1997; Riccio et al.,
reactivity in dentate granule cells at 4 h followed 1997; Senger and Campenot, 1997). Therefore,
by subsequent increases in hilus and finally in- mossy fiber-like phospho-trk immunoreactivity
creases in CA3 stratum lucidum at about 24 h could in part reflect not only distal synaptic sites
(Vezzani et al., 1999) (C. Gall, personal commu- of trk activation but also in-progress retrograde
nication). Furthermore, this anterograde ‘‘move- transport of activated trk from CA3 within the
ment’’ of BDNF immunoreactivity was abrogated mossy fibers. Thus, the increase in phospho-trk
by the axonal transport inhibitor colchicine (C. immunoreactivity observed in the dentate hilus may
Gall, personal communication). represent activated trk from mossy-fiber terminals in
These considerations lead to a model in which hilus or CA3.
CA3 stratum lucidum phospho-trk immunoreactiv-
ity is a consequence of seizure-induced BDNF
release from mossy-fiber axons activating trkB Role of BDNF in other pathologic conditions
receptors on dendrites of CA3 pyramidal cells and
hilar interneurons. Supporting a postsynaptic site Pain
for trk receptor activation is the evidence that full-
length trkB receptors are localized to the postsy- BDNF also may play an important neuromodu-
naptic density (Wu et al., 1996). Alternatively, latory role in pain transduction (Malcangio and
385

Lessmann, 2003). BDNF is synthesized by dorsal demonstrated that huntingtin normally inhibits the
horn neurons and markedly upregulated in neuron restrictive silencer element (NRSE) involved
inflammatory injury to peripheral nerves (along in tonic repression of transcription from BDNF
with NGF) (Fukuoka et al., 2001). BDNF acutely promoter II (Zuccato et al., 2003). In all of these
sensitizes nociceptive afferents and elicits hype- disorders, provision of BDNF or increasing endog-
ralgesia which is abrogated by BDNF inhibitors enous BDNF production may conceivably be ther-
(Kerr et al., 1999; Thompson et al., 1999; Pezet apeutic if applied in the appropriate spatiotemporal
et al., 2002). Central pain sensitization is an activ- context (Spires et al., 2004).
ity-dependent increase in excitability of dorsal
horn neurons leading to a clinically intractable Rett syndrome
condition termed ‘‘neuropathic pain’’ in which
normally nonpainful somatosensory stimuli (touch Rett syndrome is an X-linked postnatal neurode-
and pressure) become exquisitely painful (all- velopmental disorder that strikes approximately 1
odynia). Electrophysiological and behavioral data in 10,000 girls. It is characterized by regression of
demonstrate that inhibition of BDNF signal trans- normal development after about the age of 1 year
duction inhibits central pain sensitization (Kerr and eventually leads to several mental and physical
et al., 1999; Pezet et al., 2002). impairment, including cognitive and movement defi-
cits and breathing abnormalities. In 1999, Rett syn-
drome was linked to mutations in the MECP2 gene
Neurodegenerative diseases on the X chromosome (Amir et al., 1999). MeCP2,
the protein product of the MECP2 gene, is a me-
The idea that degenerative diseases of the nervous thyl-CpG binding protein, known to bind DNA
system may result from insufficient supply of regulatory regions to silence gene expression. In
neurotrophic factors has generated great interest in 2003, it was discovered that one of the genes nor-
BDNF as a potential therapeutic agent. Many mally turned off by MeCP2 is BDNF (Chen et al.,
reports have documented evidence of decreased 2003; Martinowich et al., 2003). Recently, a strain of
expression of BDNF in neurological disease (Murer mice missing the mouse version of MECP2 (Mecp2)
et al., 2001). Selective reduction of BDNF mRNA in has been found to have abnormally low levels of
the hippocampus has been reported in Alzheimer’s BDNF (Chang et al., 2006). These mice exhibit sev-
disease specimens (Phillips et al., 1991; Ferrer et al., eral features of human Rett syndrome. Increasing
1999), although selective upregulation appears to BDNF production in mice lacking Mecp2 restored
occur in plaque-related glial cells in an animal model mobility and extended life span (Chang et al., 2006).
(Burbach et al., 2004). Decreased BDNF protein has Neural activity triggers phosphorylation of MeCP2
been demonstrated in the substantia nigra in Par- that detaches it from the regulatory region of the
kinson’s disease (Howells et al., 2000). BDNF pro- BDNF gene and allows BDNF transcription (Zhou
motes survival of all major neuronal types affected et al., 2006). Further study of MeCP2–BDNF in-
in Alzheimer’s and Parkinson’s disease, such as hip- teractions may lead to novel insights and treatment
pocampal and neocortical neurons, cholinergic se- strategies for Rett syndrome. Interestingly, MECP2
ptal and basal forebrain neurons, and nigral abnormalities are starting to be found in other
dopaminergic neurons. neurodevelopmental disorders such as autism, sug-
Interestingly, recent work has implicated BDNF gesting that BDNF dysregulation may also have
in Huntington’s disease as well. Huntingtin, the a more widespread role in the pathophysiology of
protein mutated in Huntington’s disease, upregulates these conditions.
BDNF transcription, and loss of huntingtin-
mediated BDNF transcription leads to loss of tro- Neuropsychiatric disease
phic support to striatal neurons which subsequently
degenerate in the hallmark pathology of the dis- BDNF signaling may also be involved in affective
order (Zuccato et al., 2001). A recent study has behaviors (Altar, 1999). Environmental stresses
386

such as immobilization that induce depression also rearrangements of axons and dendrites. Basic ac-
decrease BDNF mRNA (Smith et al., 1995). Con- tivity-related changes in the CNS are thought to
versely, physical exercise is associated with depend on BDNF modulation of synaptic trans-
decreased depression and increased BDNF mRNA mission. Pathologic levels of BDNF-dependent
(Russo-Neustadt et al., 1999; Cotman and Berch- synaptic plasticity may contribute to conditions
told, 2002). Existing treatments for depression are such as epilepsy and chronic pain sensitization,
thought to act primarily by increasing endogenous whereas application of the trophic properties of
monoaminergic (i.e. serotonergic and nor- BDNF may lead to novel therapeutic options in
adrenergic) synaptic transmission, and recent stud- neurodegenerative diseases and perhaps even in
ies have shown that effective antidepressants neuropsychiatric disorders.
increase BDNF mRNA (Dias et al., 2003) and The role of BDNF in epilepsy provides a partic-
protein (Chen et al., 2001; Altar et al., 2003). Ex- ularly good example of the pleiotropic effects of
ogenous delivery of BDNF promotes the function BDNF on excitability. The hippocampus and
and sprouting of serotonergic neurons in adult rat closely associated limbic structures are thought to
brains (Mamounas et al., 1995), and BDNF-defi- be particularly important in the pro-epileptogenic
cient mice are also deficient in serotonergic inner- effects of BDNF (Binder et al., 1999a), and in-
vation (Lyons et al., 1999). Acute local BDNF creased BDNF expression in the hippocampus is
infusion has antidepressant-like effects in rats found in specimens from patients with temporal
(Shirayama et al., 2002). Thus, new pharmacolog- lobe epilepsy (Mathern et al., 1997; Takahashi et al.,
ic strategies are focused on the potential antide- 1999). It is hoped that understanding of the hyper-
pressant role of BDNF. excitability associated with BDNF in epilepsy ani-
It has also been hypothesized that BDNF may mal models may lead to novel anticonvulsant or
be involved in bipolar disorder (Tsai, 2004). In- antiepileptic therapies (Binder et al., 2001).
terestingly, lithium, a major drug for the treatment Of course, simple up- or downregulation of NTs
of bipolar disorder, increases BDNF and trkB may lead to many nonspecific effects. For ultimate
activation in cerebral cortical neurons (Hashimoto clinical application in specific conditions, it will be
et al., 2002). BDNF is an attractive candidate gene very helpful to elucidate the mechanisms of action
for susceptibility to bipolar disorder, and some of each of the effects of NT receptor activation.
(Neves-Pereira et al., 2002; Sklar et al., 2002) but Therefore, much recent research has focused on
not other (Hong et al., 2003; Nakata et al., 2003) downstream targets of the NT signaling pathways
studies suggest linkage between BDNF poly- responsible for specific phenotypic effects. For
morphisms and disease susceptibility (Green and example, BDNF activation of trkB down-regulates
Craddock, 2003). How alterations in BDNF ac- hippocampal KCC2, a K+-Cl cotransporter
tivity may relate to fluctuating bouts of mania and (Rivera et al., 2002); this suppresses chloride-
depression in bipolar disorder is still a matter of dependent fast GABAergic inhibition and may
speculation. partially account for BDNF modulation of GAB-
Aergic synapses (Wardle and Poo, 2003). In addi-
tion, BDNF phosphorylates specific subunits of
Perspective both, the NMDA receptor and the GABAA
receptor, altering their function (Suen et al.,
Since the discovery of NGF in the 1950s and 1997; Lin et al., 1998; Jovanovic et al., 2004).
BDNF in the 1980s, a great deal of evidence has Long-term effects of BDNF must take into ac-
mounted for the roles of NGF, BDNF, NT-3, and count the fact that it upregulates many other plas-
NT-4/5 in development, physiology, and pathol- ticity-related genes, such as NPY (Croll et al.,
ogy. BDNF in particular has important roles in 1994; Nawa et al., 1994). NPY, for example, may
neural development and cell survival, as well as not only modulate excitability (Baraban et al.,
appearing essential to molecular mechanisms of 1997; Reibel et al., 2000a) but also other phenom-
synaptic plasticity and larger scale structural ena such as neurogenesis (Howell et al., 2003).
387

New methods of modulation/upregulation of Altar, C.A. (1999) Neurotrophins and depression. Trends
NTs may be required to achieve translational con- Pharmacol. Sci., 20: 59–61.
Altar, C.A., Cai, N., Bliven, T., Juhasz, M., Conner, J.M.,
trol of diseases of NT deficiency. ‘‘Ampakines’’
Acheson, A.L., Lindsay, R.M. and Wiegand, S.J. (1997) An-
represent a new class of compounds that have been terograde transport of brain-derived neurotrophic factor and
shown to upregulate BDNF over a long period of its role in the brain. Nature, 389: 856–860.
time (Lauterborn et al., 2003). Of course, proper Altar, C.A. and DiStefano, P.S. (1998) Neurotrophin traffick-
dosing so as not to trigger downregulation of im- ing by anterograde transport. Trends Neurosci., 21: 433–437.
portant NT signaling pathways will be critical to Altar, C.A., Siuciak, J.A., Wright, P., Ip, N.Y., Lindsay, R.M.
and Wiegand, S.J. (1994) In situ hybridization of trkB and
avoiding deleterious side effects of these potential trkC receptor mRNA in rat forebrain and association with
new therapies. Nevertheless, these and similar high-affinity binding of [125I]BDNF, [125I]NT-4/5 and
compounds are under active clinical investigation [125I]NT-3. Eur. J. Neurosci., 6: 1389–1405.
as cognitive and memory enhancing drugs Altar, C.A., Whitehead, R.E., Chen, R., Wortwein, G. and
Madsen, T.M. (2003) Effects of electroconvulsive seizures
(Danysz, 1999; Johnson and Simmon, 2002).
and antidepressant drugs on brain-derived neurotrophic fac-
tor protein in rat brain. Biol. Psychiatry, 54: 703–709.
Acknowledgment Amir, R.E., Van den Veyver, I.B., Wan, M., Tran, C.Q.,
Francke, U. and Zoghbi, H.Y. (1999) Rett syndrome is
caused by mutations in X-linked MECP2, encoding methyl-
I would like to thank Helen Scharfman for her CpG-binding protein 2. Nat. Genet., 23: 185–188.
detailed analysis and constructive criticism. Anderson, K.D., Alderson, R.F., Altar, C.A., DiStefano, P.S.,
Corcoran, T.L., Lindsay, R.M. and Wiegand, S.J. (1995)
Differential distribution of exogenous BDNF, NGF, and
References NT-3 in the brain corresponds to the relative abundance and
distribution of high-affinity and low-affinity neurotrophin
Acheson, A., Conover, J.C., Fandl, J.P., DeChiara, T.M., Rus- receptors. J. Comp. Neurol., 357: 296–317.
sell, M., Thadani, A., Squinto, S.P., Yancopoulos, G.D. and Asztely, F., Kokaia, M., Olofsdotter, K., Ortegren, U. and
Lindsay, R.M. (1995) A BDNF autocrine loop in adult sen- Lindvall, O. (2000) Afferent-specific modulation of short-
sory neurons prevents cell death. Nature, 374: 450–453. term synaptic plasticity by neurotrophins in dentate gyrus.
Acsady, L., Kamondi, A., Sik, A., Freund, T. and Buzsaki, G. Eur. J. Neurosci., 12: 662–669.
(1998) GABAergic cells are the major postsynaptic targets of Baraban, S.C., Hollopeter, G., Erickson, J.C., Schwartzkroin,
mossy fibers in the rat hippocampus. J. Neurosci., 18: P.A. and Palmiter, R.D. (1997) Knock-out mice reveal a
3386–3403. critical antiepileptic role for neuropeptide Y. J. Neurosci., 17:
Adams, B., Sazgar, M., Osehobo, P., Van der Zee, C.E.E.M., 8927–8936.
Diamond, J., Fahnestock, M. and Racine, R.J. (1997) Nerve Barbacid, M. (1994) The trk family of neurotrophin receptors.
growth factor accelerates seizure development, enhances J. Neurobiol., 25: 1386–1403.
mossy fiber sprouting, and attenuates seizure-induced de- Barde, Y.A., Edgar, D. and Thoenen, H. (1982) Purification of
creases in neuronal density in the kindling model of epilepsy. a new neurotrophic factor from mammalian brain. EMBO J.,
J. Neurosci., 17: 5288–5296. 1: 549–553.
Adams, B., Vaccarella, L., Fahnestock, M. and Racine, R.J. Barnabe-Heider, F. and Miller, F.D. (2003) Endogenously pro-
(2002) The cholinergic system modulates kindling and kin- duced neurotrophins regulate survival and differentiation of
dling-induced mossy fiber sprouting. Synapse, 44: 132–138. cortical progenitors via distinct signaling pathways. J. Ne-
Alderson, R.F., Alterman, A.L., Barde, Y.A. and Lindsay, urosci., 23: 5149–5160.
R.M. (1990) Brain-derived neurotrophic factor increases sur- Bengzon, J., Kokaia, Z., Ernfors, P., Kokaia, M., Leanza, G.,
vival and differentiated functions of rat septal cholinergic Nilsson, O.G., Persson, H. and Lindvall, O. (1993) Regula-
neurons in culture. Neuron, 5: 297–306. tion of neurotrophin and trkA, trkB and trkC tyrosine kinase
Alderson, R.F., Curtis, R., Alterman, A.L., Lindsay, R.M. and receptor messenger RNA expression in kindling. Neurosci-
DiStefano, P.S. (2000) Truncated TrkB mediates the end- ence, 53: 433–446.
ocytosis and release of BDNF and neurotrophin-4/5 by rat Bengzon, J., Soderstrom, S., Kokaia, Z., Kokaia, M., Ernfors,
astrocytes and schwann cells in vitro. Brain Res., 871: P., Persson, H., Ebendal, T. and Lindvall, O. (1992) Wide-
210–222. spread increase of nerve growth factor protein in the rat
Alonso, M., Vianna, M.R., Depino, A.M., Mello e Souza, T., forebrain after kindling-induced seizures. Brain Res., 587:
Pereira, P., Szapiro, G., Viola, H., Pitossi, F., Izquierdo, I. 338–342.
and Medina, J.H. (2002) BDNF-triggered events in the rat Benraiss, A., Chmielnicki, E., Lerner, K., Roh, D. and Gold-
hippocampus are required for both short- and long-term man, S.A. (2001) Adenoviral brain-derived neurotrophic fac-
memory formation. Hippocampus, 12: 551–560. tor induces both neostriatal and olfactory neuronal
388

recruitment from endogenous progenitor cells in the adult Castrén, E., Pitkanen, M., Sirvio, J., Parsadanian, A., Lind-
forebrain. J. Neurosci., 21: 6718–6731. holm, D., Thoenen, H. and Riekkinen, P.J. (1993) The
Bhattacharyya, A., Watson, F.L., Bradlee, T.A., Pomeroy, induction of LTP increases BDNF and NGF mRNA but
S.L., Stiles, C.D. and Segal, R.A. (1997) Trk receptors func- decreases NT-3 mRNA in the dentate gyrus. Neuroreport,
tion as rapid retrograde signal carriers in the adult nervous 4: 895–898.
system. J. Neurosci., 17: 7007–7016. Castrén, E., Thoenen, H. and Lindholm, D. (1995) Brain-de-
Biffo, S., Offenhauser, N., Carter, B.D. and Barde, Y.A. (1995) rived neurotrophic factor messenger RNA is expressed in the
Selective binding and internalisation by truncated receptors septum, hypothalamus and in adrenergic brain stem nuclei of
restrict the availability of BDNF during development. adult rat brain and is increased by osmotic stimulation in the
Development, 121: 2461–2470. paraventricular nucleus. Neuroscience, 64: 71–80.
Binder, D.K., Croll, S.D., Gall, C.M. and Scharfman, H.E. Castrén, E., Zafra, F., Thoenen, H. and Lindholm, D. (1992)
(2001) BDNF and epilepsy: too much of a good thing? Light regulates expression of brain-derived neurotrophic fac-
Trends Neurosci., 24: 47–53. tor mRNA in rat visual cortex. Proc. Natl. Acad. Sci. U.S.A.,
Binder, D.K., Routbort, M.J. and McNamara, J.O. (1999a) 89: 9444–9448.
Immunohistochemical evidence of seizure-induced activation Cellerino, A. (1996) Expression of messenger RNA coding for
of trk receptors in the mossy fiber pathway of adult rat the nerve growth factor receptor trkA in the hippocampus of
hippocampus. J. Neurosci., 19: 4616–4626. the adult rat. Neuroscience, 70: 613–616.
Binder, D.K., Routbort, M.J., Ryan, T.E., Yancopoulos, G.D. Chang, Q., Khare, G., Dani, V., Nelson, S. and Jaenisch, R.
and McNamara, J.O. (1999b) Selective inhibition of kindling (2006) The disease progression of Mecp2 mutant mice is
development by intraventricular administration of TrkB affected by the level of BDNF expression. Neuron, 49:
receptor body. J. Neurosci., 19: 1424–1436. 341–348.
Black, I.B. (1999) Trophic regulation of synaptic plasticity. J. Chao, M.V. and Bothwell, M. (2002) Neurotrophins: to cleave
Neurobiol., 41: 108–118. or not to cleave. Neuron, 33: 9–12.
Bramham, C.R., Southard, T., Sarvey, J.M., Herkenham, M. Chao, M.V. and Hempstead, B.L. (1995) p75 and trk: a two-
and Brady, L.S. (1996) Unilateral LTP triggers bilateral in- receptor system. Trends Neurosci., 18: 321–326.
creases in hippocampal neurotrophin and trk receptor Chen, B., Dowlatshahi, D., MacQueen, G.M., Wang, J.F. and
mRNA expression in behaving rats: evidence for interhem- Young, L.T. (2001) Increased hippocampal BDNF immuno-
ispheric communication. J. Comp. Neurol., 368: 371–382. reactivity in subjects treated with antidepressant medication.
Bregola, G., Frigati, L., Zucchini, S. and Simonato, M. (2000) Biol. Psychiatry, 50: 260–265.
Different patterns of induction of fibroblast growth factor-2 Chen, W.G., Chang, Q., Lin, Y., Meissner, A., West, A.E.,
and brain-derived neurotrophic factor messenger RNAs dur- Griffith, E.C., Jaenisch, R. and Greenberg, M.E. (2003) De-
ing kindling epileptogenesis, and development of a herpes repression of BDNF transcription involves calcium-depend-
simplex vector for fibroblast growth factor-2 gene transfer in ent phosphorylation of MeCP2. Science, 302: 885–889.
vivo. Epilepsia, 41(Suppl 6): S122–S126. Cheng, B., Goodman, Y., Begley, J.G. and Mattson, M.P.
Brigadski, T., Hartmann, M. and Lessmann, V. (2005) Differ- (1994) Neurotrophin-4/5 protects hippocampal and cortical
ential vesicular targeting and time course of synaptic secre- neurons against energy deprivation- and excitatory amino
tion of the mammalian neurotrophins. J. Neurosci., 25: acid-induced injury. Brain Res., 650: 331–335.
7601–7614. Chmielnicki, E., Benraiss, A., Economides, A.N. and Goldman,
Burbach, G.J., Hellweg, R., Haas, C.A., Del Turco, D., Deicke, S.A. (2004) Adenovirally expressed noggin and brain-derived
U., Abramowski, D., Jucker, M., Staufenbiel, M. and Deller, neurotrophic factor cooperate to induce new medium spiny
T. (2004) Induction of brain-derived neurotrophic factor in neurons from resident progenitor cells in the adult striatal
plaque-associated glial cells of aged APP23 transgenic mice. ventricular zone. J. Neurosci., 24: 2133–2142.
J. Neurosci., 24: 2421–2430. Climent, E., Sancho-Tello, M., Minana, R., Barettino, D. and
Cabelli, R.J., Hohn, A. and Shatz, C.J. (1995) Inhibition of Guerri, C. (2000) Astrocytes in culture express the full-length
ocular dominance column formation by infusion of NT-4/5 Trk-B receptor and respond to brain derived neurotrophic
or BDNF. Science, 267: 1662–1666. factor by changing intracellular calcium levels: effect of
Cabelli, R.J., Shelton, D.L., Segal, R.A. and Shatz, C.J. (1997) ethanol exposure in rats. Neurosci. Lett., 288: 53–56.
Blockade of endogenous ligands of trkB inhibits formation of Conner, J.M., Lauterborn, J.C., Yan, Q., Gall, C.M. and
ocular dominance columns. Neuron, 19: 63–76. Varon, S. (1997) Distribution of brain-derived neurotrophic
Casaccia-Bonnefil, P., Carter, B.D., Dobrowsky, R.T. and factor (BDNF) protein and mRNA in the normal adult rat
Chao, M.V. (1996) Death of oligodendrocytes mediated by CNS–evidence for anterograde axonal transport. J. Neuro-
the interaction of nerve growth factor with its receptor p75. sci., 17: 2295–2313.
Nature, 383: 716–719. Conover, J.C., Erickson, J.T., Katz, D.M., Bianchi, L.M., Po-
Castrén, E., Berninger, B., Leingartner, A. and Lindholm, D. ueymirou, W.T., McClain, J., Pan, L., Helgren, M., Ip, N.Y.,
(1998) Regulation of brain-derived neurotrophic factor Boland, P., et al. (1995) Neuronal deficits, not involving mo-
mRNA levels in hippocampus by neuronal activity. Prog. tor neurons, in mice lacking BDNF and/or NT4. Nature,
Brain Res., 117: 57–64. 375: 235–238.
389

Cotman, C.W. and Berchtold, N.C. (2002) Exercise: a behavi- Elmer, E., Kokaia, Z., Kokaia, M., Carnahan, J., Nawa, H.,
oral intervention to enhance brain health and plasticity. Bengzon, J. and Lindvall, O. (1996b) Widespread increase of
Trends Neurosci., 25: 295–301. brain-derived neurotrophic factor protein in the rat forebrain
Croll, S.D., Suri, C., Compton, D.L., Simmons, M.V., after kindling-induced seizures. Soc. Neurosci. Abstr., 22: 2089.
Yancopoulos, G.D., Lindsay, R.M., Wiegand, S.J., Rudge, Elmer, E., Kokaia, Z., Kokaia, M., Carnahan, J., Nawa, H.
J.S. and Scharfman, H.E. (1999) Brain-derived neurotrophic and Lindvall, O. (1998) Dynamic changes of brain-derived
factor transgenic mice exhibit passive avoidance deficits, in- neurotrophic factor protein levels in the rat forebrain after
creased seizure severity and in vitro hyperexcitability in the single and recurring kindling-induced seizures. Neuroscience,
hippocampus and entorhinal cortex. Neuroscience, 93: 83: 351–362.
1491–1506. Ernfors, P., Bengzon, J., Kokaia, Z., Persson, H. and Lindvall,
Croll, S.D., Wiegand, S.J., Anderson, K.D., Lindsay, R.M. and O. (1991) Increased levels of messenger RNAs for neurotro-
Nawa, H. (1994) Regulation of neuropeptides in adult rat phic factors in the brain during kindling epileptogenesis.
forebrain by the neurotrophins BDNF and NGF. Eur. J. Neuron, 7: 165–176.
Neurosci., 6: 1343–1353. Ernfors, P., Wetmore, C., Olson, L. and Persson, H. (1990)
Danysz, W. (1999) CX-516 (Cortex Pharmaceuticals Inc). Identification of cells in rat brain and peripheral tissues ex-
IDrugs, 2: 814–822. pressing mRNA for members of the nerve growth factor
Dechant, G. and Barde, Y.A. (2002) The neurotrophin receptor family. Neuron, 5: 511–526.
p75(NTR): novel functions and implications for diseases of Esposito, D., Patel, P., Stephens, R.M., Perez, P., Chao, M.V.,
the nervous system. Nat. Neurosci., 5: 1131–1136. Kaplan, D.R. and Hempstead, B.L. (2001) The cytoplasmic
Dias, B.G., Banerjee, S.B., Duman, R.S. and Vaidya, V.A. and transmembrane domains of the p75 and Trk A receptors
(2003) Differential regulation of brain derived neurotrophic regulate high affinity binding to nerve growth factor. J. Biol.
factor transcripts by antidepressant treatments in the adult Chem., 276: 32687–32695.
rat brain. Neuropharmacology, 45: 553–563. Farhadi, H.F., Mowla, S.J., Petrecca, K., Morris, S.J., Seidah,
Drake, C.T., Milner, T.A. and Patterson, S.L. (1999) Ultra- N.G. and Murphy, R.A. (2000) Neurotrophin-3 sorts to the
structural localization of full-length trkB immunoreactivity constitutive secretory pathway of hippocampal neurons and
in rat hippocampus suggests multiple roles in modulating is diverted to the regulated secretory pathway by coexpres-
activity-dependent synaptic plasticity. J. Neurosci., 19: sion with brain-derived neurotrophic factor. J. Neurosci., 20:
8009–8026. 4059–4068.
Dugich-Djordjevic, M.M., Tocco, G., Lapchak, P.A., Pasinetti, Fawcett, J.P., Alonso-Vanegas, M.A., Morris, S.J., Miller,
G.M., Najm, I., Baudry, M. and Hefti, F. (1992a) Regionally F.D., Sadikot, A.F. and Murphy, R.A. (2000) Evidence that
specific and rapid increases in brain-derived neurotrophic brain-derived neurotrophic factor from presynaptic nerve
factor messenger RNA in the adult rat brain following sei- terminals regulates the phenotype of calbindin-containing
zures induced by systemic administration of kainic acid. Ne- neurons in the lateral septum. J. Neurosci., 20: 274–282.
uroscience, 47: 303–315. Fawcett, J.P., Aloyz, R., McLean, J.H., Pareek, S., Miller,
Dugich-Djordjevic, M.M., Tocco, G., Willoughby, D.A., F.D., McPherson, P.S. and Murphy, R.A. (1997) Detection
Najm, I., Pasinetti, G., Thompson, R.F., Baudry, M., Lap- of brain-derived neurotrophic factor in a vesicular fraction of
chak, P.A. and Hefti, F. (1992b) BDNF mRNA expression in brain synaptosomes. J. Biol. Chem., 272: 8837–8840.
the developing rat brain following kainic acid-induced seizure Fawcett, J.P., Bamji, S.X., Causing, C.G., Aloyz, R., Ase, A.R.,
activity. Neuron, 8: 1127–1138. Reader, T.A., McLean, J.H. and Miller, F.D. (1998) Func-
Egan, M.F., Kojima, M., Callicott, J.H., Goldberg, T.E., Ko- tional evidence that BDNF is an anterograde neuronal tro-
lachana, B.S., Bertolino, A., Zaitsev, E., Gold, B., Goldman, phic factor in the CNS. J. Neurosci., 18: 2808–2821.
D., Dean, M., Lu, B. and Weinberger, D.R. (2003) The Ferrer, I., Marin, C., Rey, M.J., Ribalta, T., Goutan, E.,
BDNF val66met polymorphism affects activity-dependent Blanco, R., Tolosa, E. and Marti, E. (1999) BDNF and full-
secretion of BDNF and human memory and hippocampal length and truncated TrkB expression in Alzheimer disease.
function. Cell, 112: 257–269. Implications in therapeutic strategies. J. Neuropathol. Exp.
Eide, F.F., Vining, E.R., Eide, B.L., Zang, K., Wang, X.Y. and Neurol., 58: 729–739.
Reichardt, L.F. (1996) Naturally occurring truncated trkB Figurov, A., Pozzo-Miller, L.D., Olafsson, P., Wang, T. and
receptors have dominant inhibitory effects on brain-derived Lu, B. (1996) Regulation of synaptic responses to high-fre-
neurotrophic factor signaling. J. Neurosci., 16: 3123–3129. quency stimulation and LTP by neurotrophins in the hippo-
Elmer, E., Kokaia, M., Ernfors, P., Ferencz, I., Kokaia, Z. and campus. Nature, 381: 706–709.
Lindvall, O. (1997) Suppressed kindling epileptogenesis and Finkbeiner, S., Tavazoie, S.F., Maloratsky, A., Jacobs, K.M.,
perturbed BDNF and trkB gene regulation in NT-3 mutant Harris, K.M. and Greenberg, M.E. (1997) CREB: a major
mice. Exp. Neurol., 145: 93–103. mediator of neuronal neurotrophin responses. Neuron, 19:
Elmer, E., Kokaia, M., Kokaia, Z., Ferencz, I. and Lindvall, O. 1031–1047.
(1996a) Delayed kindling development after rapidly recurring Frade, J.M., Rodriguez-Tebar, A. and Barde, Y.A. (1996) In-
seizures: relation to mossy fiber sprouting and neurotrophin, duction of cell death by endogenous nerve growth factor
GAP-43 and dynorphin gene expression. Brain Res., 712: 19–34. through its p75 receptor. Nature, 383: 166–168.
390

Frank, L., Ventimiglia, R., Anderson, K., Lindsay, R.M. and NGF and BDNF? Implications for the modulatory role of
Rudge, J.S. (1996) BDNF downregulates neurotrophin re- neurotrophins in activity-dependent neuronal plasticity. Mi-
sponsiveness, trkB protein and trkB mRNA levels in cultured crosc. Res. Tech., 45: 262–275.
rat hippocampal neurons. Eur. J. Neurosci., 8: 1220–1230. Grimes, M.L., Zhou, J., Beattie, E.C., Yuen, E.C., Hall, D.E.,
Frerking, M., Malenka, R.C. and Nicoll, R.A. (1998) Brain- Valletta, J.S., Topp, K.S., LaVail, J.H., Bunnett, N.W. and
derived neurotrophic factor (BDNF) modulates inhibitory, Mobley, W.C. (1996) Endocytosis of activated trkA: evidence
but not excitatory, transmission in the CA1 region of the that nerve growth factor induces formation of singaling
hippocampus. J Neurophysiol., 80: 3383–3386. endosomes. J. Neurosci., 16: 7950–7964.
Frisen, J., Verge, V.M., Fried, K., Risling, M., Persson, H., Gustafsson, E., Lindvall, O. and Kokaia, Z. (2003) Intraven-
Trotter, J., Hokfelt, T. and Lindholm, D. (1993) Character- tricular infusion of TrkB-Fc fusion protein promotes is-
ization of glial trkB receptors: differential response to injury chemia-induced neurogenesis in adult rat dentate gyrus.
in the central and peripheral nervous systems. Proc. Natl. Stroke, 34: 2710–2715.
Acad. Sci. U.S.A., 90: 4971–4975. Haapasalo, A., Koponen, E., Hoppe, E., Wong, G. and Cast-
Fryer, R.H., Kaplan, D.R., Feinstein, S.C., Radeke, M.J., ren, E. (2001) Truncated trkB.T1 is dominant negative in-
Grayson, D.R. and Kromer, L.F. (1996) Developmental and hibitor of trkB.TK+-mediated cell survival. Biochem.
mature expression of full-length and truncated trkB receptors Biophys. Res. Commun., 280: 1352–1358.
in the rat forebrain. J. Comp. Neurol., 374: 21–40. Haapasalo, A., Sipola, I., Larsson, K., Akerman, K.E., Stoilov,
Fryer, R.H., Kaplan, D.R. and Kromer, L.F. (1997) Truncated P., Stamm, S., Wong, G. and Castren, E. (2002) Regulation
trkB receptors on nonneuronal cells inhibit BDNF-induced of TRKB surface expression by brain-derived neurotrophic
neurite outgrowth in vitro. Exp. Neurol., 148: 616–627. factor and truncated TRKB isoforms. J. Biol. Chem., 277:
Fukuoka, T., Kondo, E., Dai, Y., Hashimoto, N. and Noguchi, 43160–43167.
K. (2001) Brain-derived neurotrophic factor increases in the Hall, J., Thomas, K.L. and Everitt, B.J. (2000) Rapid and se-
uninjured dorsal root ganglion neurons in selective spinal lective induction of BDNF expression in the hippocampus
nerve ligation model. J. Neurosci., 21: 4891–4900. during contextual learning. Nat. Neurosci., 3: 533–535.
Funabashi, T., Sasaki, H. and Kimura, F. (1988) Intraven- Hallbook, F., Ibanez, C.F. and Persson, H. (1991) Evolutionary
tricular injection of antiserum to nerve growth factor delays studies of the nerve growth factor family reveal a novel
the development of amygdaloid kindling. Brain Res., 458: member abundantly expressed in Xenopus ovary. Neuron, 6:
132–136. 845–858.
Gall, C., Lauterborn, J., Bundman, M., Murray, K. and Isack- Hartmann, M., Heumann, R. and Lessmann, V. (2001) Synap-
son, P. (1991) Seizures and the regulation of neurotrophic tic secretion of BDNF after high-frequency stimulation of
factor and neuropeptide gene expression in brain. Epilepsy glutamatergic synapses. EMBO J., 20: 5887–5897.
Res. Suppl., 4: 225–245. Hashimoto, R., Takei, N., Shimazu, K., Christ, L., Lu, B. and
Gall, C.M. and Isackson, P.J. (1989) Limbic seizures increase Chuang, D.M. (2002) Lithium induces brain-derived ne-
neuronal production of messenger RNA for nerve growth urotrophic factor and activates TrkB in rodent cortical neu-
factor. Science, 245: 758–761. rons: an essential step for neuroprotection against glutamate
Gall, C.M. and Lauterborn, J. (1992) In: Ribak C.E., Gall C.M. excitotoxicity. Neuropharmacology, 43: 1173–1179.
and Mody I. (Eds.), The Dentate Gyrus and its Role in Sei- He, X.P., Butler, L., Liu, X. and McNamara, J.O. (2006) The
zures. Elsevier, Amsterdam, pp. 171–185. tyrosine receptor kinase B ligand, neurotrophin-4, is not
Gall, C.M., Lauterborn, J.C., Guthrie, K.M. and Stinis, C.T. required for either epileptogenesis or tyrosine receptor kin-
(1997) Seizures and the regulation of neurotrophic factor ex- ase B activation in the kindling model. Neuroscience, 141:
pression: associations with structural plasticity in epilepsy. 515–520.
Adv. Neurol., 72: 9–24. He, X.P., Minichiello, L., Klein, R. and McNamara, J.O.
Ginty, D.D. and Segal, R.A. (2002) Retrograde neurotrophin (2002) Immunohistochemical evidence of seizure-induced ac-
signaling: Trk-ing along the axon. Curr. Opin. Neurobiol., tivation of trkB receptors in the mossy fiber pathway of adult
12: 268–274. mouse hippocampus. J. Neurosci., 22: 7502–7508.
Goggi, J., Pullar, I.A., Carney, S.L. and Bradford, H.F. (2003) Hofer, M., Pagliusi, S.R., Hohn, A., Leibrock, J. and Barde,
The control of [125I]BDNF release from striatal rat brain Y.A. (1990) Regional distribution of brain-derived neurotro-
slices. Brain Res., 967: 201–209. phic factor mRNA in the adult mouse brain. EMBO J., 9:
Gould, E. and McEwen, B.S. (1993) Neuronal birth and death. 2459–2464.
Curr. Opin. Neurobiol., 3: 676–682. Hong, C.J., Huo, S.J., Yen, F.C., Tung, C.L., Pan, G.M. and
Green, E. and Craddock, N. (2003) Brain-derived neurotrophic Tsai, S.J. (2003) Association study of a brain-derived ne-
factor as a potential risk locus for bipolar disorder: evidence, urotrophic-factor genetic polymorphism and mood disorders,
limitations, and implications. Curr. Psychiatry Rep., 5: age of onset and suicidal behavior. Neuropsychobiology, 48:
469–476. 186–189.
Griesbeck, O., Canossa, M., Campana, G., Gartner, A., Hoe- Horch, H.W. and Katz, L.C. (2002) BDNF release from single
ner, M.C., Nawa, H., Kolbeck, R. and Thoenen, H. (1999) cells elicits local dendritic growth in nearby neurons. Nat.
Are there differences between the secretion characteristics of Neurosci., 5: 1177–1184.
391

Howell, O.W., Scharfman, H.E., Herzog, H., Sundstrom, L.E., Jovanovic, J.N., Thomas, P., Kittler, J.T., Smart, T.G. and
Beck-Sickinger, A. and Gray, W.P. (2003) Neuropeptide Y is Moss, S.J. (2004) Brain-derived neurotrophic factor modu-
neuroproliferative for post-natal hippocampal precursor lates fast synaptic inhibition by regulating GABA(A) recep-
cells. J. Neurochem., 86: 646–659. tor phosphorylation, activity, and cell-surface stability. J.
Howells, D.W., Porritt, M.J., Wong, J.Y., Batchelor, P.E., Neurosci., 24: 522–530.
Kalnins, R., Hughes, A.J. and Donnan, G.A. (2000) Reduced Kafitz, K.W., Rose, C.R., Thoenen, H. and Konnerth, A.
BDNF mRNA expression in the Parkinson’s disease subst- (1999) Neurotrophin-evoked rapid excitation through TrkB
antia nigra. Exp. Neurol., 166: 127–135. receptors. Nature, 401: 918–921.
Huang, E.J. and Reichardt, L.F. (2001) Neurotrophins: roles in Kang, H. and Schuman, E.M. (1995) Long-lasting neurotro-
neuronal development and function. Annu. Rev. Neurosci., phin-induced enhancement of synaptic transmission in the
24: 677–736. adult hippocampus. Science, 267: 1658–1662.
Huang, E.J. and Reichardt, L.F. (2003) Trk receptors: roles in Kang, H., Welcher, A.A., Shelton, D. and Schuman, E.M.
neuronal signal transduction. Annu. Rev. Biochem., 72: (1997) Neurotrophins and time: different roles for trkB
609–642. signaling in hippocampal long-term potentiation. Neuron,
Hughes, P.E., Young, D., Preston, K.M., Yan, Q. and Dragu- 19: 653–664.
now, M. (1998) Differential regulation by MK801 of imme- Katoh-Semba, R., Asano, T., Ueda, H., Morishita, R., Take-
diate-early genes, brain-derived neurotrophic factor and trk uchi, I.K., Inaguma, Y. and Kato, K. (2002) Riluzole en-
receptor mRNA induced by a kindling after-discharge. Mol. hances expression of brain-derived neurotrophic factor with
Brain Res., 53: 138–151. consequent proliferation of granule precursor cells in the rat
Humpel, C., Wetmore, C. and Olson, L. (1993) Regulation of hippocampus. FASEB J., 16: 1328–1330.
brain-derived neurotrophic factor messenger RNA and pro- Katoh-Semba, R., Ichisaka, S., Hata, Y., Tsumoto, T., Eguchi,
tein at the cellular level in pentylenetetrazol-induced epileptic K., Miyazaki, N., Matsuda, M., Takeuchi, I.K. and Kato, K.
seizures. Neuroscience, 53: 909–918. (2003) NT-4 protein is localized in neuronal cells in the brain
Hyman, C., Hofer, M., Barde, Y.A., Juhasz, M., Yancopoulos, stem as well as the dorsal root ganglion of embryonic and
G.D., Squinto, S.P. and Lindsay, R.M. (1991) BDNF is a adult rats. J. Neurochem., 86: 660–668.
neurotrophic factor for dopaminergic neurons of the subst- Katoh-Semba, R., Kaisho, Y., Shintani, A., Nagahama, M.
antia nigra. Nature, 350: 230–232. and Kato, K. (1996) Tissue distribution and immunocyto-
Ip, N.Y., Ibanez, C.F., Nye, S.H., McClain, J., Jones, P.F., Gies, chemical localization of neurotrophin-3 in the brain and pe-
D.R., Belluscio, L., Le Beau, M.M., Espinosa III, R., Squinto, ripheral tissues of rats. J. Neurochem., 66: 330–337.
S.P., et al. (1992) Mammalian neurotrophin-4: structure, chro- Kernie, S.G., Liebl, D.J. and Parada, L.F. (2000) BDNF reg-
mosomal localization, tissue distribution, and receptor specifi- ulates eating behavior and locomotor activity in mice. EMBO
city. Proc. Natl. Acad. Sci. U.S.A., 89: 3060–3064. J., 19: 1290–1300.
Isackson, P.J., Huntsman, M.M., Murray, K.D. and Gall, Kerr, B.J., Bradbury, E.J., Bennett, D.L., Trivedi, P.M.,
C.M. (1991) BDNF mRNA expression is increased in adult Dassan, P., French, J., Shelton, D.B., McMahon, S.B. and
rat forebrain after limbic seizures: temporal patterns of in- Thompson, S.W. (1999) Brain-derived neurotrophic factor
duction distinct from NGF. Neuron, 6: 937–948. modulates nociceptive sensory inputs and NMDA-evoked
Ishibashi, H., Hihara, S., Takahashi, M., Heike, T., Yokota, T. responses in the rat spinal cord. J. Neurosci., 19:
and Iriki, A. (2002) Tool-use learning induces BDNF ex- 5138–5148.
pression in a selective portion of monkey anterior parietal Kim, H.G., Wang, T., Olafsson, P. and Lu, B. (1994) Ne-
cortex. Brain Res. Mol. Brain Res., 102: 110–112. urotrophin 3 potentiates neuronal activity and inhibits
Ito, H., Nakajima, A., Nomoto, H. and Furukawa, S. (2003) gamma-aminobutyratergic synaptic transmission in cortical
Neurotrophins facilitate neuronal differentiation of cultured neurons. Proc. Natl. Acad. Sci. U.S.A., 91: 12341–12345.
neural stem cells via induction of mRNA expression of basic Kim, S.Y., Smith, M.A., Post, R.M. and Rosen, J.B. (1998)
helix-loop-helix transcription factors Mash1 and Math1. J. Attenuation of kindling-induced decreases in NT-3 mRNA
Neurosci. Res., 71: 648–658. by thyroid hormone depletion. Epilepsy Res., 29: 211–220.
Jankowsky, J.L. and Patterson, P.H. (2001) The role of Knusel, B., Gao, H., Okazaki, T., Yoshida, T., Mori, N., Hefti,
cytokines and growth factors in seizures and their sequelae. F. and Kaplan, D.R. (1997) Ligand-induced down-regulation
Prog. Neurobiol., 63: 125–149. of trk messenger RNA, protein and tyrosine phosphorylation
Johnson, J.E., Barde, Y.A., Schwab, M. and Thoenen, H. in rat cortical neurons. Neuroscience, 78: 851–862.
(1986) Brain-derived neurotrophic factor supports the sur- Knusel, B., Winslow, J.W., Rosenthal, A., Burton, L.E., Seid,
vival of cultured rat retinal ganglion cells. J. Neurosci., 6: D.P., Nikolics, K. and Hefti, F. (1991) Promotion of central
3031–3038. cholinergic and dopaminergic neuron differentiation by
Johnson, S.A. and Simmon, V.F. (2002) Randomized, double- brain-derived neurotrophic factor but not neurotrophin 3.
blind, placebo-controlled international clinical trial of the Proc. Natl. Acad. Sci. U.S.A., 88: 961–965.
Ampakine CX516 in elderly participants with mild cognitive Kohara, K., Kitamura, A., Morishima, M. and Tsumoto, T.
impairment: a progress report. J. Mol. Neurosci., 19: (2001) Activity-dependent transfer of brain-derived neurotro-
197–200. phic factor to postsynaptic neurons. Science, 291: 2419–2423.
392

Kokaia, M., Asztely, F., Olofsdotter, K., Sindreu, C.B., Kull- Lessmann, V., Gottmann, K. and Heumann, R. (1994) BDNF
mann, D.M. and Lindvall, O. (1998) Endogenous neurotro- and NT-4/5 enhance glutamatergic synaptic transmission in
phin-3 regulates short-term plasticity at lateral perforant cultured hippocampal neurones. Neuroreport, 6: 21–25.
path-granule cell synapses. J. Neurosci., 18: 8730–8739. Levi-Montalcini, R. and Hamburger, V. (1951) Selective
Kokaia, M., Ernfors, P., Kokaia, Z., Elmer, E., Jaenisch, R. growth-stimulating effects of mouse sarcoma on the sensory
and Lindvall, O. (1995) Suppressed epileptogenesis in BDNF and sympathetic nervous system of the chick embryo. J. Exp.
mutant mice. Exp. Neurol., 133: 215–224. Zool., 116: 321–361.
Korte, M., Carroll, P., Wolf, E., Brem, G., Thoenen, H. and Li, S., Saragovi, H.U., Nedev, H., Zhao, C., Racine, R.J. and
Bonhoeffer, T. (1995) Hippocampal long-term potentiation is Fahnestock, M. (2005) Differential actions of nerve growth
impaired in mice lacking brain-derived neurotrophic factor. factor receptors TrkA and p75NTR in a rat model of epile-
Proc. Natl. Acad. Sci. U.S.A., 92: 8856–8860. ptogenesis. Mol. Cell Neurosci., 29: 162–172.
Korte, M., Griesbeck, O., Gravel, C., Carroll, P., Staiger, V., Li, S., Uri Saragovi, H., Racine, R.J. and Fahnestock, M.
Thoenen, H. and Bonhoeffer, T. (1996) Virus-mediated gene (2003) A ligand of the p65/p95 receptor suppresses perforant
transfer into hippocampal CA1 region restores long-term path kindling, kindling-induced mossy fiber sprouting, and
potentiation in brain-derived neurotrophic factor mutant hilar area changes in adult rats. Neuroscience, 119:
mice. Proc. Natl. Acad. Sci. U.S.A., 93: 12547–12552. 1147–1156.
Korte, M., Minichiello, L., Klein, R. and Bonhoeffer, T. (2000) Lin, S.Y., Wu, K., Levine, E.S., Mount, H.T., Suen, P.C. and
Shc-binding site in the TrkB receptor is not required for Black, I.B. (1998) BDNF acutely increases tyrosine phos-
hippocampal long-term potentiation. Neuropharmacology, phorylation of the NMDA receptor subunit 2B in cortical
39: 717–724. and hippocampal postsynaptic densities. Brain Res. Mol.
Lahteinen, S., Pitkanen, A., Saarelainen, T., Nissinen, J., Ko- Brain Res., 55: 20–27.
ponen, E. and Castren, E. (2002) Decreased BDNF signalling Lindholm, D., Castren, E., Berzaghi, M., Blochl, A. and
in transgenic mice reduces epileptogenesis. Eur. J. Neurosci., Thoenen, H. (1994) Activity-dependent and hormonal regu-
15: 721–734. lation of neurotrophin mRNA levels in the brain–implica-
Larmet, Y., Reibel, S., Carnahan, J., Nawa, H., Marescaux, C. tions for neuronal plasticity. J. Neurobiol., 25: 1362–1372.
and Depaulis, A. (1995) Protective effects of brain-derived Lindvall, O., Kokaia, Z., Bengzon, J., Elmer, E. and Kokaia,
neurotrophic factor on the development of hippocampal kin- M. (1994) Neurotrophins and brain insults. Trends Neuro-
dling in the rat. Neuroreport, 6: 1937–1941. sci., 17: 490–496.
Larsson, E., Mandel, R.J., Klein, R.L., Muzyczka, N., Lindv- Linnarsson, S., Bjorklund, A. and Ernfors, P. (1997) Learning
all, O. and Kokaia, Z. (2002) Suppression of insult-induced deficit in BDNF mutant mice. Eur. J. Neurosci., 9:
neurogenesis in adult rat brain by brain-derived neurotrophic 2581–2587.
factor. Exp. Neurol., 177: 1–8. Liu, X., Ernfors, P., Wu, H. and Jaenisch, R. (1995) Sensory
Lauterborn, J.C., Isackson, P.J. and Gall, C.M. (1994) Seizure- but not motor neuron deficits in mice lacking NT4 and
induced increases in NGF mRNA exhibit different time BDNF. Nature, 375: 238–241.
courses across forebrain regions and are biphasic in hippo- Lohof, A.M., Ip, N.Y. and Poo, M.M. (1993) Potentiation of
campus. Exp. Neurol., 125: 22–40. developing neuromuscular synapses by the neurotrophins
Lauterborn, J.C., Rivera, S., Stinis, C.T., Hayes, V.Y., Isack- NT-3 and BDNF. Nature, 363: 350–353.
son, P.J. and Gall, C.M. (1996) Differential effects of pro- Lowenstein, D.H., Seren, M.S. and Longo, F.M. (1993) Pro-
tein synthesis inhibition on the activity-dependent longed increases in neurotrophic activity associated with
expression of BDNF transcripts: evidence for immediate- kainate-induced hippocampal synaptic reorganization. Ne-
early gene responses from specific promoters. J. Neurosci., uroscience, 56: 597–604.
16: 7428–7436. Luikart, B.W., Nef, S., Shipman, T. and Parada, L.F. (2003) In
Lauterborn, J.C., Truong, G.S., Baudry, M., Bi, X., Lynch, G. vivo role of truncated trkb receptors during sensory ganglion
and Gall, C.M. (2003) Chronic elevation of brain-derived neurogenesis. Neuroscience, 117: 847–858.
neurotrophic factor by ampakines. J. Pharmacol. Exp. Ther., Lyons, W.E., Mamounas, L.A., Ricaurte, G.A., Coppola, V.,
307: 297–305. Reid, S.W., Bora, S.H., Wihler, C., Koliatsos, V.E. and
Lee, F.S., Kim, A.H., Khursigara, G. and Chao, M.V. (2001a) Tessarollo, L. (1999) Brain-derived neurotrophic factor-defi-
The uniqueness of being a neurotrophin receptor. Curr. cient mice develop aggressiveness and hyperphagia in con-
Opin. Neurobiol., 11: 281–286. junction with brain serotonergic abnormalities. Proc. Natl.
Lee, J., Duan, W. and Mattson, M.P. (2002) Evidence that Acad. Sci. U.S.A., 96: 15239–15244.
brain-derived neurotrophic factor is required for basal ne- Maisonpierre, P.C., Belluscio, L., Friedman, B., Alderson,
urogenesis and mediates, in part, the enhancement of neuro- R.F., Wiegand, S.J., Furth, M.E., Lindsay, R.M. and
genesis by dietary restriction in the hippocampus of adult Yancopoulos, G.D. (1990a) NT-3, BDNF, and NGF in the
mice. J. Neurochem., 82: 1367–1375. developing rat nervous system: parallel as well as reciprocal
Lee, R., Kermani, P., Teng, K.K. and Hempstead, B.L. (2001b) patterns of expression. Neuron, 5: 501–509.
Regulation of cell survival by secreted proneurotrophins. Maisonpierre, P.C., Belluscio, L., Squinto, S., Ip, N.Y., Furth,
Science, 294: 1945–1948. M.E., Lindsay, R.M. and Yancopoulos, G.D. (1990b)
393

Neurotrophin-3: a neurotrophic factor related to NGF and Morimoto, K., Sato, K., Sato, S., Yamada, N. and Hayabara,
BDNF. Science, 247: 1446–1451. T. (1998) Time-dependent changes in neurotrophic factor
Malcangio, M. and Lessmann, V. (2003) A common thread for mRNA expression after kindling and long-term potentiation
pain and memory synapses? Brain-derived neurotrophic fac- in rats. Brain Res. Bull., 45: 599–605.
tor and trkB receptors. Trends Pharmacol. Sci., 24: 116–121. Mowla, S.J., Farhadi, H.F., Pareek, S., Atwal, J.K., Morris,
Mamounas, L.A., Blue, M.E., Siuciak, J.A. and Altar, C.A. S.J., Seidah, N.G. and Murphy, R.A. (2001) Biosynthesis and
(1995) Brain-derived neurotrophic factor promotes the sur- post-translational processing of the precursor to brain-de-
vival and sprouting of serotonergic axons in rat brain. J. rived neurotrophic factor. J. Biol. Chem., 276: 12660–12666.
Neurosci., 15: 7929–7939. Mowla, S.J., Pareek, S., Farhadi, H.F., Petrecca, K., Fawcett,
Marksteiner, J., Ortler, M., Bellmann, R. and Sperk, G. (1990) J.P., Seidah, N.G., Morris, S.J., Sossin, W.S. and Murphy,
Neuropeptide Y biosynthesis is markedly induced in mossy R.A. (1999) Differential sorting of nerve growth factor and
fibers during temporal lobe epilepsy of the rat. Neurosci. brain-derived neurotrophic factor in hippocampal neurons. J.
Lett., 112: 143–148. Neurosci., 19: 2069–2080.
Martinowich, K., Hattori, D., Wu, H., Fouse, S., He, F., Hu, Mudo, G., Jiang, X.H., Timmusk, T., Bindoni, M. and Bell-
Y., Fan, G. and Sun, Y.E. (2003) DNA methylation-related uardo, N. (1996) Change in neurotrophins and their receptor
chromatin remodeling in activity-dependent BDNF gene reg- mRNAs in the rat forebrain after status epilepticus induced
ulation. Science, 302: 890–893. by pilocarpine. Epilepsia, 37: 198–207.
Marty, S., Berninger, B., Carroll, P. and Thoenen, H. (1996) Mudo, G., Salin, T., Condorelli, D.F., Jiang, X.H., Dell’Al-
GABAergic stimulation regulates the phenotype of hippo- bani, P., Timmusk, T., Metsis, M., Funakoshi, H. and Bell-
campal interneurons through the regulation of brain-derived uardo, N. (1995) Seizures increase trkC mRNA expression in
neurotrophic factor. Neuron, 16: 565–570. the dentate gyrus of rat hippocampus. Role of glutamate re-
Mathern, G.W., Babb, T.L., Micevych, P.E., Blanco, C.E. and ceptor activation. J. Mol. Neurosci., 6: 11–22.
Pretorius, J.K. (1997) Granule cell mRNA levels for BDNF, Murer, M.G., Yan, Q. and Raisman-Vozari, R. (2001) Brain-
NGF, and NT-3 correlate with neuron losses or supragran- derived neurotrophic factor in the control human brain, and
ular mossy fiber sprouting in the chronically damaged and in Alzheimer’s disease and Parkinson’s disease. Prog. Ne-
epileptic human hippocampus. Mol. Chem. Neuropathol., urobiol., 63: 71–124.
30: 53–76. Murphy, D.D., Cole, N.B. and Segal, M. (1998) Brain-derived
McAllister, A.K., Katz, L.C. and Lo, D.C. (1997) Opposing neurotrophic factor mediates estradiol-induced dendritic
roles for endogenous BDNF and NT-3 in regulating cortical spine formation in hippocampal neurons. Proc. Natl. Acad.
dendritic growth. Neuron, 18: 767–778. Sci. U.S.A., 95: 11412–11417.
McLean Bolton, M., Pittman, A.J. and Lo, D.C. (2000) Brain- Nakata, K., Ujike, H., Sakai, A., Uchida, N., Nomura, A.,
derived neurotrophic factor differentially regulates excitatory Imamura, T., Katsu, T., Tanaka, Y., Hamamura, T. and
and inhibitory synaptic transmission in hippocampal cul- Kuroda, S. (2003) Association study of the brain-derived ne-
tures. J. Neurosci., 20: 3221–3232. urotrophic factor (BDNF) gene with bipolar disorder. Ne-
Merlio, J.P., Ernfors, P., Jaber, M. and Persson, H. (1992) urosci. Lett., 337: 17–20.
Molecular cloning of rat trkC and distribution of cells ex- Narisawa-Saito, M. and Nawa, H. (1996) Differential regula-
pressing messenger RNAs for members of the trk family in tion of hippocampal neurotrophins during aging in rats. J.
the rat central nervous system. Neuroscience, 51: 513–532. Neurochem., 67: 1124–1131.
Merlio, J.P., Ernfors, P., Kokaia, Z., Middlemas, D.S., Ben- Nawa, H., Carnahan, J. and Gall, C. (1995) BDNF protein
gzon, J., Kokaia, M., Smith, M.L., Seisjo, B.K., Hunter, T. measured by a novel enzyme immunoassay in normal brain
and Lindvall, O. (1993) Increased production of the trkB and after seizure: partial disagreement with mRNA levels.
protein tyrosine kinase receptor after brain insults. Neuron, Eur. J. Neurosci., 7: 1527–1535.
10: 151–164. Nawa, H., Pelleymounter, M.A. and Carnahan, J. (1994) In-
Metsis, M., Timmusk, T., Arenas, E. and Persson, H. (1993) traventricular administration of BDNF increases neuropep-
Differential usage of multiple brain-derived neurotrophic tide expression in newborn rat brain. J. Neurosci., 14:
factor promoters in the rat brain following neuronal activa- 3751–3765.
tion. Proc. Natl. Acad. Sci. U.S.A., 90: 8802–8806. Neeper, S.A., Gomez-Pinilla, F., Choi, J. and Cotman, C.
Miller, F.D. and Kaplan, D.R. (2001) On Trk for retrograde (1995) Exercise and brain neurotrophins. Nature, 373: 109.
signaling. Neuron, 32: 767–770. Neves-Pereira, M., Mundo, E., Muglia, P., King, N., Mac-
Minichiello, L., Calella, A.M., Medina, D.L., Bonhoeffer, T., ciardi, F. and Kennedy, J.L. (2002) The brain-derived ne-
Klein, R. and Korte, M. (2002) Mechanism of TrkB-medi- urotrophic factor gene confers susceptibility to bipolar
ated hippocampal long-term potentiation. Neuron, 36: disorder: evidence from a family-based association study.
121–137. Am. J. Hum. Genet., 71: 651–655.
Minichiello, L., Korte, M., Wolfer, D., Kuhn, R., Unsicker, K., Nibuya, M., Morinobu, S. and Duman, R.S. (1995) Regulation
Cestari, V., Rossi-Arnaud, C., Lipp, H.P., Bonhoeffer, T. of BDNF and trkB mRNA in rat brain by chronic electro-
and Klein, R. (1999) Essential role for TrkB receptors in convulsive seizure and antidepressant drug treatments. J.
hippocampus-mediated learning. Neuron, 24: 401–414. Neurosci., 15: 7539–7547.
394

Nibuya, M., Nestler, E.J. and Duman, R.S. (1996) Chronic Reibel, S., Larmet, Y., Carnahan, J., Marescaux, C. and De-
antidepressant administration increases the expression of paulis, A. (2000a) Endogenous control of hippocampal epile-
cAMP response element binding protein (CREB) in rat hip- ptogenesis: a molecular cascade involving brain-derived
pocampus. J. Neurosci., 16: 2365–2372. neurotrophic factor and neuropeptide Y. Epilepsia, 41(Sup-
Osehobo, P., Adams, B., Sazgar, M., Verdi, J., Racine, R. and pl 6): S127–S133.
Fahnestock, M. (1996) Effects of in vivo BDNF infusion on Reibel, S., Larmet, Y., Le, B.T., Carnahan, J., Marescaux, C.
amygdala kindling, sprouting, and hilar area. Soc. Neurosci. and Depaulis, A. (2000b) Brain-derived neurotrophic factor
Abstr., 22: 995. delays hippocampal kindling in the rat. Neuroscience, 100:
Patapoutian, A. and Reichardt, L.F. (2001) Trk receptors: me- 777–788.
diators of neurotrophin action. Curr. Opin. Neurobiol., 11: Riccio, A., Pierchala, B.A., Ciarallo, C.L. and Ginty, D.D.
272–280. (1997) An NGF-trkA-mediated retrograde signal to tran-
Patel, M.N. and McNamara, J.O. (1995) Selective enhancement scription factor CREB in sympathetic neurons. Science, 277:
of axonal branching of cultured dentate gyrus neurons by 1097–1100.
neurotrophic factors. Neuroscience, 69: 763–770. Rios, M., Fan, G., Fekete, C., Kelly, J., Bates, B., Kuehn, R.,
Patterson, S.L., Abel, T., Deuel, T.A., Martin, K.C., Rose, J.C. Lechan, R.M. and Jaenisch, R. (2001) Conditional deletion
and Kandel, E.R. (1996) Recombinant BDNF rescues deficits of brain-derived neurotrophic factor in the postnatal brain
in basal synaptic transmission and hippocampal LTP in leads to obesity and hyperactivity. Mol. Endocrinol., 15:
BDNF knockout mice. Neuron, 16: 1137–1145. 1748–1757.
Patterson, S.L., Grover, L.M., Schwartzkroin, P.A. and Rivera, C., Li, H., Thomas-Crusells, J., Lahtinen, H., Viitanen,
Bothwell, M. (1992) Neurotrophin expression in rat hip- T., Nanobashvili, A., Kokaia, Z., Airaksinen, M.S., Voipio,
pocampal slices: a stimulus paradigm inducing LTP in CA1 J., Kaila, K. and Saarma, M. (2002) BDNF-induced TrkB
evokes increases in BDNF and NT-3 mRNAs. Neuron, 9: activation down-regulates the K+-Cl- cotransporter KCC2
1081–1088. and impairs neuronal Cl- extrusion. J. Cell Biol., 159:
Pencea, V., Bingaman, K.D., Wiegand, S.J. and Luskin, M.B. 747–752.
(2001) Infusion of brain-derived neurotrophic factor into the Roback, J.D., Marsh, H.N., Downen, M., Palfrey, H.C. and
lateral ventricle of the adult rat leads to new neurons in the Wainer, B.H. (1995) BDNF-activated signal transduction in
parenchyma of the striatum, septum, thalamus, and hypo- rat cortical glial cells. Eur. J. Neurosci., 7: 849–862.
thalamus. J. Neurosci., 21: 6706–6717. Rocamora, N., Welker, E., Pascual, M. and Soriano, E. (1996)
Pezet, S., Malcangio, M., Lever, I.J., Perkinton, M.S., Thomp- Upregulation of BDNF mRNA expression in the barrel cor-
son, S.W., Williams, R.J. and McMahon, S.B. (2002) Nox- tex of adult mice after sensory stimulation. J. Neurosci., 16:
ious stimulation induces Trk receptor and downstream ERK 4411–4419.
phosphorylation in spinal dorsal horn. Mol. Cell Neurosci., Rose, C.R., Blum, R., Pichler, B., Lepier, A., Kafitz, K.W. and
21: 684–695. Konnerth, A. (2003) Truncated TrkB-T1 mediates neurotro-
Phillips, H.S., Hains, J.M., Armanini, M., Laramee, G.R., phin-evoked calcium signalling in glia cells. Nature, 426:
Johnson, S.A. and Winslow, J.W. (1991) BDNF mRNA is 74–78.
decreased in the hippocampus of individuals with Al- Roux, P.P., Colicos, M.A., Barker, P.A. and Kennedy, T.E.
zheimer’s disease. Neuron, 7: 695–702. (1999) p75 neurotrophin receptor expression is induced in
Pioro, E.P. and Cuello, A.C. (1990) Distribution of nerve apoptotic neurons after seizure. J. Neurosci., 19: 6887–6896.
growth factor receptor-like immunoreactivity in the adult rat Rudge, J.S., Mather, P.E., Pasnikowski, E.M., Cai, N., Corco-
central nervous system. Effect of colchicine and correlation ran, T., Acheson, A., Anderson, K., Lindsay, R.M. and
with the cholinergic system–I. Forebrain. Neuroscience, 34: Wiegand, S.J. (1998) Endogenous BDNF protein is increased
57–87. in adult rat hippocampus after a kainic acid induced excito-
Poo, M.M. (2001) Neurotrophins as synaptic modulators. Nat. toxic insult but exogenous BDNF is not neuroprotective.
Rev. Neurosci., 2: 24–32. Exp. Neurol., 149: 398–410.
Qiao, X., Hughes, P.E., Venero, J.L., Dugich-Djordjevic, Russo-Neustadt, A., Beard, R.C. and Cotman, C.W. (1999)
M.M., Nichols, N.R., Hefti, F. and Knusel, B. (1996) NT- Exercise, antidepressant medications, and enhanced brain
4/5 protects against adrenalectomy-induced apoptosis of rat derived neurotrophic factor expression. Neuropsychophar-
hippocampal granule cells. Neuroreport, 7: 682–686. macology, 21: 679–682.
Qiao, X., Suri, C., Knusel, B. and Noebels, J.L. (2001) Absence Ryden, M. and Ibanez, C.F. (1996) Binding of neurotrophin-3
of hippocampal mossy fiber sprouting in transgenic mice to p75LNGFR, TrkA, and TrkB mediated by a single func-
overexpressing brain-derived neurotrophic factor. J. Neuro- tional epitope distinct from that recognized by trkC. J. Biol.
sci. Res., 64: 268–276. Chem., 271: 5623–5627.
Rashid, K., Van der Zee, C.E., Ross, G.M., Chapman, C.A., Rylett, R.J. and Williams, L.R. (1994) Role of neurotrophins in
Stanisz, J., Riopelle, R.J., Racine, R.J. and Fahnestock, M. cholinergic-neurone function in the adult and aged CNS.
(1995) A nerve growth factor peptide retards seizure devel- Trends Neurosci., 17: 486–490.
opment and inhibits neuronal sprouting in a rat model of Saarelainen, T., Hendolin, P., Lucas, G., Koponen, E., Sa-
epilepsy. Proc. Natl. Acad. Sci. U.S.A., 92: 9495–9499. iranen, M., MacDonald, E., Agerman, K., Haapasalo, A.,
395

Nawa, H., Aloyz, R., Ernfors, P. and Castren, E. (2003) Ac- Senger, D.L. and Campenot, R.B. (1997) Rapid retrograde ty-
tivation of the TrkB neurotrophin receptor is induced by rosine phosphorylation of trkA and other proteins in rat
antidepressant drugs and is required for antidepressant-in- sympathetic neurons in compartmented cultures. J. Cell Biol.,
duced behavioral effects. J. Neurosci., 23: 349–357. 138: 411–421.
Saarelainen, T., Lukkarinen, J.A., Koponen, S., Grohn, O.H., Shieh, P.B. and Ghosh, A. (1999) Molecular mechanisms un-
Jolkkonen, J., Koponen, E., Haapasalo, A., Alhonen, L., derlying activity-dependent regulation of BDNF expression.
Wong, G., Koistinaho, J., Kauppinen, R.A. and Castren, E. J. Neurobiol., 41: 127–134.
(2000a) Transgenic mice overexpressing truncated trkB ne- Shieh, P.B., Hu, S.C., Bobb, K., Timmusk, T. and Ghosh, A.
urotrophin receptors in neurons show increased susceptibility (1998) Identification of a signaling pathway involved in
to cortical injury after focal cerebral ischemia. Mol. Cell Ne- calcium regulation of BDNF expression. Neuron, 20:
urosci., 16: 87–96. 727–740.
Saarelainen, T., Pussinen, R., Koponen, E., Alhonen, L., Shirayama, Y., Chen, A.C., Nakagawa, S., Russell, D.S. and
Wong, G., Sirvio, J. and Castren, E. (2000b) Transgenic mice Duman, R.S. (2002) Brain-derived neurotrophic factor pro-
overexpressing truncated trkB neurotrophin receptors in duces antidepressant effects in behavioral models of depres-
neurons have impaired long-term spatial memory but nor- sion. J. Neurosci., 22: 3251–3261.
mal hippocampal LTP. Synapse, 38: 102–104. Simonato, M., Bregola, G., Armellin, M., Del Piccolo, P., Rodi,
Sato, K., Kashihara, K., Morimoto, K. and Hayabara, T. D., Zucchini, S. and Tongiorgi, E. (2002) Dendritic targeting
(1996) Regional increases in brain-derived neurotrophic fac- of mRNAs for plasticity genes in experimental models of
tor and nerve growth factor mRNAs during amygdaloid temporal lobe epilepsy. Epilepsia, 43(Suppl 5): 153–158.
kindling, but not in acidic and basic fibroblast growth factor Sklar, P., Gabriel, S.B., McInnis, M.G., Bennett, P., Lim,
mRNAs. Epilepsia, 37: 6–14. Y.M., Tsan, G., Schaffner, S., Kirov, G., Jones, I., Owen,
Scharfman, H., Goodman, J., Macleod, A., Phani, S., Anton- M., Craddock, N., DePaulo, J.R. and Lander, E.S. (2002)
elli, C. and Croll, S. (2005) Increased neurogenesis and the Family-based association study of 76 candidate genes in bi-
ectopic granule cells after intrahippocampal BDNF infusion polar disorder: BDNF is a potential risk locus. Brain-derived
in adult rats. Exp. Neurol., 192: 348–356. neutrophic factor. Mol. Psychiatry, 7: 579–593.
Scharfman, H.E. (1997) Hyperexcitability in combined ent- Smith, M.A., Makino, S., Kvetnansky, R. and Post, R.M.
orhinal/hippocampal slices of adult rat after exposure to (1995) Stress and glucocorticoids affect the expression of
brain-derived neurotrophic factor. J. Neurophysiol., 78: brain-derived neurotrophic factor and neurotrophin-3
1082–1095. mRNAs in the hippocampus. J. Neurosci., 15: 1768–1777.
Scharfman, H.E. (2004) Functional implications of seizure-in- Smith, M.A., Zhang, L.-X., Lyons, W.E. and Mamounas, L.A.
duced neurogenesis. Adv. Exp. Med. Biol., 548: 192–212. (1997) Anterograde transport of endogenous brain-derived
Scharfman, H.E., Goodman, J.H. and Sollas, A.L. (1999) Ac- neurotrophic factor in hippocampal mossy fibers. Neurore-
tions of brain-derived neurotrophic factor in slices from rats port, 8: 1829–1834.
with spontaneous seizures and mossy fiber sprouting in the Sobreviela, T., Clary, D.O., Reichardt, L.F., Brandabur, M.M.,
dentate gyrus. J. Neurosci., 19: 5619–5631. Kordower, J.H. and Mufson, E.J. (1994) TrkA-immunore-
Scharfman, H.E., Goodman, J.H., Sollas, A.L. and Croll, S.D. active profiles in the central nervous system: colocalization
(2002) Spontaneous limbic seizures after intrahippocampal with neurons containing p75 nerve growth factor receptor,
infusion of brain-derived neurotrophic factor. Exp. Neurol., choline acetyltransferase, and serotonin. J. Comp. Neurol.,
174: 201–214. 350: 587–611.
Scharfman, H.E., Mercurio, T.C., Goodman, J.H., Wilson, Spires, T.L., Grote, H.E., Varshney, N.K., Cordery, P.M., van
M.A. and MacLusky, N.J. (2003) Hippocampal excitability Dellen, A., Blakemore, C. and Hannan, A.J. (2004) Envi-
increases during the estrous cycle in the rat: a potential role ronmental enrichment rescues protein deficits in a mouse
for brain-derived neurotrophic factor. J. Neurosci., 23: model of Huntington’s disease, indicating a possible disease
11641–11652. mechanism. J. Neurosci., 24: 2270–2276.
Schinder, A.F. and Poo, M. (2000) The neurotrophin hy- Spruston, N., Lubke, J. and Frotscher, M. (1997) Interneurons
pothesis for synaptic plasticity. Trends Neurosci., 23: in the stratum lucidum of the rat hippocampus: an anatom-
639–645. ical and electrophysiological characterization. J. Comp. Ne-
Schmidt-Kastner, R., Humpel, C., Wetmore, C. and Olson, L. urol., 385: 427–440.
(1996) Cellular hybridization for BDNF, trkB, and NGF Suen, P.-C., Wu, K., Levine, E.S., Mount, H.T.J., Xu, J.-L.,
mRNAs and BDNF-immunoreactivity in rat forebrain after Lin, S.-Y. and Black, I.B. (1997) Brain-derived neurotrophic
pilocarpine-induced status epilepticus. Exp. Brain Res., 107: factor rapidly enhances phosphorylation of the postsynaptic
331–347. N-methyl-D-aspartate receptor subunit 1. Proc. Natl. Acad.
Schmidt-Kastner, R. and Olson, L. (1995) Decrease of ne- Sci. U.S.A., 94: 8191–8195.
urotrophin-3 mRNA in adult rat hippocampus after pilocar- Takahashi, M., Hayashi, S., Kakita, A., Wakabayashi, K.,
pine seizures. Exp. Neurol., 136: 199–204. Fukuda, M., Kameyama, S., Tanaka, R., Takahashi, H. and
Segal, R.A. (2003) Selectivity in neurotrophin signaling: theme Nawa, H. (1999) Patients with temporal lobe epilepsy show
and variations. Annu. Rev. Neurosci., 26: 299–330. an increase in brain-derived neurotrophic factor protein and
396

its correlation with neuropeptide Y. Brain Res., 818: Van der Zee, C.E., Rashid, K., Le, K., Moore, K.A., Stanisz, J.,
579–582. Diamond, J., Racine, R.J. and Fahnestock, M. (1995) Intra-
Tanaka, T., Saito, H. and Matsuki, N. (1997) Inhibition of ventricular administration of antibodies to nerve growth fac-
GABAa synaptic responses by brain-derived neurotrophic tor retards kindling and blocks mossy fiber sprouting in adult
factor (BDNF) in rat hippocampus. J. Neurosci., 17: rats. J. Neurosci., 15: 5316–5323.
2959–2966. Vezzani, A., Ravizza, T., Moneta, D., Conti, M., Borroni,
Tao, X., Finkbeiner, S., Arnold, D.B., Shaywitz, A.J. and A., Rizzi, M., Samanin, R. and Maj, R. (1999) Brain-
Greenberg, M.E. (1998) Ca2+ influx regulates BDNF tran- derived neurotrophic factor immunoreactivity in the limbic
scription by a CREB family transcription factor-dependent system of rats after acute seizures and during spontaneo
mechanism. Neuron, 20: 709–726. us convulsions: temporal evolution of changes as
Tao, X., West, A.E., Chen, W.G., Corfas, G. and Greenberg, compared to neuropeptide Y. Neuroscience, 90:
M.E. (2002) A calcium-responsive transcription factor, 1445–1461.
CaRF, that regulates neuronal activity-dependent expression Von Bartheld, C.S., Byers, M.R., Williams, R. and Bothwell,
of BDNF. Neuron, 33: 383–395. M. (1996a) Anterograde transport of neurotrophins and
Thakker-Varia, S., Alder, J., Crozier, R.A., Plummer, M.R. axodendritic transfer in the developing visual system. Nature,
and Black, I.B. (2001) Rab3A is required for brain-derived 379: 830–833.
neurotrophic factor-induced synaptic plasticity: transcrip- Von Bartheld, C.S., Williams, R., Lefcort, F., Clary, D.O.,
tional analysis at the population and single-cell levels. J. Ne- Reichardt, L.F. and Bothwell, M. (1996b) Retrograde trans-
urosci., 21: 6782–6790. port of neurotrophins from the eye to the brain in chick
Thoenen, H. (1995) Neurotrophins and neuronal plasticity. embryos: roles of the p75NTR and trkB receptors. J. Ne-
Science, 270: 593–598. urosci., 16: 2995–3008.
Thompson, S.W., Bennett, D.L., Kerr, B.J., Bradbury, E.J. and Wardle, R.A. and Poo, M.M. (2003) Brain-derived neurotro-
McMahon, S.B. (1999) Brain-derived neurotrophic factor is phic factor modulation of GABAergic synapses by postsy-
an endogenous modulator of nociceptive responses in the naptic regulation of chloride transport. J. Neurosci., 23:
spinal cord. Proc. Natl. Acad. Sci. U.S.A., 96: 7714–7718. 8722–8732.
Timmusk, T., Belluardo, N., Metsis, M. and Persson, H. Wetmore, C., Ernfors, P., Persson, H. and Olson, L. (1990)
(1993a) Widespread and developmentally regulated expres- Localization of brain-derived neurotrophic factor mRNA to
sion of neurotrophin-4 mRNA in rat brain and peripheral neurons in the brain by in situ hybridization. Exp. Neurol.,
tissues. Eur. J. Neurosci., 5: 605–613. 109: 141–152.
Timmusk, T., Palm, K., Metsis, M., Reintam, T., Paalme, V., Wu, K., Xu, J., Suen, P., Levine, E., Huang, Y., Mount, H.T.J.,
Saarma, M. and Persson, H. (1993b) Multiple promoters di- Lin, S. and Black, I.B. (1996) Functional trkB neurotrophin
rect tissue-specific expression of the rat BDNF gene. Neuron, receptors are intrinsic components of the adult brain postsy-
10: 475–489. naptic density. Mol. Brain Res., 43: 286–290.
Tolwani, R.J., Buckmaster, P.S., Varma, S., Cosgaya, J.M., Xu, B., Gottschalk, W., Chow, A., Wilson, R.I., Schnell, E.,
Wu, Y., Suri, C. and Shooter, E.M. (2002) BDNF overex- Zang, K., Wang, D., Nicoll, R.A., Lu, B. and Reichardt, L.F.
pression increases dendrite complexity in hippocampal dent- (2000) The role of brain-derived neurotrophic factor recep-
ate gyrus. Neuroscience, 114: 795–805. tors in the mature hippocampus: modulation of long-term
Tønder, N., Kragh, J., Finsen, B., Bolwig, T.G. and Zimmer, J. potentiation through a presynaptic mechanism involving
(1994) Kindling induces transient changes in neuronal ex- TrkB. J. Neurosci., 20: 6888–6897.
pression of somatostatin, neuropeptide Y, and calbindin in Xu, B., Goulding, E.H., Zang, K., Cepoi, D., Cone, R.D.,
adult rat hippocampus and fascia dentata. Epilepsia, 35: Jones, K.R., Tecott, L.H. and Reichardt, L.F. (2003) Brain-
1299–1308. derived neurotrophic factor regulates energy balance down-
Tonra, J.R., Curtis, R., Wong, V., Cliffer, K.D., Park, J.S., stream of melanocortin-4 receptor. Nat. Neurosci., 6:
Timmes, A., Nguyen, T., Lindsay, R.M., Acheson, A. and 736–742.
DiStefano, P.S. (1998) Axotomy upregulates the anterograde Xu, B., Michalski, B., Racine, R.J. and Fahnestock, M. (2002)
transport and expression of brain-derived neurotrophic fac- Continuous infusion of neurotrophin-3 triggers sprouting,
tor by sensory neurons. J. Neurosci., 18: 4374–4383. decreases the levels of TrkA and TrkC, and inhibits epile-
Tsai, S.J. (2004) Is mania caused by overactivity of central ptogenesis and activity-dependent axonal growth in adult
brain-derived neurotrophic factor? Med. Hypotheses, 62: rats. Neuroscience, 115: 1295–1308.
19–22. Xu, B., Michalski, B., Racine, R.J. and Fahnestock, M. (2004)
Tyler, W.J., Perrett, S.P. and Pozzo-Miller, L.D. (2002) The The effects of brain-derived neurotrophic factor (BDNF)
role of neurotrophins in neurotransmitter release. Neurosci- administration on kindling induction, Trk expression and
entist, 8: 524–531. seizure-related morphological changes. Neuroscience, 126:
Urfer, R., Tsoulfas, P., O’Connell, L., Shelton, D.L., Parada, 521–531.
L.F. and Presta, L.G. (1995) An immunoglobulin-like do- Yacoubian, T.A. and Lo, D.C. (2000) Truncated and full-
main determines the specificity of neurotrophin receptors. length TrkB receptors regulate distinct modes of dendritic
EMBO J., 14: 2795–2805. growth. Nat. Neurosci., 3: 342–349.
397

Yamada, K. and Nabeshima, T. (2003) Brain-derived neurotro- Zhou, X.-F. and Rush, R.A. (1996) Endogenous brain-derived
phic factor/TrkB signaling in memory processes. J. Pharma- neurotrophic factor is anterogradely transported in primary
col. Sci., 91: 267–270. sensory neurons. Neuroscience, 74: 945–951.
Yan, Q., Matheson, C., Sun, J., Radeke, M.J., Feinstein, S.C. Zhou, Z., Hong, E.J., Cohen, S., Zhao, W.N., Ho, H.Y., Sch-
and Miller, J.A. (1994) Distribution of intracerebral ven- midt, L., Chen, W.G., Lin, Y., Savner, E., Griffith, E.C., Hu,
tricularly administered neurotrophins in rat brain and its L., Steen, J.A., Weitz, C.J. and Greenberg, M.E. (2006)
correlation with trk receptor expression. Exp. Neurol., 127: Brain-specific phosphorylation of MeCP2 regulates activity-
23–36. dependent Bdnf transcription, dendritic growth, and spine
Yan, Q., Radeke, M.J., Matheson, C.R., Talvenheimo, J., maturation. Neuron, 52: 255–269.
Welcher, A.A. and Feinstein, S.C. (1997a) Immunocyto- Zigova, T., Pencea, V., Wiegand, S.J. and Luskin, M.B. (1998)
chemical localization of trkB in the central nervous system of Intraventricular administration of BDNF increases the
the adult rat. J. Comp. Neurol., 378: 135–157. number of newly generated neurons in the adult olfactory
Yan, Q., Rosenfeld, R.D., Matheson, C.R., Hawkins, N., bulb. Mol. Cell Neurosci., 11: 234–245.
Lopez, O.T., Bennett, L. and Welcher, A.A. (1997b) Ex- Zuccato, C., Ciammola, A., Rigamonti, D., Leavitt, B.R.,
pression of brain-derived neurotrophic factor protein in the Goffredo, D., Conti, L., MacDonald, M.E., Friedlander,
adult rat central nervous system. Neuroscience, 78: R.M., Silani, V., Hayden, M.R., Timmusk, T., Sipione, S.
431–448. and Cattaneo, E. (2001) Loss of huntingtin-mediated BDNF
Zaccaro, M.C., Ivanisevic, L., Perez, P., Meakin, S.O. and gene transcription in Huntington’s disease. Science, 293:
Saragovi, H.U. (2001) p75 Co-receptors regulate ligand- 493–498.
dependent and ligand-independent Trk receptor activation, Zuccato, C., Tartari, M., Crotti, A., Goffredo, D., Valenza, M.,
in part by altering Trk docking subdomains. J. Biol. Chem., Conti, L., Cataudella, T., Leavitt, B.R., Hayden, M.R.,
276: 31023–31029. Timmusk, T., Rigamonti, D. and Cattaneo, E. (2003) Hunt-
Zhou, X.F. and Rush, R.A. (1994) Localization of neurotro- ingtin interacts with REST/NRSF to modulate the tran-
phin-3-like immunoreactivity in the rat central nervous sys- scription of NRSE-controlled neuronal genes. Nat. Genet.,
tem. Brain Res., 643: 162–172. 35: 76–83.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 23

Sex steroids and the dentate gyrus

Tibor Hajszan1,2,, Teresa A. Milner3,4 and Csaba Leranth1,5

1
Department of Obstetrics, Gynecology, and Reproductive Sciences, Yale University School of Medicine, 333 Cedar Street,
FMB 312, New Haven, CT 06520, USA
2
Department of Biophysics, Biological Research Center, Hungarian Academy of Sciences, Szeged, Hungary
3
Department of Neurology and Neuroscience, Division of Neurobiology, Weill-Cornell Medical College,
411 East 69th Street, New York, NY 10021, USA
4
Harold and Milliken Hatch Laboratory of Neuroendocrinology, The Rockefeller University, New York, NY, USA
5
Department of Neurobiology, Yale University School of Medicine, New Haven, CT, USA

Abstract: In the late 1980s, the finding that the dentate gyrus contains more granule cells in the male than
in the female of certain mouse strains provided the first indication that the dentate gyrus is a significant
target for the effects of sex steroids during development. Gonadal hormones also play a crucial role in
shaping the function and morphology of the adult brain. Besides reproduction-related processes, sex ster-
oids participate in higher brain operations such as cognition and mood, in which the hippocampus is a
critical mediator. Being part of the hippocampal formation, the dentate gyrus is naturally involved in these
mechanisms and as such, this structure is also a critical target for the activational effects of sex steroids.
These activational effects are the results of three major types of steroid-mediated actions. Sex steroids
modulate the function of dentate neurons under normal conditions. In addition, recent research suggests
that hormone-induced cellular plasticity may play a larger role than previously thought, particularly in the
dentate gyrus. Specifically, the regulation of dentate gyrus neurogenesis and synaptic remodeling by sex
steroids received increasing attention lately. Finally, the dentate gyrus is influenced by gonadal hormones
in the context of cellular injury, and the work in this area demonstrates that gonadal hormones have
neuroprotective potential. The expression of estrogen, progestin, and androgen receptors in the dentate
gyrus suggests that sex steroids, which could be of gonadal origin and/or synthesized locally in the dentate
gyrus, may act directly on dentate cells. In addition, gonadal hormones could also influence the dentate
gyrus indirectly, by subcortical hormone-sensitive structures such as the cholinergic septohippocampal
system. Importantly, these three sex steroid-related themes, functional effects in the normal dentate gyrus,
mechanisms involving neurogenesis and synaptic remodeling, as well as neuroprotection, have substantial
implications for understanding normal cognitive function, with clinical importance for epilepsy,
Alzheimer’s disease and mental disorders.

Keywords: androgen; estrogen; progesterone; sex difference; electrophysiology; neurogenesis; synaptic


remodeling; neuroprotection

Introduction

Corresponding author. Tel.: +1 203 785 4748; In 1989, Wimer and Wimer (1989) reported that in
Fax: +1 203 785 7684; E-mail: tibor.hajszan@yale.edu certain mouse strains, the dentate gyrus contains

DOI: 10.1016/S0079-6123(07)63023-4 399


400

more granule cells in the male than in the female. The finding that prepubescent female rats in-
A few years later, the dentate granule cell layer was jected neonatally with testosterone develop a male-
shown to be larger in males relative to females, both like dentate gyrus and perform better in the Morris
in adult and prepubescent rats, which is well cor- water maze (Roof, 1993) suggests that the ‘‘gen-
related with performance in spatial memory tasks der’’ of the dentate gyrus and many other sexually
(Roof and Havens, 1992; Roof, 1993). Later, in the dimorphic brain functions and structures could be
hilus of the dentate gyrus, the number of synapses readily manipulated via the hormonal milieu. The
formed by mossy fibers, the axons of dentate gran- window during which these hormonal manipula-
ule cells, was demonstrated to be higher in male rats tions can affect dentate gyrus morphology seems
than in females, consistent with the idea that more to be restricted to the first few postnatal days
granule cells in males provide a more robust input (Roof, 1993), because no morphological responses
to CA3 pyramidal neurons (Parducz and Garcia- to sex steroid treatment in the rat dentate gyrus
Segura, 1993). Using rigorous stereology tech- were found before or after this period (Isgor and
niques, however, two studies performed later have Sengelaub, 1998). The findings of Roof (1993)
found no sexual dimorphism in the volume of the support the so-called ‘‘aromatization hypothesis,’’
dentate granule cell layer in Sprague-Dawley (Isgor i.e., the brain is masculinized by estrogen that is
and Sengelaub, 1998) and Long-Evans rats (Jones produced by aromatase from testosterone, which
and Watson, 2005). Paying particular attention to is readily available in the developing male but not
rodent strains in this respect is emphasized by an- in the female (Naftolin et al., 1975). Also consist-
other mouse study, demonstrating that the overall ent with this hypothesis, estrogen receptor (ER)
volume of the dentate granule cell layer in A/J mice mRNA levels in the dentate gyrus increase signifi-
is larger in males than in females, while there is cantly between birth and postnatal day 4, and then
no such sexual dimorphism in C57Bl/6J mice decline by postnatal day 10, while adult male rat
(Tabibnia et al., 1999). These key developments in ER mRNA levels are similar to those found in
our understanding of sex differences in the dentate newborn and postnatal day 10 animals (O’Keefe
gyrus were followed by other examples of sexual et al., 1995). On the other hand, it has been
dimorphism in certain aspects of dentate gyrus reported that the dentate granule cell layer is sig-
function and morphology. For example, female nificantly larger in testicular feminization mutant
rats produce more newly born cells than males in (Tfm) male rats than in wild-type females (Jones
the dentate gyrus, but not in the subventricular and Watson, 2005). Because Tfm rats express a
zone (Tanapat et al., 1999). dysfunctional androgen receptor (AR), this finding
Expression of proteins regulated by estrogen, implicates the AR in the development of the
such as brain-derived neurotrophic factor (BDNF) dentate gyrus, which contradicts the aromatization
(Sohrabji et al., 1995), also differs in males vs. hypothesis. It should be noted, however, that Tfm
females. Using immunocytochemistry, it appears rats retain a considerable portion of AR activity,
that the mossy fibers contain the vast majority of whereas the dentate granule cell layer is not
BDNF protein (Conner et al., 1997). When a pro- sexually dimorphic in Tfm mice with complete
estrous or estrous female rat with high estrogen deletion of AR function (Tabibnia et al., 1999).
levels was compared to a metestrous female,
ovariectomized female, or male, which have low
estrogen blood concentrations, BDNF protein Sex steroids and dentate physiology
expression in the mossy fibers was relatively low
(Scharfman et al., 2003). These studies are The findings of sexual dimorphism strongly
consistent with BDNF expression in mossy fiber- suggest that sex steroids, and their receptors, are
containing micropunches of CA3 assayed by important regulators of dentate gyrus organiza-
ELISA, demonstrating a higher level of BDNF tion, and emphasize their important role in devel-
in the female relative to the male rat (Franklin and opment. The rest of this review, however, will
Perrot-Sinal, 2006). focus on more recent and exciting research on the
401

activational effects of gonadal hormones in the Considering androgen, intrahippocampal mi-


adult brain. In adulthood, sex steroids participate croinjection of neither dehydroepiandrosterone-
not only in reproduction-related central actions, sulfate (DHEAS) nor trilostane, an inhibitor of the
but also in higher brain functions such as cogni- enzyme that metabolizes DHEAS, alters dentate
tion and mood regulation (Korol and Kolo, 2002; field excitatory postsynaptic potential slopes or
Seidman, 2003; Steiner et al., 2003; MacLusky population spike amplitudes, but increases the
et al., 2006). Because the hippocampal formation amplitude of a late component of the postsynaptic
plays an essential role in declarative, spatial, and potential. Both DHEAS and trilostane abolishes
contextual memory, as well as in the regulation GABA-mediated paired-pulse inhibition. In addi-
of mood and the hypothalamic-pituitary-adrenal tion, both DHEAS and trilostane markedly
axis, this limbic structure is a critical target of increases the spontaneous firing rate of dentate
hormone action (McEwen and Alves, 1999; hilar interneurons and synchronizes their firing
McEwen, 2003). As a part of the hippocampal during hippocampal theta rhythm induced by tail
formation, it is highly likely that the dentate gyrus pinch (Steffensen, 1995).
also plays an important role in cognitive function Many studies indicate that sex steroid-induced
and mood regulation. Although research has made electrophysiological changes may be due to mod-
progress in associating hippocampal subregions ulation of dentate glutamate and GABA recep-
with specific functional roles in complex topics tors. Indeed, ovariectomy in rats decreases [3H]
such as cognition and mood, proving that any glutamate binding to N-methyl-D-aspartate
specific aspect of these behaviors is dependent on (NMDA) receptors in the dentate gyrus, while
the dentate gyrus has been more difficult, probably hormone replacement with estradiol, tamoxifen,
because complex hippocampal operations require or raloxifene prevents this decrease (Cyr et al.,
uncompromised signal flow throughout the entire 2000, 2001). On the other hand, [3H] MK-801
hippocampal formation, rather than only in select binding shows that the density of noncompetitive
areas. NMDA antagonist sites is significantly increased
Electrophysiological studies may provide the in the dentate gyrus of ovariectomized compared
most specific insights into the ways sex steroids to sham-operated rats (El-Bakri et al., 2004).
influence the adult dentate gyrus. Working with 17b-estradiol returns [3H] MK-801 binding to the
rat hippocampal slices in the presence or absence normal levels, while progesterone has no effect
of 17b-estradiol, Kim et al. (2006) have reported (Weiland, 1992; El-Bakri et al., 2004). In addi-
that 17b-estradiol significantly potentiates the am- tion, estradiol treatment of ovariectomized rats
plitude and slope of field excitatory postsynaptic significantly increases NMDA R1 subunit protein
potentials in dentate gyrus directly, as well as in levels in granule cell somata, in comparison with
CA3 following mossy fiber stimulation. Repetitive nontreated animals, without concomitant changes
hilar stimuli frequently evoke multiple population in the corresponding mRNA hybridization signal
spikes in CA3 at proestrus and estrus, but only (Gazzaley et al., 1996). Regarding other receptor
rarely at other cycle stages, and never in slices of types, ovarian steroids have no effect on the
ovariectomized rats (Scharfman et al., 2003). This density of kainate or a-amino-3-hydroxy-5-
hyperexcitability in CA3 at proestrus was blocked methylisoxazole-4-propionic acid (AMPA) recep-
by exposure to the high-affinity neurotrophin re- tors (Weiland, 1992; Cyr et al., 2000), while est-
ceptor antagonist K252a, or by an antagonist of radiol-benzoate increases [3H] muscimol binding
the a7 nicotinic cholinergic receptor, whereas it in the dentate gyrus (Schumacher et al., 1989). In
was induced at metestrus by the addition of BDNF situ hybridization showed that progesterone sup-
to hippocampal slices (Scharfman et al., 2003). presses mRNA levels of the a1 GABAA receptor
These findings indicate that an estrogen-induced subunit in the dentate gyrus of animals that were
interaction of BDNF and a7 nicotinic receptors is pretreated with estradiol (Weiland and Orchinik,
important for estrous cycle-related changes in CA3 1995). Finally, there is a significant negative
and dentate gyrus (Scharfman et al., 2003). correlation between testosterone levels and the
402

mRNA level for the a1 GABAA receptor subunit efficacy. Indeed, Ormerod et al. (2003) have
(Orchinik et al., 1995). reported that relative to vehicle-treated rats, the
number of new cells increases following a 4-h ex-
posure but decreases following a 48-h exposure to
Sex steroids and dentate plasticity
estrogen in ovariectomized animals. This decrease
at 48 h is abolished by adrenalectomy, suggesting a
Although numerous findings support the view that
role of adrenal activity (Ormerod et al., 2003).
molecular mechanisms, including receptor
In addition to being effective under normal
changes, play a critical role in alterations of neu-
conditions, estrogen is capable of influencing ne-
ronal activity and functional plasticity (see above),
urogenic potential when neurogenesis is examined
recent evidence suggests that sex steroids may also
in other contexts. For example, a strong reduction
influence long-term potentiation (Lynch, 2004)
in cell proliferation occurs in the dentate gyrus and
and learning/memory via mediating aspects of
subventricular zone of mice sacrificed 20 days after
structural plasticity, such as neurogenesis and
streptozotocin administration, which induces a
synaptic remodeling (Kandel, 2001; Kasai et al.,
diabetic state. This reduction is completely re-
2003).
lieved by 10 days of estradiol pellet implantation,
which increases the circulating estrogen levels
Sex steroids and neurogenesis 30-fold (Saravia et al., 2006). In addition, there is
a striking effect of aging on cell proliferation
Due to the recent confirmation that neurogenesis that appears to be influenced by estradiol. Perez-
continues throughout the lifespan in the adult Martin et al. (2005) have reported that treatment
dentate gyrus (Altman and Das, 1965), the sub- of 22-month-old ovariectomized animals for 10
granular zone, where progenitors are primarily weeks with a weekly subcutaneous injection of
generated, has drawn considerable attention. estradiol-valerianate, or with soy extract added to
Translational implications are one reason: dentate the drinking water, reverses the age-associated
neurogenesis may be a critical factor in the patho- decline in dentate granule cell production (Perez-
physiology of depression and in the mechanism of Martin et al., 2005). Similarly, estrogen also
antidepressant action (Santarelli et al., 2003). The normalizes the deficient granule cell proliferation
first indication that sex steroids may influence in the dentate gyrus of aging mice (De Nicola
adult neurogenesis came from Tanapat et al. et al., 2006).
(1999), who demonstrated that female rats Several studies have provided insight into the
produce more newly born cells than males in the mechanisms underlying the influence of estrogen
dentate gyrus (but not in the subventricular zone). on neurogenesis. Mazzucco et al. (2006) have
They have also reported a fluctuation in cell shown that both diarylpropionitrile, an ERb ago-
proliferation during the estrous cycle: females pro- nist, and propyl-pyrazole triol, an ERa agonist,
duce more newly born cells during proestrus (when significantly enhances cell proliferation in the
estrogen levels are highest) compared with estrus dentate gyrus of female rats. Other findings sug-
and diestrus. Ovariectomy diminishes, while acute gest that ERs are involved in the induction of
treatment with estrogen rapidly increases, cell adult neurogenesis by an interaction with insulin-
proliferation in ovariectomized rats, an effect that like growth factor-1 (IGF-1), as estradiol and
is reversed by the administration of progesterone IGF-1 have a cooperative effect to promote
(Tanapat et al., 1999, 2005; Falconer and Galea, neurogenesis (Perez-Martin et al., 2003). Admin-
2003). Both the prolonged absence of ovarian istration of IGF-1 significantly increases dentate
hormones and chronic treatment decreases the neurogenesis compared to rats treated with
potential of estrogen to stimulate cell proliferation vehicle; and rats treated with both IGF-1 and
(Tanapat et al., 2005), suggesting that both dose- estradiol show a higher level of cell proliferation
response and temporal characteristics of estrogen than rats treated with IGF-1 or estradiol alone
treatment may critically influence its neurogenic (Perez-Martin et al., 2003).
403

Serotonin has also been linked to the effect of unparalleled synaptogenic power in the adult hip-
estradiol on dentate neurogenesis. Administration pocampus. For example, hormone replacement
of 5-hydroxytryptophan, a precursor to serotonin, induces changes on the order of 50–100% in the
restores cell proliferation that was decreased by number of CA1 spine synapses of rats (MacLusky
ovariectomy, whereas estradiol is unable to reverse et al., 2006; Parducz et al., 2006). This synapto-
this change in ovariectomized rats treated with genic efficacy seems to be rivaled only by antide-
p-chlorophenylalanine, an inhibitor of serotonin pressant drugs (Hajszan et al., 2005). Moreover,
synthesis (Banasr et al., 2001). These data impli- estrogen-induced remodeling of hippocampal
cate the central serotonergic system in the medi- spine synapses is remarkably rapid, similar to the
ation of estrogen effects on dentate neurogenesis. time course required for long-term potentiation
Indeed, several studies indicate that estrogen in- induction (MacLusky et al., 2005). The temporal
fluences the dentate serotonergic system (Bowman analogy suggests that formation of spine synapses
et al., 2002). In case of 5-HT1A receptors, estra- may be involved in sex steroid-modulated
diol treatment reduces 5-HT1A gene expression in cognitive functions. Indeed, a great deal of evi-
the dentate gyrus (Birzniece et al., 2001), while dence has accumulated which suggests that rapid
ovariectomy increases 5-HT1A receptor stimula- remodeling and stabilization of small spines
tion, which is reversed by estradiol (Le Saux and (and their associated synaptic contacts) may rep-
Di Paolo, 2005). resent a mechanism of memory formation and
Besides the plethora of data with respect of storage (Sorra and Harris, 2000; Kandel, 2001;
estrogen and neurogenesis, there are almost no Kasai et al., 2003).
published work that address the neurogenic effect Unfortunately, several studies support the view
of androgen except an initial study in songbird that the dentate gyrus may miss this ‘‘synaptogenic
suggesting that testosterone promotes neurogene- party.’’ Woolley et al. (1990) have reported no
sis (Louissaint et al., 2002). Another study has significant changes in dendritic spine density
demonstrated that in a neuronal stem cell culture across the estrous cycle in CA3 pyramidal cells
stimulated with epidermal growth factor, nandro- or dentate granule cells of the rat. Using a similar
lone, a synthetic androgen reduces cell prolifera- Golgi-impregnation technique, Gould et al. (1990)
tion (Brannvall et al., 2005). The decrease is have shown that ovariectomy or gonadal steroid
abolished by flutamide, an AR antagonist. Nand- replacement do not affect spine density of CA3
rolone also reduces new cell production in the pyramidal cells or granule cells of the dentate
dentate gyrus, an effect observed in both female gyrus. In a later study, Miranda et al. (1999) have
and male rats (Brannvall et al., 2005). For more demonstrated that there may be effects in the
details on gonadal hormone modulation of hippo- dentate gyrus, but they are likely to be dependent
campal neurogenesis in the adult, some excellent on age and the temporal pattern of estradiol re-
reviews are available (Gould et al., 2000; Galea placement. In addition, Szymczak et al. (2006)
et al., 2006). For more details on adult neurogen- suggest that ERb expression is negatively corre-
esis and its role in mood regulation, we refer the lated with synapse formation.
reader to other chapters within this volume. Another approach to the topic of synaptic re-
modeling is to address the expression of proteins
that are associated with the pre- or postsynaptic
Sex steroids and synaptic remodeling apparatus. This approach has also failed to lead to
a compelling body of evidence that hormonal fluc-
In 1992, the discovery that estradiol mediates fluc- tuations modulate dentate synaptic remodeling.
tuation in hippocampal CA1 spine synapse density For example, immunoreactivity for spinophilin, a
during the estrous cycle in the adult rat marked the marker of dendritic spines, is increased in the hilar
beginning of a new era in the research of sex ster- region of the dentate gyrus, as well as in CA3, of
oids (Woolley and McEwen, 1992). Subsequent ovariectomized rats treated with estrogen for 2
extensive work has shown that sex steroids hold an days (Brake et al., 2001). However, levels of
404

syntaxin and synaptophysin (presynaptic proteins dentate area examined, the chosen methodological
associated with the transmitter release machinery), approach alone may be critical. What measures
as well as spinophilin, are unaltered by hormone most reliably the remodeling of synaptic connec-
treatment in the dentate gyrus of rhesus monkeys tions, i.e., synaptogenesis or loss of synapses, is
(Choi et al., 2003). Although young female rhesus debatable. Above, we list several light microscopic
monkeys show a trend toward an estrogen-induced approaches, such as Golgi-based estimation of
increase in immunoreactive spines in the dent- dendritic spine density and histochemical detection
ate gyrus outer molecular layer, this effect of pre- and/or postsynaptic marker molecules,
appears to be statistically insignificant (Hao which are widely applied, mainly due to their
et al., 2003). relative methodological simplicity. However, it is
Glia have also been associated with synaptic impossible to decide at the light microscopic level
remodeling, as expansion of the dendritic tree and what proportion of the measured molecular synap-
spine growth may occur at the expense of shrink- tic markers is actually associated with synapses,
ing glial, primarily astroglial volume. As a result, and what proportion represents extrasynaptic mol-
presumably, abundance of astroglial processes ecules that are processed and/or stored in different
and markers is negatively correlated with dendri- cellular compartments. Therefore, levels of synap-
tic spine density. However, the ways sex steroids tic marker molecules may and do change without
alter dentate gyrus glia do not appear to be con- alterations in the number of synapses (Li et al.,
sistent. Luquin et al. (1993) have shown that the 2004). Although Golgi-based estimation of den-
surface density of astroglial cells is positively in- dritic spine density is a less controversial ap-
fluenced by estrogen and progesterone. The sur- proach, it also has several limitations. The most
face density of astroglial cells was significantly important is that spines may or may not form
increased over control values by 5 h after the in- synapses, and the Golgi method is incapable of
jection of estrogen to ovariectomized rats, and as differentiating between spines that have synapses
early as 1 h after the administration of progester- and spines that do not. Thus, similar to synaptic
one; it reached maximal values by 24 h and re- marker molecules, measures of dendritic spines do
turned to control levels by 48 h (Luquin et al., not necessarily reflect the true number of spine
1993). In contrast, levels of glial fibrillary acidic synapses.
protein (GFAP) intron 1, a molecular marker of The most important point in this debate is that
adult astrocytes, shows that GFAP transcription the number of actual spine synapses is more rel-
and mRNA are both decreased in the outer mo- evant to the functional status of neurons than the
lecular layer of the dentate gyrus on the afternoon number of dendritic spines or the levels of any
of proestrus, when plasma estradiol levels are molecular markers. Thus, when the true number of
highest (Stone et al., 1998b). In vitro, astrocytes synapses is questioned, one should count synapses
show interesting bidirectional responses, such that themselves using electron microscopic stereologi-
estrogen treatment increases GFAP transcription cal techniques, because the above-discussed light
in monotypic astrocytic cultures but decreases microscopic markers are not reliable. Treating
GFAP transcription in astrocytes cocultured with ovariectomized rats with 10 mg/day subcutaneous
neurons (Stone et al., 1998b). In mice, Lei et al. estradiol-bezoate for 2 days, a similar schedule
(2003) have reported similar findings, i.e., long- that has been used in previous studies (Gould
term 17b-estradiol treatment in aged female mice et al., 1990), estrogen increased the number of
significantly lowered the number of astrocytes spine synapses in the CA1 stratum radiatum by
in the dentate gyrus and CA1 compared with 71.6% over oil-treated control values, by 50.1% in
placebo. the CA3 stratum radiatum, and by 99.1% in the
What may be in the background of such vari- molecular layer of the dentate gyrus (Fig. 1). It is
able findings? Besides confounding variables such noteworthy that the number of granule cell spine
as the age of animals, strain, dose and temporal synapses in the dentate gyrus doubles after estro-
characteristics of hormone treatment, and/or the gen administration.
405

to trigger excitotoxic cell death in the hilus of the


dentate gyrus, and patients with temporal lobe
epilepsy often exhibit neuronal loss in the hilus
also (Margerison and Corsellis, 1966). Kainic acid
is commonly used as a convulsant to elicit severe
seizures (status epilepticus) and excitotoxic dam-
age in the dentate gyrus of rats (Ben-Ari and
Cossart, 2000). Estrogen administration is capable
of preventing kainic acid-induced degeneration
(Azcoitia et al., 1998; Veliskova et al., 2000).
Estrogen may also mediate the protective actions
of other steroids. For example, the neuroster-
oids pregnenolone and DHEA showed a dose-
dependent protective effect of hilar neurons
against kainic acid. The administration of the aro-
Fig. 1. Effects of ovariectomy and estrogen replacement on the matase inhibitor fadrozole, that blocks the con-
number of hippocampal spine synapses. Young adult female version of these steroids into estrogen, prevented
Sprague-Dawley rats (250 g) were ovariectomized and one week this effect (Veiga et al., 2003). Interestingly,
later, they received either 10 mg/rat/day estradiol-benzoate (EB, 2-methoxyestradiol, an estradiol metabolite, in-
solid columns) or 200 ml/rat/day sesame oil vehicle (oil, open
columns) subcutaneously for 2 days. Two days after the last
duced significant neuronal loss in the hilus, de-
injection, the animals were sacrificed by transcardial perfusion tected 96 h after the treatment with this steroid.
of fixative, and their brains were processed for electron micro- This finding suggests that endogenous metabolism
scopic stereological analysis. Spine synapses were counted in of 17b-estradiol to 2-methoxyestradiol may coun-
the CA1 and CA3 strata radiata, and in the molecular layer of terbalance the neuroprotective effects of estrogen
the dentate gyrus (DG). Significantly different from the cor-
responding Oil group (t-test, po0.001 in CA1 and DG, po0.01
(Picazo et al., 2003).
in CA3). Regarding mechanisms for the neuroprotective
effects of estrogen in studies of seizure-induced
neuronal damage, Veliskova et al. (2000) as well as
Neuroprotective effects Haynes et al. (2003) suggest that intracellular ERs
mediate the neuroprotective effect of estrogen, be-
Due to the vulnerability of some types of neurons cause tamoxifen pretreatment effectively abolished
in the dentate gyrus to insults, this area is a com- estrogen-induced neuroprotection. An interaction
mon subject of neurodegenerative/neuroprotective of ER and IGF-1 receptor signaling may also be
experiments. The work summarized below is important (Azcoitia et al., 1999a). Furthermore,
focused around two topics for which neuroprotec- GABAB receptors are likely to play a role, because
tion is particularly germane: epilepsy and there was a loss of GABAB receptor-mediated in-
Alzheimer’s disease. hibition after kainic acid-induced status epilepticus
Effects of estrogen on seizures vary, depending in the rat dentate gyrus, and pretreatment with
on the experimental approach, and many other estrogen could prevent it (Velisek and Veliskova,
factors. Estrogen may increase neuronal excitabil- 2002).
ity and thus mediate proconvulsant effects that Other steroids besides estrogen are also likely to
have been reported in the past, but reviews of the reduce seizure-induced damage, and the proges-
clinical and animal data show that estrogen may terone metabolite 3a,5a-tetrahyroprogesterone
also have no effect or even anticonvulsant effects (allopregnanolone) has been shown to be one ex-
(Scharfman et al., 2003; Hajszan and MacLusky, ample. Blocking progesterone’s metabolism to
2006; Veliskova, 2006). The protective role of 3a,5a-tetrahyroprogesterone reduced progester-
estrogen in seizure-induced damage is more one’s protective effects in the dentate gyrus
straightforward. Severe seizures have been shown (Rhodes et al., 2004). In the kainate model,
406

3a,5a-tetrahyroprogesterone was able to protect of the lateral part of the entorhinal cortex in-
the hilus from kainic acid (Ciriza et al., 2004). creased axonal sprouting in the outer one-third of
Another metabolite was also effective: 5a-hydro- the molecular layer of the dentate gyrus. Ovariec-
xyprogesterone (Ciriza et al., 2004). tomized mice receiving high and moderate estro-
Other models of injury, which use adrenalec- gen supplementation displayed the same sprouting
tomy to examine neuronal loss in the dentate response. In ovariectomized nontreated mice,
gyrus, focus on the granule cells, because adrenal- however, the sprouting response was significantly
ectomy selectively kills granule cells (see chapter reduced (to nearly nothing) (Kadish and van
by M. Joels in this volume). In this model, estra- Groen, 2002). Finally, Stone et al. (1998a) have
diol treatment reduced pyknotic cell number com- shown that in wild-type ECX mice, ovariectomy
pared to vehicle administration (Frye, 2001). decreases commissural/associational sprouting to
Interestingly, a synthetic glucocorticoid, dexamet- the inner molecular layer of the dentate gyrus,
hasone can also induce apoptosis in the dentate which is reversed by estradiol replacement. In
gyrus, and pretreatment with estrogen substan- ECX apolipoprotein E-knockout mice, however,
tially attenuated the dexamethasone-induced neu- estradiol did not enhance sprouting, suggesting
ronal damage (Haynes et al., 2003). Colchicine, a that sprouting may be stimulated by estrogen
microtubule polymerization inhibitor, also selec- through its up-regulation of apolipoprotein E
tively kills granule cells, an effect that is increased expression, leading to increased recycling of mem-
by ovariectomy and ameliorated by 17b-estradiol brane lipids for use by sprouting neurons (Stone
(Liu et al., 2001). et al., 1998a). Estrogen and apolipoprotein E may
Regarding other sex steroids, treatment of fe- therefore interact in their modulation of both
male or male rats with progesterone or its met- Alzheimer’s disease risk and recovery from
abolites, 5a-dihydroprogesterone and 3a,5a- neuronal injury.
tetrahyroprogesterone similarly reduced the total
number of adrenalectomy-induced pyknotic cells
Mediation of sex steroid effects
in the dentate gyrus compared with vehicle ad-
ministration. In case of androgen, testosterone and
Effects of sex steroids in the dentate gyrus depend
its metabolites, 5a-dihydrotestosterone and 5a-
on the location of their action, and the sites of
androstane-3a,17b-diol significantly reduced the
steroid synthesis. Below we discuss the distribution
number of pyknotic cells in the dentate gyrus
of receptors, which is summarized in Fig. 2. Sex
compared to vehicle-administered, adrenalectomi-
steroids are capable of acting directly on granule
zed female rats (Frye and McCormick, 2000a, b).
cells, as well as indirectly via nongranule cells
Estrogen also has been found to play a critical
within the dentate gyrus (interneurons, mossy
neuroprotective role in Azlheimer’s disease. Uni-
cells). The influence of sex steroids could also be
lateral entorhinal cortex lesion (ECX) is frequently
mediated by subcortical, hormone-sensitive struc-
used as a model of Alzheimer’s disease-like deaff-
tures, such as the septohippocampal cholinergic
erentation in the dentate gyrus. ECX elicits sprout-
neurons. Sites of sex steroid synthesis are usually
ing in the molecular layer, which is affected by
thought to be peripheral, but local synthesis is also
gonadectomy and hormone replacement, but only
possible, and is discussed further below.
in female rats: ovariectomy reduces fiber out-
growth and estrogen restores it (Stone et al., 2000).
However, testosterone replacement had no effect Distribution of ERa in the dentate gyrus
on sprouting in castrated ECX males (Morse et al.,
1986). Sprouting in hippocampal cultures of Estrogen binding, as well as mRNA and immuno-
C57Bl/6J mice was increased by 75% after treat- reactivity for ERa, have been detected in nuclei
ment with 17b-estradiol, which was blocked by an of scattered GABAergic interneurons, located pre-
antagonist of nuclear receptors, tamoxifen (Teter dominantly in the subgranular region of the dent-
et al., 1999). In intact female mice in vivo, lesions ate gyrus (Loy et al., 1988; Shughrue et al., 1997;
407

Fig. 2. Subcellular localization of estrogen (ER), androgen (AR), and progestin (PR) receptors in the dentate gyrus. A subset of
GABAergic interneurons contains nuclear ERa (dark pink). Granule cells, newly born cells (identified by DCX) and some GABAergic
interneurons contain cytosolic and plasma membrane-associated ERb (blue). Dendritic spines, many originating from granule cells
contain ERa, ERb, AR (dark green), and PR (purple). A few dendritic spines in the hilus, likely originating from mossy cells, contain
ERa and ERb. ERa, ERb, AR, and PR are found in axons and axon terminals. Some ERa-containing terminals are cholinergic
(acetylcholine, orange); some ERb-containing terminals resemble monoaminergic boutons. Lot of astrocytes (stars), mostly in the
molecular layer, also contain ERa, ERb, AR, and PR. (See Color Plate 23.2 in color plate section.)

Weiland et al., 1997; Perlman et al., 2005). Nuclear at several extranuclear sites in the dentate gyrus
ERa-immunoreactive interneurons co-express ne- (Milner et al., 2001). Specifically, ERa-immunore-
uropeptide Y, calbindin-D28k and calretinin, but activity is affiliated with the cytoplasmic plasma-
not cholecystokinin or parvalbumin (Nakamura lemma of select hilar interneurons and
and McEwen, 2005). In addition to nuclear recep- with endosomes of a few granule cell perikarya.
tors, ultrastructural studies have revealed ERa Moreover, ERa-labeled profiles are dispersed
408

throughout the dentate gyrus. Approximately half been found in the perikarya of granule cells as well
of these labeled profiles are unmyelinated axons as cells in the dentate subgranular layer (Li et al.,
and axon terminals that contain numerous small, 1997; Shughrue et al., 1997; Mitra et al., 2003;
synaptic vesicles. ERa-labeled terminals form both Milner et al., 2005). Szymczak et al. (2006) have
symmetric and asymmetric synapses on dendritic found that ERb mRNA and protein are displayed
shafts and spines, suggesting that ERa-positive in high levels in the estrus and in low levels in the
axons arise from sources in addition to inhibitory proestrus phase. Recently, robust mRNA expres-
interneurons. Dual labeling revealed that ERa- sion for both the a and b subtypes of ERs has been
immunoreactivity is contained in axons and ter- found in proliferating and differentiating cells of
minals labeled with vesicular acetylcholine trans- neuronal phenotype in the subgranular zone of the
porter (Towart et al., 2003), suggesting that dentate gyrus (Isgor and Watson, 2005). Further-
estrogen could rapidly and directly affect the lo- more, ERb-immunoreactive glia has been ob-
cal release and/or uptake of acetylcholine. About served in the hilus of the dentate gyrus of male
one-quarter of the ERa-immunoreactive profiles and female rats. ERb-immunoreactivity has been
are dendritic spines, many originating from gran- localized in glial processes and perikarya and, in
ule cells. In dendritic spines, ERa-immunoreactiv- some cases, in glial cell nuclei. Double
ity is often associated with the spine apparatus, immunocytochemical labeling of ERb and the
suggesting that estrogen might act locally through specific astroglial marker, GFAP revealed that the
ERa to influence protein synthesis during synaptic ERb-immunoreactive glial cells are astrocytes
remodeling. The remaining one-quarter of ERa- (Azcoitia et al., 1999b). Ultrastructural analysis
labeled profiles are from glial origin that resemble showed ERb-immunoreactivity at several extranu-
astrocytes and are often located near the spines of clear sites in the dentate gyrus (Milner et al., 2005).
granule cells. Vesicular acetylcholine transporter- ERb-immunoreactivity is affiliated with cytoplas-
containing terminals often abut ERa-positive pre- mic organelles, especially endomembranes and mi-
synaptic and glial profiles and unlabeled terminals tochondria, and with the membranes primarily of
that contact ERa-immunoreactive spines (Towart granule cell perikarya and proximal dendrites. Re-
et al., 2003), suggesting that acetylcholine release cent studies revealed that neuronal perikarya and
might play a critical role in estrogen-modulated dendrites labeled with doublecortin, a marker of
structural plasticity. Collectively, these results im- newly generated cells, also contain extranuclear
ply that ERa may serve as both a genomic and ERb-immunoreactivity in both the adult and neo-
nongenomic transducer of estrogen action in the natal dentate gyrus (Herrick et al., 2006). ERb-
dentate gyrus. labeled dendritic shafts and spines have mostly
been found in the molecular layer. In dendritic
processes, ERb-immunoreactivity is near the per-
Distribution of ERb in the dentate gyrus isynaptic zone adjacent to synapses formed by un-
labeled terminals. The ERb protein can also be
The cellular and subcellular locations of ERb- found in preterminal axons and axon terminals,
immunoreactivity in the dentate gyrus are similar associated with clusters of small, synaptic vesicles.
yet distinct from ERa. In monkey, dense ERb ERb-labeled axons are particularly dense in the
hybridization signal has been seen in the dentate hilus and outer molecular layer, forming both
gyrus, CA1, CA2, CA3, CA4, and the pros- asymmetric and symmetric synapses with dend-
ubiculum/subiculum areas of the hippocampus rites. Finally, ERb-immunoreactivity has been de-
(Gundlah et al., 2000). In rodents, cells in or near tected in glial profiles throughout the dentate
the dentate granule cell layer transiently express gyrus, some of which appose doublecortin-labeled
high levels of estrogen binding and ERa protein in perikarya and dendrites (Herrick et al., 2006).
the nucleus during the first two postnatal weeks These results suggest that ERb may serve prima-
(O’Keefe et al., 1995; Solum and Handa, 2001). In rily as a nongenomic transducer of estrogen ac-
adult rats and mice, ERb mRNA and protein has tions in the dentate gyrus.
409

Distribution of progestin receptor (PR) in the AR-immunoreactivity has also been detected in
dentate gyrus astrocytic profiles; many of them apposing termi-
nals that synapse on unlabeled dendritic spines or
Cells containing PR mRNA have been detected in forming gap junctions with other AR-positive
the dentate subgranular zone (Hagihara et al., or unlabeled astrocytes (Tabori et al., 2005).
1992). By light microscopy, nuclear PR-immuno- Together, these results suggest that ARs may serve
reactivity is undetectable in the dentate gyrus; as both a genomic and nongenomic transducer of
however, ultrastructural analysis revealed that the androgen action in the dentate gyrus.
PR protein is found at several extranuclear sites
(Waters et al., 2005). In the molecular layer and
hilus, PR-immunoreactivity is present in dendritic Role of local steroid synthesis
spines, closely associated with the postsynaptic
density. The PR protein is expressed in axons and Recent studies (Rune et al., 2006) suggest that lo-
axon terminals that contain small synaptic vesicles. cally synthesized steroids may contribute to hip-
PR-positive terminals and en passant axonal pocampal activational effects. Using slice cultures,
boutons form synapses with dendritic spines. PR- Rune and colleagues have reported that the
immunoreactivity has also been found in glia, number of dentate proliferative cells decreases,
many resembling astrocytes and some forming whereas the number of apoptotic cells increases
presumed gap junctions with other astrocytic pro- dose-dependently, in response to reduced estradiol
files. The considerable lack of nuclear PR labeling release into the medium after treatment with let-
may indicate that progesterone uses nongenomic rozole, an aromatase enzyme inhibitor (Fester
signaling mechanism in the dentate gyrus to di- et al., 2006). This also holds true for cell cultures
rectly affect dendritic spine morphology and transfected with siRNA against steroidogenic
synaptic plasticity. acute regulatory protein (StAR). StAR transports
cholesterol to the inner mitochondrial membrane,
where it is converted by the cytochrome P-450 en-
Distribution of AR in the dentate gyrus zyme complex, and as such, it is the first step in the
cascade of estrogen synthesis. Application of est-
Previous light microscopic studies have shown that radiol to the medium had no effect on prolifera-
AR mRNA, immunoreactivity, and binding are tion and apoptosis, whereas the antiproliferative
present in pyramidal cell nuclei but not granule and pro-apoptotic effects of StAR knockdown and
cells (Commins and Yahr, 1985; Sar et al., 1990; letrozole administration were restored by treat-
Simerly et al., 1990; Kerr et al., 1995). However, ment of the cultures with estradiol (Fester et al.,
AR-immunoreactivity is present in disperse, punc- 2006). The data of Rune and colleagues are also
tuate processes that are most dense in the pyram- supported by other studies. In situ hybridization
idal cell layer and diffusely distributed in the revealed that StAR and aromatase are highly ex-
mossy fiber pathway (Tabori et al., 2005). Electron pressed in neuronal cells of the dentate gyrus. In
microscopic analysis revealed AR-immunoreactiv- addition, StAR- and aromatase-positive cells are
ity at several extranuclear sites in the dentate strictly correlated with steroidogenic factor-1, a
gyrus. AR labeling has been found in dendritic regulator of steroid biosynthesis, as shown by
spines, many arising from granule cell dendrites. computer-assisted confocal microscopy in double
AR is affiliated with clusters of small, synaptic labeling experiments (Wehrenberg et al., 2001).
vesicles within preterminal axons and axon termi- Similarly, Hojo et al. (2004) have reported in
nals, the majority of these being in the central hi- a rather comprehensive study that in the granule
lus. AR-immunoreactive preterminal axons are cells of the adult male rat dentate gyrus, signifi-
most prominent in the CA3 stratum lucidum. AR- cant colocalization is seen for both cytochromes
labeled terminals exclusively form asymmetric P45017a (dehydroepiandrosterone synthase)
synapses. Throughout the dentate gyrus, and P450 aromatase by means of
410

immunohistochemical staining of slices. Only a partially dependent on input from basal forebrain
weak immunoreaction of these P450s has been cholinergic neurons (Rudick et al., 2003). More
observed in astrocytes and oligodendrocytes evidence is available for the cholinergic role in
(Hojo et al., 2004). More importantly, stimulation dentate neurogenesis. Selective neurotoxic lesion
of hippocampal neurons with NMDA induced a of the forebrain cholinergic input with 192
significant net production of estradiol. The anal- IgG-saporin reduced dentate cell proliferation
ysis of radioactive metabolites demonstrated (Mohapel et al., 2005). Conversely, systemic ad-
the conversion from [3H] pregnenolone to [3H] ministration of the cholinergic agonist physostig-
estradiol through dehydroepiandrosterone and mine increased dentate neurogenesis. The neuro-
testosterone. This activity was abolished by the genic effect of acetylcholine appears to involve
application of specific inhibitors of cytochrome nicotinic receptors containing the b2 subunit
P450s (Hojo et al., 2004). In summary, although (Harrist et al., 2004), as well as m2, m3, and m4
these findings seem compelling, considering the muscarinic receptors (Ma et al., 2000; Mohapel
fact that the majority of the above-mentioned et al., 2005). Consistent with these findings, ova-
data have been obtained from cultures, which riectomy upregulated m4 receptors in the dentate
may be quite different from conditions in vivo, gyrus, whereas estrogen treatment restored
one should exercise caution in interpreting such m4 binding to the level of the sham group
results. (El-Bakri et al., 2002). However, other results
suggest no septohippocampal involvement in the
synaptogenic action of estrogen, because rats
Subcortical mediation of sex steroid effects that received estrogen implants into the medial
septum did not exhibit changes in astroglial proc-
In addition to direct effects of sex steroids on ess density in the dentate gyrus (Lam and Leranth,
dentate cells, indirect actions may influence the 2003).
dentate gyrus. We already mentioned above that
the raphe serotonergic system mediates the dentate
neurogenic effect of estrogen (Banasr et al., 2001).
Consistent with this line of argument, raphe se- Concluding remarks
rotonergic neurons express ERa (Leranth et al.,
1999). Another structure that appears to be critical The studies discussed in this review clearly dem-
is the septohippocampal cholinergic system. Septal onstrate that both the organization and function-
cholinergic neurons express nuclear ERs (Shugh- ing of the dentate gyrus is a significant target of sex
rue et al., 2000) and hippocampal cholinergic steroids. The gonadal hormone modulation of
axons and terminals contain extranuclear ERa physiological activity, neurogenesis, synaptic re-
(Towart et al., 2003). Moreover, estrogen affects modeling, and neurodegeneration/neuroprotection
septohippocampal cholinergic neurons both gen- mechanisms are likely to be relevant to higher
omically and nongenomically (Gibbs and Aggar- brain functions such as cognition and mood, in
wal, 1998; Rudick et al., 2003). Androgen also can addition to their clinical implications in epilepsy,
influence the expression of cholinergic markers in Alzheimer’s disease and mental disorders. Relative
the dentate gyrus. Specifically, gonadectomy re- to hippocampal subfield CA1, however, the role of
duced the density of choline acetyltransferase gonadal hormones in the dentate gyrus has not
immunoreactive fibers in the dentate gyrus, which received substantial attention. In particular, our
was reversed by the addition of testosterone pro- understanding of androgen effects on the dentate
pionate (Nakamura et al., 2002). Although there is gyrus is extremely limited relative to estrogen. The
no direct data from the dentate gyrus, elsewhere in available data suggest potent, complex, and po-
the hippocampus estrogen modulates the inhibi- tentially important effects, however, and therefore
tion by specific GABAergic interneurons, which is merit more attention in the future.
411

Abbreviations nuclei after gonadal hormone manipulation in female rats.


Neuroendocrinology, 74(2): 135–142.
Bowman, R.E., Ferguson, D. and Luine, V.N. (2002) Effects of
AR androgen receptor
chronic restraint stress and estradiol on open field activity,
BDNF brain-derived neurotrophic factor spatial memory, and monoaminergic neurotransmitters in
DHEAS dehydroepiandrosterone-sulfate ovariectomized rats. Neuroscience, 113(2): 401–410.
ECX entorhinal cortex lesion Brake, W.G., Alves, S.E., Dunlop, J.C., Lee, S.J., Bulloch, K.,
ER estrogen receptor Allen, P.B., Greengard, P. and McEwen, B.S. (2001) Novel
GFAP glial fibrillary acidic protein target sites for estrogen action in the dorsal hippocampus: an
IGF-1 insulin-like growth factor-1 examination of synaptic proteins. Endocrinology, 142(3):
1284–1289.
NMDA N-methyl-D-aspartate Brannvall, K., Bogdanovic, N., Korhonen, L. and Lindholm,
PR progestin receptor D. (2005) 19-Nortestosterone influences neural stem cell pro-
StAR steroidogenic acute regulatory protein liferation and neurogenesis in the rat brain. Eur. J. Neurosci.,
Tfm testicular feminization mutant 21(4): 871–878.
Choi, J.M., Romeo, R.D., Brake, W.G., Bethea, C.L., Rose-
nwaks, Z. and McEwen, B.S. (2003) Estradiol increases
pre- and post-synaptic proteins in the CA1 region of the
hippocampus in female rhesus macaques (Macaca mulatta).
Endocrinology, 144(11): 4734–4738.
Acknowledgements Ciriza, I., Azcoitia, I. and Garcia-Segura, L.M. (2004) Reduced
progesterone metabolites protect rat hippocampal neurones
This work was supported by NIH grants from kainic acid excitotoxicity in vivo. J. Neuroendocrinol.,
16(1): 58–63.
MH074021 (T.H.); DA08259, NS07080 and Commins, D. and Yahr, P. (1985) Autoradiographic localiza-
HL18974 (T.A.M.); MH060858 and NS042644 tion of estrogen and androgen receptors in the sexually
(C.L.); as well as by a Hungarian National Office dimorphic area and other regions of the gerbil brain. J.
for Research and Technology grant RET-08/04. Comp. Neurol., 231(4): 473–489.
Conner, J.M., Lauterborn, J.C., Yan, Q., Gall, C.M. and
Varon, S. (1997) Distribution of brain-derived neurotrophic
factor (BDNF) protein and mRNA in the normal adult rat
References CNS: evidence for anterograde axonal transport. J. Neuro-
sci., 17(7): 2295–2313.
Altman, J. and Das, G.D. (1965) Autoradiographic and histo- Cyr, M., Ghribi, O. and Di Paolo, T. (2000) Regional and
logical evidence of postnatal hippocampal neurogenesis in selective effects of oestradiol and progesterone on NMDA
rats. J. Comp. Neurol., 124(3): 319–335. and AMPA receptors in the rat brain. J. Neuroendocrinol.,
Azcoitia, I., Sierra, A. and Garcia-Segura, L.M. (1998) Estra- 12(5): 445–452.
diol prevents kainic acid-induced neuronal loss in the rat Cyr, M., Thibault, C., Morissette, M., Landry, M. and Di
dentate gyrus. Neuroreport, 9(13): 3075–3079. Paolo, T. (2001) Estrogen-like activity of tamoxifen and
Azcoitia, I., Sierra, A. and Garcia-Segura, L.M. (1999a) raloxifene on NMDA receptor binding and expression of its
Neuroprotective effects of estradiol in the adult rat hippo- subunits in rat brain. Neuropsychopharmacology, 25(2):
campus: interaction with insulin-like growth factor-I signal- 242–257.
ling. J. Neurosci. Res., 58(6): 815–822. De Nicola, A.F., Saravia, F.E., Beauquis, J., Pietranera, L. and
Azcoitia, I., Sierra, A. and Garcia-Segura, L.M. (1999b) Ferrini, M.G. (2006) Estrogens and neuroendocrine hypo-
Localization of estrogen receptor beta-immunoreactivity in thalamic-pituitary-adrenal axis function. Front. Horm. Res.,
astrocytes of the adult rat brain. Glia, 26(3): 260–267. 35: 157–168.
Banasr, M., Hery, M., Brezun, J.M. and Daszuta, A. (2001) El-Bakri, N.K., Adem, A., Suliman, I.A., Mulugeta, E., Karls-
Serotonin mediates oestrogen stimulation of cell proliferation son, E., Lindgren, J.U., Winblad, B. and Islam, A. (2002)
in the adult dentate gyrus. Eur. J. Neurosci., 14(9): Estrogen and progesterone treatment: effects on muscarinic
1417–1424. M(4) receptor subtype in the rat brain. Brain Res., 948(1–2):
Ben-Ari, Y. and Cossart, R. (2000) Kainate, a double agent that 131–137.
generates seizures: two decades of progress. Trends Neuro- El-Bakri, N.K., Islam, A., Zhu, S., Elhassan, A., Mohammed,
sci., 23(11): 580–587. A., Winblad, B. and Adem, A. (2004) Effects of estrogen and
Birzniece, V., Johansson, I.M., Wang, M.D., Seckl, J.R., progesterone treatment on rat hippocampal NMDA recep-
Backstrom, T. and Olsson, T. (2001) Serotonin 5-HT(1A) tors: relationship to Morris water maze performance. J. Cell.
receptor mRNA expression in dorsal hippocampus and raphe Mol. Med., 8(4): 537–544.
412

Falconer, E.M. and Galea, L.A. (2003) Sex differences in cell Hajszan, T., MacLusky, N.J. and Leranth, C. (2005) Short-
proliferation, cell death and defensive behavior following term treatment with the antidepressant fluoxetine triggers
acute predator odor stress in adult rats. Brain Res., 975(1–2): pyramidal dendritic spine synapse formation in rat hippo-
22–36. campus. Eur. J. Neurosci., 21(5): 1299–1303.
Fester, L., Ribeiro-Gouveia, V., Prange-Kiel, J., von Schassen, Hao, J., Janssen, W.G., Tang, Y., Roberts, J.A., McKay, H.,
C., Bottner, M., Jarry, H. and Rune, G.M. (2006) Prolifer- Lasley, B., Allen, P.B., Greengard, P., Rapp, P.R., Kordow-
ation and apoptosis of hippocampal granule cells require er, J.H., Hof, P.R. and Morrison, J.H. (2003) Estrogen
local oestrogen synthesis. J. Neurochem., 97(4): 1136–1144. increases the number of spinophilin-immunoreactive spines
Franklin, T.B. and Perrot-Sinal, T.S. (2006) Sex and ovarian in the hippocampus of young and aged female rhesus mon-
steroids modulate brain-derived neurotrophic factor (BDNF) keys. J. Comp. Neurol., 465(4): 540–550.
protein levels in rat hippocampus under stressful and non- Harrist, A., Beech, R.D., King, S.L., Zanardi, A., Cleary,
stressful conditions. Psychoneuroendocrinology, 31(1): M.A., Caldarone, B.J., Eisch, A., Zoli, M. and Picciotto,
38–48. M.R. (2004) Alteration of hippocampal cell proliferation in
Frye, C.A. (2001) Estradiol tends to improve inhibitory avoid- mice lacking the beta 2 subunit of the neuronal nicotinic
ance performance in adrenalectomized male rats and reduces acetylcholine receptor. Synapse, 54(4): 200–206.
pyknotic cells in the dentate gyrus of adrenalectomized male Haynes, L.E., Lendon, C.L., Barber, D.J. and Mitchell, I.J.
and female rats. Brain Res., 889(1–2): 358–363. (2003) 17 Beta-oestradiol attenuates dexamethasone-induced
Frye, C.A. and McCormick, C.M. (2000a) Androgens are lethal and sublethal neuronal damage in the striatum and
neuroprotective in the dentate gyrus of adrenalectomized hippocampus. Neuroscience, 120(3): 799–806.
female rats. Stress, 3(3): 185–194. Herrick, S.P., Waters, E.M., Drake, C.T., McEwen, B.S. and
Frye, C.A. and McCormick, C.M. (2000b) The neurosteroid, Milner, T.A. (2006) Extranuclear estrogen receptor beta
3alpha-androstanediol, prevents inhibitory avoidance deficits immunoreactivity is on doublecortin-containing cells in the
and pyknotic cells in the granule layer of the dentate gyrus adult and neonatal rat dentate gyrus. Brain Res., 1121:
induced by adrenalectomy in rats. Brain Res., 855(1): 46–58.
166–170. Hojo, Y., Hattori, T.A., Enami, T., Furukawa, A., Suzuki, K.,
Galea, L.A., Spritzer, M.D., Barker, J.M. and Pawluski, J.L. Ishii, H.T., Mukai, H., Morrison, J.H., Janssen, W.G.,
(2006) Gonadal hormone modulation of hippocampal ne- Kominami, S., Harada, N., Kimoto, T. and Kawato, S.
urogenesis in the adult. Hippocampus, 16(3): 225–232. (2004) Adult male rat hippocampus synthesizes estradiol
Gazzaley, A.H., Weiland, N.G., McEwen, B.S. and Morrison, from pregnenolone by cytochromes P45017alpha and P450
J.H. (1996) Differential regulation of NMDAR1 mRNA and aromatase localized in neurons. Proc. Natl. Acad. Sci.
protein by estradiol in the rat hippocampus. J. Neurosci., U.S.A., 101(3): 865–870.
16(21): 6830–6838. Isgor, C. and Sengelaub, D.R. (1998) Prenatal gonadal steroids
Gibbs, R.B. and Aggarwal, P. (1998) Estrogen and basal fore- affect adult spatial behavior, CA1 and CA3 pyramidal cell
brain cholinergic neurons: implications for brain aging and morphology in rats. Horm. Behav., 34(2): 183–198.
Alzheimer’s disease-related cognitive decline. Horm. Behav., Isgor, C. and Watson, S.J. (2005) Estrogen receptor alpha and
34(2): 98–111. beta mRNA expressions by proliferating and differentiating
Gould, E., Tanapat, P., Rydel, T. and Hastings, N. (2000) cells in the adult rat dentate gyrus and subventricular zone.
Regulation of hippocampal neurogenesis in adulthood. Biol. Neuroscience, 134(3): 847–856.
Psychiatry, 48(8): 715–720. Jones, B.A. and Watson, N.V. (2005) Spatial memory perform-
Gould, E., Woolley, C.S., Frankfurt, M. and McEwen, B.S. ance in androgen insensitive male rats. Physiol. Behav., 85(2):
(1990) Gonadal steroids regulate dendritic spine density in 135–141.
hippocampal pyramidal cells in adulthood. J. Neurosci., Kadish, I. and van Groen, T. (2002) Low levels of estrogen
10(4): 1286–1291. significantly diminish axonal sprouting after entorhinal cor-
Gundlah, C., Kohama, S.G., Mirkes, S.J., Garyfallou, V.T., tex lesions in the mouse. J. Neurosci., 22(10): 4095–4102.
Urbanski, H.F. and Bethea, C.L. (2000) Distribution of Kandel, E.R. (2001) The molecular biology of memory storage:
estrogen receptor beta (ERbeta) mRNA in hypothalamus, a dialogue between genes and synapses. Science, 294(5544):
midbrain and temporal lobe of spayed macaque: continued 1030–1038.
expression with hormone replacement. Brain Res. Mol. Brain Kasai, H., Matsuzaki, M., Noguchi, J., Yasumatsu, N. and
Res., 76(2): 191–204. Nakahara, H. (2003) Structure-stability-function relation-
Hagihara, K., Hirata, S., Osada, T., Hirai, M. and Kato, J. ships of dendritic spines. Trends Neurosci., 26(7): 360–368.
(1992) Distribution of cells containing progesterone receptor O’Keefe, J.A., Li, Y., Burgess, L.H. and Handa, R.J. (1995)
mRNA in the female rat di- and telencephalon: an in situ Estrogen receptor mRNA alterations in the developing rat
hybridization study. Brain Res. Mol. Brain Res., 14(3): hippocampus. Brain Res. Mol. Brain Res., 30(1): 115–124.
239–249. Kerr, J.E., Allore, R.J., Beck, S.G. and Handa, R.J. (1995)
Hajszan, T. and MacLusky, N.J. (2006) Neurologic links Distribution and hormonal regulation of androgen receptor
between epilepsy and depression in women: is hippocampal (AR) and AR messenger ribonucleic acid in the rat hippo-
neuroplasticity the key? Neurology, 66(6 Suppl 3): S13–S22. campus. Endocrinology, 136(8): 3213–3221.
413

Kim, M.T., Soussou, W., Gholmieh, G., Ahuja, A., Tanguay, proliferation in vitro via muscarinic receptor activation
A., Berger, T.W. and Brinton, R.D. (2006) 17beta-Estradiol and MAP kinase phosphorylation. Eur. J. Neurosci., 12(4):
potentiates field excitatory postsynaptic potentials within 1227–1240.
each subfield of the hippocampus with greatest potentiation MacLusky, N.J., Hajszan, T., Prange-Kiel, J. and Leranth, C.
of the associational/commissural afferents of CA3. Neuro- (2006) Androgen modulation of hippocampal synaptic plas-
science, 141(1): 391–406. ticity. Neuroscience, 138(3): 957–965.
Korol, D.L. and Kolo, L.L. (2002) Estrogen-induced changes MacLusky, N.J., Luine, V.N., Hajszan, T. and Leranth, C.
in place and response learning in young adult female rats. (2005) The 17alpha and 17beta isomers of estradiol both in-
Behav. Neurosci., 116(3): 411–420. duce rapid spine synapse formation in the CA1 hippocampal
Lam, T.T. and Leranth, C. (2003) Gonadal hormones act subfield of ovariectomized female rats. Endocrinology,
extrinsic to the hippocampus to influence the density of hip- 146(1): 287–293.
pocampal astroglial processes. Neuroscience, 116(2): Margerison, J.H. and Corsellis, J.A. (1966) Epilepsy and the
491–498. temporal lobes. A clinical, electroencephalographic and ne-
Le Saux, M. and Di Paolo, T. (2005) Changes in 5-HT1A uropathological study of the brain in epilepsy, with particular
receptor binding and G-protein activation in the rat brain reference to the temporal lobes. Brain, 89(3): 499–530.
after estrogen treatment: comparison with tamoxifen and Mazzucco, C.A., Lieblich, S.E., Bingham, B.I., Williamson,
raloxifene. J. Psychiatry Neurosci., 30(2): 110–117. M.A., Viau, V. and Galea, L.A. (2006) Both estrogen recep-
Lei, D.L., Long, J.M., Hengemihle, J., O’Neill, J., Manaye, tor alpha and estrogen receptor beta agonists enhance cell
K.F., Ingram, D.K. and Mouton, P.R. (2003) Effects of est- proliferation in the dentate gyrus of adult female rats.
rogen and raloxifene on neuroglia number and morphology Neuroscience, 141: 1793–1800.
in the hippocampus of aged female mice. Neuroscience, McEwen, B.S. (2003) Mood disorders and allostatic load. Biol.
121(3): 659–666. Psychiatry, 54(3): 200–207.
Leranth, C., Shanabrough, M. and Horvath, T.L. (1999) McEwen, B.S. and Alves, S.E. (1999) Estrogen actions in the
Estrogen receptor-alpha in the raphe serotonergic and sup- central nervous system. Endocr. Rev., 20(3): 279–307.
ramammillary area calretinin-containing neurons of the Milner, T.A., Ayoola, K., Drake, C.T., Herrick, S.P., Tabori,
female rat. Exp. Brain Res., 128(3): 417–420. N.E., McEwen, B.S., Warrier, S. and Alves, S.E. (2005)
Li, C., Brake, W.G., Romeo, R.D., Dunlop, J.C., Gordon, M., Ultrastructural localization of estrogen receptor beta
Buzescu, R., Magarinos, A.M., Allen, P.B., Greengard, P., immunoreactivity in the rat hippocampal formation. J.
Luine, V. and McEwen, B.S. (2004) Estrogen alters hippo- Comp. Neurol., 491(2): 81–95.
campal dendritic spine shape and enhances synaptic protein Milner, T.A., McEwen, B.S., Hayashi, S., Li, C.J., Reagan, L.P.
immunoreactivity and spatial memory in female mice. Proc. and Alves, S.E. (2001) Ultrastructural evidence that hippo-
Natl. Acad. Sci. U.S.A., 101(7): 2185–2190. campal alpha estrogen receptors are located at extranuclear
Li, X., Schwartz, P.E. and Rissman, E.F. (1997) Distribution of sites. J. Comp. Neurol., 429(3): 355–371.
estrogen receptor-beta-like immunoreactivity in rat fore- Miranda, P., Williams, C.L. and Einstein, G. (1999) Granule
brain. Neuroendocrinology, 66(2): 63–67. cells in aging rats are sexually dimorphic in their response to
Liu, Z., Gastard, M., Verina, T., Bora, S., Mouton, P.R. and estradiol. J. Neurosci., 19(9): 3316–3325.
Koliatsos, V.E. (2001) Estrogens modulate experimentally Mitra, S.W., Hoskin, E., Yudkovitz, J., Pear, L., Wilkinson,
induced apoptosis of granule cells in the adult hippocampus. H.A., Hayashi, S., Pfaff, D.W., Ogawa, S., Rohrer, S.P.,
J. Comp. Neurol., 441(1): 1–8. Schaeffer, J.M., McEwen, B.S. and Alves, S.E. (2003)
Louissaint, A.J., Rao, S., Leventhal, C. and Goldman, S.A. Immunolocalization of estrogen receptor beta in the mouse
(2002) Coordinated interaction of neurogenesis and brain: comparison with estrogen receptor alpha. Endocrino-
angiogenesis in the adult songbird brain. Neuron, 34(6): logy, 144(5): 2055–2067.
945–960. Mohapel, P., Leanza, G., Kokaia, M. and Lindvall, O. (2005)
Loy, R., Gerlach, J.L. and McEwen, B.S. (1988) Autoradio- Forebrain acetylcholine regulates adult hippocampal neuro-
graphic localization of estradiol-binding neurons in the rat genesis and learning. Neurobiol. Aging, 26(6): 939–946.
hippocampal formation and entorhinal cortex. Brain Res., Morse, J.K., Scheff, S.W. and DeKosky, S.T. (1986) Gonadal
467(2): 245–251. steroids influence axon sprouting in the hippocampal dentate
Luquin, S., Naftolin, F. and Garcia-Segura, L.M. (1993) Nat- gyrus: a sexually dimorphic response. Exp. Neurol., 94(3):
ural fluctuation and gonadal hormone regulation of astrocyte 649–658.
immunoreactivity in dentate gyrus. J. Neurobiol., 24(7): Naftolin, F., Ryan, K.J., Davies, I.J., Reddy, V.V., Flores, F.,
913–924. Petro, Z., Kuhn, M., White, R.J., Takaoka, Y. and Wolin, L.
Lynch, M.A. (2004) Long-term potentiation and memory. (1975) The formation of estrogens by central neuroendocrine
Physiol. Rev., 84(1): 87–136. tissues. Recent Prog. Horm. Res., 31: 295–319.
Ma, W., Maric, D., Li, B.S., Hu, Q., Andreadis, J.D., Grant, Nakamura, N., Fujita, H. and Kawata, M. (2002) Effects of
G.M., Liu, Q.Y., Shaffer, K.M., Chang, Y.H., Zhang, L., gonadectomy on immunoreactivity for choline acetyltransf-
Pancrazio, J.J., Pant, H.C., Stenger, D.A. and Barker, J.L. erase in the cortex, hippocampus, and basal forebrain of
(2000) Acetylcholine stimulates cortical precursor cell adult male rats. Neuroscience, 109(3): 473–485.
414

Nakamura, N.H. and McEwen, B.S. (2005) Changes in intern- Santarelli, L., Saxe, M., Gross, C., Surget, A., Battaglia, F.,
euronal phenotypes regulated by estradiol in the adult rat Dulawa, S., Weisstaub, N., Lee, J., Duman, R., Arancio, O.,
hippocampus: a potential role for neuropeptide Y. Neuro- Belzung, C. and Hen, R. (2003) Requirement of hippocampal
science, 136(1): 357–369. neurogenesis for the behavioral effects of antidepressants.
Orchinik, M., Weiland, N.G. and McEwen, B.S. (1995) Chronic Science, 301(5634): 805–809.
exposure to stress levels of corticosterone alters GABAA Sar, M., Lubahn, D.B., French, F.S. and Wilson, E.M. (1990)
receptor subunit mRNA levels in rat hippocampus. Brain Immunohistochemical localization of the androgen receptor
Res. Mol. Brain Res., 34(1): 29–37. in rat and human tissues. Endocrinology, 127(6): 3180–3186.
Ormerod, B.K., Lee, T.T. and Galea, L.A. (2003) Estradiol Saravia, F.E., Beauquis, J., Revsin, Y., Homo-Delarche, F.,
initially enhances but subsequently suppresses (via adrenal de Kloet, E.R. and De Nicola, A.F. (2006) Hippocampal
steroids) granule cell proliferation in the dentate gyrus of neuropathology of diabetes mellitus is relieved by estrogen
adult female rats. J. Neurobiol., 55(2): 247–260. treatment. Cell. Mol. Neurobiol., 26: 943–957.
Parducz, A. and Garcia-Segura, L.M. (1993) Sexual differences Scharfman, H.E., Mercurio, T.C., Goodman, J.H., Wilson,
in the synaptic connectivity in the rat dentate gyrus. Neuro- M.A. and MacLusky, N.J. (2003) Hippocampal excitability
sci. Lett., 161(1): 53–56. increases during the estrous cycle in the rat: a potential role
Parducz, A., Hajszan, T., Maclusky, N.J., Hoyk, Z., Csakvari, for brain-derived neurotrophic factor. J. Neurosci., 23(37):
E., Kurunczi, A., Prange-Kiel, J. and Leranth, C. (2006) 11641–11652.
Synaptic remodeling induced by gonadal hormones: neuronal Schumacher, M., Coirini, H. and McEwen, B.S. (1989) Regu-
plasticity as a mediator of neuroendocrine and behavioral lation of high-affinity GABAA receptors in the dorsal hip-
responses to steroids. Neuroscience, 138(3): 977–985. pocampus by estradiol and progesterone. Brain Res., 487(1):
Perez-Martin, M., Azcoitia, I., Trejo, J.L., Sierra, A. and 178–183.
Garcia-Segura, L.M. (2003) An antagonist of estrogen Seidman, S.N. (2003) The aging male: androgens, erectile dysfunc-
receptors blocks the induction of adult neurogenesis by tion, and depression. J. Clin. Psychiatry, 64(Suppl 10): 31–37.
insulin-like growth factor-I in the dentate gyrus of adult Shughrue, P.J., Lane, M.V. and Merchenthaler, I. (1997) Com-
female rat. Eur. J. Neurosci., 18(4): 923–930. parative distribution of estrogen receptor-alpha and -beta
Perez-Martin, M., Salazar, V., Castillo, C., Ariznavarreta, C., mRNA in the rat central nervous system. J. Comp. Neurol.,
Azcoitia, I., Garcia-Segura, L.M. and Tresguerres, J.A. 388(4): 507–525.
(2005) Estradiol and soy extract increase the production of Shughrue, P.J., Scrimo, P.J. and Merchenthaler, I. (2000) Est-
new cells in the dentate gyrus of old rats. Exp. Gerontol., rogen binding and estrogen receptor characterization (ERal-
40(5): 450–453. pha and ERbeta) in the cholinergic neurons of the rat basal
Perlman, W.R., Tomaskovic-Crook, E., Montague, D.M., forebrain. Neuroscience, 96(1): 41–49.
Webster, M.J., Rubinow, D.R., Kleinman, J.E. and Weickert, Simerly, R.B., Chang, C., Muramatsu, M. and Swanson, L.W.
C.S. (2005) Alteration in estrogen receptor alpha mRNA levels (1990) Distribution of androgen and estrogen receptor
in frontal cortex and hippocampus of patients with major mRNA-containing cells in the rat brain: an in situ hybrid-
mental illness. Biol. Psychiatry, 58(10): 812–824. ization study. J. Comp. Neurol., 294(1): 76–95.
Picazo, O., Azcoitia, I. and Garcia-Segura, L.M. (2003) Sohrabji, F., Miranda, R.C. and Toran-Allerand, C.D. (1995)
Neuroprotective and neurotoxic effects of estrogens. Brain Identification of a putative estrogen response element in the
Res., 990(1–2): 20–27. gene encoding brain-derived neurotrophic factor. Proc. Natl.
Rhodes, M.E., McCormick, C.M. and Frye, C.A. (2004) Acad. Sci. U.S.A., 92(24): 11110–11114.
3alpha,5alpha-THP mediates progestins’ effects to protect Solum, D.T. and Handa, R.J. (2001) Localization of estrogen
against adrenalectomy-induced cell death in the dentate gyrus receptor alpha (ER alpha) in pyramidal neurons of the de-
of female and male rats. Pharmacol. Biochem. Behav., 78(3): veloping rat hippocampus. Brain Res. Dev. Brain Res.,
505–512. 128(2): 165–175.
Roof, R.L. (1993) The dentate gyrus is sexually dimorphic in Sorra, K.E. and Harris, K.M. (2000) Overview on the structure,
prepubescent rats: testosterone plays a significant role. Brain composition, function, development, and plasticity of hippo-
Res., 610(1): 148–151. campal dendritic spines. Hippocampus, 10(5): 501–511.
Roof, R.L. and Havens, M.D. (1992) Testosterone improves Steffensen, S.C. (1995) Dehydroepiandrosterone sulfate sup-
maze performance and induces development of a male hip- presses hippocampal recurrent inhibition and synchronizes neu-
pocampus in females. Brain Res., 572(1–2): 310–313. ronal activity to theta rhythm. Hippocampus, 5(4): 320–328.
Rudick, C.N., Gibbs, R.B. and Woolley, C.S. (2003) A role for Steiner, M., Dunn, E. and Born, L. (2003) Hormones and
the basal forebrain cholinergic system in estrogen-induced mood: from menarche to menopause and beyond. J. Affect.
disinhibition of hippocampal pyramidal cells. J. Neurosci., Disord., 74(1): 67–83.
23(11): 4479–4490. Stone, D.J., Rozovsky, I., Morgan, T.E., Anderson, C.P. and
Rune, G.M., Lohse, C., Prange-Kiel, J., Fester, L. and Frotsc- Finch, C.E. (1998a) Increased synaptic sprouting in response
her, M. (2006) Synaptic plasticity in the hippocampus: effects to estrogen via an apolipoprotein E-dependent mechanism:
of estrogen from the gonads or hippocampus? Neurochem. implications for Alzheimer’s disease. J. Neurosci., 18(9):
Res., 31(2): 145–155. 3180–3185.
415

Stone, D.J., Rozovsky, I., Morgan, T.E., Anderson, C.P., Velisek, L. and Veliskova, J. (2002) Estrogen treatment protects
Lopez, L.M., Shick, J. and Finch, C.E. (2000) Effects of age GABA(B) inhibition in the dentate gyrus of female rats after
on gene expression during estrogen-induced synaptic sprout- kainic acid-induced status epilepticus. Epilepsia, 43(Suppl 5):
ing in the female rat. Exp. Neurol., 165(1): 46–57. 146–151.
Stone, D.J., Song, Y., Anderson, C.P., Krohn, K.K., Finch, Veliskova, J. (2006) The role of estrogens in seizures and ep-
C.E. and Rozovsky, I. (1998b) Bidirectional transcription ilepsy: the bad guys or the good guys? Neuroscience, 138(3):
regulation of glial fibrillary acidic protein by estradiol in vivo 837–844.
and in vitro. Endocrinology, 139(7): 3202–3209. Veliskova, J., Velisek, L., Galanopoulou, A.S. and Sperber, E.F.
Szymczak, S., Kalita, K., Jaworski, J., Mioduszewska, B., (2000) Neuroprotective effects of estrogens on hippocampal
Savonenko, A., Markowska, A., Merchenthaler, I. and cells in adult female rats after status epilepticus. Epilepsia,
Kaczmarek, L. (2006) Increased estrogen receptor beta ex- 41(Suppl 6): S30–S35.
pression correlates with decreased spine formation in the rat Waters, E.M., Herrick, S.P., McEwen, B.S. and Milner, T.A.
hippocampus. Hippocampus, 16(5): 453–463. (2005) Subcellular localization of progestin receptors in the
Tabibnia, G., Cooke, B.M. and Breedlove, S.M. (1999) Sex rat hippocampus. Society for Neuroscience Abstracts,
difference and laterality in the volume of mouse dentate gyrus pp. 403–412.
granule cell layer. Brain Res., 827(1–2): 41–45. Wehrenberg, U., Prange-Kiel, J. and Rune, G.M. (2001) Ster-
Tabori, N.E., Stewart, L.S., Znamensky, V., Romeo, R.D., oidogenic factor-1 expression in marmoset and rat hippo-
Alves, S.E., McEwen, B.S. and Milner, T.A. (2005) Ultra- campus: co-localization with StAR and aromatase. J.
structural evidence that androgen receptors are located at Neurochem., 76(6): 1879–1886.
extranuclear sites in the rat hippocampal formation. Neuro- Weiland, N.G. (1992) Estradiol selectively regulates agonist
science, 130(1): 151–163. binding sites on the N-methyl-D-aspartate receptor complex
Tanapat, P., Hastings, N.B. and Gould, E. (2005) Ovarian in the CA1 region of the hippocampus. Endocrinology,
steroids influence cell proliferation in the dentate gyrus of 131(2): 662–668.
the adult female rat in a dose- and time-dependent manner. Weiland, N.G. and Orchinik, M. (1995) Specific subunit
J. Comp. Neurol., 481(3): 252–265. mRNAs of the GABAA receptor are regulated by proges-
Tanapat, P., Hastings, N.B., Reeves, A.J. and Gould, E. (1999) terone in subfields of the hippocampus. Brain Res. Mol.
Estrogen stimulates a transient increase in the number of new Brain Res., 32(2): 271–278.
neurons in the dentate gyrus of the adult female rat. J. Ne- Weiland, N.G., Orikasa, C., Hayashi, S. and McEwen, B.S.
urosci., 19(14): 5792–5801. (1997) Distribution and hormone regulation of estrogen
Teter, B., Harris-White, M.E., Frautschy, S.A. and Cole, G.M. receptor immunoreactive cells in the hippocampus of male
(1999) Role of apolipoprotein E and estrogen in mossy fiber and female rats. J. Comp. Neurol., 388(4): 603–612.
sprouting in hippocampal slice cultures. Neuroscience, 91(3): Wimer, C.C. and Wimer, R.E. (1989) On the sources of strain
1009–1016. and sex differences in granule cell number in the dentate area
Towart, L.A., Alves, S.E., Znamensky, V., Hayashi, S., McE- of house mice. Brain Res. Dev. Brain Res., 48(2): 167–176.
wen, B.S. and Milner, T.A. (2003) Subcellular relationships Woolley, C.S., Gould, E., Frankfurt, M. and McEwen, B.S.
between cholinergic terminals and estrogen receptor-alpha in (1990) Naturally occurring fluctuation in dendritic spine den-
the dorsal hippocampus. J. Comp. Neurol., 463(4): 390–401. sity on adult hippocampal pyramidal neurons. J. Neurosci.,
Veiga, S., Garcia-Segura, L.M. and Azcoitia, I. (2003) Neuro- 10(12): 4035–4039.
protection by the steroids pregnenolone and dehydroepiand- Woolley, C.S. and McEwen, B.S. (1992) Estradiol mediates
rosterone is mediated by the enzyme aromatase. J. fluctuation in hippocampal synapse density during the est-
Neurobiol., 56(4): 398–406. rous cycle in the adult rat. J. Neurosci., 12(7): 2549–2554.
Plate 23.2. Subcellular localization of estrogen (ER), androgen (AR), and progestin (PR) receptors in the dentate gyrus. A subset of
GABAergic interneurons contains nuclear ERa (dark pink). Granule cells, newly born cells (identified by DCX) and some GABAergic
interneurons contain cytosolic and plasma membrane-associated ERb (blue). Dendritic spines, many originating from granule cells
contain ERa, ERb, AR (dark green), and PR (purple). A few dendritic spines in the hilus, likely originating from mossy cells, contain
ERa and ERb. ERa, ERb, AR, and PR are found in axons and axon terminals. Some ERa-containing terminals are cholinergic
(acetylcholine, orange); some ERb-containing terminals resemble monoaminergic boutons. Lot of astrocytes (stars), mostly in the
molecular layer, also contain ERa, ERb, AR, and PR. (For B/W version, see page 407 in the volume.)
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 24

Plastic processes in the dentate gyrus:


a computational perspective

Brian E. Derrick

Department of Biology, The Cajal Neuroscience Research Institute, The University of Texas at San Antonio,
6900 N. Loop 1604 West, San Antonio, TX 78249-0662, USA

Abstract: The dentate gyrus has the capacity for numerous types of synaptic plasticity that use diverse
mechanisms and are thought essential for the storage of information in the hippocampus. Here we review
the various forms of synaptic plasticity that involve afferents and efferents of the dentate gyrus, and, from a
computational perspective, relate how these plastic processes might contribute to sparse, orthogonal en-
coding, and the selective recall of information within the hippocampus.

Keywords: dentate gyrus; LTP; LTD; metaplasticity; pattern separation

Plastic processes of the dentate gyrus hippocampal formation, has rekindled interest in
the function of the dentate gyrus, as well as indi-
The dentate gyrus occupies an unusual place in cating a mechanism of plasticity besides LTP and
both the history and anatomy of the hippocampal long-term depression (LTD) that also might con-
formation. Historically, it was the center of excite- tribute to memory storage (Shors and Matzel,
ment, being the site where the phenomenon of long- 1997).
term potentiation (LTP), an activity-dependent in- Anatomically, the dentate occupies an unusual
crease in synaptic efficacy, was first identified (Bliss place in the hippocampal formation. It is not part
and Gardner-Medwin, 1973; Bliss and Lomo, of the hippocampus (the dentate gyrus, the hip-
1973). However, in terms of synaptic plasticity, pocampus proper — areas CA1–CA3 — together
the dentate gyrus has been relegated to ‘‘second with the hilar region — are structures of the ‘‘hip-
place’’ behind the CA1 region, upon which the bulk pocampal formation’’). In terms of evolution, it is
of studies LTP have been focused for the past two a recent structure, attaining its present form with
decades. Recently, the dentate has again become a the emergence of mammals, the structure itself
focus of much attention in that it is one of the few appears to have been almost an afterthought,
neural sites that display neurogenesis after devel- tagged on to the hippocampus relatively late in
opment — dentate granule cells are continually evolution (Treves and Roudi, 2005). From the
generated well into adulthood (Altman and Das, perspective of hippocampal function, here it has
1965; Bayer, 1982). This rather remarkable process, remained — simply the first relay of a three-part
which is unique to the dentate gyrus of the ‘‘trisynaptic circuit,’’ where it has been suggested
to act as a ‘‘gate’’ to the hippocampus due prima-
Corresponding author. Tel.: +1 210 458 4273; rily to its modulation of output with behavioral
Fax: +1 210 458 5658; E-mail: bderrick@utsa.edu state (Winson and Abzug, 1978; Bramham and

DOI: 10.1016/S0079-6123(07)63024-6 417


418

Srebro, 1989). However, we know now that infor- archicortex (it is a surprise to many that the hip-
mation flow through the trisynaptic circuit may be pocampus is actually cortex, albeit the evolution-
only one (and not even the primary) pathway of ary older three-layered archicortex; Amaral and
input flow through the hippocampus (Mizumori et Witter, 1989). The principal cell of the dentate
al., 1989; Yeckel and Berger, 1990; Do et al., gyrus is the granule cell. In the rat, the dentate
2002). Such findings offer new clues that have re- contains 1–2 million granule cells (Boss et al.,
defined the function and operation of both the 1985), a very large number of cells given that cor-
dentate gyrus and the hippocampus. tical input to the dentate (300,000–400,000 cells
More surprising is that the dentate displays nu- of the entorhinal cortex in the rat) is, by compar-
merous forms of neural plasticity. In addition to ison, relatively small. The projection of a structure
the phenomenon of LTP, other plastic changes are to one with a larger number of neurons, referred to
observed at synapses to and from the dentate, such as ‘‘fan out,’’ ‘‘input expansion,’’ or ‘‘input recod-
as LTD and metaplasticity. In addition, long-term ing’’ (Rolls and Treves, 1998), offers clues to the
changes in granule cell function, characterized by function of the dentate. It is an aspect of the dent-
long-term alterations in neurotransmitter and ne- ate gyrus thought crucial for ‘‘pattern separation,’’
uromodulator synthesis (White et al., 1987; Hong a process now thought to be one of the primary
et al., 1988), the formation of new synaptic con- functions of the dentate, and essential for the for-
tacts on granule cell targets (mossy fiber synapto- mation of discrete ‘‘memories’’ within the hippo-
genesis, Ben-Ari and Represa, 1990; Cavazos et campus proper (Marr, 1971; Morris and
al., 1991), and the addition of new granule cells McNaughton, 1987; Treves and Rolls, 1992, 1994).
well into adulthood (adult granule cell neurogen- Classically, the dentate gyrus was first charac-
esis, Altman and Das, 1965) are other plastic terized as the initial stage of the three-stage ‘trisy-
processes displayed by the dentate gyrus that, to- naptic circuit’ of the hippocampal formation
gether, have the potential to contribute to accurate (Andersen et al., 1966). As such, the dentate gyrus
and efficient information storage within the hip- represents the first stage of this circuit and the
pocampal formation. Because many of these plas- primary target of highly processed cortical infor-
tic changes involve the expression of numerous mation relayed from the entorhinal cortex via the
genes, and result in long-lasting morphological main synaptic input to the dentate, the perforant
changes (such as synaptogenesis and neurogene- pathway. The subsequent second stage of the so-
sis), it seems odd that the one structure within the called ‘trisynaptic circuit’ is the output from the
hippocampal formation that shows the most pro- dentate granule cells, the mossy fibers. The mossy
nounced and sustained genetic, morphological, fibers then project to the most proximal dendritic
and functional plastic changes would be overshad- region of pyramidal cells within the CA3 region.
owed by the CA1 region of the hippocampus. The mossy fiber axons of the granule cells are rel-
The purpose of this review is to introduce the atively thin and unmyelinated, making them par-
myriad of plastic processes that are displayed by ticularly slow as compared with the third and final
synapses of the dentate gyrus, and to describe their stage of the trisynaptic circuit, the CA3 projection
possible contribution to information storage on the to CA1 mediated by CA3 ‘‘Schaffer collaterals.’’
basis of current information-theoretical views of in- Crucial to the more recent views of hippocampal
formation processing and memory storage that are operation is the finding that the primary input to
thought operative in the hippocampal formation. CA3 pyramidal cells is not from the dentate gyrus.
Rather, the primary synaptic input to CA3 py-
ramidal cells is recurrent input arising from other
The dentate gyrus: A brief description of anatomy CA3 pyramidal cells (Amaral et al., 1993). This
and architecture recurrent excitatory input (termed the commis-
sural/associational CA3 pathway) is of particular
As is the case for most of the hippocampal for- interest computationally, as such recurrent net-
mation, the dentate is essentially three-layered works are capable of performing autoassociative
419

Fig. 1. Schematic representation of the dentate–CA3 interface. Perforant path inputs to the CA3 region arrive directly via synapses on
the distal CA3 dendrites, and indirectly by a disynaptic relay from the dentate gyrus via the sparse but powerful mossy fiber–CA3
projections. Recurrent connections constitute the greatest input to CA3 pyramidal cells, and are thought to serve as the substrate for
autoassociative storage and recall. Adapted with permission from Amaral et al. (1990) and Rolls and Treves (1998).

function and thus represent a substrate for ‘‘con- of discharging pyramidal cells, but also display LTP
tent addressable’’ memories (or CAMs), a more (Do et al., 2002) as well as associative LTP (Marti-
general term for distributed memory devices nez et al., 2002), a feature crucial to many of the
(Kohonen, 1984). models of CA3 computation (Treves and Rolls,
The view of the hippocampal formation as a 1994). A more realistic ‘‘schematic’’ view of dent-
‘trisynaptic circuit’ is, however, an oversimplifica- ate–CA3 hippocampal circuitry is shown in Fig. 1.
tion of a much more elaborate and parallel input An underappreciated feature of the cortical input
system within the hippocampus (Yeckel and Berger, to the hippocampal formation is that not only can
1998). Anatomical studies indicate that there are the direct perforant path inputs activate CA3 py-
substantial direct perforant path projections to ramidal cells, but CA3 pyramidal cells are also the
both the CA3 and CA1 regions of the hippocam- first cells to respond to perforant path input (Yeckel
pus (Andersen et al., 1966; Amaral et al., 1990; and Berger, 1990; Do et al., 2002). Extracellular
Claiborne et al., 1993). The dentate and CA3 re- measures of the summed output of the CA3 region
gions receive direct input from the same popula- and the dentate gyrus, as reflected in population
tions of neurons in layers II of the medial and spike latency, indicate that the CA3 region responds
lateral entorhinal cortex (Amaral and Witter, 1989). faster than the dentate gyrus, with spike discharges
In addition, layer III neurons of the lateral and in vivo preceding granule cell spikes by 0.5–3 ms
medial entorhinal cortex project directly to apical (Fig. 2). Thus, it is the CA3 region, rather than the
dendrites of CA1 pyramidal cells and these projec- dentate, that is the first population of principal cells
tions to the CA1 region constitute a separate ‘‘tem- to respond to cortical input relayed by the perforant
poroammonic pathway.’’ These inputs, once pathway. As noted in the original study (Yeckel and
ignored and deemed insignificant, were recognized Berger, 1990), this finding is in line with a number of
as potentially important by neuroanatomists studies indicating early evoked responses in the CA3
(Hjorth-Simonsen and Jeune, 1972; Steward 1976; region that precede dentate responses (Segal, 1972).
Steward and Scoville, 1976; Buzsaki, 1988; Witter Because the CA3 region is the first region to respond
et al., 1988). Subsequent physiological studies in with spike output to perforant path activity, the
vivo (Yeckel and Berger, 1990; Berzhanskaya et al., primacy of CA3 responses is likely relevant to the
1998) demonstrated that the direct perforant path function of the rather unusual ‘‘backward’’ projec-
projections to the CA3 region are not only capable tions from CA3 pyramidal cells to hilar neurons
420

Fig. 2. Comparison of perforant path responses evoked in the dentate gyrus and CA3 region. Shown are field potentials recorded in
the dentate gyrus (hatched lines) or the CA3 pyramidal layer (solid lines) following activation of the medial (top) or lateral perforant
pathway (bottom traces). Responses were recorded in vivo from the same animals under pentobarbital anesthesia. Calibration: 0.5 mV,
5 ms (CA3); 1 mV, 5 ms (dentate).

(Scharfman, 1996). Because these interneurons can layer (Amaral and Witter, 1989). The dentate asso-
exert both excitatory and inhibitory actions on ciational projections reflect yet another recurrent
dentate granule cells, the projection from the CA3 excitatory pathway within the hippocampal forma-
region to these interneurons may serve to modulate tion, albeit one that is relayed disynaptically via
dentate granule cell output in response to direct mossy cells (Amaral, 1978; Amaral and Witter,
perforant path–CA3 input. 1995). The granule cells themselves display features
The granule cells of the dentate gyrus dentate that are unique as well: they have a high threshold
appear, by many accounts, unremarkable. They are for activation, and granule cell bursting occurs
small (6–10 mm) and have extensive, conical, spin- rarely in vivo (Jung and McNaughton, 1993), al-
ous dendrites that extend almost 1 mm into the though granule cells can fire sustained bursts during
stratum moleculare (or the ‘‘molecular layer’’) of theta rhythm (Munoz et al., 1990), a 4–12 Hz os-
the dentate (Steward, 1976; Amaral and Witter, cillation that is a prominent feature of hippocampal
1989). It is here in the molecular layer that the activity during exploration and, presumably, during
dentate gyrus receives its input from three primary learning (Buzsaki, 2002).
sources: the lateral entorhinal cortex (LEC), which Granule cells are notoriously silent, as evidenced
targets the outer one third of the dentate molecular by the difficulty in obtaining ‘‘place fields’’ —
layer; the medial entorhinal cortex (MEC), whose single unit firing that correlates with a specific
projections occupy the middle one third of the place field in an environment (O’Keefe and Nadel,
granule cell dendrites in the molecular layer; and the 1978) — and they often fire at very low rates
commissural/associational projections, originating (approximately 0.5 Hz, Jung and McNaughton,
from hilar glutamatergic ‘‘mossy’’ cells and termi- 1993). As discussed further below, these tight con-
nating in the innermost one third of the molecular straints on granule cell activity may be viewed as
421

another feature crucial for ‘‘pattern separation’’ connections of granule cells to CA3 are sparse,
and the formation of sparse outputs by the dentate with each granule cell making contact with only
gyrus (Skaggs and McNaughton, 1992). 8–10 CA3 pyramidal cells, they form unusual
Another unusual aspect of the dentate is that the synapses on CA3 cells, with each mossy fiber
V-shaped blades of this structure appear not to be bouton essentially engulfing a highly elaborated
homogeneous, actually differing in a number of dendritic CA3 spine with multiple synaptic con-
substantial ways. They appear to be under separate tacts, often referred to as ‘‘thorny excrescences’’
genetic control, as the lower (infrapyramidal) blade (Gonzales et al., 2001). The mossy fiber synapse
and the mossy fiber (suprapyramidal) layer differ itself is unusual in several respects, with the pri-
among strains of mice (Schwegler et al., 1990). In mary difference being that LTP observed at these
addition, mossy fiber efferents arising from granule synapses is of a type that does not require the ac-
cells in the internal (suprapyramidal) layer ramify tivation of N-methyl-D-aspartate (NMDA) recep-
and extend through the stratum lucidum of the tors (Harris and Cotman, 1986).
CA3 region, whereas the mossy fibers arising from In addition to the CA3 pyramidal cells, the
the outer (infrapyramidal) blade of the dentate mossy fibers also target numerous neurons within
ramify extensively in the hilus and terminate pri- the hilar region. The array of neuronal subtypes
marily in the infrapyramidal layer of the CA3 re- within the hilar region of the dentate is impressive
gion lying just below the CA3 pyramidal cell layer, (Amaral, 1978); however the most studied neurons
or stratum pyramidale (Claiborne et al., 1986). The in the hilus are the so-called ‘‘mossy cells.’’ These
two blades may also have functional differences, cells are innervated by the mossy fibers and, by
because the infrapyramidal region develops greater virtue of their excitatory glutamatergic projections,
mossy fiber sprouting following seizures or high- both ipsilaterally and contralaterally to the inner
frequency mossy fiber stimulation (Cavazos et al., molecular layer, comprise the commissural/
1991; Escobar et al., 1997; Scharfman et al., 2002), associational system of the dentate gyrus. This re-
as well as spatial learning (Ramirez-Amaya et al., current excitatory input has been scrutinized re-
2001). In support of possible ‘‘divisions of labor’’ cently (Kleschevnikov and Routtenberg, 2003), and
among the dentate blades, a greater degree of ex- studies have revealed that these projections show
citability is observed in the lower (infrapyramidal) plasticity both at the mossy fiber–mossy cell
blade, possibly due to differential innervation of synapse and at their contacts with granule cells in
granule cells by inhibitory interneurons (Scharfman the inner molecular layer. Those familiar with the
et al., 2002). Moreover, dendrites on granule cells in major target of the dentate granule cells, the CA3
the infrapyramidal layer show fewer spines (Clai- region, may notice a similarity between the dentate
borne et al., 1986). Interestingly, while both blades and CA3 region in that both have extensive recur-
show the expression of Arc, an immediate-early rent excitatory inputs that also project contralater-
gene associated with synaptic plasticity, following ally. Thus, both the dentate and CA3 region have
stimulation of the perforant path that induces LTP, extensive, bilateral, and recurrent excitatory feed-
Arc expression in behaving animals associated with back. Thus these two structures can be thought of
learning is seen only in granule cells in the supra- as a single, bilateral, highly re-entrant structure.
pyramidal (inner) blade lying between the hilus and The functional implications of this feature remain
the CA1 region (Chawla et al., 2005). However, the to be defined.
functional relevance of these differences among the
blades of the dentate gyrus is unknown.
The efferents of the granule cells, the mossy fib- Information/theoretic views of the dentate gyrus
ers, also display unusual features: these granule
cell efferents are thin, unmyelinated fibers that, Although the dentate gyrus defies any single ac-
upon their exit from granule cells, form many bi- count of function based on its operation (such as its
furcations, projecting primarily to the proximal early conceptualization as ‘‘gating’’ input to the
spines of CA3 pyramidal neurons. Although the CA3 region (Winson and Abzug, 1978), clues to its
422

functions may be provided by current information- become associated with subsequent CA3 inputs
theoretic models of hippocampal information over time (Levy, 1996). The view of autoassocia-
processing. To more fully appreciate the contribu- tive processes in the CA3 region, first formalized
tion of the dentate to hippocampus-based memory, by Marr (1971), remains a feature of most models
it is necessary to consider information-theoretic of hippocampal function (Morris and McNaugh-
models of information storage in the CA3 region of ton, 1987; Treves and Rolls, 1994; Hasselmo et al.,
the hippocampus. 1995; McClelland et al., 1995; Rolls and Treves,
The late David Marr’s original conceptualizat- 1998) and is supported by both models of the CA3
ion of hippocampal function remains central to region and behavioral studies (Nakazawa et al.,
most information-theoretic theories of hippocam- 2002).
pal operation today (Marr, 1971). As suggested by As noted above, the recurrent CA3–CA3 con-
Marr, the CA3 region may serve as an ‘‘autoasso- nections constitute the majority of excitatory in-
ciator’’ by virtue of its extensive recurrent puts to CA3 pyramidal cells, and these fibers also
(CA3–CA3) excitatory connections (Morris and display Hebbian LTP (Harris and Cotman, 1986).
McNaughton, 1987; Treves and Rolls, 1994; Rolls It is thought that representations within the CA3
and Treves, 1998). Neural network models of autoassociative system involve the formation of
autoassociators were pioneered by theoreticians attractor states (simply relatively stable ensembles
such as Tuevo Kohonen (Kohonen et al., 1977; of active, recurrently connected CA3 cells) as a
Kohonen, 1984) and later re-formalized by John result of plastic changes at the recurrent CA3–CA3
Hopfield (Hopfield, 1982). In such models auto- synapses. These plastic changes are thought to oc-
associators are networks which by virtue of their cur by potentiation of active CA3–CA3 synapses
extensive recurrent input are capable of unique whose activity is initiated by the direct perforant
computational operations. Representations in au- path–CA3 inputs, and ‘‘fixed’’ by strong postsy-
toassociative systems involve the formation of naptic activity initiated by the mossy fiber
‘‘attractors’’ (or stable ensembles of active, recur- synapses from dentate granule cells. In this
rently connected cells as a result of plastic changes scheme, the mossy fiber synapses are thought to
among recurrent synapses). Thus autoassociative act as ‘‘detonator’’ synapses (Morris and McN-
architectures and their recurrent connections can aughton, 1987) that direct associative plasticity,
allow for the formation of stable ‘‘attractor’’ and thus encoding, in this system. Due to the rel-
states. This architecture also can enable the recall ative proximity of these synapses to the CA3 py-
of an entire pattern of CA3 activity in response to ramidal cell body, activation of these ‘‘detonator’’
the presentation of only a subset of the original synapses can initiate CA3 cell firing and, via
pattern. This feature, termed ‘‘pattern comple- postsynaptic depolarization, induce LTP among
tion,’’ is a fundamental computational feature of other synapses to the CA3 pyramidal cell that were
autoassociators (Kohonen et al., 1977). If one recently active (Levy and Steward, 1979, 1983).
thinks of an ensemble of many recurrently con- Thus, encoding of information in the CA3 region
nected CA3 neurons with strengthened synapses is thought to result from the activation of CA3
among them as a representation, one can see that recurrent collaterals by perforant path input and
the activation of a subset of this ensemble could, the subsequent potentiation of recently active per-
via recurrent activation of other CA3 neurons via forant path–CA3 and CA3–CA3 synapses by
strengthened synapses, recruit the entire original mossy fiber ‘‘detonator synapses.’’
ensemble. Thus recurrent inputs provide the subst- Although autoassociators are capable of pattern
rate of CA3 ‘‘autoassociation’’ and pattern com- completion given a partial input of the original
pletion from partial inputs, a property that may pattern, they have their limitations. Crucial for
encode not only single stable ensembles of co-ac- efficient encoding and accurate recall in autoasso-
tive CA3 pyramidal cells, but also sequences of ciative systems is a sufficient separation of patterns
stable states, provided that the output from the that are encoded within the CA3 region, such that
CA3 system can, via the CA3 recurrent inputs, stored representations in the CA3 system have
423

minimal overlap with other, previously encoded consists of 1–2 million granule cells that receive
ensembles (orthogonalization). This separation of input from only 300,000 cells in the entorhinal
patterns is necessary because accurate encoding cortex (Boss et al., 1985). This ‘‘fan out’’ of cor-
and retrieval within autoassociative memory sys- tical input to a structure with many more elements
tems can occur only if correlations and redundan- is one way to enforce pattern separation (referred
cies among discrete attractor states (ensembles of to as ‘‘input expansion’’ or ‘‘expansion recoding;’’
active CA3 neurons) are reduced (see Rolls and Rolls and Treves, 1998). In addition, both the rel-
Treves, 1998). Thus, encoding information in an ative quiescence of granule cells in the dentate and
autoassociator demands special considerations the sparse connectivity of the mossy fiber projec-
and requires the formation of orthogonal, or un- tions (so that each granule cell forms mossy fiber
correlated, ensembles of CA3 attractors. How this contacts with only 10–15 CA3 pyramidal cells;
is done is the subject of much speculation, al- Amaral et al., 1990) together make it unlikely that
though a feature common to many models is that any given CA3 pyramidal cell can be activated by
orthogonalization (sometimes referred to more more than one granule cell. Thus the anatomically
generally as ‘‘pattern separation’’) requires sparse sparse dentate–CA3 connectivity by the mossy fib-
encoding, a process that is crucial for encoding in ers is an important feature for the formation of
the CA3 region. This sparsification is accom- sparse, orthogonal CA3 representations. Together,
plished primarily by transformation of perforant these aspects of the dentate are ideally suited for
path inputs to the dentate into a sparse code that is pattern separation, and the formation of sparse,
then relayed to the CA3 region (Morris and McN- orthogonal outputs to the CA3 region that allow
aughton, 1987). The formation of sparse dentate the encoding of discrete, stable ensembles or ‘‘at-
outputs can increase the capacity of a recurrent tractors’’ in the CA3 autoassociative system.
autoassociative system, and by virtue of a fewer While important, the anatomical features of the
number of outputs (and plasticity among them) mossy fiber output are but one way the dentate
can also serve to reduce any similarities among may relay sparse transforms to the CA3 autoas-
ensembles formed within the CA3 region. Sparse sociative system. As we will discuss later, when
encoding can serve as one process that minimizes taken in a computational context, the plastic fea-
similarity and interference among CA3 attractors, tures of the dentate gyrus, including LTP, the var-
because representations using a minimal number ious forms of LTD, and metaplastic processes,
of elements are less likely to overlap with other together, not only may ensure sparse, orthogonal
stored patterns (Rolls and Treves, 1998). outputs to the CA3 autoassociative system, but
In many models of hippocampal function, the also may allow for the dentate to encode discrete
task of ‘‘pattern separation’’ and orthogonalizat- attractors for similar, or even identical inputs that
ion of cortical inputs relayed to the CA3 region is are encoded over time. In this view, the combina-
ascribed to the dentate gyrus (Morris and McN- tion of plastic processes in the dentate that allow
aughton, 1987; Treves and Rolls, 1994; Rolls and the formation of sparse, orthogonal CA3 inputs
Treves, 1998). The dentate gyrus can serve in the also may allow partial inputs to converge on dis-
formation of sparse, orthogonal CA3 ensembles tinct stable states encoded by very similar inputs
via several mechanisms. As discussed extensively over time — a feature characteristic of episodic
by Rolls and Treves (1998), the dentate gyrus is memory, long thought of as the principal type
thought to form sparse outputs to the CA3 region, of memory involving the human hippocampus
a process that can be facilitated in a variety of (Tulving, 1983).
ways by features of the dentate gyrus. Foremost
among these features are the anatomical charac-
teristics of the dentate gyrus–mossy fiber system. Synaptic plasticity in the dentate gyrus — LTP
This system appears ideally suited to ensure the
formation of sparse, uncorrelated input patterns Foremost among the plastic processes studied in
relayed to the CA3 region. The rat dentate gyrus the dentate gyrus is LTP. The hypothesis that
424

Fig. 3. Perforant path–dentate LTP and heterosynaptic LTD. Plot of field EPSP slopes of medial (K) and lateral (J) perforant
path–dentate responses following high-frequency stimulation of the medial perforant pathway. In addition to the potentiation ob-
served at medial perforant path synapses, a sustained depression of inactive lateral perforant path synaptic responses (‘‘heterosynaptic
LTD’’) is observed. Calibration: 0.5 mV, 5 ms. Villarreal and Derrick (2006).

memory involves changes in synaptic efficacy has This synapse specificity is in contrast to the non-
remained a primary assumption for quite some associative phenomena, such as sensitization,
time (Ramon y Cajal, 1937; Hebb, 1949; James, wherein a general increase in responsivity of the
1890). LTP is a rapidly induced and long-lasting postsynaptic cell is observed. Importantly, input
change in synaptic strength that is observed fol- specificity is a feature essential for any kind of
lowing high-frequency afferent activity. LTP was memory that requires both high capacity and high
first described formally in studies of the rabbit fidelity, as this feature provides sparsity by in-
perforant path projections to the dentate gyrus creasing the number of modifiable inputs that can
using an in vivo preparation (Bliss and Lomo, aid in encoding and retrieval in autoassociative
1973), and the first studies that addressed LTP systems.
longevity also were conducted using perforant Another crucial feature of LTP making it an
path–dentate recordings (Barnes, 1979). However, attractive potential mechanism for information
since its discovery, and the subsequent discovery storage is that it is remarkably persistent. Prior to
of similar plastic processes in other brain struc- the discovery of LTP, the only activity-dependent
tures (from the amygdala to spinal cord), the bulk electrophysiological change of substantial dura-
of studies of LTP, particularly in vitro, have fo- tion was post-tetanic potentiation (PTP), a prima-
cused on the Schaffer–CA1 synapse (Fig. 3). rily presynaptic phenomenon lasting from seconds
LTP has features that make it a compelling to minutes (Zucker and Regehr, 2002). By con-
candidate mechanism that may contribute to in- trast, in the hippocampal formation, LTP can per-
formation storage. LTP is induced rapidly after its sist from hours to weeks or months, depending on
induction, which, in the laboratory, requires a the stimulation parameters. In an intact animal,
sufficient postsynaptic depolarization that usually LTP usually decays within several weeks (Barnes,
is provided by high frequency stimulation of a 1979).
sufficient number of afferents. Dentate LTP also is A number of investigators have dismissed LTP
‘‘input specific,’’ that is, LTP is confined to as a long-term memory mechanism because its
synapses of afferents activated with stimulation. duration is far too short to account for memory
425

that lasts years (Shors and Matzel, 1997). While Mody, 2003). Consequently, background synaptic
plasticity that lasts only days to weeks is too brief activity completely unrelated to learning may be
to store long-term memory, the limits of LTP lon- sufficient to mediate LTP decay at perforant
gevity may be an important feature of LTP in the path–granule cell synapses, suggesting that synap-
hippocampal formation. One reason is that LTP, tic activity unrelated to learning also may contrib-
at least within the hippocampus, need not be per- ute to LTP decay. Regardless, because LTP decay
manent. This is because, as is thought to be the requires activation of NMDA receptors, LTP de-
case in humans (Scoville and Milner, 1957), the rat cay still reflects an active process (Villarreal et al.,
hippocampus also is thought to have a time-de- 2002). Thus perforant path–dentate LTP, at least
limited role in memory, with persistent, ‘‘long theoretically, is potentially permanent, and could
term’’ memory being gradually consolidated in last as long as the synapse itself, provided that it is
neocortical areas over time (Marr, 1971; Squire, not ‘‘erased’’ by subsequent synaptic activity and
1992). Currently, the most popular view is that NMDA receptor activation.
memories formed by the hippocampus are trans- LTP also is ‘‘associative.’’ Specifically, if activity
ferred to and consolidated in the neocortex, pos- in a set of afferents to a common postsynaptic
sibly during slow wave (NREM) sleep states target induces LTP, then individual active
(Wilson and McNaughton, 1993; Cartwright, synapses can be recruited to express LTP, pro-
2004), a process that usually requires 2–3 weeks vided that the synapse is coactive within a delim-
in the rat, as indicated by both lesion and imaging ited window of time (Levy and Steward, 1983).
data (Bontempi et al., 1999). Thus, if the hippo- The phenomenon of associativity is derived from
campus does indeed serve as a temporary repos- the requirements for activation of the NMDA re-
itory, or ‘‘buffer’’ for information, LTP may not ceptor (specifically, both presynaptic release of
need to persist for extended periods of time. In this glutamate and postsynaptic depolarization, with
scheme, any long-term retention of information the latter being essential for relieving the magne-
within the hippocampus beyond the a few weeks sium block of the NMDA channel). Associativity
might even be detrimental to hippocampal-based is a property derived directly from, and is essen-
memory, resulting in interference with previously tially identical to, the property of ‘‘cooperativity’’
stored patterns of synaptic activity (Rosenzweig (McNaughton et al., 1978). Cooperativity indi-
et al., 2002). cates LTP induction has a threshold, and activa-
How long, then, can LTP last? We reported tion of a threshold number of afferents is essential
that, in perforant path projections to the dentate to induce LTP. Because LTP is induced in a ‘‘co-
gyrus, LTP decay is an active process mediated by operative’’ manner, it follows that associativity is
NMDA receptors, and blocking these receptors an inherent property of ‘‘cooperativity.’’ Although
can prevent decay (Villarreal et al., 2002). There- it has been suggested that ‘‘cooperativity’’ and
fore it appears that LTP doesn’t simply ‘‘fade;’’ ‘‘associativity’’ are distinct, with cooperativity re-
instead, it is actively ‘‘erased’’ via processes that flecting a threshold of afferent activity necessary to
require NMDA receptor activation. A likely can- induce LTP and associativity involving distinct
didate mechanism for such ‘‘erasure’’ is LTD. As sets of afferents, it should be noted that the first
discussed more fully below, dentate LTD, like demonstration of cooperativity used two distinct
LTP, requires the activation of NMDA receptors afferent systems to the dentate gyrus, the medial
(Desmond et al., 1991; Christie and Abraham, and lateral perforant pathways (McNaughton et
1992b). Thus, NMDA receptor activation and a al., 1978). However, because interactions among
reversal of LTP by the induction of LTD (also active afferent fibers could not be ruled out, the
referred to as ‘‘depotentiation,’’ both discussed term cooperativity, rather than associativity, may
below) may mediate LTP decay. However, have been chosen out of rigor, as the term ass-
NMDA receptors on dentate granule cells can be ociativity, as envisaged by Hebb, tacitly implies a
activated by low frequency perforant path synaptic postsynaptic integration of presynaptic activity,
activity and even miniature EPSPs (Dalby and something that could not be ruled out with
426

‘‘cooperative’’ effects. It should be noted here that from medial perforant path afferents. This may be
these studies (McNaughton et al., 1978; Levy and due, at least in part, to the fact that the medial and
Steward, 1983) were conducted at a time when the lateral projections are stratified in the angular
NMDA receptor was unheard of. Subsequently, a bundle, so that only stimulation of fibers deep in
convincing demonstration of Hebbian-like ass- the angular bundle evokes lateral perforant path
ociativity was provided by studies showing that responses (McNaughton and Barnes, 1977). Lat-
pairing of low intensity, low frequency afferent eral perforant path–dentate responses differ from
activity with postsynaptic depolarization is capa- medial responses in a number of ways, including a
ble of inducing LTP (Gustafsson and Wigstrom, slower rising phase of dentate field EPSPs (due to
1986; Gustafsson et al., 1987). the more distal regions of the granule cell dendrite
It is often assumed that mechanisms underlying where lateral perforant path synapses terminate).
LTP can be generalized. However, what has In addition, medial perforant path responses show
emerged from recent studies is that the mechanisms a pronounced frequency depression of synaptic
underlying the induction, and possibly the mainte- responses, whereas lateral perforant path re-
nance and expression, of LTP not only differ at sponses show robust frequency facilitation (McN-
different synapses within the hippocampal forma- aughton and Barnes, 1977). The reasons for these
tion, but may even involve a number of distinct differences in afferent responses are not clear, al-
mechanisms, even at a single synapse. These differ- though it likely involves differences among the
ences can arise from numerous variables, including afferents themselves, as depression and facilitation
what stimulation parameters are used, and even the are also observed when paired pulses are delivered
animal’s behavioral state at the time of LTP induc- to medial and lateral perforant path projections to
tion (Davis et al., 2004; see below). Thus there may the CA3 region, (Do et al., 2002).
be a multitude of activity-dependent mechanisms It may come as no surprise that LTP in these
that serve to increase synaptic strength, all of which two pathways differ in a number of important
have been labeled ‘‘LTP.’’ Yet these processes may ways. The medial perforant pathway is a glut-
have distinct mechanisms of induction (conditions amatergic afferent system and displays LTP that
necessary to initiate LTP) and/or maintenance depends on activation of the NMDA glutamate
(mechanisms initiated by LTP induction that un- receptor. Thus LTP at medial perforant path
derlie its expression for sustained periods of time). synapses may reflect a form of LTP induction
Such differences in both induction and maintenance similar to the more widely-studied type at the
mechanisms may be indicative of distinct ‘‘forms’’ Schaffer–CA1 synapse, as well as among neocor-
of LTP. tical regions (Kirkwood et al., 1993; Trepel and
The heterogeneity and complexity of LTP are Racine, 1998), and appears to be the primary re-
illustrated by the features of LTP induction in the ceptor involved in the induction of LTD as well
primary input to the dentate — the perforant (Desmond et al., 1991; Christie and Abraham,
pathway. As noted above, the perforant pathway 1992b). NMDA receptors are activated by gluta-
projections originate from layer II neurons within mate released by the presynaptic terminal. NMDA
the medial and the lateral entorhinal cortices, and receptors are also voltage-dependent, such that
they target, respectively, the middle one third of current flow is prevented at hyperpolarized mem-
the molecular layer for medial perforant path brane potentials due to a cationic block of the
(MPP) afferents, and the outermost third of the channel by magnesium. However, the combination
molecular layer for the lateral perforant path of presynaptic activity with postsynaptic depolari-
(LPP) inputs (Steward, 1976; Amaral and Witter, zation sufficient to remove the cationic block of
1989). the NMDA receptor pore allows current flow (cal-
Responses to the medial perforant path projec- cium influx via NMDA receptor channels), allow-
tion are, for a variety of reasons, the most robust, ing for the induction of LTP or LTD. This
and many studies that employ stimulation of the property of NMDA receptors, the requirement
perforant path often are likely to evoke responses for both presynaptic activity together with a
427

sufficient level of postsynaptic activity, is essen- regulate the expression of other ‘‘late effector’’
tially a physical instantiation of the synaptic learn- genes. The major IEGs implicated in LTP include
ing rule formalized by Donald Hebb in the 1940s Arc (Steward et al., 1998), Homer (Kato et al.,
(Hebb, 1949). It is now referred to as the ‘‘Hebb 1997), and Zif268 (Davis et al., 2000; Jones et al.,
rule,’’ this simple local learning rule requiring cor- 2001). Arc is a very interesting IEG in that the
related pre- and postsynaptic activity has signifi- mRNA for this gene is transported into granule
cant computational power when employed in cells dendrites, where the synthesis of Arc proteins
various neural network architectures. Networks occurs (Steward et al., 1998; Guzowski et al.,
with elements that use the Hebb rule display im- 2000). Similarly, the transport of CaMKII
portant computational features, and therefore it is mRNAs to synaptic regions, and enhanced syn-
thought to be a principal ‘‘coincidence detector’’ thesis of CaMKII, also is observed (Havik et al.,
that mediates synaptic plasticity based on corre- 2003). Finally, there is a protracted increase in the
lated pre- and postsynaptic activity. synthesis of NR2B subunits of the NMDA recep-
As with many synapses in the hippocampal for- tor (Thomas et al., 1996; Williams et al., 2003).
mation, LTP at the medial perforant path–dentate Together, these proteins are suggested to partici-
synapse is blocked by antagonists selective for pate in the formation of elaborate modifications to
NMDA receptors. As with most known forms of the postsynaptic density (PSD) of granule cell
LTP, the influx of calcium postsynaptically as a synapses, a process thought to involve NR2B
result of NMDA receptor activation is thought NMDA receptor subunits and CaMKII via inter-
to be the crucial first step in LTP induction actions with actin in PSD. Here, elaboration of the
(Barrionuevo et al., 1980). The subsequent steps postsynaptic element and localization of CaMKII
following calcium entry are not yet defined in detail, may mediate the insertion of new receptors and/or
but as with LTP at other synapses, a concerted receptor subunits (Lisman and McIntyre, 2001),
effort has been made to identify important molec- also referred to as ‘‘AMPA receptor trafficking’’
ular and biochemical events that eventually increase (see Malinow, 2003, for review). These genes also
synaptic strength. Here a number of protein kinases may mediate the morphological changes reported
have been implicated, including a-calcium/calmod- to occur at axospinous synapses in the molecular
ulin kinase II (CaMKII; Thomas et al., 1994; Wu et layer (Trommald et al., 1996), such as changes in
al., 2006), protein kinase A (Wu et al., 2006), and postsynaptic structure (Desmond and Levy, 1983,
the ERK/MAPK cascade (Maguire et al., 1999; 1986). However, few data exist for the role of
Davis et al., 2000; Sweatt, 2001; Wu et al., 2006). AMPA trafficking in dentate granule cells, as most
The subsequent phosphorylation of the cAMP re- data supporting this process in LTP were obtained
sponse element binding protein (CREB) is likely the in studies of LTP in the Schaffer–CA1 pathway.
first transcription factor activated by these kinases Many of the genes thought to be involved in dent-
(Ying et al., 2002), and this is thought to initiate the ate LTP induction (genes that also code for pro-
expression of immediate-early genes (IEGs; Schulz teins that can associate with the PSD and that are
et al., 1999; Davis et al., 2000). The later phase implicated in AMPA receptor trafficking) are also
(>3 h) of LTP maintenance in the dentate requires thought to be involved in LTP at the Schaf-
the synthesis of proteins (Krug et al., 1984; Nguyen fer–CA1 synapses. Thus NMDA receptor-depend-
and Kandel, 1996) and transcription of new ent LTP as observed at medial perforant
mRNAs (Frey et al., 1996), and appears to involve path–dentate and Schaffer–CA1 synapses may re-
a later activation of protein kinase C (Colley et al., flect a common, prototypical ‘‘form’’ of LTP in-
1990; Thomas et al., 1994). duction and expression.
Several IEGs are implicated in the expression of Induction of LTP at the lateral perforant path
LTP. These genes are induced rapidly in a protein synapse is quite different from the medial perfo-
synthesis-independent manner, and generate a rant path projections. As first demonstrated by
number of proteins that, together with other IEG Bramham et al. (1988), naloxone, an antagonist of
proteins, can form transcription factors that the opioid receptor, blocks lateral perforant
428

path–dentate LTP in vivo. This was verified by while effective in inducing LTP in the medial per-
studies in vitro, where naloxone blocked lateral forant–dentate synapses, is not effective in induc-
perforant path–dentate LTP, and m opioid recep- ing LTP in lateral perforant path–dentate synapses
tor-selective agonists facilitated LTP induction (Colino and Malenka, 1993). A similar lack of
(Xie and Lewis, 1991). However, in these studies, ‘‘single pulse associativity’’ also is observed at
naloxone had little or no effect on low frequency mossy fiber synapses on CA3 pyramidal cells, a
lateral perforant path–dentate responses. The lack synapse that also contains and releases opioid
of effect of this opioid receptor antagonist on re- peptides (Chavkin et al., 1985) that are thought
sponses evoked at low frequencies suggests that essential for Hebbian LTP induction at the mossy
endogenous opioid peptides play a frequency-de- fiber synapse (Derrick et al., 1994a; Williams and
pendent role in LTP induction at lateral perforant Johnston, 1996). While this effect has been inter-
path–dentate synapses (Bramham et al., 1988; preted as reflecting a lack of associativity at mossy
Matthies et al., 2000; Krug et al., 2001). Although fiber synapses (Zalutsky and Nicoll, 1990), the lack
the requirement for opioid peptides in LTP induc- of single pulse mossy fiber associativity may sim-
tion may seem unusual, it may be less surprising ply reflect this requirement for mossy fiber activity
when one considers that the lateral entorhinal cor- necessary for the frequency-dependent release of
tex projections co-release opioid peptides with its opioid peptides (Neumaier and Chavkin, 1989;
principal transmitter, glutamate (Chavkin et al., Wagner et al., 1990, 1991; Derrick and Martinez,
1985). These peptides, derived primarily from pro- 1994a). Thus the requirement for high frequency
enkephalin, with [Met]- and [Leu]-enkephalin be- activity for LTP induction at lateral perforant
ing the primary opioid peptides produced by pro- path–dentate synapses similarly may reflect a re-
enkephalin and released by the lateral perforant quirement for the frequency-dependent release of
path, are stored in dense-core vesicles and are re- opioid peptides for LTP induction (Colino and
leased in an activity-dependent manner (Neumaier Malenka, 1993). There are a number of other ac-
and Chavkin, 1989; Wagner et al., 1990, 1991). tions of both pro-enkephalin and pro-dynorphin-
The frequency dependence of opioid peptide re- derived opioid peptides released by dentate gran-
lease by the lateral perforant path is common to ule cells. Most notable is the rather surprising
peptide co-transmitters; principally, their release finding that dynorphins, which act at kappa re-
often requires intense or repetitive activity (Hong ceptors and exert primarily inhibitory effects, at-
et al., 1988). It is thought that such activity is es- tenuate perforant path synaptic responses and
sential for the release of neuropeptides that are LTP induction, most likely via dendritic release of
stored in dense core vesicles, as they often are re- dynorphin-derived peptides (Wagner et al., 1993),
mote from both release sites and plasma mem- adding a different dimension to the roles of opioid
brane presynaptic terminals, and thus may require peptides in hippocampal information processing.
large amounts of calcium for their release. This The actions of opioids within the hippocampus
aspect of peptide storage and release may also ex- are varied and often depend on the receptors in-
plain why opioid receptor blockade has little effect volved as well as the site of the receptors. How-
on synaptic responses evoked in opioidergic pro- ever, the primary effect of many opioid peptides in
jections at low frequencies. Tacit in this interpre- the hippocampus is that they elicit excitation (Fry
tation is the idea that high-frequency lateral et al., 1979; Corrigall, 1983), which is observed
perforant path activity is necessary for LTP in- with activation of both the m and d opioid receptor
duction in order to achieve both postsynaptic de- types (Piguet and North, 1983). The application of
polarization and the frequency-dependent release exogenous opioids can even elicit seizures, and re-
of opioid neuropeptides. peated application can eventually elicit kindling
A similar dependence of lateral perforant path (Cain and Corcoran, 1985). These excitatory
LTP on high-frequency afferent activity is also effects are unusual for m and d opioid receptors,
observed in vitro, as pairing low frequency presy- as these opioid receptor types are usually coupled
naptic activity with postsynaptic depolarization, to the superfamily of GTPase-binding proteins
429

that inhibit cellular responses (Piguet and North, path–dentate synapses, is not likely to be related
1983). Why would neuromodulators normally as- to NMDA receptors. Perhaps the enhanced
sociated with inhibitory actions elicit excitation postsynaptic depolarization afforded by opioid
within the hippocampus? This paradox was clar- peptides permits calcium influx, which is necessary
ified in subsequent studies indicating that one for the induction of virtually all forms of LTP,
principal mechanism by which opioids mediate from other sources. For example, voltage-depend-
their excitatory effects is their attenuation of ent calcium channels can respond to the depolari-
GABA release (Cohen et al., 1992). Thus the ex- zation afforded by the disinhibitory effects of
citatory effects of proenkephalin-derived opioid opioid peptides. In addition, d opioid receptors are
peptides result from an attenuation of GABAergic suggested to activate phospholipase C and the
inhibition, an effect thought to result from the m or formation of IP3, a second messenger that often
d opioid receptor-mediated enhancement of po- regulates intracellular calcium release (Bramham,
tassium conductances in GABAergic interneurons, 1992). Thus direct actions of opioid peptides on
which inhibit the release of GABA (North et al., principal cells may be necessary for LTP induction
1987). by enhancing intracellular Ca++ levels by en-
How could these disinhibitory effects be impor- hancing release of Ca++ from intracellular stores.
tant in LTP induction? Agents that attenuate The suggestion of a distinct, NMDA receptor-
GABAergic inhibition generally facilitate LTP in- independent, opioid receptor-dependent form of
duction (Wigstrom and Gustaffson, 1985). This LTP in lateral perforant path afferents is also sup-
effect is thought to arise from the enhanced depo- ported by our own studies in area CA3, where
larization, and presumably, enhanced NMDA re- projections from the lateral perforant path termi-
ceptor activation that follows a reduction in nates in the stratum lacunosum-moleculare. This
GABAergic inhibition. In fact, the disinhibitory direct perforant path projection to CA3 pyramidal
effect of opioid peptides likely underlies their re- cells is appreciable: the average CA3 pyramidal
quirement for the induction of lateral perforant cell receives the same number of perforant path
path–dentate LTP, as the block of lateral perfo- synapses as does the average granule cell (Amaral
rant path LTP following naloxone can be pre- et al., 1990). Here, the story is quite similar to the
vented by co-application of a GABAA receptor dentate gyrus, in that while m, but not d, opioid
antagonist (Bramham and Sarvey, 1996). This the- receptor antagonists are effective in blocking LTP
ory of the contribution of opioid peptides to LTP at lateral perforant path–CA3 synapses, NMDA
induction has been longstanding (Swearengen and receptor antagonists have no effect on LTP induc-
Chavkin, 1987). tion (Do et al., 2002; Kosub et al., 2005). This is in
Although its modulation by opioids has been line with studies of opioid peptide receptor occu-
clarified, the mechanisms underlying lateral perfo- pation, which show a high density of m opioid re-
rant path LTP induction are not as simple. Studies ceptors in the stratum lacunosum-moleculare of
by Bramham and colleagues have shown that, the rat CA3 region (Neumaier and Chavkin, 1989).
while the activation of both m and d-type opioid Conversely, NMDA receptor antagonists are
receptors is crucial for LTP induction (Bramham effective in blocking LTP at medial perforant
et al., 1991a; Bramham and Sarvey, 1996), the fa- path–CA3 synapses (Do et al., 2002), whereas op-
cilitation of LTP by opioid peptides is unlikely to ioid receptor antagonists have little or no effect
arise via a facilitation of NMDA receptor activa- (Breindl et al., 1994; Do et al., 2002). In addition,
tion. This is because LTP at the lateral perforant stimulation of perforant path fibers often results in
path–dentate synapse was found to be independent a depression of responses at neighboring, inactive
from NMDA receptors in vivo, as competitive synapses, a phenomenon termed ‘‘heterosynaptic
NMDA receptor antagonists were ineffective in LTD.’’ In our studies, we found that when LTP is
blocking its induction (Bramham et al., 1991b). induced in a lateral perforant pathway, the
Thus, the way that opioid receptor activation fa- NMDA receptor antagonist CPP was ineffective
cilitates LTP induction at lateral perforant in blocking lateral perforant path LTP induction,
430

CPP was effective in blocking the heterosynaptic CPP or AP-5 (Snyder et al., 2001). As uncompeti-
LTD of medial perforant path–CA3 responses tive antagonists generally are effective at most
normally observed following lateral perforant path NMDA receptors, it remains possible that the
stimulation (Kosub et al., 2005). Thus, like their NMDA receptors in the terminal fields of lateral
synapses in the dentate gyrus, synapses from the perforant path projections may be a heteromeric
lateral perforant pathway to the CA3 region dis- variety (such as NR2B multimers, Massey et al.,
play LTP induction that is insensitive to NMDA 2004) that are unaffected by typical NMDA re-
receptor antagonists, but is sensitive to antagonists ceptor antagonists. In this case, the disinhibitory
of the m opioid receptor. effects of opioid peptides might contribute to LTP
Although these results are consistent with the induction via enhanced postsynaptic depolariza-
high density of m opioid receptors in the termina- tion and a facilitation of the activation of NMDA
tion site of the lateral perforant path in area CA3, receptor heteromers that may be insensitive to
physiological data indicate that NMDA receptors traditional competitive NMDA receptor antago-
are expressed in both stratum lacunosum-molecul- nists.
are of area CA3 and the outer molecular layer of It remains unknown whether the LTP at medial
the dentate, although to a lesser degree than the and lateral perforant path terminals reflect distinct
regions where medial perforant path afferents ter- ‘‘forms’’ of LTP that display distinct mechanisms
minate (see Monaghan and Cotman, 1985; Bram- of both induction and maintenance, or if the LTP
ham, 1992). If LTP induction in both the dentate in the lateral and medial perforant pathways sim-
and the CA3 region is independent from NMDA ply reflect differences in the mechanisms of induc-
receptors, what is the function of NMDA recep- ing a common form of LTP. The issue is of
tors located in their synaptic regions (Milner and interest, as there may be as many forms of LTP as
Drake, 2001)? One possibility may be the induc- there are forms of LTP induction. However, it is
tion of heterosynaptic LTD. Supporting this view, also quite possible that perforant path–dentate
the NMDA receptors found in the molecular layer LTP may differ simply in its induction require-
of the dentate may be extrasynaptic and found on ments, yet share common, downstream molecular
dendrites and dendritic shafts (Monaghan and processes of LTP maintenance. Nonetheless, if
Cotman, 1985; Bramham, 1992). Because recent medial and lateral perforant path synapses display
evidence implicates extrasynaptic NMDA recep- a similar form LTP that differs only in its induc-
tors, possibly NMDA receptors containing the tion mechanism, this also is of importance, as it
NR2B subunit, in LTD induction (Massey et al., suggests that distinct rules, or ‘‘constraints,’’ gov-
2004), it is possible that the principal process in- ern LTP induction among different synaptic pop-
volving NMDA receptors at lateral perforant path ulations of the hippocampal formation. Such
targets (and likely localized to extrasynaptic sites) difference may serve to restrict LTP in specific
is the induction of various forms of perforant path synaptic population to specific types of synaptic
LTD (see below). activity. For example, because opioid receptor ac-
It is also important to note the independence of tivation is crucial for the induction of lateral per-
LTP at the lateral perforant path synapse from forant path LTP, and because opioid peptides
NMDA receptors is not unchallenged. Zhang and display frequency-dependent release (Neumaier
Levy (1992) showed that lateral perforant and Chavkin, 1989; Wagner et al., 1990, 1991;
path–DG LTP could be blocked by systemic ad- Derrick et al., 1994a, b), it would be expected that
ministration of ketamine, an uncompetitive LTP at this synapse would only occur with high
NMDA receptor antagonist. A possible explana- frequency presynaptic activity. Thus repetitive pre-
tion for this effect may be the type of NMDA synaptic activity, such as is observed with bursting
receptor involved in LTP induction. It is known (in this case, bursting in cells of the lateral ent-
that not all NMDA receptor heteromers (such as orhinal cortex giving rise to the lateral perforant
NMDA receptors that contain NR2B subunit) are pathway) may be absolutely required for LTP in-
sensitive to competitive NMDA antagonists like duction in these projections. Thus, the additional
431

requirement for opioid peptides, and their require- of specific monoamines often associated with
ment for high frequency presynaptic activity or arousal, such as dopamine, norepinephrine, and
bursting for their release, may be viewed as a fea- serotonin (Li et al., 2003; Straube et al., 2003;
ture that imposes tighter ‘‘constraints’’ on the Sanberg et al., 2006).
types of synaptic activity that are effective in in- When LTP is induced either with theta bursts or
ducing lateral perforant path LTP. Satisfaction of by novel environments, LTP displays a distinct,
this additional constraint might only be provided extended time course. The implication of these
by certain types of presynaptic activity (such as findings is that distinct molecular mechanisms, or
bursting), thereby restricting LTP induction in the modulation or modification of a common mo-
these synaptic populations to states, such as theta lecular mechanism, may underlie LTP mainte-
states, that are associated with bursting of granule nance when induced under different conditions.
cells (Munoz et al., 1990) or cells of the entorhinal Thus, not only can different synapses display
cortex (Jeffery et al., 1995). different ‘‘forms’’ of LTP, but a given synapse may
While no data are available on whether or not also display distinct ‘‘forms’’ of LTP depending on
medial and lateral perforant path synapses display the animal’s behavioral state during its induction.
distinct ‘‘forms’’ of LTP, it should be noted that Thus, the present challenge is two-fold: to deter-
different ‘‘forms’’ of LTP may occur even within a mine the necessary and sufficient mechanisms that
single synaptic population. Recent evidence sug- may be shared by different ‘‘forms’’ of LTP, if they
gests that different forms of LTP, reflected in exist, and to determine the functional significance
differences in the longevity of LTP, and therefore of the variations of LTP induction and mainte-
LTP maintenance, can be regulated by a number nance. This approach has now become crucial, as
of variables, such as the pattern of afferent activity differences in LTP forms or LTP induction mech-
used to induce LTP, or by the behavioral state of anisms among the afferent populations and sub-
the animal during LTP induction. In the case of regions in the hippocampal formation are likely to
LTP induced with different types of stimulation, reveal important clues regarding the role of these
high-frequency stimulation that mimics naturally- distinct hippocampal afferents and hippocampal
occurring theta rhythm is suggested to induce a subregions in the processing and storage of infor-
‘‘non-decremental’’ form of LTP at Schaffer–CA1 mation in the hippocampal formation.
synapses (Staubli and Lynch, 1987), possibly by
inducing a number of forms of LTP (Morgan and
Teyler, 2001). In the case of different behavioral Plasticity of mossy fibers
states, the longevity of LTP also can depend on
behavioral state of the animal when LTP is in- It should also be mentioned that opioid peptides are
duced. At perforant path–dentate synapses, LTP also thought to play a role in LTP in efferents of the
induced following a single stimulation session typ- granule cells, the mossy fibers. In the first report by
ically persists for 5–7 days (Barnes, 1979; Davis Harris and Cotman (1986) that LTP at the mossy
et al., 2004). However, the duration of LTP is ex- fiber–CA3 synapse is insensitive to NMDA recep-
tended to more than 2 weeks if perforant path LTP tor antagonists, these investigators also suggested
is induced while the animal is actively engaged in that the insensitivity of mossy fiber LTP to NMDA
learning (such as during the initial exploration of a receptor antagonists may reflect a distinct ‘‘form’’
novel environment; Davis et al., 2004). Thus, the of LTP, although these investigators suggested that
production of LTP with a sustained time course is closer scrutiny of mossy fiber LTP maintenance
highly suggestive of distinct mechanisms of LTP might be important in order to verify this possibil-
maintenance, and thus distinct ‘‘forms’’ of LTP. ity. In addition, mossy fiber LTP induced in vivo
The mechanisms underlying these effects remain to displays an unusual, incremental, or slowly devel-
be elucidated, although the latter effect of novelty oping LTP that decay little over hours following
on enhancing the longevity of perforant induction, a feature not shared by commissural/
path–dentate LTP appears to require the presence associational CA3 inputs that display NMDA
432

receptor-dependent LTP. These differences in both postsynaptic factor that apparently is provided
the mechanism of LTP induction and the incre- by NMDA receptor activation is crucial for the
mental time course of LTP expression led us to induction of mossy fiber LTP.
suggest mossy fiber–CA3 LTP reflects a distinct What role might opioid peptides play in the in-
form of LTP (Derrick and Martinez, 1989). Our duction of mossy fiber LTP? It seems likely that, as
subsequent studies (Derrick et al., 1991, 1992) with lateral perforant path LTP, opioid receptor
showed that mossy fiber LTP induced in vivo re- activation may facilitate LTP induction as a result
quires the activation of opioid receptors, replicating of their disinhibitory effects. However, tacit in this
an original report by Martin (1983). Subsequent view is that mossy fiber LTP induction involves
studies revealed that the activation of the m opioid postsynaptic depolarization. Although there is am-
receptor is crucial for the induction of mossy fiber ple evidence that mossy fiber LTP depends on
LTP (Derrick et al., 1992). Interestingly, subsequent postsynaptic factors, including postsynaptic depo-
studies revealed that induction and development of larization (Jaffe and Johnston, 1990; Derrick et al.,
mossy fiber LTP is blocked by protein synthesis 1994b), m opioid receptors may have entirely
inhibitors (Barea-Rodriguez et al., 2000). As such different actions at this synapse. Johnston and
inhibitors usually only alter the late phase (>3 h) of colleagues (Williams and Johnston, 1996) demon-
NMDA receptor-dependent LTP, this result offers strate that while naloxone blocks mossy fiber LTP
further proof of distinct mechanisms of mainte- in vitro, the addition of GABAA receptor antag-
nance, and thus a distinct form of LTP. A 1991 onists does not obviate the block of mossy fiber
report by Nicoll and colleagues came to a similar LTP induction by naloxone. Thus, unlike the lat-
conclusion, but for different reasons. In this study, eral perforant pathway, reducing GABAergic in-
the induction of mossy fiber LTP was not only hibition does not obviate the dependence of mossy
NMDA receptor-independent, but displayed induc- fiber LTP on m opioid receptor activation. This
tion mechanisms that were exclusively presynaptic suggests another role for m opioid receptors in
(Zalutsky and Nicoll, 1990). mossy fiber LTP induction that is distinct from
However, the report by Zalutsky and Nicoll their actions on GABAergic inhibition.
(1990) was in conflict with earlier reports indicat- This same conclusion can be reached a priori
ing a dependence of mossy fiber LTP on both from previous studies in vitro; while the mossy
postsynaptic calcium (Williams and Johnston, fiber–CA3 synapse normally does not display sin-
1989) and depolarization (Jaffe and Johnston, gle-pulse associativity, single-pulse associativity
1990). While at that time, our data favored a pre- can be observed at mossy fiber synapses when m
synaptic locus for mossy fiber LTP induction opioid receptor agonists (which normally would be
(Derrick et al., 1992), our subsequent studies re- released with repetitive activity) are provided (Der-
vealed an intensity-dependent threshold for mossy rick et al., 1994a). However, single mossy fiber
fiber LTP induction (Derrick and Martinez, pulses delivered in conjunction with postsynaptic
1994b). Moreover, associative mossy fiber LTP depolarization of CA3 pyramidal cells still fails to
could be observed when weak mossy fiber synapses induce single-pulse associative mossy fiber LTP.
were co-active with intense activation of CA3 This effect is difficult to reconcile from the per-
associational/commissural inputs. Importantly, spective of opioid peptide-mediated disinhibition;
this effect could be blocked by an NMDA recep- if opioid receptor activation were necessary solely
tor antagonist (Derrick et al., 1994b). This was a for disinhibitory effects (and enhanced postsynap-
crucial finding, because NMDA antagonists, while tic depolarization), it would be expected that in-
effective in blocking LTP among commissural/ duced postsynaptic depolarization of CA3
associational inputs, normally do not block mossy pyramidal cells would obviate the need for any
fiber LTP. That NMDA receptor antagonists further depolarization afforded by the disinhibito-
blocked mossy fiber LTP induced associatively ry effects of opioid peptides. However, this ap-
by coactivation of an NMDA receptor dependent pears not to be the case, leaving open the
pathway offers evidence that a common possibility that m opioid receptors may contribute
433

to mossy fiber LTP via other mechanisms. Among region activate CA3 pyramidal cells 2–5 ms earlier
the possible effects of mu receptors are an en- than granule cells of the dentate (Fig. 2). Direct
hancement of presynaptic activity mediated by re- perforant path afferents are capable of discharging
duced GABAB receptor activation (Mott and CA3 pyramids (which occurs in vivo with the same
Lewis, 1992; Jin and Chavkin, 1999), or a m op- perforant path stimulation intensities that activate
ioid receptor-mediated activation of second mes- granule cells; see Do et al., 2002; Fig. 2). Thus the
senger cascades. For instance, while m opioid activation of CA3 region by perforant path would
receptors typically decrease calcium influx and en- initiate recurrent CA3–CA3 activity rapidly, as the
hance potassium efflux in most neuronal popula- commissural/associational CA3–CA3 fibers have
tions, m opioid receptors can activate the MAP and conduction velocities of 1.5 m/s in vivo (Do et al.,
ERK cascades in some preparations (Zimprich et 2002). By contrast, the parallel and simultaneous
al., 1995), and are even reported to facilitate tran- perforant path input to the granule cells elicits
scription via the translocation of calmodulin to granule cell firing 1–3 ms later than the CA3 py-
nuclear sites (Wang et al., 2000). The latter effect is ramidal cells. In addition, because the conduction
of particular interest, as it appears that the slowly velocity of the mossy fibers is quite slow,
developing, ‘‘incremental’’ potentiation seen with 0.4–0.8 m/s (Derrick et al., 1994a), the disynaptic
mossy fiber LTP in vivo requires protein synthesis activation of CA3 pyramidal cells by the mossy
(Barea-Rodriguez et al., 2000). fibers would occur after both PP–CA3 and
Nonetheless, the data suggesting exclusively pre- CA3–CA3 recurrent activity. This sequence of
synaptic mossy fiber LTP is nonetheless compel- CA3 afferent activation (PP–CA3, then CA3–CA3,
ling (Zalutsky and Nicoll, 1990; Weisskopf et al., followed by mossy fiber–CA3 synapses) places
1993; Mellor and Nicoll, 2001; Castillo et al., 2002; mossy fiber–CA3 synaptic activity in the position
Calixto et al., 2003). Computationally, if any of being the last CA3 synapse to be activated di-
synapse could operate effectively with a purely synaptically by perforant path input. This sequence
presynaptic form of LTP induction, particularly if is crucial, because the induction of associative LTP
its role were simply a ‘‘detonator’’ synapse, then at recently active synapses is asymmetric, and is
the mossy fiber synapse would be it. It is already observed only when presynaptic activity precedes
sparse anatomically, and the various features of postsynaptic depolarization (Levy and Steward,
dentate granule cells, in combination with com- 1983; Gustafsson and Wigstrom, 1986; Gustafsson
petitive learning exemplified by LTP and LTD at et al., 1987). This asymmetric feature, characterized
perforant path–dentate synapses, together may more recently as asymmetric LTP mediated by
provide sufficient constraints for the formation of spike timing-dependent plasticity (STDP), suggests
sparse mossy fiber outputs to the CA3 region. that perforant path–CA3 and CA3–CA3 synapses
These effects may obviate any need for coincident activated prior to, or simultaneously with postsy-
Hebbian mechanisms at mossy fiber synapses. naptic depolarization resulting from mossy fib-
Moreover, given the presumed ‘‘detonator’’ role er–CA3 ‘‘detonators’’ would induce associative
of this synapse, mossy fiber synapses need not dis- LTP at the recently activated CA3 synapses. As
play ‘‘associativity;’’ rather, mossy fiber synaptic mossy fiber ‘‘detonators’’ fire after PP–CA3 and
activity and its ‘‘detonator’’ properties are thought CA3–CA3 activity, this delay in detonator activity
crucial for inducing associative LTP at other, co- would be optimal for the induction of associative
active CA3 synapses. LTP at recently activated PP–CA3 and CA3–CA3
The view that mossy fiber activity plays a crucial synapses. A recent study suggests such an effect or
role by virtue of its ‘‘detonator’’ properties and the mossy fiber activity of commissural/associational
associative induction of LTP at other, coactive CA3 CA3 fibers, although it appears to follow rules
synapses is supported by the temporal sequence of atypical of those usually associated with STDP
afferent activation of CA3 pyramids following per- (Kobayashi and Poo, 2004).
forant path activity. As noted above, the perforant A primary assumption in this view of encoding in
path projections to the dentate gyrus and the CA3 the CA3 region is that mossy fiber synaptic activity
434

can induce associative LTP at perforant path and of mossy fiber LTP may occur in distinct behavi-
CA3–CA3 synapses that are located in more distal oral states, during particular patterns of afferent
dendritic regions of the CA3 pyramids. While this activity, which then may determine if pre- or
effect is plausible in the CA1 region, given the abil- postsynaptic forms of mossy fiber LTP are in-
ity of action potentials to ‘‘backpropagate’’ to den- duced. For instance, it is possible that the postsy-
dritic regions following CA1 pyramidal cell naptic, Hebbian form of mossy fiber LTP may be
discharge (Magee and Johnston, 1997), the CA3 essential during encoding, whereas a strictly pre-
pyramids do not show a large apical dendrite that synaptic, non-associative, and short-lived form of
extends into the stratum radiatum, as do the CA1 mossy fiber LTP may operate during recall. This
pyramids. However, data obtained in vitro suggest occurrence of distinct forms of mossy fiber LTP
that the mossy fibers are indeed able to induce LTP may serve to prevent Hebbian associativity during
at more distal CA3–CA3 and PP–CA3 synapses recall states, an event that can have catastrophic
(Chatterji et al., 1989; McMahon and Barrionuevo, effects on previously encoded patterns. This prob-
2002; Kobayashi and Poo, 2004). In addition, while lem, often referred to as the ‘‘stability plasticity
there is evidence for the mossy fibers having ‘‘det- dilemma’’ (Grossberg, 1987) is a critical, but often
onator’’ aspects and being able to initiate firing of ignored problem inherent in most neural network
CA3 pyramidal cell (Henze et al., 2002), this effect learning schemes (Hertz et al., 1991). Competitive
is frequency dependent. Thus repetitive activation learning and self-organization, as is suggested to
of mossy fiber–CA3 synapses is necessary for mossy occur in both perforant path–dentate (Hasselmo
fiber synapses to exhibit their ‘‘detonator’’ effects. et al., 1995; Rolls and Treves, 1998) and mossy
This requirement presents an additional problem fiber–CA3 synapses (Derrick and Martinez, 1996),
because, as noted above, repetitive activation of the would, over time, result in continual changing of
mossy fiber synapse would necessitate bursting in synaptic strength with each input, making any
granule cells, and such bursting appears to be a rare stable learning or categorization impossible (Hertz
event (Jung and McNaughton, 1993). However, et al., 1991). Thus the expression of a distinct,
granule cell bursting is observed in the intact animal presynaptic and non-associative form of mossy
during theta rhythm (Munoz et al., 1990). This ob- fiber LTP limited to periods of recall would limit
servation suggests that granule cell bursting, essen- the associative influence of mossy fiber synapses
tial for initiating the ‘‘detonator’’ features of the during recall. Thus a presynaptic, non-associative
mossy fiber synapse, may be limited to periods form of mossy fiber LTP could serve as a novel
during periods of theta rhythm. Thus while the solution to the ‘‘stability–plasticity dilemma,’’
mossy fiber synapse may display detonator proper- which is a problem inherent in all devices that re-
ties, these properties may be limited to behavioral quire both encoding and retrieval (Hertz et al.,
states that generate granule cell bursting, such as 1991).
that observed during theta rhythm which occurs Subsequent studies of the different types of
during exploratory behaviors. mossy fiber LTP nonetheless suggest that postsy-
What might explain the differences in results naptic factors are operative even with these ap-
from different laboratories regarding pre- and parently distinct forms of LTP induction (Yeckel
postsynaptic mechanisms of mossy fiber LTP? et al., 1999). Furthermore, the participation of a
Barrionuevo and colleagues (Urban and Bar- number of factors thought to act transsynaptically
rionuevo, 1996) suggested that mossy fiber LTP at mossy fiber synapses (such as the transsynaptic
itself may express different forms of LTP. Perhaps, activation of Eph receptors (Contractor et al.,
as observed with novelty-induced facilitation of 2002; Armstrong et al., 2006) and the possible re-
LTP that occurs at medial perforant path–dentate cruitment of extant, but ‘‘silent’’ mossy fiber
synapses (Davis et al., 2004), this may indicate the synapses with LTP induction (Reid et al., 2004)
induction of a different form of mossy fiber LTP also suggest a requirement for postsynaptic proc-
that depends on conditions that co-occur with esses in many forms of mossy fiber plasticity, and
LTP induction. Thus, pre- or postsynaptic forms with other studies, together support the view that
435

postsynaptic factors are involved in the induction regulators of the kind of mossy fiber LTP that will
of both mossy fiber LTP and mossy fiber LTD be induced. Alternatively, the latter result would
(Derrick et al., 1994b; Derrick and Martinez, 1996; suggest distinct forms of LTD that also may have
Yeckel et al., 1999; Kapur et al., 2001; Contractor specialized functional roles in scaling, sparsificati-
et al., 2002; Lei et al., 2003; Wang et al., 2004; on, depotentiation, or information storage.
Armstrong et al., 2006). The evidence for both Hebbian and non-
Although the above characterization of LTP in- Hebbian processes at the mossy fiber synapse
duction at the mossy fiber–CA3 synapse, as well as gives rise to an important question: what would be
lateral perforant path inputs to both the dentate the utility of having ‘‘detonator’’ synapses that are
and CA3 regions, together suggest a mechanism of plastic? What function could mossy fibers have
LTP induction involving opioid peptides, there is that would make Hebbian plasticity at a detonator
good evidence that a multiplicity of LTP and LTD synapse important? The dentate gyrus is thought
forms may exist at the mossy fiber synapse (Lei et crucial for the encoding of new information
al., 2003). Perhaps distinct forms of LTP can be (McNaughton et al., 1989), and shows the great-
induced at mossy fiber synapses, and which form is est activity in humans with novelty (Zeineh et al.,
induced may depend critically on the initial condi- 2003). Yet most models of autoassociative recall
tions during its induction. As we showed earlier in do not indicate that plasticity among detonator
the dentate gyrus, LTP with an enhanced time synapses is essential (Rolls and Treves, 1998; Kali
course can be observed in vivo depending on the and Dayan, 2000). Moreover, behavioral studies
conditions during LTP induction (Davis et al., indicate that the dentate gyrus is not essential for
2004; Sanberg et al., 2006). In this view, multiple recollection of spatial memory or the activity of
forms of LTP are possible at a synapse, with the place fields (McNaughton et al., 1989; Mizumori et
form of LTP that is expressed depending critically al., 1989), whereas others indicate the absence of
on the conditions during (and shortly after) LTP mossy fiber LTP has little effect on spatial memory
induction. Given that the mossy fiber–CA3 synapse (Huang et al., 1995). Thus current evidence sug-
displays the three known forms of LTP (homosy- gests that the dentate gyrus is crucial for the en-
naptic, heterosynaptic and associative LTD; coding, but not for the recall, of information.
Derrick and Martinez, 1996), and given that a va- However, these behavioral finding must be inter-
riety of cellular mechanisms (both pre- and postsy- preted carefully, as plasticity in dentate–CA3 in-
naptic) underlie what appears to be distinct forms puts may contribute to aspects of hippocampal
of homosynaptic mossy fiber LTD (Manabe, 1997; memory that may not be readily apparent when
Lei et al., 2003; Wang et al., 2004), the possibility using simple behavioral paradigms that are used to
that there are different forms of mossy fiber LTP assess only spatial memory. Given that the hippo-
remains the most parsimonious and compelling ex- campus is found to participate in more elaborate
planation for the disparate results seen with mossy tasks (Agster et al., 2002;), including those that
fiber LTP. Further work hopefully will elaborate on assess directly episodic-like memory (Fortin et al.,
the differences in the molecular machinery under- 2002; Kart-Teke et al., 2006), it remains possible
lying and the maintenance and expression that are that the contribution of the mossy fibers to mem-
associated with the distinct induction mechanisms ory may become evident in more elaborate and
of mossy fiber LTP and LTD. The findings from demanding tasks.
such studies may reveal a common cascade among Given that the dentate gyrus itself displays place
the different induction mechanisms, or reveal func- fields (Jung and McNaughton, 1993) and elaborate
tionally distinct forms of plasticity at this synapse. mechanisms of synaptic plasticity, as do the mossy
The former result would make us reconsider the fiber–CA3 outputs (Derrick and Martinez, 1996), it
classification of LTP and its ‘‘forms’’ based simply seems curious that elaborate plastic processes
upon differences in their induction mechanisms, would even exist in the dentate and their mossy
and may also suggest that these distinct induction fiber efferents unless recall was an important aspect
mechanisms may have functional significance, being of mossy fiber function. The possible contribution
436

of the dentate gyrus to information storage may serving to increase synaptic strength cannot oper-
best be understood from a theoretical view and the ate alone; otherwise, the strength of synapses
presumed role of the direct perforant path inputs to could only increase, eventually reaching a point of
CA3 to initiate recall (Treves and Rolls, 1992, saturation. Other mechanisms that permit either
1994). However, such a view may be problematic, the reversal or the inverse of LTP are likely to be
because activation of a subset of the direct perfo- necessary. Such a phenomenon, termed LTD, is
rant path–CA3 lines, which are thought to provide observed at the same synapses that display LTP.
the partial input that initiates pattern completion in LTD was noted in early studies, although its pos-
the CA3 region (Treves and Rolls, 1994), might fail sible role in information storage was only sug-
to select accurately among the many potential CA3 gested (Barrionuevo et al., 1980). As it became
attractors if they happen to share a critical subset of apparent that any memory device that serves as a
direct perforant path–CA3 inputs. Perhaps plastic- temporary repository for information must have
ity at perforant path–CA3, together with mossy some means to both increase and decrease synaptic
fiber LTP may cooperate during recall, particularly strength, the mechanisms underlying LTD became
recollection of more elaborate aspects of hippo- a primary focus for many studies in the 1990s
campal memory not readily detectable by current (Bear and Abraham, 1996).
behavioral methods that only assess spatial learn- In contrast to LTP, individual ‘‘forms’’ of LTD
ing. In this view, perforant path–CA3 and dent- in the dentate, as evidenced by their distinct mech-
ate–CA3 inputs operate together during recall to anisms of induction, were noted early. Homosy-
select among specific attractors within the CA3 au- naptic LTD, the term used to describe LTD that
toassociative system, with plasticity in both the di- follows synaptic activity, is typically input specific
rect perforant path–CA3 and mossy fiber–CA3 and induced experimentally in many hippocampal
inputs acting as a simple pattern associator that is synapses by repetitive low frequency (0.5–5 Hz)
embedded within the CA3 autoassociative system. stimulation (Huang et al., 1992). At most dentate
In this way, both mossy fiber and perforant path synapses, homosynaptic LTD, like LTP, is cooper-
inputs could together select a smaller subset of CA3 ative (Kerr and Abraham, 1995), associative
pyramidal cells. Thus the additional input from (Christie and Abraham, 1992a), and also requires
modified mossy fiber–CA3 synapses thus might al- postsynaptic calcium, although the role of NMDA
low for a partial input arising from direct perforant receptors is controversial; although some studies
path–CA3 synapses to select among potential CA3 indicate that homosynaptic LTD is NMDA recep-
attractors. Such a feature would allow for the se- tor-dependent (Thiels et al., 1996), other studies
lection of CA3 attractors that may share high de- indicate that both homosynaptic LTD and depot-
grees of similarity among their perforant path–CA3 entiation appear to involve metabotropic glutamate
inputs. In this view, plasticity at mossy fiber ‘‘det- receptors (O’Mara et al., 1995a; Manahan-
onators,’’ and subsequent mossy fiber activity dur- Vaughan, 1998; Kulla et al., 1999; Wu et al.,
ing recall, may be crucial for the selective recall of 2004) and calcium influx from intracellular stores,
specific patterns or sequences of events that may rather than NMDA receptors (O’Mara et al.,
share similar features. Implicit in this view is that 1995b).
mossy fiber plasticity, and the dentate gyrus in For afferents to the dentate gyrus, studying ho-
general, may be crucial for both the formation and mosynaptic LTD is problematic, as the induction
recall of discrete episodic-like memories. requirements vary among laboratories, and it is
notoriously difficult to elicit, both in vivo and in
vitro in the slice preparation (Errington et al.,
LTD 1995; Abraham, 1996; Abraham et al., 1996), al-
though some laboratories have found that partic-
LTP is noteworthy in that its induction follows the ular paradigms aid in its induction in vitro
rule of pre- and postsynaptic associativity as for- (Abraham, 1996) and in vivo (Manahan-Vaughan,
malized by Hebb (1949). However, a mechanism 1998). Not only does this characteristic make LTD
437

difficult to study, but it also illustrates an impor- homosynaptic LTD in the dentate can be induced
tant limitation of the in vitro preparation, as al- in an associative manner. However, associative
terations in dentate function may be an inevitable LTD appears to involve cellular mechanisms of
consequence of slice preparation. The viability of expression that are similar to heterosynaptic LTD,
studying plastic processes in dentate slices is not although their mechanisms of induction differ in
too surprising when one considers that prepara- that associative LTD in the dentate is reportedly
tion of hippocampal slices involves widespread NMDA receptor-independent (Christie et al.,
deafferentation, activity, and such trauma can 1995; Abraham and Tate, 1997; Philpot et al.,
cause sustained plastic changes in dentate func- 1999).
tion. For example, preparation of hippocampal Depotentiation is related to LTD, but refers to a
slices induce a number of IEGs such as c-fos reversal of LTP (Levy and Steward, 1979; Bar-
(Dragunow et al., 1989). As IEGs often show a rionuevo et al., 1980; Fujii et al., 1991) and may be
refractory period persisting for hours after their thought of as homosynaptic LTD of potentiated
induction (Sheng and Greenberg, 1990), the dent- synapses. However, depotentiation is distinct from
ate slice preparation may be of limited utility in homosynaptic LTP in two important ways. First,
dissecting the molecular events that mediate the depotentiation has a narrow window of induction;
variety of molecular events that underlie the var- in the dentate, stimulation must occur minutes fol-
ious forms of plasticity in the dentate gyrus. lowing LTP induction, otherwise LTP becomes re-
LTD can be elicited at synapses adjacent to sistant to depotentiation (Martin, 1998; Kulla et al.,
those that are active or potentiated. This form of 1999). Second, depotentiation and homosynaptic
LTD is referred to as ‘‘heterosynaptic’’ and is ob- LTD appear to involve distinct cellular mechanisms
served at synapses that were not activated by the (Lee et al., 2000; Soderling and Derkach, 2000).
stimulus used to induce potentiation. Heterosy- The early (o3 h) (Otani et al., 1989) phase of LTP
naptic LTD is robust at the perforant path–dent- is suggested to involve phosphorylation of the
ate projections following the induction of LTP in GluR1 AMPA subunit at both CaMKII and PKA
one set of afferents (such as the medial perforant sites on this receptor subunit (Soderling and
path), which then induces heterosynaptic LTD of Derkach, 2000). Studies suggest that whereas de-
responses at inactive synapses (in this case, either potentiation involves dephosphorylation at PKA
unstimulated medial perforant path synapses, or serine/threonine residues, homosynaptic LTD at
lateral perforant path synapses), and vice versa. naive synapses may involve dephosphorylation of
Here, LTD induction appears to involve to both CaMKII serine/threonine residues (Lee et al.,
NMDA receptors (Desmond et al., 1991) and the 2000).
L-type voltage-dependent calcium channels Just as the mechanisms underlying the induction
(VDCCs; Wickens and Abraham, 1991; Christie of heterosynaptic LTD, homosynaptic LTD, and
and Abraham, 1994; Camodeca et al., 1998), sug- depotentiation appear distinct, so are the treat-
gesting that the low levels of calcium necessary for ments that can affect their induction. For example,
heterosynaptic LTD may be provided by VDCCs 5-HT4 agonists (Kulla and Manahan-Vaughan,
activated in response to NMDA receptors, per- 2002) and D1/D5 dopamine antagonists block de-
haps via distinct NMDA receptors at distinct, ex- potentiation, but not LTP or heterosynaptic LTD
trasynaptic locations (such as the NR2B subunits (Otmakhova and Lisman, 1998; Kulla and Man-
of the NMDA receptor, or similar receptors on ahan-Vaughan, 2000). Conversely, drugs, such as
dendrites or the spine base and neck; Massey et al., nimodipine, an L-type calcium channel blocker,
2004). can block heterosynaptic LTD in the dentate gyrus
The third type of LTD, associative LTD, was (Wickens and Abraham, 1991) without altering
once regarded as a less than replicable phenome- LTP (Christie and Abraham, 1994). The number
non (Bear and Malenka, 1994); however, studies of studies of LTD in the dentate gyrus are limited,
showing homosynaptic LTD displays cooper- however, as LTD in the dentate is not easily in-
ativity (Kerr and Abraham, 1995) suggest duced in the slice preparation (Bear and Abraham,
438

1996), and may indicate the importance of extrin- provide sparse, orthogonal inputs to the CA3 re-
sic neuromodulators in the induction of LTD in gion (Rolls and Treves, 1998).
the dentate gyrus (Pang et al., 1993). Another interpretation is that LTD, operating
The diversity of the mechanisms that induce in conjunction with LTP, together serves in infor-
LTD and depotentiation suggest that different mation storage (Bear and Abraham, 1996; Bear,
types of LTD may serve distinct functions (Kerr 1999; Braunewell and Manahan-Vaughan, 2001).
and Abraham, 1996). As noted earlier, mechanisms In this way, LTP and LTD may act in concert,
that serve to increase synaptic strength would be enhancing the dynamic range of plasticity at a
expected to be accompanied by other mechanisms given synapse, as bidirectional plasticity can en-
that can reverse increases in synaptic strength; oth- hance the capacity, fidelity, and flexibility of the
erwise, interference would follow from an eventual networks (Dayan and Willshaw, 1991). Bidirec-
saturation of synaptic plasticity. Thus LTD may tional plasticity observed at the perforant
play a role in reversing LTP (‘‘depotentiation’’). In path–dentate synapse also is thought to be crucial
line with this view, heterosynaptic LTD can reverse in competitive learning schemes (Rolls and Treves,
established dentate LTP (Doyere et al., 1997) a 1998). Thus, while competitive learning is often
process likely to be essential in structures that have implicated in self-organization (Hertz et al., 1991),
a transient role in information storage, such as the it also can be viewed simply as another way to
hippocampal formation (Bontempi et al., 1999). remove redundancies and to sparsify dentate out-
Normalization processes are thought to be a crucial puts to the CA3 region (Rolls and Treves, 1998).
aspect of neuronal homeostasis — maintaining In this view, heterosynaptic LTD and homosy-
neuronal excitability within a dynamic range of naptic LTD have the potential to contribute to
neuronal output, which is essential to maintain normalization and sparse encoding (Morris, 1989).
maximal variance in neuronal output. Thus LTP
induced in a set of synapses on a neuron may result
in a concomitant, equivalent net decrease in the Metaplasticity
strength of other, inactive, synapses on the same
neuron. Metaplasticity refers to an alteration in the thresh-
Other forms of LTD may play a role in sparse old or magnitude of LTP or LTD induction in a
coding (Skaggs and McNaughton, 1992), ensuring neuron as a result of prior neuronal or synaptic
that only the most active synapses increase in activity (see Abraham and Tate, 1997; Philpot
strength in response to a given input, whereas et al., 1999). Metaplastic processes are thought to
other less active synapses are depressed, preserving regulate subsequent changes in synaptic strength
the sparse encoding essential for distributed mem- depending on the prior neuronal ‘history’ of
ory systems that employ the Hebb rule (Skaggs synaptic activity. Thus, metaplastic effects are
and McNaughton, 1992). For example, both ho- used to describe activity-dependent and sustained
mosynaptic LTD and associative LTD may facil- effects that can regulate the ability of a synapse, or
itate sparse encoding by depressing synapses that synapses on a given neuron, to change synaptic
show ill-timed or only modest activity. In this strength in response to subsequent synaptic activ-
view, these forms of LTD may filter out back- ity. As such, metaplasticity reflects a distinct
ground synaptic ‘‘noise,’’ reducing and optimizing mechanism of information storage (Philpot et al.,
the number of modifiable elements that contribute 1999).
to a given representation. Thus LTD may con- Metaplastic effects follow predictions of several
tribute to information storage by removing redun- theories of information processing that account for
dancies, and providing a sparse inverted code, or activity-dependent self-organization during devel-
‘‘transform,’’ of the same perforant path input that opment. In particular, the Bienenstock, Cooper and
is relayed directly to distal dendrites of CA3 neu- Munroe (BCM) theory (Bienenstock et al., 1982;
rons (Morris and McNaughton, 1987). This fea- Bear et al., 1987; Shouval et al., 1997), originally
ture may be crucial during encoding in order to used to account for visual cortex development,
439

employs processes that regulate synaptic strength


within an entire neuron as a consequence of prior
synaptic activity, and serves as a useful framework
with which to predict metaplastic effects. Generally,
the BCM theory predicts that if postsynaptic activ-
ity is insufficient for LTP induction, LTD will oc-
cur. This transition point where postsynaptic
activity induces LTD or LTP, in the BCM algo-
rithm, is denoted by ym (Fig. 4). Importantly, ym is
not fixed; instead, it changes as a function of
postsynaptic activity. For example, activity that is
not sufficient for LTP induction not only induces
LTD, but also ‘‘shifts’’ the threshold for LTP in-
duction (indicated by ym) to the left, so that the
threshold of postsynaptic activity necessary to in-
duce LTP at synapses on that neuron is subse-
quently reduced. Conversely, if stimulation is
sufficient to induce LTP, ym shifts to the right, in-
creasing the threshold for inducing LTP, and favo-
ring the induction of LTD with subsequent synaptic
activity.
An important aspect of changes in ym is that
this process, as envisaged by Bienenstock et al.
(1982), is ‘‘cell wide,’’ and any change in the
Fig. 4. The BCM rule and sliding modifications of LTP and
threshold for synaptic plasticity that follows ac- LTP induction thresholds. (A) The induction of LTP or LTD
tivity would affect all synapses of a neuron. Such a depends on the level of postsynaptic activity. Postsynaptic de-
feature can serve a normalization, or scaling, func- polarization (PSD) at or beyond the threshold for LTP induc-
tion by maintaining a constant net excitatory input tion (ym) results in the induction of LTP. (B) Prior neuronal
to a given neuron so that output remains within activity sufficient for LTP induction shifts ym to the right,
making subsequent postsynaptic activity more likely to induce
the neuron’s maximal dynamic range (Bienenstock LTD. (C) Prior activity insufficient to induce LTP shifts ym to
et al., 1982; Barrionuevo and Brown, 1983). It is the left, lowering the threshold for LTP induction. Note that a
important to note that metaplastic effects are dis- fixed level of postsynaptic activity that is initially the threshold
tinct from normalization or synaptic scaling, as for LTP induction (——) in (A) subsequently induces LTD (B)
metaplasticity refers specifically to the modifica- or LTP (C) depending on prior activity and the resultant shift in
ym.
tion of thresholds necessary to induce LTP or
LTD, whereas scaling reflects modifications in
synaptic drive. (shifting ym to the left). However, exposure of
When the rules of ‘‘cell wide’’ metaplasticity that these rats to light for various periods of time re-
alter LTP and LTD thresholds are employed in sulted in a shift of LTD/LTP thresholds to values
models of visual cortex, these rules can allow for similar to controls, confirming activity-based al-
self-organization, categorization, and a selective terations in LTP and LTD threshold in the visual
‘tuning’ of specific cells within a network to spe- cortex. This simple but powerful local learning
cific inputs (Shouval et al., 1997). These predic- rule, which appears operative in the visual cortex,
tions have been confirmed to some degree in the can allow for emergent properties in neuronal net-
developing visual cortex, where rats reared in dark works, including self-organization and stimulus
environments showed a greater propensity to dis- selectivity (Shouval et al., 1997).
play LTP than LTD. Thus, reduced input to the Within the hippocampus, metaplastic phenom-
visual cortex favored subsequent induction of LTP ena also are observed, as previous studies indicate
440

that low-frequency stimulation that is ineffective stimulation may underlie these discrepancies. By
in inducing either LTP or LTD (‘priming’ stimu- contrast, when stimulation is sufficient to induce
lation) can subsequently lead to a cell-wide de- LTP, subsequent perforant path–dentate potentiat-
crease (Holland and Wagner, 1998) in LTD ion with further perforant path stimulation is re-
threshold and enhance stimulation-induced re- duced (Frey et al., 1995). In addition, stimulation of
versal of LTP (or depotentiation). However, most lateral perforant path–dentate inputs that induces
hippocampal studies in vitro that utilize priming LTP in behaving animals reveals that subsequent
stimulation also report input-specific effects lim- stimulation at the same intensity induces LTP of a
ited to stimulated synapses (Huang et al., 1992). smaller magnitude for several weeks after initial
Thus, the validity of the BCM theory is compli- LTP induction. Conversely, when LTD follow lat-
cated by priming studies that report metaplastic eral perforant path stimulation, stimulation 1–3
effects that are not ‘‘cell wide,’’ but rather, are weeks later at this same intensity results in a robust
limited to activated synapses (‘‘input specific meta- LTP (Villarreal and Derrick, 2000). Perhaps synap-
plasticity’’). tic activation that does not induce plasticity, as ob-
Within the dentate gyrus, input specific effects served with priming, may lead to a modification of
are also observed with priming stimulation. Christie metaplastic rules such that patterned synaptic ac-
and Abraham (1992a) report that priming the lat- tivity that fails to induce LTP will always favor
eral perforant path with low frequency (5 Hz) theta LTD (Christie and Abraham, 1992a), whereas low
burst stimulation resulted in a facilitation of the frequency activity that is random or not patterned
induction of LTD. At first glance, these results seem might decrease the threshold for both LTD and
to be in opposition to the BCM rule (as stimulation LTP induction (Christie et al., 1995).
that does not evoke LTP should slide the LTD/LTP Recent studies suggest that the dentate gyrus
threshold to the left, favoring LTP). However, it also can display ‘‘cell wide’’ metaplastic effects.
should be noted that priming stimulation itself did This is observed following non-synaptic (anti-
not induce a change in synaptic strength. Thus dromic) activation of dentate granule cells in vivo
these findings are consonant with the BCM model, by hilar stimulation (Abraham et al., 2001). Per-
as it would be expected that low frequency stimu- haps the level of postsynaptic depolarization is the
lation that was not effective in inducing either LTP crucial factor in setting LTP and LTD thresholds,
or LTD may still result in a leftward shift in thresh- an effect that may not be effective when priming
olds. Thus, both the threshold for LTD and LTP stimulation is used.
would decrease. However, in both studies of CA1 Given that there is now evidence for both ‘‘input
and the dentate, the threshold for LTP induction specific’’ and ‘‘cell-wide’’ metaplasticity in the
was increased (Fujii et al., 1991; Wexler and Stan- dentate gyrus, how might the two forms of meta-
ton, 1993), suggesting that the threshold for LTP plasticity be important? Perhaps ‘‘cell wide’’ and
made a rightward, rather than a leftward shift. ‘‘input specific’’ forms of metaplasticity serve func-
Subsequent studies by the same investigators found tionally distinct roles at neuronal and synaptic
that modest lateral perforant pathway stimulation levels, respectively. For example, input specific
resulted in a facilitation of subsequent LTP induc- metaplasticity may operate during the initial stages
tion (Christie et al., 1995). This still is problematic, of synaptic plasticity to ensure sparse encoding. In
as stronger stimulation would be expected to shift this scheme, synapses that have been recently po-
the BCM curve to the right, rather than the left, and tentiated may be refractory to further potentiat-
increase the threshold for LTP induction, favoring ion, while those neurons receiving postsynaptic
LTD induction. input insufficient to induce LTP would be partic-
While these findings suggest metaplastic effects ularly sensitive to LTP induction with subsequent
governed by priming do not always follow the pat- synaptic activity. Such an effect would ensure that
ter predicted by the BCM algorithm, or its cell wide recently potentiated synapses are ‘‘opted out’’ of
functions, it seems likely that the use of priming potentiation by later inputs, further aiding in
441

dentate pattern separation. By contrast, cell-wide (Rycroft and Gibb, 2002), and the duration of
metaplastic effects that follow BCM rules may re- metaplastic effects in vivo (Villarreal and Derrick,
quire synaptic plasticity (either LTD or LTP). 2000) parallels LTP longevity (Davis et al., 2004),
Thus a BCM-like process that occurs with synaptic alterations in the multimeric composition of
changes can allow for long-term metaplastic NMDA receptors that involve the NR2B subunit
effects as predicted by the BCM theory, and al- of the NMDA receptor may reflect an important
low for sustained stimulus selectivity, a feature mechanism underlying metaplasticity. Subsequent
that would be particularly advantageous in a com- down stream mechanisms are suggested to involve
petitive learning device that performs categoriza- CaMKII (Zhang and Levy, 1992) and protein kin-
tion functions. ase C (Gisabella et al., 2003).
Interestingly, many priming effects that appear Given that the primary function of the dentate is
to follow BCM rules are observed primarily in the thought to be pattern separation and the genera-
lateral perforant pathway and the mossy fiber tion of a sparse, inverted code that allows for or-
pathway, pathways that contain and release pro- thogonal outputs to the CA3 region, input specific
enkephalin-derived opioid peptides. However, no metaplasticity may play a role in the pattern sep-
studies have addressed metaplasticity in medial aration processes of the dentate. For example, cells
perforant path afferents, or the contribution of that were recently active and potentiated will dis-
opioid peptides to the induction or expression of play a rightward shift in ym, and thus an increase
metaplastic phenomena in opioidergic lateral per- in LTP induction threshold, making these potenti-
forant path afferents (Christie and Abraham, ated synapses refractory to any further induction
1992a; Christie et al., 1995; Francesconi et al., of LTP, and even favor the induction of LTD with
1997). The current evidence suggests it is a likely subsequent synaptic activity. By contrast, synapses
role for opioid peptides, given that the disinhibi- that are activated to a degree that is insufficient to
tory effects of opioid peptides can greatly influence induce LTP would show a leftward shift in y,
the levels of postsynaptic activity, which, in turn, greatly reducing LTP induction threshold for sub-
can influence NMDA receptor activation, as well sequent LTP induction. Therefore, synapses that
as ym, and thus permit the differential induction of have been recently activated or modified may thus
either LTP or LTD (Francesconi et al., 1997; be ‘‘opted out’’ and resistant LTP induction or to
Wagner et al., 2001). any further increase in potentiation. Thus, analo-
The mechanisms underlying metaplastic effects gous to competitive learning, input-specific forms
remain unknown, although alterations in glutamate of metaplasticity observed in perforant path–dent-
receptors (Abraham and Tate, 1997; Castellani ate synapses may also contribute to the ‘‘pattern
et al., 2001) and activation of group II (Manahan- separation’’ functions of the dentate gyrus, favo-
Vaughan, 1998; Gisabella et al., 2003) and Group I ring LTP induction at synapses activated below
(Wu et al., 2004) metabotropic glutamate receptors LTP threshold, allowing for formation of sparse,
are suggested to mediate metaplastic effects, and non-overlapping patterns of granule cell activity, a
the induction of LTD and depotentiation in general function that is likely crucial for encoding infor-
(Manahan-Vaughan, 1998). Recent theories also mation in the CA3 autoassociative system.
implicate alterations in NMDA receptors and their
subunit composition as a likely means of regulating
postsynaptic responses to synaptic activity (Caste- Conclusions
llani et al., 2001). Several studies suggest common
mechanisms of metaplasticity among lateral and The dentate gyrus displays plastic processes that
medial inputs to the dentate (Abraham et al., 2001) may underlie learning and memory in the hippo-
that may involve alterations in NMDA receptors. campus. However, the same plastic processes that
As NR2B NMDA receptors show a decreased cal- are thought to contribute to information storage
cium conductance once bound to calmodulin also might serve to ensure the formation of sparse,
442

orthogonal transforms that are relayed to the CA3 and directing storage in, the CA3 autoassociative
region that can promote the encoding of discrete system over time. In this view, the sustained plastic
CA3 attractors, because these same plastic proc- processes displayed by the dentate gyrus, rather
esses also can aid in the formation of sparse, or- than subserving ‘‘information storage’’ in a formal
thogonal outputs to the CA3 region. First, both the sense, maintains a ‘‘memory’’ for past inputs in
anatomically sparse mossy fiber projections and the order to maintain a continued generation of
‘‘input expansion’’ seen in the relay from the ent- sparse, orthogonal patterns that are relayed to
orhinal cortex to the dentate gyrus are anatomical the CA3 region, where the actual information is
features that can facilitate dentate pattern separa- both stored and recalled. Alternatively, bidirec-
tion. In addition, both competitive learning (as ex- tional plasticity and metaplasticity in the dentate
emplified by LTP and LTD) and metaplasticity, may play a more direct role in memory storage,
often implicated in self-organization, also may con- rather than simply serving to maintain separation
tribute in ensuring sparse orthogonal dentate out- among patterns relayed and stored in the CA3
puts relayed to the CA3 region, principally by system. The dentate gyrus itself displays place
removing redundancies from dentate outputs and fields (Jung and McNaughton, 1993) and elaborate
by limiting synaptic plasticity at previously potenti- mechanisms of long-lasting synaptic potentiation
ated synapses, and favoring LTP only at synapses and depression, as does the mossy fiber–CA3
that have not been potentiated previously. Moreo- synapse (Derrick and Martinez, 1996). Thus it
ver, the requirement of opioid peptides for plasticity seems odd that such elaborate plastic processes
and their requirement for specific types of would exist unless recall also was an important
presynaptic activity for their release may constrain aspect of dentate–mossy fiber function. Perhaps
conditions during which plasticity can occur, the plasticity observed in both perforant path
further contributing to sparse encoding. Together, inputs and mossy fiber outputs of the dentate are
these mechanisms may insure that the dentate crucial for both encoding and recall. In this case,
gyrus relays sparse, orthogonal transforms of the sustained changes at both perforant path–dentate
same perforant path input arriving distally on and dentate–CA3 synapses, in conjunction with
CA3 pyramids and aiding in the encoding of activity initiated by direct perforant path–CA3 in-
discrete attractors within the CA3 autoassociative puts, may allow for a more ‘‘fine tuned’’ selection
matrix. of attractors in the CA3 region. This is because
Recent evidence suggests that the dentate gyrus, partial inputs arriving distally at perforant
while essential for encoding, is not necessary for path–CA3 inputs and subsequent pattern comple-
accurate recall (Mizumori et al., 1989; Lee and tion have the possibility of converging on any
Kesner, 2004; Jerman et al., 2006). Why would number of possible CA3 attractors that share these
there be synaptic plasticity in both afferents and same perforant path–CA3 inputs. Thus, sustained
efferents of granule cells, and elaborate mecha- perforant path–dentate and mossy fiber–CA3 plas-
nisms regulating it if the output is only crucial for ticity might contribute to recall, with the mossy
encoding? One interpretation is that the ‘‘plastic’’ fiber inputs acting as an additional ‘‘cue’’ that
processes of LTP, LTD, and metaplasticity may allows for the selective activation of CA3 attrac-
operate primarily to enforce pattern separation by tors among many that may share common
the dentate gyrus and sparse encoding in the CA3 features. In this view, plasticity in the dentate
system. Thus, the plastic processes in the dentate, gyrus and its CA3 outputs are critical for accurate
rather than participating in memory and informa- recall among a number of potential CA3 attractor
tion storage per se, actually may serve primarily to states. In this case, it would be predicted that
preserve pattern separation functions of the dent- damage to the dentate gyrus, or altering mecha-
ate. These plastic processes could then provide nisms of LTP at either perforant path–dentate of
mechanisms that allow for sustained pattern mossy fiber–CA3 synapses, would display behavi-
separation of dentate input patterns relayed to, oral deficits, but primarily in tasks that require
443

finer discriminations, such as disambiguation of potentiation of mossy fibers in the hippocampus. J. Neuro-
encoded events that have common features (such sci., 26(13): 3474–3481.
Barea-Rodriguez, E.J., Rivera, D.T., Jaffe, D.B. and Martinez
as context). Thus the dentate gyrus, and plastic
Jr., J.L. (2000) Protein synthesis inhibition blocks the induc-
processes observed in the dentate CA3 relay, may tion of mossy fiber long-term potentiation in vivo.
be crucial in the formation of similar, but discrete J. Neurosci., 20(22): 8528–8532.
events over time, a feature that may be crucial for Barnes, C.A. (1979) Memory deficits associated with senes-
accurate encoding and recollection of episodic cence: a neurophysiological and behavioral study in the rat.
memory. J. Comp. Physiol. Psychol., 93(1): 74–104.
Barrionuevo, G. and Brown, T.H. (1983) Associative long-term
potentiation in hippocampal slices. Proc. Natl. Acad. Sci., 80:
7347–7351.
References Barrionuevo, G., Schottler, F. and Lynch, G. (1980) The effects
of repetitive low frequency stimulation on control and
Abraham, W.C. (1996) Induction of heterosynaptic and ‘‘potentiated’’ synaptic responses in the hippocampus. Life
homosynaptic LTD in hippocampal subregions in vivo. Sci., 27(24): 2385–2391.
J. Physiol. Paris, 90: 305–306. Bayer, S.A. (1982) Changes in the total number of dentate
Abraham, W.C., Mason-Parker, S.E., Bear, M.F., Webb, S. granule cells in juvenile and adult rats: a correlated
and Tate, W.P. (2001) Heterosynaptic metaplasticity volumetric and thymidine study. Exp. Brain Res., 46:
in the hippocampus in vivo: a BCM-like modifiable 315–323.
threshold for LTP. Proc. Natl. Acad. Sci., 98(19): Bear, M. and Abraham, W.C. (1996) Long-term depression in
10924–10929. hippocampus. Annu. Rev. Neurosci., 19: 437–462.
Abraham, W.C., Mason-Parker, S.E. and Logan, B. (1996) Bear, M.F. (1999) Homosynaptic long-term depression: a
Low-frequency stimulation does not readily cause long-term mechanism for memory? Proc. Natl. Acad. Sci. U.S.A.,
depression or depotentiation in the dentate gyrus of awake 96(17): 9457–9458.
rats. Brain Res., 722(1–2): 217–221. Bear, M.F., Cooper, L.N. and Ebner, F.F. (1987) A physio-
Abraham, W.C. and Tate, W.P. (1997) Metaplasticity: a new logical basis for a theory of synapse modification. Science,
vista across the field of synaptic plasticity. Prog. Brain Res., 237(4810): 42–48.
52(4): 303–323. Bear, M.F. and Malenka, R.C. (1994) Synaptic plasticity: LTP
Agster, K.L., Fortin, N.J. and Eichenbaum, H. (2002) The and LTD. Curr. Opin. Neurobiol., 4(3): 389–399.
hippocampus and disambiguation of overlapping sequences. Ben-Ari, Y. and Represa, A. (1990) Brief seizure episodes
J. Neurosci., 22(13): 5760–5768. induce long-term potentiation and mossy fibre sprou
Altman, J. and Das, G.D. (1965) Autoradiological and histo- ting in the hippocampus. Trends Neurosci., 13(8): 312–318.
logical evidence of postnatal hippocampal neurogenesis in Berzhanskaya, J., Urban, N. and Barrionuevo, G. (1998)
rats. J. Comp. Neurol., 124: 319–335. Electrophysiological and pharmacological characterization
Amaral, D.G. (1978) A Golgi study of cell types in the hilar of the direct perforant path input to hippocampal area CA3.
region of the hippocampus in the rat. J. Comp. Neurol., J. Neurophysiol., 79(4): 2111–2118.
182(4 Pt 2): 851–914. Bienenstock, E.L., Cooper, L.N. and Munro, P.W. (1982)
Amaral, D.G., Ishizuka, N. and Claiborne, B. (1990) Neurons, Theory for the development of neuron selectivity: orientation
numbers and the hippocampal network. Prog. Brain Res., 83: specificity and binocular interaction in visual cortex. J. Ne-
1–11. urosci., 2(1): 32–48.
Amaral, D.G. and Witter, M.P. (1989) The three-di Bliss, T. and Gardner-Medwin, A. (1973) Long lasting potent-
mensional organization of the hippocampal formation: iation of synaptic-transmission in the dentate area of the
a review of anatomical data. Neuroscience, 31(3): unanesthetized rabbit following stimulation of the perforant
571–591. path. J. Physiol., 232(2): 357–374.
Amaral, D.G. and Witter, M.P. (1995) Hippocampal Bliss, T.V.P. and Lomo, T. (1973) Long-lasting potentiation of
formation. In: Paxinos G. (Ed.), The Rat Nervous synaptic transmission in the dentate area of the anaesthetized
System (4th ed.). Academic Press, San Diego, CA, pp. rabbit following stimulation of the perforant path.
443–493. J. Physiol., 232: 331–356.
Andersen, P., Holmqvist, B. and Voorhoeve, P.E. (1966) Bontempi, B., Laurent-Demir, C., Destrade, C. and Jaffard, R.
Excitatory synapses on hippocampal apical dendrites (1999) Time-dependent reorganization of brain circuitry
activated by entorhinal stimulation. Acta Physiol. Scand., underlying long-term memory storage. Nature, 400(6745):
66: 461–472. 671–675.
Armstrong, J.N., Saganich, M.J., Xu, N.J., Henkemeyer, M., Boss, B.D., Peterson, G.M. and Cowan, W.M. (1985) On the
Heinemann, S.F. and Contractor, A. (2006) B-ephrin reverse number of neurons in the dentate gyrus of the rat. Brain Res.,
signaling is required for NMDA-independent long-term 338: 144–150.
444

Bramham, C.R. (1992) Opioid receptor-dependent long- plasticity: dependence on AMPA and NMDA receptors.
term potentiation: peptidergic regulation of synaptic Proc. Natl. Acad. Sci., 98(22): 12772–12777.
plasticity in the hippocampus. Neurochem. Int., 20(4): Castillo, P.E., Schoch, S., Schmitz, F., Sudhof, T.C. and
441–455. Malenka, RC. (2002) RIM1alpha is required for pre-
Bramham, C.R., Errington, M.L. and Bliss, T.V. (1988) synaptic long-term potentiation. Nature, 415(6869): 327–330.
Naloxone blocks the induction of long-term potentia Cavazos, J.E., Golarai, G. and Sutula, T.P. (1991) Mossy fiber
tion in the lateral but not in the medial perforant synaptic reorganization induced by kindling: time course of
pathway in the anesthetized rat. Brain Res., 449(1–2): development, progression, and permanence. J. Neurosci.,
352–356. 11(9): 2795–2803.
Bramham, C.R., Milgram, N.W. and Srebro, B. (1991a) Delta Chatterji, S., Stanton, P.K. and Sejnowski, T.J. (1989) Com-
opioid receptor activation is required to induce LTP of missural synapses, but not mossy fiber synapses, in hippo-
synaptic transmission in the lateral perforant path in vivo. campal field CA3 exhibit associative long-term potentiation.
Brain Res., 567(1): 42–50. Brain Res., 495: 145–150.
Bramham, C.R., Milgram, N.W. and Srebro, B. (1991b) Chavkin, C., Shoemaker, W.J., McGinty, J.F., Bayon, A. and
Activation of AP5-sensitive NMDA receptors is not required Bloom, F.E. (1985) Characterization of pro-dynorphin and
to induce LTP of synaptic transmission in the lateral perfo- proenkephalin neuropeptide systems in the rat hippocampus.
rant path. Eur. J. Neurosci., 3: 1300–1308. J. Neurosci., 5: 808–816.
Bramham, C.R. and Sarvey, J.M. (1996) Endogenous activa- Chawla, M.K., Guzowski, J.F., Ramirez-Amaya, V., Lipa, P.,
tion of mu and delta-1 opioid receptors is required for long- Hoffman, K.L., Marriot, L.K., Worley, P.F., McNaughton,
term potentiation induction in the lateral perforant path: B.L. and Barnes, C.A. (2005) Sparse, environmentally-
dependence on GABAergic inhibition. J. Neurosci., 16: selective expression of Arc RNA in the upper
8123–8131. blade of the rodent fascia dentate. Hippocampus, 15(5):
Bramham, C.R. and Srebro, B. (1989) Synaptic plasticity in the 579–586.
hippocampus is modulated by behavioral state. Brain Res., Christie, B.R. and Abraham, W.C. (1992a) Priming of associ-
493(1): 74–86. ative LTD by theta frequency synaptic activity. Neuron, 8:
Braunewell, K.H. and Manahan-Vaughan, D. (2001) Long- 79–84.
term depression: a cellular basis for learning? Rev. Neurosci., Christie, B.R. and Abraham, W.C. (1992b) NMDA-dependent
12(2): 121–140. heterosynaptic long-term depression in the dentate gyrus of
Breindl, A., Derrick, B.E., Rodriguez, S.B. and Martinez Jr., anaesthetized rats. Synapse, 10(1): 1–6.
J.L. (1994) Opioid receptor-dependent long-term potentiat- Christie, B.R. and Abraham, W.C. (1994) L-type voltage-
ion at the lateral perforant path-CA3 synapse in rat hippo- sensitive calcium channel antagonists block heterosynaptic
campus. Brain Res. Bull., 33(1): 17–24. long-term depression in the dentate gyrus of anaesthetized
Buzsaki, G. (1988) Polysynaptic long-term potentiation: a rats. Neurosci. Lett., 167(1-2): 41–45.
physiological role of the perforant path-CA3/CA1 pyrami- Christie, B.R., Stellwagen, D. and Abraham, W.C. (1995)
dal cell synapse. Brain Res., 455: 192–195. Evidence for common expression mechanisms underlying
Buzsaki, G. (2002) Theta oscillations in the hippocampus. heterosynaptic and associative long-term depression in the
Neuron, 33: 325–340. dentate gyrus. J. Neurophysiol., 74(3): 1244–1247.
Cain, D.P. and Corcoran, M.E. (1985) Epileptiform effects of Claiborne, B.J., Amaral, D.G. and Cowan, W.M. (1986)
met-enkephalin, beta-endorphin and morphine: kindling of A light and electron microscopic analysis of the mossy
generalized seizures and potentiation of epileptiform effects fibers of the rat dentate gyrus. J. Comp. Neurol., 246(4):
by handling. Brain Res., 338(2): 327–336. 435–458.
Calixto, E., Thiels, E., Klann, E. and Barrionuevo, G. (2003) Claiborne, B.J., Xiang, Z. and Brown, T.H. (1993) Hippo-
Early maintenance of hippocampal mossy fiber — long-term campal circuitry complicates analysis of long-term poten
potentiation depends on protein and RNA synthesis and tiation in mossy fiber synapses. Hippocampus, 3(2):
presynaptic granule cell integrity. J. Neurosci., 23(12): 115–121.
4842–4849. Cohen, G.A., Doze, V.A. and Madison, D.V. (1992) Opioid
Camodeca, N., Rowan, M.J. and Anwyl, R. (1998) Induction inhibition of GABA release from presynaptic terminals of rat
of LTD by increasing extracellular Ca2+ from a low hippocampal interneurons. Neuron, 9: 325–335.
level in the dentate gyrus in vitro. Neurosci. Lett., 255(1): Colino, A. and Malenka, R.C. (1993) Mechanisms underlying
53–56. the induction of long-term potentiation in rat medial and
Cartwright, R.D. (2004) The role of sleep in changing our lateral perforant paths in vitro. J. Neurophysiol., 69:
minds: a psychologist’s discussion of papers on memory re- 1150–1159.
activation and consolidation in sleep. Learn. Mem., 11(6): Colley, P.A., Sheu, F.S. and Routtenberg, A. (1990) Inhibition
660–663. of protein kinase C blocks two components of LTP persist-
Castellani, G.C., Quinlan, E.M., Cooper, L.N. and Shouval, ence, leaving initial potentiation intact. J. Neurosci., 10(10):
H.Z. (2001) A biophysical model of bidirectional synaptic 3353–3360.
445

Contractor, A., Rogers, C., Maron, C., Henkemeyer, M., dentate gyrus in the freely moving rat. J. Neurophysiol.,
Swanson, G.T. and Heinemann, S.F. (2002) Trans-synaptic 77(2): 571–578.
Eph receptor-ephrin signaling in hippocampal mossy fiber Dragunow, M., Abraham, W.C., Goulding, M., Mason, S.E.,
LTP. Science, 296(5574): 1864–1869. Robertson, H.A. and Faull, R.L. (1989) Long-term potent-
Corrigall, W.A. (1983) Opiates and the hippocampus: a review iation and the induction of c-fos mRNA and proteins in the
of the functional and morphological evidence. Pharmacol. dentate gyrus of unanesthetized rats. Neurosci. Lett., 101(3):
Biochem. Behav., 18: 255–262. 274–280.
Dalby, N.O. and Mody, I. (2003) Activation of NMDA recep- Errington, M.L., Bliss, T.V., Richter-Levin, G., Yenk, K.,
tors in rat dentate gyrus granule cells by spontaneous and Doyere, V. and Laroche, S. (1995) Stimulation at 1–5 Hz
evoked transmitter release. J. Neurophysiol., 90(2): 786–797. does not produce long-term depression or depotentiation in
Davis, C., Jones, F. and Derrick, B.E. (2004) Novelty enhances the hippocampus of the adult rat in vivo. J. Neurophysiol.,
the induction and maintenance of LTP in the in the dentate 74(4): 1793–1799.
gyrus. J. Neurosci., 24(29): 6497–6506. Escobar, M., Barea-Rodriguez, E.J., Derrick, B.E. and Martinez
Davis, S., Vanhoutte, P., Pages, C., Caboche, J. and Laroche, S. Jr., J.L. (1997) Mossy fiber synaptogenesis is induced by
(2000) The MAPK/ERK cascade targets both Elk-1 and high frequency stimulation of mossy fibers. Brain Res., 751:
cAMP response element-binding protein to control long-term 330–335.
potentiation-dependent gene expression in the dentate gyrus Fortin, N.J., Agster, K.L. and Eichenbaum, H.B. (2002)
in vivo. J. Neurosci., 20(12): 4563–4572. Critical role of the hippocampus in memory for sequences
Dayan, P. and Willshaw, D.J. (1991) Optimising synaptic of events. Nat. Neurosci., 5(5): 458–462.
learning rules in linear associative memories. Biol. Cybern., Francesconi, W., Berton, F., Demuro, A., Madamba, S.G. and
65(4): 253–265. Siggins, G.R. (1997) Naloxone blocks long-term depression
Derrick, B.E. and Martinez Jr., J.L. (1989) A unique opioid of excitatory transmission in rat CA1 hippocampus in vitro.
peptide-dependent form of long-term potentiation is found in Arch. Ital. Biol., 135(1): 37–48.
the CA3 region of the rat hippocampus. Adv. Biosci., 75: Frey, U., Frey, S., Schollmeier, F. and Krug, M. (1996)
213–216. Influence of actinomycin D, a RNA synthesis inhibitor, on
Derrick, B.E. and Martinez Jr., J.L. (1994a) Opioid receptor long-term potentiation in rat hippocampal neurons in vivo
activation is one factor underlying the frequency depen- and in vitro. J. Physiol., 490(Pt 3): 703–711.
dence of mossy fiber LTP induction. J. Neurosci., 14(7): Frey, U., Schollmeier, K., Reymann, K.G. and Seidenbecher,
4359–4367. T. (1995) Asymptotic hippocampal long-term potentiation in
Derrick, B.E. and Martinez Jr., J.L. (1994b) Frequency- rats does not preclude additional potentiation at later phases.
dependent associative long-term potentiation at the hippo- Neuroscience, 67(4): 799–807.
campal mossy fiber-CA3 synapse. Proc. Natl. Acad. Sci. Fry, J.P., Zieglgansberger, W. and Herz, A. (1979) Specific
U.S.A., 91: 10290–10294. versus non-specific actions of opioids on hippocampal neu-
Derrick, B.E. and Martinez Jr., J.L. (1996) Associative, bidi- rons in the rat brain. Brain Res., 163: 295–305.
rectional modifications at the hippocampal mossy fibre-CA3 Fujii, S., Saito, K., Miyakawa, H., Ito, K. and Kato, H. (1991)
synapse. Nature, 381: 429–434. Reversal of long-term potentiation (depotentiation)
Derrick, B.E., Rodriguez, S.B., Lieberman, D.N. and Martinez induced by tetanus stimulation of the input to CA1 neurons
Jr., J.L. (1992) Mu opioid receptors are associated with of guinea pig hippocampal slices. Brain Res., 555(1):
the induction of LTP at hippocampal mossy fiber synapses. 112–122.
J. Pharmacol. Exp. Ther., 263: 725–733. Gisabella, B., Rowan, M.J. and Anwyl, R. (2003) Mechanisms
Desmond, N.L., Colbert, C.M., Zhang, D.X. and Levy, W.B. underlying the inhibition of long-term potentiation by
(1991) NMDA receptor antagonists block the induction of preconditioning stimulation in the hippocampus in vitro.
long-term depression in the hippocampal dentate gyrus of the Neuroscience, 121(2): 297–305.
anesthetized rat. Brain Res., 552: 93–98. Gonzales, R.B., DeLeon Galvan, C.J., Rangel, Y.M. and
Desmond, N.L. and Levy, W.B. (1983) Synaptic correlates of Claiborne, B.J. (2001) Distribution of thorny excrescences on
associative potentiation/depression: an ultrastructural study CA3 pyramidal neurons in the rat hippocampus. J. Comp.
in the hippocampus. Brain Res., 265(1): 21–30. Neurol., 430(3): 357–368.
Desmond, N.L. and Levy, W.B. (1986) Changes in the postsy- Grossberg, S. (1987) Competitive learning: from interactive
naptic density with long-term potentiation in the dentate activation to adaptive resonance. Cogn. Sci., 11: 23–63.
gyrus. J. Comp. Neurol., 253(4): 476–482. Gustafsson, B. and Wigstrom, H. (1986) Hippocampal long-
Do, V.H., Martinez, C.O., Martinez Jr., J.L. and Derrick, B.E. lasting potentiation produced by pairing single volleys
(2002) Long-term potentiation in direct perforant path pro- and brief conditioning tetani evoked in separate afferents.
jections to hippocampal area CA3 in vivo. J. Neurophysiol., J. Neurosci., 6(6): 1575–1582.
87: 669–674. Gustafsson, B., Wigstrom, H., Abraham, W.C. and Huang,
Doyere, V., Srebro, B. and Laroche, S. (1997) Heterosynaptic Y.Y. (1987) Long-term potentiation in the hippocampus us-
LTD and depotentiation in the medial perforant path of the ing depolarizing current pulses as the conditioning stimulus
446

to single volley synaptic potentials. J. Neurosci., 7(3): Jaffe, D. and Johnston, D. (1990) Induction of long-term
774–780. potentiation at hippocampal mossy fibers follow a Hebbian
Guzowski, J.F., Lyford, G.L., Stevenson, G.D., Houston, F.P., rule. J. Neurophysiol., 64: 948–960.
McGaugh, J.L., Worley, P.F. and Barnes, C.A. (2000) James, W. (1890) Association. In: Psychology: A Briefer
Inhibition of activity-dependent arc protein expression in Course. Holt, New York, pp. 253–279.
the rat hippocampus impairs the maintenance of long-term Jeffery, K.J., Donnett, J.G. and O’Keefe, J. (1995)
potentiation and the consolidation of long-term memory. Medial septal control of theta-correlated unit firing in the
J. Neurosci., 20(11): 3993–4001. entorhinal cortex of awake rats. Neuroreport, 6(16):
Harris, E.W. and Cotman, C.W. (1986) Long-term potentiation 2166–2170.
of guinea pig mossy fiber responses is not blocked by Jerman, T., Kesner, R.P. and Hunsaker, M.R. (2006) Discon-
N-methyl D-aspartate antagonists. Neurosci. Lett., 70: nection analysis of CA3 and DG in mediating encoding but
132–137. not retrieval in a spatial maze learning task. Learn. Mem.,
Hasselmo, M.E., Schnell, E. and Barkai, E. (1995) Dynamics of 13(4): 458–464.
learning and recall at excitatory recurrent synapses and Jin, W. and Chavkin, C. (1999) Mu opioids enhance mossy fiber
cholinergic modulation in rat hippocampal region CA3. synaptic transmission indirectly by reducing GABAB recep-
J. Neurosci., 15(7 Pt 2): 5249–5262. tor activation. Brain Res., 821(2): 286–293.
Havik, B., Rokke, H., Bardsen, K., Davanger, S. and Jones, M.W., Errington, M.L., French, P.J., Fine, A., Bliss,
Bramham, C.R. (2003) Bursts of high-frequency stimulation T.V., Garel, S., Charnay, P., Bozon, B., Laroche, S. and
trigger rapid delivery of pre-existing alpha-CaMKII mRNA Davis, S. (2001) A requirement for the immediate early gene
to synapses: a mechanism in dendritic protein synthesis Zif268 in the expression of late LTP and long-term memories.
during long-term potentiation in adult awake rats. Eur. Nat. Neurosci., 4(3): 289–296.
J. Neurosci., 17(12): 2679–2689. Jung, M.W. and McNaughton, B.L. (1993) Spatial selectivity of
Hebb, D.O. (1949) The Organization of Behavior. Wiley, unit activity in the hippocampal granular layer. Hippocam-
New York. pus, 3: 165–182.
Henze, D.A., Wittner, L. and Buzsaki, G. (2002) Single granule Kali, S. and Dayan, P. (2000) The involvement of
cells reliably discharge targets in the hippocampal CA3 recurrent connections in area CA3 in establishing the
network in vivo. Nat. Neurosci., 5(8): 790–795. properties of place fields: a model. J. Neurosci., 20(19):
Hertz, J., Krogh, A. and Palmer, R.G. (1991) Introduction 7463–7477.
to the Theory of Neural Computation. Addison-Wesley Kapur, A., Yeckel, M.F. and Johnston, D. (2001) Hippocampal
Publishing, Redwood City, CA. mossy fiber activity evokes Ca2+ release in CA3 pyramidal
Hjorth-Simonsen, A. and Jeune, B. (1972) Origin and ter- neurons via a metabotropic glutamate receptor. J. Ne-
mination of the hippocampal perforant path in the rat urophysiol., 79(4): 2181–2190.
studied by silver impregnation. J. Comp. Neurol., 144: Kart-Teke, E., De Souza Silva, M.A., Huston, J.P. and
215–232. Dere, E. (2006) Wistar rats show episodic-like memory
Holland, L.L. and Wagner, J.J. (1998) Primed facilitation of for unique experiences. Neurobiol. Learn. Mem., 85(2):
homosynaptic long-term depression and depotentiation in rat 173–182.
hippocampus. J. Neurosci., 18(3): 887–894. Kato, A., Ozawa, F., Saitoh, Y., Hirai, K. and Inokuchi, K.
Hong, J.S., McGinty, J.F., Grimes, L., Kanamatsu, T., Obie, J. (1997) vesl, a gene encoding VASP/Ena family related pro-
and Mitchell, C.L. (1988) Seizure-induced alterations in the tein, is upregulated during seizure, long-term potentiation
metabolism of hippocampal opioid peptides suggest opioid and synaptogenesis. FEBS Lett., 412(1): 183–189.
modulation of seizure-induced behaviors. In: McGinty, J.F., O’Keefe, J. and Nadel, L. (1978) Hippocampus as a Cognitive
Friedman, D.P. (Eds.), Opioids in the Hippocampus. NIDA, Map. Oxford University Press, New York.
Rockville, MD, pp. 48–66. Kerr, D.S. and Abraham, W.C. (1995) Cooperative interactions
Hopfield, J.J. (1982) Neural networks and physical systems with among afferents govern the induction of homosynaptic long-
emergent collective computational abilities. Proc. Natl. Acad. term depression in the hippocampus. Proc. Natl. Acad. Sci.,
Sci. U.S.A., 79(8): 2554–2558. 92(25): 11637–11641.
Huang, Y.Y., Colino, A., Selig, D.K. and Malenka, R.C. Kerr, D.S. and Abraham, W.C. (1996) LTD: many means to
(1992) The influence of prior synaptic activity on the how many ends? Hippocampus, 6(1): 30–34.
induction of long-term potentiation. Science, 255(5045): Kirkwood, A., Dudek, S.M., Gold, J.T., Aizenman, C.D. and
730–733. Bear, M.F. (1993) Common forms of synaptic plasticity in
Huang, Y.Y., Kandel, E.R., Varshavsky, L., Brandon, the hippocampus and neocortex in vitro. Science, 260(5113):
E.P., Qi, M., Idzerda, R.L., McKnight, G.S. and 1518–1521.
Bourtchouladze, R. (1995) A genetic test of the effects Kleschevnikov, A.M. and Routtenberg, A. (2003) Long-term
of mutations in PKA on mossy fiber LTP and its potentiation recruits a trisynaptic excitatory associative net-
relation to spatial and contextual learning. Cell, 83(7): work within the mouse dentate gyrus. Eur. J. Neurosci.,
1211–1222. 18(6): 1717.
447

Kobayashi, K. and Poo, M.M. (2004) Spike train timing- Li, S., Cullen, W.K., Anwyl, R. and Rowan, M.J. (2003)
dependent associative modification of hippocampal Dopamine-dependent facilitation of LTP induction in hippo-
CA3 recurrent synapses by mossy fibers. Neuron, 41(3): campal CA1 by exposure to spatial novelty. Nat. Neurosci.,
445–454. 6(5): 526–531.
Kohonen, T. (1984) Self organization and Associative Memory. Lisman, J.E. and McIntyre, C.C. (2001) Synaptic plasticity:
Springer-Verlag, Berlin. a molecular memory switch. Curr. Biol., 11(19): R788–
Kohonen, T., Lehtio, P., Rovamo, J., Hyvarinen, J., Bry, K. R791.
and Vainio, L. (1977) A principle of neural associative mem- Magee, J.C. and Johnston, D. (1997) A synaptically-controlled,
ory. Neuroscience, 2: 1065–1076. associative signal for Hebbian plasticity in hippocampal neu-
Kosub, K., Do, V. and Derrick, B.E. (2005) NMDA antago- rons. Science, 275: 209–213.
nists block heterosynaptic LTD, but not LTP, following Maguire, C., Casey, M., Kelly, A., Mullany, P.M. and Lynch,
lateral perforant path stimulation. Neurosci. Lett., 374: M.A. (1999) Activation of tyrosine receptor kinase plays a
29–34. role in expression of long-term potentiation in the rat dentate
Krug, M., Brodemann, R. and Wagner. (2001) Simultaneous gyrus. Hippocampus, 9(5): 519–526.
activation and opioid modulation of long-term potentiation Malinow, R. (2003) AMPA receptor trafficking and long-term
in the dentate gyrus and the hippocampal CA3 region after potentiation. Philos. Trans. R. Soc. Lond. B Biol. Sci.,
stimulation of the perforant pathway in freely moving rats. 358(1432): 707–714.
Brain Res., 913(1): 68–77. Manabe, T. (1997) Two forms of hippocampal long-term
Krug, M., Lossner, B. and Ott, T. (1984) Anisomycin depression, the counterpart of long-term potentiation. Rev.
blocks the late phase of long-term potentiation in the Neurosci., 8(3–4): 179–193.
dentate gyrus of freely moving rats. Brain Res. Bull., 13(1): Manahan-Vaughan, D. (1998) Priming of group 2 metabotrop-
39–42. ic glutamate receptors facilitates induction of long-term de-
Kulla, A. and Manahan-Vaughan, D. (2000) Depotentiation pression in the dentate gyrus of freely moving rats.
in the dentate gyrus of freely moving rats is modulated Neuropharmacology, 37(12): 1459–1464.
by D1/D5 dopamine receptors. Cereb. Cortex, 10(6): O’Mara, S.M., Rowan, M.J. and Anwyl, R. (1995a) Metabo-
614–620. tropic glutamate receptor-induced homosynaptic long-term
Kulla, A. and Manahan-Vaughan, D. (2002) Modulation by depression and depotentiation in the dentate gyrus of the
serotonin 5-HT(4) receptors of long-term potentiation and rat hippocampus in vitro. Neuropharmacology, 34(8):
depotentiation in the dentate gyrus of freely moving rats. 983–989.
Cereb. Cortex, 12(2): 150–162. O’Mara, S.M., Rowan, M.J. and Anwyl, R. (1995b) Dantrolene
Kulla, A., Reymann, K.G. and Manahan-Vaughan, D. (1999) inhibits long-term depression and depotentiation of synaptic
Time-dependent induction of depotentiation in the dentate transmission in the rat dentate gyrus. Neuroscience, 68(3):
gyrus of freely moving rats: involvement of group 2 metabo- 621–624.
tropic glutamate receptors. Eur. J. Neurosci., 11(11): Marr, D. (1971) Simple Memory: a theory for archicorticortex.
3864–3872. Proc. R. Soc. Lond. B, 262: 23–81.
Lee, H.K., Barbarosie, M., Kameyama, K., Bear, M.F. and Martin, M.R. (1983) Naloxone and long-term potentiation of
Huganir, R.L. (2000) Regulation of distinct AMPA receptor hippocampal CA3 field potentials in vitro. Neuropeptides,
phosphorylation sites during bidirectional synaptic plasticity. 4(1): 45–50.
Nature, 405(6789): 955–959. Martin, S.J. (1998) Time-dependent reversal of dentate LTP by
Lee, I. and Kesner, R.P. (2004) Encoding versus retrieval of 5 Hz stimulation. Neuroreport, 9(17): 3775–3781.
spatial memory: double dissociation between the dentate Martinez, C.O., Do, V.H. and Derrick, B.E. (2002) Associative
gyrus and the perforant path inputs into CA3 in the dorsal LTP Among afferents to the CA3 region in vivo. Brain Res.,
hippocampus. Hippocampus, 14(1): 66–76. 94: 86–94.
Lei, S., Pelkey, K.A., Topolnik, L., Congar, P., Lacaille, J.C. Massey, P.V., Johnson, B.E., Moult, P.R., Auberson, Y.P.,
and McBain, C.J. (2003) Depolarization-induced long-term Brown, M.W., Molnar, E., Collingridge, G.L. and Bashir,
depression at hippocampal mossy fiber-CA3 pyramidal neu- Z.I. (2004) Differential roles of NR2A and NR2B-
ron synapses. J. Neurosci., 23(30): 9786–9795. containing NMDA receptors in cortical long-term pot-
Levy, W.B. (1996) A sequence predicting CA3 is a flexible entiation and long-term depression. J. Neurosci., 24:
associator that learns and uses context to solve hippocampal- 7821–7828.
like tasks. Hippocampus, 6(6): 579–590. Matthies, H., Schroeder, H., Becker, A., Loh, H., Hollt, V. and
Levy, W.B. and Steward, O. (1979) Synapses as associative Krug, M. (2000) Lack of expression of long-term potentiat-
memory elements in the hippocampal formation. Brain Res., ion in the dentate gyrus but not in the CA1 region of the
175: 233–245. hippocampus of mu-opioid receptor-deficient mice. Neuro-
Levy, W.B. and Steward, O. (1983) Temporal contiguity re- pharmacology, 39(6): 952–960.
quirements for long-term associative potentiation/depression McClelland, J.L., McNaughton, B.L. and O’Reilly, R.C. (1995)
in the hippocampus. Neuroscience, 8(4): 791–797. Why there are complementary learning systems in the
448

hippocampus and neocortex: insights from the successes and Nguyen, P.V. and Kandel, E.R. (1996) A macromolecular
failures of connectionist models of learning and memory. synthesis-dependent late phase of long-term potentia-
Psychol. Rev., 102: 419–457. tion requiring cAMP in the medial perforant pathway
McMahon, D.B. and Barrionuevo, G. (2002) Short- and long- of rat hippocampal slices. J. Neurosci., 16(10):
term plasticity of the perforant path synapse in hippocampal 3189–3198.
area CA3. J. Neurophysiol., 88(1): 528–533. North, R.A., Williams, J.T., Surprenant, A. and Christie, M.J.
McNaughton, B.L. and Barnes, C.A. (1977) Physiological iden- (1987) Mu and delta receptors belong to a family of receptors
tification and analysis of dentate granule cell responses to that are coupled to potassium channels. Proc. Natl. Acad.
stimulation of the medial and lateral perforant pathways in Sci. U.S.A., 84(15): 5487–5491.
the rat. J. Comp. Neurol., 175: 439–454. Otani, S., Marshall, C.J., Tate, W.P., Goddard, G.V. and
McNaughton, B.L., Barnes, C.A., Meltzer, J. and Sutherland, Abraham, W.C. (1989) Maintenance of long-term potentiat-
R.J. (1989) Hippocampal granule cells are necessary for nor- ion in rat dentate gyrus requires protein synthesis but not
mal spatial learning but not for spatially-selective pyramidal messenger RNA synthesis immediately post-tetanization.
cell discharge. Exp. Brain Res., 76: 485–496. Neuroscience, 28(3): 519–526.
McNaughton, B.L., Douglas, R.M. and Goddard, D.V. (1978) Otmakhova, N.A. and Lisman, J.E. (1998) D1/D5 dopamine
Synaptic enhancement in fascia dentata: cooperativity among receptors inhibit depotentiation at CA1 synapses via
coactive afferents. Brain Res., 157: 277–293. cAMP-dependent mechanism. J. Neurosci., 18(4):
Mellor, J. and Nicoll, R.A. (2001) Hippocampal mossy fiber 1270–1279.
LTP is independent of postsynaptic calcium. Nat. Neurosci., Pang, K., Williams, M.J. and Olton, D.S. (1993) Activation
4(2): 125–126. of the medial septal area attenuates LTP of the lateral
Milner, T.A. and Drake, C.T. (2001) Ultrastructural perforant path and enhances heterosynaptic LTD of the
evidence for presynaptic mu opioid receptor modulation medial perforant path in aged rats. Brain Res., 632(1–2):
of synaptic plasticity in NMDA-receptor-containing dend- 150–160.
rites in the dentate gyrus. Brain Res. Bull., 54(2): 131–140. Philpot, B., Bear, M.F. and Abraham, W.C. (1999) Metaplas-
Mizumori, S.J., McNaughton, B.L., Barnes, C.A. and Fox, ticity: the plasticity of synaptic plasticity. In: Katz P.S. (Ed.),
K.B. (1989) Preserved spatial coding in hippocampal CA1 Beyond Neurotransmission: Neuromodulation and its
pyramidal cells during reversible suppression of CA3c Importance for Information processing. Oxford, London,
output: evidence for pattern completion in hippocampus. pp. 161–197.
J. Neurosci., 9(11): 3915–3928. Piguet, P. and North, R.A. (1983) Opioid actions at mu and
Monaghan, D.T. and Cotman, C.W. (1985) Distribution of delta receptors in the rat dentate gyrus in vitro. J. Pharmacol.
N-methyl-D-aspartate sensitive l-[3H]glutamate-binding sites Exp. Ther., 266(2): 1139–1149.
in rat brain. J. Neurosci., 5: 2909–2919. Ramirez-Amaya, V., Balderas, I., Sandoval, J., Escobar, M.L.
Morgan, S.L. and Teyler, T.J. (2001) Electrical stimuli pat- and Bermudez-Rattoni, F. (2001) Spatial long-term memory
terned after the theta-rhythm induce multiple forms of LTP. is related to mossy fiber synaptogenesis. J. Neurosci., 21(18):
J. Neurophysiol., 86(3): 1289–1296. 7340–7348.
Morris, R.G.M. (1989) Computational neuroscience: modeling Ramon y Cajal, S. (1937) Recuerdos de mi Vida. MIT Press,
the brain. In: Morris R.G.M. (Ed.), Parallel-Distributed Cambridge.
Processing: Implication for Psychology and Neurobiology. Reid, C.A., Dixon, D.B., Takahashi, M., Bliss, T.V. and Fine,
Oxford University Press, NY, pp. 203–212. A. (2004) Optical quantal analysis indicates that long-term
Morris, R.G.M. and McNaughton, B.L. (1987) Memory stor- potentiation at single hippocampal mossy fiber synapses is
age in a distributed model of hippocampal formation. Trends expressed through increased release probability, recruitment
Neurosci., 10: 408–414. of new release sites, and activation of silent synapses. J. Ne-
Mott, D.D. and Lewis, D.V. (1992) GABAB receptors mediate urosci., 24(14): 3618–3626.
disinhibition and facilitate long-term potentiation in the Rolls, E.T. and Treves, A. (1998) Neural Networks and Brain
dentate gyrus. Epilepsy Res. Suppl., 7: 119–134. Function. Oxford University Press, New York.
Munoz, M.D., Nunez, A. and Garcia-Austt, E. (1990) In vivo Rosenzweig, E.S., Barnes, C.A. and McNaughton, B.L.
intracellular analysis of rat dentate gyrus granule cells. Brain (2002) Making room for new memories. Nat. Neurosci., 5:
Res., 509: 91–98. 6–8.
Nakazawa, K., Quirk, M.C., Chitwood, R.A., Watanabe, M., Rycroft, B.K. and Gibb, A.J. (2002) Direct effects of
Yeckel, M.F., Sun, L.D., Kato, A., Carr, C.A., Johnston, D., calmodulin on NMDA receptor single channel gating
Wilson, M.A. and Tonegawa, S. (2002) Requirement for in rat hippocampal granule cells. J. Neurosci., 22: 8860–
hippocampal CA3 NMDA receptors in associative memory 8868.
recall. Science, 297(5579): 211–218. Sanberg, C.-D., Jones, F., Do, V., Dieguez, D. and Derrick,
Neumaier, J.F. and Chavkin, C. (1989) Release of endogenous B.E. (2006) Antagonists of the 5-HT1a receptor block dent-
opioid peptides displaces [3H]diprenorphine binding in rat ate LTP induction in novel, but not familiar environments.
hippocampal slices. Brain Res., 493: 292–302. Learn. Mem., 13(1): 52–62.
449

Scharfman, H.E. (1996) Conditions required for polysynaptic for the IEG Arc to localize selectively near activated
excitation of dentate granule cells by area CA3 pyra- postsynaptic sites on dendrites. Neuron, 21(4): 741–
midal cells in rat hippocampal slices. Neuroscience, 72(3): 751.
655–658. Straube, T., Korz, V., Balschun, D. and Frey, J.U. (2003)
Scharfman, H.E., Sollas, A.L., Smith, K.L., Jackson, M.B. and Requirement of beta-adrenergic receptor activation and
Goodman, J.H. (2002) Structural and functional asymmetry protein synthesis for LTP-reinforcement by novelty in rat
in the normal and epileptic rat dentate gyrus. J. Comp. dentate gyrus. J. Physiol., 552(Pt 3): 953–960.
Neurol., 454(4): 424–439. Swearengen, E. and Chavkin, C. (1987) NMDA receptor
Schulz, S., Siemer, H., Krug, M. and Hollt, V. (1999) Direct antagonist D-APV depresses excitatory activity produced
evidence for biphasic cAMP responsive element-binding by normorphine in rat hippocampal slices. Neurosci. Lett.,
protein phosphorylation during long-term potentiation 78: 80–84.
in the rat dentate gyrus in vivo. J. Neurosci., 19(13): Sweatt, J.D. (2001) The neuronal MAP kinase cascade: a
5683–5692. biochemical signal integration system subserving synaptic
Schwegler, H., Crusio, W.E. and Brust, I. (1990) Hippocampal plasticity and memory. J. Neurochem., 76(1): 1–10.
mossy fibers and radial-maze learning in the mouse: a Thiels, E., Xie, X., Yeckel, M.F., Barrionuevo, G. and Berger,
correlation with spatial working memory but not with T.W. (1996) NMDA receptor-dependent LTD in different
non-spatial reference memory. Neuroscience, 34(2): subfields of hippocampus in vivo and in vitro. Hippocampus,
293–298. 6(1): 43–51.
Scoville, W.B. and Milner, B. (1957) Loss of recent memory by Thomas, K.L., Davis, S., Hunt, S.P. and Laroche, S. (1996)
bilateral hippocampal lesions. J. Neurol. Neurosurg. Psychi- Alterations in the expression of specific glutamate receptor
atry, 20: 11–21. subunits following hippocampal LTP in vivo. Learn. Mem.,
Segal, M. (1972) Hippocampal unit responses to perforant path 3(2–3): 197–208.
stimulation. Exp. Neurol., 35: 541–546. Thomas, K.L., Laroche, S., Errington, M.L., Bliss, T.V.
Sheng, M. and Greenberg, M.E. (1990) The regulation and and Hunt, S.P. (1994) Spatial and temporal changes in sig-
function of c-fos and other immediate early genes in the nal transduction pathways during LTP. Neuron, 13(3):
nervous system. Neuron, 4(4): 477–485. 737–745.
Shors, T.J. and Matzel, L.D. (1997) Long-term potentiation: Trepel, C. and Racine, R.J. (1998) Long-term potentiation in
what’s learning got to do with it? Behav. Brain Sci., 20(4): the neocortex of the adult, freely moving rat. Cereb. Cortex,
597–614. 8(8): 719–729.
Shouval, H., Intrator, N. and Cooper, L.N. (1997) BCM Treves, A. and Rolls, E.T. (1992) Computational con-
network develops orientation selectivity and ocular straints suggest the need for two distinct input systems
dominance in natural scene environment. Vision Res., to the hippocampal CA3 network. Hippocampus, 2(2):
37(23): 3339–3342. 189–199.
Skaggs, W.E. and McNaughton, B.L. (1992) Computational Treves, A. and Rolls, E.T. (1994) Computational analysis of the
approaches to hippocampal function. Curr. Opin. Neuro- role of the hippocampus in memory. Hippocampus, 4:
biol., 2(2): 209–211. 374–391.
Snyder, J.S., Kee, N. and Wojtowicz, J.M. (2001) Effects of Treves, A. and Roudi, Y. (2005) On the evolution of the brain.
adult neurogenesis on synaptic plasticity in the rat dentate In: Chow C.C., Gutkin B., Hansel D. and Meunier C. (Eds.),
gyrus. J. Neurophysiol., 85(6): 2423–2431. Methods and Models in Neurophysics. Elsevier, Holland,
Soderling, T.R. and Derkach, V.A. (2000) Postsynaptic pp. 641–690.
protein phosphorylation and LTP. Trends Neurosci., 23(2): Trommald, M., Hulleberg, G. and Andersen, P. (1996) Long-
75–80. term potentiation is associated with new excitatory spine
Squire, L. (1992) Memory and the hippocampus: a synthesis synapses on rat dentate granule cells. Learn. Mem., 3(2–3):
from findings with rats, monkeys, and humans. Psychol. 218–228.
Rev., 99: 195–231. Tulving, E. (1983) Elements of Episodic Memory. Clarendon
Staubli, U. and Lynch, G. (1987) Stable hippocampal long-term Press, Oxford.
potentiation elicited by ‘theta’ pattern stimulation. Brain Urban, N.N. and Barrionuevo, G. (1996) Induction of Hebbian
Res., 435(1–2): 227–234. and non-Hebbian mossy fiber long-term potentiation by dis-
Steward, O. (1976) Topographic organization of the projections tinct patterns of high-frequency stimulation. J. Neurosci.,
from the entorhinal area to the hippocampal formation of the 16(13): 4293–4299.
rat. J. Comp. Neurol., 167: 285–314. Villarreal, D., Do, V., Haddad, E. and Derrick, B.E. (2002)
Steward, O. and Scoville, S.A. (1976) Cells of origin of ent- NMDA antagonists sustain LTP and spatial memory:
orhinal cortical afferents to the hippocampus and fascia den- evidence for active processes underlying LTP decay. Nat.
tata of the rat. J. Comp. Neurol., 169: 347–370. Neurosci., 5: 48–52.
Steward, O., Wallace, C.S., Lyford, G.L. and Worley, Villarreal, D.M. and Derrick, B.E. (2000) Long-term
P.F. (1998) Synaptic activation causes the mRNA metaplasticity: correspondence with LTP longevity
450

and the BCM Theory. Abst. Soc. for Neurosci., # 134.1 hippocampal synaptic plasticity. Neuroscience, 118(4):
(online). 1003–1013.
Villarreal, D.M. and Derrick, B.E. (2006) Input-specific Williams, S. and Johnston, D. (1989) Long-term potentiation
long-term potentiation reinduction at the rat perforant of hippocampal mossy fiber synapses is blocked by
path to dentate gyrus. Soc. for Neurosci. Abst., # 731.2 postsynaptic injection of calcium chelators. Neuron, 3:
(online). 583–588.
Wagner, J.J., Caudle, R.M., Neumaier, J.F. and Chavkin, C. Williams, S.H. and Johnston, D. (1996) Actions of endogenous
(1990) Stimulation of endogenous opioid release displaces opioids on NMDA receptor-independent long-term potent-
mu receptor binding in rat hippocampus. Neuroscience, 37: iation in area CA3 of the hippocampus. J. Neurosci., 16:
45–53. 3652–3660.
Wagner, J.J., Etemad, L.R. and Thompson, A.M. (2001) Wilson, M.A. and McNaughton, B.L. (1993) Dynamics of the
Opioid-mediated facilitation of long-term depression hippocampal ensemble code for space. Science, 261(5124):
in rat hippocampus. J. Pharmacol. Exp. Ther., 296(3): 1055–1058.
776–781. Winson, J. and Abzug, C. (1978) Dependence upon behavior
Wagner, J.J., Evans, C.J. and Chavkin, C. (1991) Focal stim- of neuronal transmission from perforant pathway through
ulation of the mossy fibers releases endogenous dynorphins entorhinal cortex. Brain Res., 147(2): 422–427.
that bind to k1-opioid receptors in guinea pig hippocampus. Witter, M.P., Griffioen, A.W., Jorritsma-Byham, B. and
J. Neurochem., 57: 333–343. Krijnen, J.L. (1988) Entorhinal projections to the hippocam-
Wagner, J.J., Terman, G.W. and Chavkin, C. (1993) Endog- pal CA1 region in the rat: an underestimated pathway.
enous dynorphins inhibit excitatory neurotransmission and Neurosci. Lett., 85(2): 193–198.
block LTP induction in the hippocampus. Nature, 363(6428): Wu, J., Rowan, M.J. and Anwyl, R. (2004) Synaptically
451–454. stimulated induction of group I metabotropic glutamate
Wang, D., Tolbert, L.M., Carlson, K.W. and Sadee, W. receptor-dependent long-term depression and depotentiation
(2000) Nuclear Ca2+/calmodulin translocation activated is inhibited by prior activation of metabotropic glutamate
by mu-opioid (OP3) receptor. J. Neurochem., 74(4): receptors and protein kinase C. Neuroscience, 123(2):
1418–1425. 507–514.
Wang, J., Yeckel, M.F., Johnston, D. and Zucker, R.S. Wu, J., Rowan, M.J. and Anwyl, R. (2006) Long-term potent-
(2004) Photolysis of postsynaptic caged Ca2+ can potenti- iation is mediated by multiple kinase cascades involv-
ate and depress mossy fiber synaptic responses in rat ing CaMKII or either PKA or p42/44 MAPK in the
hippocampal CA3 pyramidal neurons. J. Neurophysiol., adult rat dentate gyrus in vitro. J. Neurophysiol., 95(6):
91(4): 1596–1607. 3519–3527.
Weisskopf, M.G., Zalutsky, R.A. and Nicoll, R.A. (1993) Xie, C.-W. and Lewis, D.V. (1991) Opioid mediated facilitation
The opioid peptide dynorphin mediates heterosynaptic of long-term potentiation at the lateral perforant path-
depression of hippocampal mossy fibre synapses dentate granule cell synapse. J. Pharmacol. Exp. Ther., 256:
and modulates long-term potentiation. Nature, 362: 289–296.
423–427. Yeckel, M.F. and Berger, T.W. (1990) Feedforward excitation
Wexler, E.M. and Stanton, P.K. (1993) Priming of homosy- of the hippocampus by afferents from the entorhinal cortex:
naptic long-term depression in hippocampus by previous redefinition of the role of the trisynaptic pathway. Proc. Natl.
synaptic activity. Neuroreport, 4(5): 591–594. Acad. Sci. U.S.A., 87: 5832–5836.
White, J.D., Gall, C.M. and McKelvy, J.F. (1987) Enke- Yeckel, M.F. and Berger, T.W. (1998) Spatial distribution of
phalin biosynthesis and enkephalin gene expression potentiated synapses in hippocampus: dependence on cellular
are increased in hippocampal mossy fibers following mechanisms and network properties. J. Neurosci., 18:
a unilateral lesion of the hilus. J. Neurosci., 7(3): 438–450.
753–759. Yeckel, M.F., Kapur, A. and Johnston, D. (1999) Multiple
Wickens, J.R. and Abraham, W.C. (1991) The involvement forms of LTP in hippocampal CA3 neurons use a common
of L-type calcium channels in heterosynaptic long-term postsynaptic mechanism. Nat. Neurosci., 2(7): 625–633.
depression in the hippocampus. Neurosci. Lett., 130(1): Ying, S.W., Futter, M., Rosenblum, K., Webber, M.J., Hunt,
128–132. S.P., Bliss, T.V. and Bramham, C.R. (2002) Brain-derived
Wigstrom, H. and Gustaffson, B. (1985) Facilitation of hippo- neurotrophic factor induces long-term potentiation in intact
campal long-lasting potentiation by GABA antagonists. Acta adult hippocampus: requirement for ERK activation coupled
Physiol. Scand., 125: 159–172. to CREB and upregulation of Arc synthesis. J. Neurosci.,
Williams, J.M., Guevremont, D., Kennard, J.T., Mason- 22(5): 1532–1540.
Parker, S.E., Tate, W.P. and Abraham, W.C. (2003) Zalutsky, R.A. and Nicoll, R.A. (1990) Comparison of two
Long-term regulation of N-methyl-D-aspartate receptor forms of long-term potentiation in single hippocampal neu-
subunits and associated synaptic proteins following rons. Science, 248: 1619–1624.
451

Zeineh, M.M., Engel, S.A., Thompson, P.M. and Bookheimer, Zimprich, A., Simon, T. and Hollt, V. (1995) Transfected rat
S.Y. (2003) Dynamics of the hippocampus during encoding mu opioid receptors (rMOR1 and rMOR1B) stimulate
and retrieval of face-name pairs. Science, 299(5606): 577–580. phospholipase C and Ca2+ mobilization. Neuroreport, 7(1):
Zhang, D.X. and Levy, W.B. (1992) Ketamine blocks the 54–56.
induction of LTP at the lateral entorhinal cortex-dentate Zucker, R.S. and Regehr, W.G. (2002) Short-term synaptic
gyrus synapses. Brain Res., 593(1): 124–127. plasticity. Annu. Rev. Physiol., 64: 355–405.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 25

Control of synaptic consolidation in the dentate


gyrus: mechanisms, functions, and therapeutic
implications

Clive R. Bramham

Department of Biomedicine and Bergen Mental Health Research Center, University of Bergen, Jonas Lies vei 91,
N-5009 Bergen, Norway

Abstract: Synaptic consolidation refers to the development and stabilization of protein synthesis-depend-
ent modifications of synaptic strength as observed during long-term potentiation (LTP) and long-term
depression (LTD). Activity-dependent changes in synaptic strength are thought to underlie memory storage
and other adaptive responses of the nervous systems of importance in mood stability, reward behavior, and
pain control. This chapter focuses on the mechanisms and functions of synaptic consolidation in the
dentate gyrus, a critical structure not only in hippocampal memory function, but also in regulation of stress
responses and cognitive aspects of depression. Recent evidence suggests that synaptic consolidation at
excitatory medial perforant path-granule cell synapses requires brain-derived neurotrophic factor (BDNF)
signaling and induction of the immediate early gene activity-regulated cytoskeleton-associated protein
(Arc). Arc mRNA is strongly induced and transported to dendritic processes following high-frequency
stimulation (HFS) that induces LTP in the rat dentate gyrus in vivo. Sustained synthesis of Arc during a
surprisingly protracted time-window is required for hyperphosphorylation of actin depolymerizing factor/
cofilin and local expansion of the actin cytoskeleton in vivo. Furthermore, this process of Arc-dependent
synaptic consolidation is activated in response to brief infusion of BDNF. Microarray expression profiling
has revealed a panel of BDNF-regulated genes that may cooperate with Arc during synaptic consolidation.
In addition to regulating gene expression, BDNF signaling modulates the fine localization and biochemical
activation of the translation machinery. By modulating the spatial and temporal translation of newly
induced (Arc) and constitutively-expressed mRNA in dendrites, BDNF may effectively control the window
of synaptic consolidation. Dysregulation of BDNF synthesis and Arc function, specifically within the
dentate gyrus, is linked to behavioral symptoms and cognitive deficits in animal models of depression and
Alzheimer’s disease. Therapeutics strategies targeting synaptic consolidation hold promise for the future.

Keywords: long-term potentiation; brain-derived neurotrophic factor; BDNF; Neurotrophin; hippocampus;


Arc; gene expression; Depression; memory

Corresponding author. Tel.: +47 55 58 60 32; Fax: +47 55 58


64 10; E-mail: clive.bramham@biomed.uib.no

DOI: 10.1016/S0079-6123(07)63025-8 453


454

Control of synaptic consolidation in the dentate triggers synaptic consolidation at adult excitatory
gyrus synpases. BDNF, unlike other members of the neu-
rotrophin family (nerve growth factor, neurotro-
Synaptic consolidation phin-3, and neurotrophin-4), is widely distributed
across the adult brain and in hippocampal princi-
Persistent activity-dependent changes in synaptic pal cells. BDNF signals through catalytic TrkB
strength are believed to underlie a range of adaptive receptors present on pre- and postsynaptical
brain responses, including memory formation, elements of glutamatergic synapses (Drake et al.,
mood stability, and drug addiction (Bliss and 1999). Postsynaptic TrkB receptors are found
Collingridge, 1993; Nestler et al., 2002). However, in the postsynaptic density (PSD) and TrkB
our understanding of the mechanisms by which co-immunoprecipitates with NMDAR complex
altered activity patterns trigger lasting changes proteins (Wu et al., 1996; Drake et al., 1999; Aoki
in synaptic efficacy, exemplified by long-term et al., 2000; Husi et al., 2000). Postsynaptic release
potentiation (LTP) and long-term depression of BDNF from secretogranin II-immunoreactive
(LTD), is far from complete. A critical factor in vesicles occurs in response to HFS, and is sensitive
high-frequency stimulation (HFS)-induced LTP is to NMDAR blockade. BDNF may therefore be
postsynaptic activation of N-methyl-D-aspartate viewed as a co-neurotransmitter acting in tandem
receptors (NMDARs) and voltage-dependent with glutamate at excitatory synapses. Two addi-
calcium channels. The resulting activation of tional features important to mention in the context
calcium-sensitive kinases and other calcium-sensi- of synaptic consolidation are: (1) BDNF regulates
tive effectors underlies a transient, early phase of protein synthesis through both transcriptional and
LTP maintained by protein phosphorylation and post-transcriptional mechanisms, and (2) BDNF
membrane insertion of AMPA-type glutamate is capable of stimulating its own release, possibly
receptors. The generation of stable, late phase allowing sustained, regenerative signaling at
LTP requires at least one period of new gene synapses (Santi et al., 2006).
expression and protein synthesis (Bliss and Col- Genetic and pharmacological studies establish-
lingridge, 1993; Frey et al., 1996; Nguyen and ing the role of BDNF-TrkB in LTP have been
Kandel, 1996). Analogous to memory consolida- reviewed in detail elsewhere (Bramham and
tion, the process of generating late LTP is referred Messaoudi, 2005). The contributions of BDNF
to as synaptic consolidation. signaling to LTP can be classified as permissive or
Late phase LTP is operationally defined on the instructive. Permissive mechanisms make synapses
basis of sensitivity of LTP maintenance to broad competent for LTP. For example, constitutive
spectrum inhibitors of global gene expression and (non-evoked) presynaptic BDNF signaling in-
protein synthesis. Progress in unraveling the mo- creases the fraction of neurotransmitter vesicles
lecular basis of synaptic consolidation has been docked in the active zone. This increase in the
hampered by the lack of evidence that expression of readily releasable pool of vesicles enables sustained
a specific gene product or group of gene products glutamate release during trains of action poten-
mediates LTP. Alterations in gene expression in re- tials, and facilitates LTP induction (Figurov et al.,
sponse to changes in neuronal activity are common, 1996). Instructive BDNF signaling, in contrast, is
and, as pointed out by several authors (Sanes and initiated in response to HFS and is required for
Lichtman, 1999; Routtenberg and Rekart, 2005), it LTP development. Acute release of BDNF during
is likely that many regulated genes have no role or HFS modulates the induction and early mainte-
only a subsidiary role in generating stable LTP. nance of LTP (Kossel et al., 2001). A series of
studies based on focal application of BDNF sug-
BDNF as a trigger of synaptic consolidation gests a mechanism involving activation of a
tetrodotoxin (TTX)-insensitive voltage-dependent
Several lines of evidence suggest that BDNF, act- sodium channel leading to rapid depolarization
ing through its receptor tyrosine kinase (TrkB) and enhanced calcium influx into dendritic spines
455

during LTP induction (Blum and Konnerth, polymerization (Matsuzaki et al., 2004). Insertion
2005). The mechanism of TrkB coupling to of glutamate receptors at postsynaptic mem-
the TTX-insensitive channel is unclear at the branes, increase in PSD diameter, and structural
moment. remodeling of spines are all connected to regula-
Formation of late LTP is coupled to a period of tion of actin dynamics (Geinisman, 2000; Lisman
sustained BDNF release and TrkB receptor acti- and Zhabotinsky, 2001; Weeks et al., 2001; Zhou
vation (Kang et al., 1997; Aicardi et al., 2004). et al., 2001; Fukazawa et al., 2003; Harris et al.,
Extracellular concentrations of BDNF are ele- 2003; Matsuzaki et al., 2004; Okamoto et al., 2004;
vated for less than a minute following weak HFS, Zito et al., 2004; Oertner and Matus, 2005).
leading to decremental early LTP; whereas Fukazawa et al. (2003) showed that LTP at
stimulus protocols that induce late LTP produce MPP-granule cell synapses in anesthetized rats
BDNF elevation lasting at least 5 min. Pharmaco- is associated with an increase in filamentous actin
logical inhibition of TrkB receptor activation (F-actin) content at activated synapses and
during the first hour after HFS blocks synaptic enhanced phosphorylation of cofilin, a major
consolidation in hippocampal region CA1 (Kang regulator of actin dynamics. Phosphorylation of
et al., 1997). Furthermore, bath application of cofilin on Ser3 inhibits activity and promotes actin
BDNF induces a protein synthesis-dependent polymerization. LIM domain kinase (LIMK), one
increase in synaptic efficacy termed BDNF-LTP of the major cofilin kinases in brain, modulates the
(Kang and Schuman, 1996; Messaoudi et al., morphology of dendritic spines through regulation
1998). Induction of BDNF-LTP at medial perfo- of cofilin activity. Mice lacking the LIMK1 gene
rant path (MPP)-granule cell synapses in the dent- exhibit small, actin-deficient spines. Although the
ate gyrus is blocked by inhibitors of RNA initial amplitude of LTP is actually larger, LIMK
synthesis and occluded by prior expression of knockout mice fail to sustain late phase LTP
late phase, but not early phase, of HFS-LTP (Meng et al., 2002). Taken together, this suggests a
(Messaoudi et al., 2002; Ying et al. 2002). BDNF- major role for cofilin regulation in actin-dependent
LTP lasts at least 1 day in awake rats and, like synapse expansion underlying late LTP.
HFS-LTP, is associated with increases in both How is gene expression coupled to actin dy-
synaptic efficacy and EPSP-spike coupling. namics and synapse expansion? LTP is associated
BDNF-LTP induction is not affected by NMDA with the induction of a number of immediate-early
receptor blockade, consistent with the hypothesis genes that encode a variety of transcription factors
that NMDA receptor activation stimulates and other proteins, unassociated with transcrip-
BDNF release but is not involved in BDNF-LTP tion, and some of these have been causally linked
induction or expression. Like HFS-LTP, BDNF- to late phase LTP and long-term memory. Arc is
LTP induction requires extracellular signal- unique because it encodes the only mRNA known
regulated protein kinase (ERK) activation and is to undergo transport to distal dendritic processes
coupled to upregulation, dendritic transport, and of granule cells within the first hour of LTP in-
translation of the immediate early gene activity- duction (Link et al., 1995; Lyford et al., 1995). Arc
regulated cytoskeleton-associated protein (Arc, mRNA is enriched at stimulated synapses, and Arc
aka Arg 3.1). These results suggest that endog- protein is transiently elevated in dendritic spines
enous BDNF activates synaptic consolidation and following LTP induction (Steward and Worley,
that exogenous BDNF mimics this effect. 2001; Moga et al., 2004; Rodriguez et al., 2005).
Arc co-precipitates with crude F-actin and is
found in the PSD of excitatory synapses (Lyford
Arc as mediator of synapse consolidation et al., 1995; Husi et al., 2000). Arc is dynamically
expressed in principal neurons of many cortical
Persistent LTP is thought to occur when small and limbic structures during behavioral training,
spines are converted to large mushroom-shaped and this expression is necessary for long-term
spines through a mechanism dependent on actin memory in a variety of memory tasks (Guzowski
456

et al., 1999, 2000; Plath et al., 2006; Vazdarjanova only transiently inhibited when Arc antisense
et al., 2006). is administered shortly before (5 min) HFS, sug-
Early work of Guzowski et al. using Arc anti- gesting that sustained synthesis of Arc during a
sense oligodeoxynucleotides (AS) suggested a role time-window lasting at least 2 h is necessary to
for Arc in LTP maintenance in the dentate gyrus. consolidate LTP. These results suggest a pivotal
This issue has now been examined in more detail role for activity-induced Arc synthesis in control
(Messaoudi et al., 2007). These authors show that of synaptic consolidation and local expansion of
Arc antisense application at 2 h (but not 4 h) after the actin cytoskeleton. Furthermore, Arc AS
LTP induction leads to rapid and complete re- infusion blocks BDNF-LTP and reverses mainte-
versal of LTP. LTP reversal is coupled to rapid nance of BDNF-LTP over a critical time-window.
knockdown of newly synthesized Arc mRNA and Thus, Arc-dependent consolidation is directly ac-
protein, dephosphorylation of hyperphosphory- tivated by local BDNF application. Figure 1
lated cofilin, and loss of a discrete band of phal- presents a working model of Arc function in LTP.
loidin staining in the dentate molecular layer, Although the exact role of actin polymerization
indicating disappearance of new F-actin at MPP in synapse expansion is unknown, this process
synapses. Introduction of the F-actin stabilizer must involve the addition of newly synthesized or
jasplakinolide during LTP maintenance blocks the translocated proteins to the postsynaptic special-
reversal of LTP by Arc AS. Interestingly, LTP is ization. Rather than working alone, Arc is likely to

Fig. 1. BDNF as a trigger of synaptic consolidation. The mechanism of stable LTP formation at glutamatergic synapses is presented as
a two-stage process: translation activation and Arc-dependent consolidation. In the translation activation stage, patterned high-
frequency stimulation leads to sustained postsynaptic release of BDNF and activation of TrkB receptors pre- and postsynaptically.
Postsynaptic TrkB leads to (1) rapid activation and translocation of the translation machinery in dendritic spines, and (2) Arc
transcription in granule cell bodies. Translation activation is mediated by phosphorylation of the cap-binding protein eIF4E, and
possibly by more mRNA-specific mechanisms such as relief of miRNA-mediated translation repression. Spines activated in this way
may effectively capture and translate local mRNA pools. In this model, transcripts liberated from local RNA storage granules are
translated first, followed by dendritic transport and sustained translation of newly synthesized Arc mRNA. During Arc-dependent
consolidation, sustained translation of Arc is necessary for cofilin phosphorylation, local F-actin expansion, and formation of stable
LTP. This is a working model; several points require further experimental validation (see text). Adapted from Soule et al. (2006) with
permission from the Biochemical Society.
457

be part of a coordinate process of synapse growth This raises the intriguing possibility that zif268
involving rapid activity-induced regulation of stimulates the prolonged increase in Arc induction
many of genes. Wibrand et al. (2006) used cDNA following HFS-LTP in the dentate gyrus. The fact
microarray expression profiling to screen for genes that zif268 is not induced during BDNF-LTP
that are co-upregulated with Arc mRNA 3 h after suggests that Arc-dependent consolidation is in-
BDNF-LTP induction in the dentate gyrus of dependent of zif268 (Messaoudi et al., 2002;
anesthetized rats. Of nine genes upregulated Ying et al., 2002; Wibrand et al., 2006). Endog-
more than fourfold, five were selected for inde- enous BDNF may nonetheless contribute to zif268
pendent confirmation. Real-time PCR confirmed expression during HFS-LTP.
robust upregulation of all five genes: neuronal BDNF gene expression is enhanced following
activity-regulated pentraxin (Narp), neuritin, HFS-LTP induction in the CA1 region and dent-
ADP-ribosylation factor-like protein-4 (ARL4L), ate gyrus (Patterson et al., 1992; Castren et al.,
TGF-b-induced immediate early gene-1 (TIEG1), 1993; Bramham et al., 1996; Lee et al., 2005b).
and calcium/calmodulin kinase-related peptide LTP at MPP-granule cell synapses of awake rats
(CARP). In situ hybridization histochemistry is associated with NMDA receptor-dependent in-
further revealed upregulation of these genes in creases in TrkB and BDNF mRNA expression
somata of postsynaptic granule cells following (Bramham et al., 1996). TrkB mRNA is elevated
BDNF-LTP and HFS-LTP. Consistent with a unilaterally in dentate granule cells 2 h after LTP
process of synaptic growth and remodeling, Narp, induction, whereas BDNF mRNA is elevated
neuritin, and TIEG1 are involved in synaptogen- bilaterally at 6 and 24 h after LTP induction. The
esis, axon guidance, and glutamate receptor bilateral increase in BDNF expression is coupled
clustering during development. to unilateral LTP induction, as commissural re-
Another immediate early gene critical for late sponses and contralateral MPP-granule cell re-
phase LTP in the dentate gyrus and long-term sponses were unaffected. Conceivably, such
memory is zif286, one of the four members of the bilateral effects are due to altered network prop-
early growth response (egr) family of zinc-finger erties following LTP (Dragoi et al., 2003). Unilat-
transcription factors. Mice with constitutive eral increases in BDNF gene expression occur in
deletion of zif268 (aka egr1) show impaired dentate granule cells 3 h after BDNF-LTP induc-
LTP maintenance one day after HFS as well as tion. Much of this response is due to regulation of
impaired long-term memory consolidation and the highly calcium-responsive exon III BDNF
reconsolidation (Jones et al., 2001; Bozon et al., promoter (Tao et al., 1998; Wibrand et al.,
2003). Other studies examining the effects of an- 2006). This suggests that BDNF signaling regu-
tisense injection into the hippocampus concluded lates BDNF transcription during long-term synap-
that synthesis of BDNF, but not zif268, is involved tic plasticity. However, the key question, whether
in initial memory consolidation (Lee et al., 2004). new BDNF transcription actually contributes to
Conversely, zif268, but not BDNF, synthesis was synaptic strengthening, remains to be addressed.
implicated in reconsolidation. It is just as likely that BDNF transcription
Immediate early gene responses are typically serves only to replenish depleted stores of protein
rapid and protein synthesis-independent. In addi- following intensive BDNF secretion.
tion to the classical protein synthesis-independent
induction, Arc mRNA is induced through a
delayed protein synthesis-dependent mechanism. BDNF and translation control in dendrites
Interestingly, recent work has identified Arc as a
direct transcriptional target of zif268 and egr-3 Synaptic consolidation requires transcription,
(Li et al., 2005). Protein synthesis-dependent localization, and translation of mRNA. Each of
expression of Arc following exploration of a novel these mechanisms is modulated by BDNF signa-
environment or seizures is at least partly depend- ling (Klann and Dever, 2004; Richter and
ent on egr-transcription factors including zif268. Sonenberg, 2005; Bramham and Wells, 2007).
458

The discovery of mRNA, ribosomes, and transla- eIF4E is controlled by several binding proteins
tion factors within dendrites and even within (BPs), most notably eIF4E-binding proteins
the spine itself suggested that synapses could be (4E-BPs). Signaling through receptor-coupled
modified directly — and perhaps individually — phosphoinositide 3-kinase (PI3K) kinase and
through regulation of local protein synthesis. mammalian target of rapamycin (mTOR) leads
While dendritic protein synthesis is regulated by to phosphorylation of 4E-BP and liberation of
glutamate, acetylcholine, norepinephrine, dopa- eIF4E. BDNF stimulates cap-dependent transla-
mine, and probably other neurotransmitters, the tion in dendrites through TrkB-coupled PI3K-
BDNF-TrkB system stands out as a potentially mTOR and Ras-ERK (Takei et al., 2001,
autonomous modulator of translation at gluta- 2002; Kelleher et al., 2004; Schratt et al., 2004;
mate synapses. In the CA1 region of adult hippo- Takei et al., 2004; Kanhema et al., 2006). Genetic
campal slices, BDNF-LTP is induced in synaptic or pharmacological inhibition of mTOR or
regions isolated from the pyramidal cell somata ERK impairs LTP maintenance and abolishes
(Kang and Schuman, 1996). In primary hippo- BDNF-induced LTP (Rosenblum et al., 2002;
campal neuronal cultures, BDNF induces expres- Tang et al., 2002; Ying et al., 2002; Cammalleri
sion of a reporter construct in isolated dendrites et al., 2003; Kelleher et al., 2004). BDNF-LTP
(Aakalu et al., 2001). In synaptodendrosome (SD) in the dentate gyrus is coupled to transient ERK-
preparations, subcellular fractions containing dependent phosphorylation of eIF4E and en-
resealed terminal-spine contacts, BDNF rapidly hanced expression of eIF4E protein (Kanhema
stimulates translation of several mRNAs cou- et al., 2006).
pled to LTP or spine morphogenesis, including Protein synthesis in synaptic plasticity is also
a-calcium/calmodulin-dependent protein kinase II controlled at the level of peptide chain elongation.
(aCaMKII), Arc, LIMK1, and GluR1 (Yin et al., Eukaryotic elongation factor 2 (eEF2) is a GTP-
2002; Schratt et al., 2004). Regulation of transla- binding protein that mediates translocation of
tion by BDNF has been shown in SDs prepared peptidyl-tRNAs from the A-site to the P-site on
from forebrain and hippocampus of young (PN the ribosome. Phosphorylation of eEF2 on Thr56
15–22 day) rats (Yin et al., 2002; Schratt et al., inhibits eEF2-ribosome binding and arrests elon-
2004; Kanhema et al., 2006), as well as dentate gation (Browne and Proud, 2002). eEF2 phos-
gyrus from 2–4 month old rats (Kanhema et al., phorylation observed during LTP could therefore
2006). explain decreased synthesis of some proteins in 2D
gel studies. Paradoxically, however, some tran-
scripts important for synaptic plasticity such as
Translation factor regulation Arc and aCaMKII undergo enhanced translation
under conditions of eEF2 phosphorylation and
Recent work has examined the biochemical mech- global suppression of protein synthesis (Scheetz
anisms by which BDNF modulates local protein et al., 2000; Chotiner et al., 2003; Kanhema et al.,
synthesis. Phosphorylation of the eukaryotic initi- 2006). BDNF-LTP in vivo is coupled to ERK-
ation factor 4E (eIF4E) is considered the rate-lim- dependent phosphorylation of eEF2 and enhanced
iting step for translation of most mRNAs (those Arc expression in whole dentate gyrus homogen-
with a 7-methyl-guanosine ‘‘cap’’ at the 50 end). ates (Kanhema et al., 2006). Consolidation of taste
Phosphorylation of eIF4E on Ser209 is correlated memory in the neocortex is also associated with
with enhanced rates of translation, whereas hypo- ERK activation and transient eEF2 phosphorylat-
phosphorylation is associated with decreased ion. (Belelovsky et al., 2005). In constrast, BDNF
translation. eIF4E is phosphorylated by mitogen- treatment of SDs does not alter eEF2 phosphor-
activated integrating kinase 1 (MNK1), whose ylation state, but leads to eIF4E phosphorylation
activity is regulated by ERK and p38 mitogen ac- and rapid synthesis of Arc, aCaMKII, LIMK, and
tivated protein kinase (MAPK). The availability of other proteins (Yin et al., 2002; Schratt et al., 2004;
459

Kanhema et al., 2006). This suggests a compart- miRNA


ment-specific regulation in which initiation is
selectively enhanced by BDNF at synapses, while Another area bound to be important for regu-
both initiation and elongation are modulated at lated protein synthesis in neuronal plasticity is
synaptic sites. microRNAs (miRNAs). miRNAs are small, non-
eEF2 is phosphorylated by eEF2 kinase, which coding RNAs that bind to the 30 UTR of target
itself is subject to tight regulation by calcium/ mRNAs and either block translation or induce
calmodulin, mTOR, ERK, and PKA through transcript degradation (Schratt et al., 2006). As
multiple phosphorylation sites. Recent studies miRNAs generally bind numerous target
have shown bidirectional effects of BDNF on mRNAs, and most mRNAs are regulated by
eEF2 phosphorylation depending on the prepara- more than one species of miRNA, the possibilities
tion studied (Inamura et al., 2005). At synapses, for fine regulation of translation are enormous.
eEF2 is phosphorylated in response to NMDA Schratt et al. showed that miR-134, a brain-
(Scheetz et al., 2000), but not BDNF treatment specific miRNA found in the synaptodendritic
(Kanhema et al., 2006). In primary cortical neu- compartment, negatively regulates dendritic spine
rons, BDNF induces dephosphorylation of eEF2 morphogenesis in cultured hippocampal neurons
and increases elongation rates (Inamura et al., by repressing translation of LIMK1 mRNA. Ap-
2005). Serotonin similarly reduces eEF2 phos- plication of BDNF relieves miR134-mediated re-
phorylation in Aplysia synaptosomes. In this case, pression of LIMK1 translation and promotes
serotonin appears to offset an increase in eEF2 spine morphogenesis. Interestingly, the RNA in-
phosphorylation triggered by calcium influx terference complex (RISC) proteins that mediate
(Carroll et al., 2004). Thus, the direction of the effects of miRNA are themselves subject to
eEF2 phosphorylation is context-dependent and regulation. Ashraf et al. recently demonstrated
controlled by multiple transmitters. While eEF2 that RISC pathway proteins regulate dendritic
appears to be important in transcript-specific and mRNA transport and local protein synthesis
compartment-specific translation control in synap- during acquisition of long-term memory in flies
tic plasticity, further work is needed to resolve (Ashraf et al., 2006). According to their model,
these mechanisms. training in olfactory/electric shock paradigm trig-
In addition to cap-dependent mechanisms, gers proteosomal degradation of the SDE3 heli-
mRNA translation may be initiated at internal case Armitage, leading to relief of RISC-mediated
ribosomal entry sites (IRES). A large percentage suppression of several mRNAs, including aCaM-
of total translation persists even when cap- KII, kinesin heavy chain, and staufen. Adding to
dependent mechanisms are almost completely this complexity, regulation of miRNA transcrip-
abolished. Evidence suggests that IRES-depend- tion may be coupled to specific physiological
ent translation is enhanced or maintained in the effects. Impey et al. showed that calcium/cyclic
context of eIF4E hyperphosphorylation and AMP responsive element binding protein
saturation of cap-dependent mechanisms, and (CREB)-dependent transcription of miR132 pro-
during conditions such as ischemia when eIF2a motes neuronal morphogenesis by decreasing
and eEF2 are phosphorylated (Dyer et al., synthesis of the GTPase activating protein,
2003). A number of dendritic mRNA species, p250GAP. Like many CREB-dependent effects
including Arc and aCAMKII, contain IRES se- in neurons, this mechanism was activated by
quences and are capable of IRES-dependent BDNF (Vo et al., 2005). Furthermore, a recent
translation in cultured hippocampal neurons and study employing microarray expression profiling
non-neuronal cell lines (Pinkstaff et al., 2001). and PCR analysis has revealed specific upregula-
However, IRES-mediated translation has not tion and downregulation of miRNAs following
been explored in the context of synaptic LTP induction in the dentate gyrus in vivo
plasticity (Fig. 2). (Wibrand et al., 2006).
460

Fig. 2. TrkB and translation control in dendritic spines. The cartoon depicts some of the major signaling pathways coupling TrkB to
regulation of eIF4E and eEF2. TrkB activation of PI3K-mTOR and Ras-ERK promotes eIF4E phosphorylation and enhances
translation initiation. Phosphorylation of eEF2 stalls ribosomes and arrests peptide chain elongation. BDNF-TrkB signaling has
bidirectional effects on eEF2 phosphorylation. Adapted from Soule et al. (2006) with permission from the Biochemical Society.

Translocation also induces a redistribution of eIF4E to an


mRNA granule-rich cytoskeletal fraction (Smart
Translocation and positioning of the translation et al., 2003).
machinery is another feature of activity-dependent
regulation. RNA storage granules in dendrites
discharge mRNA in response to strong depolari- Model of BDNF-controlled synaptic consolidation
zation (Krichevsky and Kosik, 2001; Kanai et al.,
2004). During LTP, ribosomes move from dendri- To summarize, current evidence supports a model
tic shafts to spines (Ostroff et al., 2002), and for synaptic consolidation triggered by postsynap-
pre-existing a-CaMKII mRNA moves to the tic BDNF signaling and involving regulation
synaptodendritic compartment (Havik et al., of gene expression and local mRNA translation
2003). BDNF treatment of synaptoneurosomes (Fig. 1). According to this model, bursts of
461

excitatory synaptic activity trigger sustained re- translation of PKMz, which is thought to sustain
lease of BDNF and activation of postsynaptic LTP by increasing the number of active postsy-
TrkB receptors. TrkB signaling rapidly activates naptic a-amino-3-hydroxy-5-methyl-4-isoxazole-
translation in spines and induces transcription of propionic acid receptors (AMPAR) (Ling et al.,
Arc mRNA in cell bodies. Translational activation 2006). Interestingly, the F-actin destabilizing agent
in spines consists of global (eIF4E), and probably latrunculin B blocks new synthesis of PKMz and
more mRNA-specific (miRNA), mechanisms. attenuates LTP maintenance (Kelly et al., 2007).
Spines activated in this way may effectively Taken together, this suggests a sequential mecha-
capture and translate mRNAs liberated from nism of LTP maintenance in the dentate gyrus in
local RNA storage granules, as well as newly in- which Arc-dependent consolidation couples to
duced Arc mRNA transport from the soma. PKCz-dependent expression at the level of actin
Translation of Arc during a critical time-window polymerization.
is necessary for cofilin phosphorylation, local
F-actin expansion, and formation of stable LTP.
The model predicts that translation of Arc under- Extrinsic modulation of synaptic consolidation
lies actin polymerization-dependent enlargement
of synapses. Modulatory transmitters such as norepinephrine,
serotonin, dopamine, and acetylcholine modulate
LTP induction and/or maintenance (Stanton and
Stages in synaptic consolidation: from Arc Sarvey, 1985a, b; Bramham and Srebro, 1989;
to PKCz? Frey et al., 1991; Bramham et al., 1997; Swanson-
Park et al., 1999; Graves et al., 2001; Kulla and
Arc is likely to interact with other core mecha- Manahan-Vaughan, 2002; Straube and Frey, 2003;
nisms of synaptic consolidation. The list of critical Harley et al., 2005). These extrinsic neurotrans-
players includes N-cadherin, members of the inte- mitter systems have diffuse, global patterns of
grin receptor family, matrix metalloproteinase-9, innervation, and communicate through spatially
and the atypical protein kinase C isoform PKMz, dispersed, volume transmission. Neuronal firing of
(Bahr et al., 1997; Bozdagi et al., 2000; Ling et al., modulatory inputs is a function of the animal’s
2002; Chan et al., 2003; Nagy et al., 2006). behavioral or attentional state, with changes in
Of these, PKMz warrants special note. PKMz is activity dictating the functional modes of target
constitutively active because it consists of a PKMz networks (i.e., local rhythmic activity, timing of
catalytic domain, but lacks the autoinhibitory synaptic events, frequency and duration of action
domain of most PKC isoforms. Evidence suggests potential firing). The classical modulatory trans-
that persistent PKMz activity is sufficient and mitters may also set the biochemical tone in target
necessary for a major component of late phase neurons by modulating PKA and CREB activity.
LTP. Thus, bath application of the myristoylated Acetylcholine, dopamine, and norepinephrine
z-pseudosubstrate inhibititory peptide (ZIP) to have all been found to modulate local protein
hippocampal slices blocks potentiation produced synthesis in mammalian principal neurons
by intracellular perfusion of PKMz and reverses (Feig and Lipton, 1993; Gelinas and Nguyen,
the maintenance of LTP in the CA1 region (Ling 2005; Smith et al., 2005), and serotonergic regula-
et al., 2002; Serrano et al., 2005). A recent study tion of local protein synthesis is well known from
further reported that maintenance of LTP in the studies of Aplysia sensory neurons (Casadio et al.,
dentate gyrus of awake rats is rapidly reversed by 1999).
intrahippocampal injection of ZIP 24 h after While these actions are spatially dispersed, they
LTP induction (Pastalkova et al., 2006). More- are likely to exert broad influence and act in
over, injection of ZIP produces persistent loss of concert with specific mechanisms triggered by
1-day-old spatial information. LTP is associated glutamate and BDNF signaling. b-adrenergic re-
with transcription and enhanced dendritic ceptor activation facilitates late LTP through a
462

mechanism involving local protein synthesis, but receptors activates retrograde signaling pathways
not transcription (Straube and Frey, 2003; Gelinas in axons leading to activation of nuclear sub-
and Nguyen, 2005). Locus coeruleus activation strates, such as CREB, and modulation of gene
induces a delayed protein synthesis-dependent expression. HFS of the perforant pathway leads to
LTP in the dentate gyrus (Walling and Harley, CREB phosphorylation in the entorhinal cortex
2004). b-Adrenergic receptor activation is also and this effect is blocked by intracerebroventricu-
necessary for protein synthesis underlying novelty- lar application of the Trk inhibitor K252a (Kelly
induced facilitation of late LTP (Straube et al., et al., 2000; Gooney and Lynch, 2001). BDNF-
2003), and PKA activity is necessary for synaptic LTP, induced by local infusion into the dentate
tagging and input specificity of late LTP in the gyrus, similarly leads to retrograde activation of
CA1 region (Young et al., 2006). One intriguing CREB in entorhinal cortex located some 4 mm
possibility is that modulatory inputs are able to away (Gooney et al., 2004). Such retrograde nu-
contract or expand the time-window of synaptic clear signaling could contribute to sustained
consolidation through effects on local mRNA expansion of presynaptic specializations and
translation. enhanced neurotransmitter release. BDNF treat-
ment of cell cultures is associated with enhanced
vesicle docking, expression of Rab3a, a small
Presynaptic mechanisms in synaptic consolidation GTP-binding proteins involved in vesicle traffick-
ing, and enhanced expression of SNARE proteins
Quantal analysis and biochemical studies support involved in vesicle exocytosis.
a contribution of enhanced glutamate release to
LTP expression at MPP-granule cell synapses
(Min et al., 1998; Errington et al., 2003). Evidence Lateral perforant path
supporting a role for BDNF in enhanced gluta-
mate transmitter release during LTP includes the The lateral perforant input (LPP) projecting from
following: (1) the maintenance phase of BDNF- layer II stellate neurons in the lateral entorhinal
LTP, like HFS-LTP, is associated with a lasting cortex to the outer-third of the dentate gyrus mo-
increase in potassium-evoked glutamate release lecular layer has been less extensively studied in
from synaptosomes (Gooney et al., 2004), (2) TrkB the context of LTP mechanisms. The LPP contains
receptors are autophosphorylated in synaptosomes proenkephalin-derived peptides that are released
collected during the maintenance phase of both in a frequency-dependent manner. Pharmacologi-
HFS- and BDNF-induced LTP (Gooney et al., cal studies show that m and d opioid receptor ac-
2002, 2004), (3) the Trk inhibitor, K252a, blocks tivation is necessary for LTP induction in the
the sustained enhancement in neurotransmitter lateral perforant pathway, while antagonists for
release. Finally, LTP maintenance is associated these receptors have no effect on LTP induction
with enhanced, depolarization-evoked release of or maintenance in the MPP (Bramham, 1992).
BDNF from dentate gyrus tissue (Gooney and Several lines of evidence suggest that opioid pep-
Lynch, 2001). tides facilitate LTP induction by dampening GAB-
The kinetics of these events suggest that BDNF Aergic inhibition (Madison and Nicoll, 1988; Xie
acts rapidly on terminals to trigger a lasting in- and Lewis, 1991; Hanse and Gustafsson, 1992;
crease in glutamate release. However, more de- Bramham and Sarvey, 1996; Milner and Drake,
layed mechanisms involving retrograde nuclear 2001). Relief from GABAA receptor-mediated in-
signaling could contribute. The classic hypothesis hibition facilitates activation of NMDARs and
of target-derived trophic support involves signa- voltage-dependent calcium channels. In transverse
ling from the nerve terminal to the nucleus (Ginty hippocampal slices, LPP LTP and spike-timing-
and Segal, 2002; Delcroix et al., 2003; Campenot dependent LTP and LTD is NMDAR-dependent,
and MacInnis, 2004). In sympathetic and sensory whereas de-potentiation and de-depression re-
neurons, neurotrophin binding to presynaptic Trk quire group 1 metabotropic glutamate receptor
463

activation (Xie and Lewis, 1991; Hanse and elevated 24 h after LTP induction of the MPP,
Gustafsson, 1992; Colino and Malenka, 1993; but not LPP (Mezey et al., 2004).
Bramham and Sarvey, 1996; Lin et al., 2006).
In anesthetized rats, however, normal LPP LTP is
Functions and clinical implications of synaptic
obtained in the presence of NMDAR antagonists
consolidation in the dentate gyrus
that abolish MPP LTP (Bramham et al., 1991).
In vivo LTP induction in the LPP input to CA3
Perturbations in dentate gyrus synaptic plasticity
and is also opioid receptor-dependent and
are thought to contribute to a range of clinical
NMDAR-independent (Do et al., 2002; Kosub
conditions including memory loss, Alzheimer’s
et al., 2005). The discrepancy between the hippo-
disease, and depression. There are many potential
campal slice and in vivo preparations with regard
mechanisms involved, including those described
to the NMDAR dependence of LPP LTP remains
above for synaptic consolidation, including
unresolved, although loss of fine inhibitory control
BDNF, Arc, and other critical mediators. In
in the slice preparation due to transection of
addition, the role of newly generated granule cells
the perpendicularly oriented interneurons during
and the synaptic plasticity of these new cells are
slice preparation may favor NMDA receptor-
likely to contribute.
dependent mechanisms.
The role of BDNF and other neurotrophic
factors in the LPP is not well known. Studies of Alzheimer’s disease
paired-pulse plasticity have revealed a remarkable
pathway-specificity, such that NT-3 modulates Modulation of BDNF synthesis would be expected
only LPP responses and BDNF modulates only to universally enhance the actions of BDNF.
MPP responses (Kokaia et al., 1998; Asztely et al., Given the diverse effects of BDNF, even within
2000). Conditional NT-3 mutant mice have deficits processes like LTP, it may be advantageous to
in the LPP LTP, but not MPP, LTP induction target specific effector mechanisms such as Arc-
(Shimazu et al., 2006). dependent synaptic consolidation. Recent animal
The molecular mechanisms of late phase LTP in studies connect Arc expression in dentate granule
the LPP have similarly received scant attention. cells to Alzheimer’s disease. In transgenic mice
However, the available evidence reveals certain expressing human amyloid precursor protein
similarities with MPP LTP mechanisms. For ex- (hAPP) and hAPP-derived amyloid-b (Ab), Arc
ample: (1) Arc mRNA localizes selectively in the expression is reduced primarily in granule cells
LPP termination zone following long (>15 min) of the dentate gyrus (Palop et al., 2005). Other
sessions of repetitive HFS of the LPP (Steward neuronal populations expressing hAPP, including
et al., 1998), (2) LPP LTP induced by standard CA1 pyramidal cells, show normal basal Arc
short bursts of HFS is associated with input- expression and induction following exposure to a
specific accumulation of F-actin (Fukazawa et al., novel environment. In neuronal cell cultures,
2003), and (3) Like MPP LTP (Villarreal et al., BDNF-mediated induction of Arc is suppressed
2002), stable LPP LTP in awake rats is degraded even at sublethal levels of Ab (Wang et al., 2006).
by NMDA receptor activation (Abraham et al., Downregulation of Arc expression in granule cells
2006). This NMDAR-dependent decay of LPP may therefore contribute to deficits in LTP and
LTP is induced by high-frequency heterosynaptic learning observed in rodent models of Alzheimer’s
stimulation of the MPP. Electron microscopic disease (Rowan et al., 2003; Gureviciene et al.,
studies of synapse morphometry show that LTP 2004; Jacobsen et al., 2006).
of both the lateral and medial peforant path are Although progress has been made in elucidating
associated with increases in the number of unper- the computational role of the entorhinal cortex
forated axospinous synapses. However, differences (which provides input to the dentate gyrus through
at the ultrastructural level probably exist as the the perforant path) and CA3 region (which re-
number of perforated axospinous synapses is ceives output from granule cells), the function of
464

the dentate gyrus remains enigmatic. Recent work through classic mechanism of target-derived sup-
suggests that the dentate gyrus allows fine spatio- port, with active synapses competing for limiting
temporal separation of novel and complex cues, amounts of growth factor (Castren, 2005). An-
thereby disambiguating stimuli to allow sparse other possibility is that BDNF, as a direct conse-
encoding of information (Kesner et al., 2004; quence of NMDAR activation, drives synaptic
Lee et al., 2005a). Arc mRNA levels in the dent- consolidation of nascent contacts as it does at
ate gyrus are elevated several hours following mature synapses.
LTP induction or exploration of a novel environ-
ment, whereas Arc mRNA induction in CA1
region and other brain areas is only short-lived Depression
(minutes) (Ramirez-Amaya et al., 2005). It is
tempting to speculate that the protracted time- Alterations in BDNF signaling figure prominently
window of Arc-dependent consolidation is in current theories of depression (Castren, 2004;
uniquely associated with the function of the dent- Monteggia et al., 2004; Berton and Nestler, 2006;
ate gyrus in disambiguating information for Kuipers and Bramham, 2006). Chronic stress
encoding in the entorhinal-hippocampal circuitry. leading to depression-like behavior in animals
Given the pivotal role of ongoing Arc synthesis, is associated with downregulation of BDNF
it’s possible that this window has a dynamic range transcription and TrkB signaling, while chronic
that is regulated at the level of Arc mRNA trans- antidepressant treatment increases BDNF synthe-
lation. If so, it may prove possible to contract sis reverses the effects of stress on BDNF expres-
or expand the window of synaptic consolidation sion and behavior. Furthermore, the dentate gyrus
therapeutically. appears to be key structure in cognitive aspects of
depression. First, chronic mild stress leading to
anhedonia-like behavior in rats is associated with
Neurogenesis reductions in BDNF expression and CREB activ-
ity in the dentate gyrus, but not hippocampus
Among the unique features of the dentate gyrus is proper (Gronli et al., 2006). Second, acute bilateral
the lifelong production of new granule cells from infusion of the BDNF into the dentate gyrus,
progenitor cells located in the subgranular zone using a protocol similar to that which induced
(Gould and Gross, 2002; Aimone et al., 2006). BDNF-LTP, has antidepressant-like behavioral
Although the function of neurogenesis is still un- effects (Shirayama et al., 2002). Third, injection
clear, it seems likely that the unique contributions of virus expressing histone deacetylase, which
of the dentate gyrus are tied to the availability of modulates BDNF promoter activity, blocks the
new granule cells. One hypothesis is that new behavioral effects of antidepressants (Tsankova
granule cell populations are selectively engaged in et al., 2006).
new information storage, the unique associations It is noteworthy that several important events
of successive populations providing a kind of associated with BDNF-LTP are also induced
timestamp for episodic memory and recall by chronic antidepressant treatment. Chronic
(Aimone et al., 2006). Recent work indicates that treatment with the selective serotonin reuptake
HFS-LTP promotes the survival of newborn gran- inhibitor fluoxetine leads to enhanced TrkB signa-
ule cells (Bruel-Jungerman et al., 2006), and ling, CREB activation, and induction of several
NMDAR activation promotes integration of these BDNF-LTP associated genes including BDNF,
cells within the dentate gyrus circuitry (Tashiro Arc, neuritin, TIEG1, and CARP (Alme et al.,
et al., 2006). BDNF-TrkB signaling is similarly 2006). Dagestad et al. (2006) recently reported that
implicated in survival of adult born granule cells chronic, but not acute, treatment with fluoxetine
(Sairanen et al., 2005; Scharfman et al., 2005). results in region-specific changes in translation
Castren has suggested that BDNF promotes the factor activity. In the dentate gyrus, a pattern of
survival and maturation of new connections dual eIF4E and eEF2 phosphorylation is seen
465

similar to that observed following BDNF-LTP in- References


duction (Dagestad et al., 2006). Taken together
this suggests a direct connection between BDNF- Aakalu, G., Smith, W.B., Nguyen, N., Jiang, C. and Schuman,
induced synaptic strengthening in the dentate E.M. (2001) Dynamic visualization of local protein synthesis
in hippocampal neurons. Neuron, 30(2): 489–502.
gyrus and antidepressant drug action. It will be Abraham, W.C., Mason-Parker, S.E., Irvine, G.I., Logan, B.
important to determine whether BDNF-LTP and Gill, A.I. (2006) Induction and activity-dependent re-
actually develops during antidepressant treatment versal of persistent LTP and LTD in lateral perforant path
and whether specific inhibition of BDNF-LTP synapses in vivo. Neurobiol. Learn. Mem., 86(1): 82–90.
blocks the behavioral effects of local BDNF Aicardi, G., Argilli, E., Cappello, S., Santi, S., Riccio, M.,
Thoenen, H. and Canossa, M. (2004) Induction of long-term
infusion. potentiation and depression is reflected by corresponding
changes in secretion of endogenous brain-derived neurotro-
phic factor. Proc. Natl. Acad. Sci. U.S.A., 101(44):
15788–15792.
Abbreviations Aimone, J.B., Wiles, J. and Gage, F.H. (2006) Potential role for
adult neurogenesis in the encoding of time in new memories.
Nat. Neurosci., 9(6): 723–727.
AMPAR a-amino-3-hydroxy-5-methyl-4- Alme, M., Wibrand, K., Kuipers, S., Dagestad, G. and
isoxazolepropionic acid (AMPA) Bramham, C.R. (2006) Chronic fluoxetine treatment en-
receptor hances expression of BDNF-regulated genes in prefrontal
Arc activity-regulated cytoskeleton- cortex. Soc. Neurosci., Abstr., 314.10.
Aoki, C., Wu, K., Elste, A., Len, G., Lin, S., McAuliffe, G. and
associated protein
Black, I.B. (2000) Localization of brain-derived neurotrophic
BDNF brain-derived neurotrophic factor and TrkB receptors to postsynaptic densities of adult
factor rat cerebral cortex. J. Neurosci. Res., 59(3): 454–463.
a-CaMKII calcium/calmodulin-dependent Ashraf, S.I., McLoon, A.L., Sclarsic, S.M. and Kunes, S. (2006)
protein kinase II Synaptic protein synthesis associated with memory is regu-
lated by the RISC pathway in Drosophila. Cell, 124(1):
CREB calcium/cyclic AMP responsive
191–205.
element binding protein Asztely, F., Kokaia, M., Olofsdotter, K., Ortegren, U. and
eEF2 eukaryotic elongation factor 2 Lindvall, O. (2000) Afferent-specific modulation of short-
eIF4E eukaryotic initiation factor 4E term synaptic plasticity by neurotrophins in dentate gyrus.
4E-BP eIF4E-binding protein Eur. J. Neurosci., 12(2): 662–669.
Bahr, B.A., Staubli, U., Xiao, P., Chun, D., Ji, Z.X., Esteban,
ERK extracellular signal-regulated
E.T. and Lynch, G. (1997) Arg-Gly-Asp-Ser-selective
protein kinase adhesion and the stabilization of long-term potentiation:
HFS high-frequency stimulation pharmacological studies and the characterization of a can-
LIMK LIM domain kinase didate matrix receptor. J. Neurosci., 17(4): 1320–1329.
LTP long-term potentiation Belelovsky, K., Elkobi, A., Kaphzan, H., Nairn, A.C. and
Mnk1 mitogen-activated integrating Rosenblum, K. (2005) A molecular switch for translational
control in taste memory consolidation. Eur. J. Neurosci.,
kinase 1 22(10): 2560–2568.
mTOR mammalian target of rapamycin Berton, O. and Nestler, E.J. (2006) New approaches to
NMDAR N-methyl-D-aspartate receptor antidepressant drug discovery: beyond monoamines. Nat.
PI3K phosphoinositide 3-kinase Rev. Neurosci., 7(2): 137–151.
PSD postsynaptic density Bliss, T.V. and Collingridge, G.L. (1993) A synaptic model of
memory: long-term potentiation in the hippocampus. Nature,
SD synaptodendrosome 361(6407): 31–39.
Blum, R. and Konnerth, A. (2005) Neurotrophin-mediated
rapid signaling in the central nervous system: mechanisms
and functions. Physiology (Bethesda), 20: 70–78.
Bozdagi, O., Shan, W., Tanaka, H., Benson, D.L. and Huntley,
Acknowledgments
G.W. (2000) Increasing numbers of synaptic puncta during
late-phase LTP: N-cadherin is synthesized, recruited to
Supported by the Norwegian Research Council synaptic sites, and required for potentiation. Neuron, 28(1):
and European Union grant 504231. 245–259.
466

Bozon, B., Davis, S. and Laroche, S. (2003) A requirement for Castren, E. (2005) Is mood chemistry? Nat. Rev. Neurosci.,
the immediate early gene zif268 in reconsolidation of recog- 6(3): 241–246.
nition memory after retrieval. Neuron, 40(4): 695–701. Castren, E., Pitkanen, M., Sirvio, J., Parsadanian, A., Lind-
Bramham, C.R. (1992) Opioid receptor dependent long-term holm, D., Thoenen, H. and Riekkinen, P.J. (1993) The in-
potentiation: peptidergic regulation of synaptic plasticity in duction of LTP increases BDNF and NGF mRNA but
the hippocampus [see comments]. Neurochem. Int., 20(4): decreases NT-3 mRNA in the dentate gyrus. Neuroreport,
441–455. 4(7): 895–898.
Bramham, C.R., Bacher-Svendsen, K. and Sarvey, J.M. (1997) Chan, C.S., Weeber, E.J., Kurup, S., Sweatt, J.D. and Davis,
LTP in the lateral perforant path is beta-adrenergic receptor- R.L. (2003) Integrin requirement for hippocampal synaptic
dependent. Neuroreport, 8(3): 719–724. plasticity and spatial memory. J. Neurosci., 23(18):
Bramham, C.R. and Messaoudi, E. (2005) BDNF function in 7107–7116.
adult synaptic plasticity: the synaptic consolidation hypoth- Chotiner, J.K., Khorasani, H., Nairn, A.C., O’Dell, T.J. and
esis. Prog. Neurobiol., 76: 99–125. Watson, J.B. (2003) Adenylyl cyclase-dependent form of
Bramham, C.R., Milgram, N.W. and Srebro, B. (1991) Acti- chemical long-term potentiation triggers translational regu-
vation of AP5-sensitive NMDA receptors is not required to lation at the elongation step. Neuroscience, 116(3): 743–752.
induce LTP of synaptic transmission in the lateral perforant Colino, A. and Malenka, R.C. (1993) Mechanisms underlying
path. Eur. J. Neurosci., 3: 1300–1308. induction of long-term potentiation in rat medial and lateral
Bramham, C.R. and Sarvey, J.M. (1996) Endogenous activa- perforant paths in vitro. J. Neurophysiol., 69(4): 1150–1159.
tion of mu and delta-1 opioid receptors is required for long- Dagestad, G., Kuipers, S.D., Messaoudi, E. and Bramham,
term potentiation induction in the lateral perforant path: de- C.R. (2006) Chronic fluoxetine induces region-specific
pendence on GABAergic inhibition. J. Neurosci., 16(24): changes in translation factor eIF4E and eEF2 activity in
8123–8131. the rat brain. Eur. J. Neurosci., 23: 2814–2818.
Bramham, C.R., Southard, T., Sarvey, J.M., Herkenham, M. Delcroix, J.D., Valletta, J.S., Wu, C., Hunt, S.J., Kowal, A.S.
and Brady, L.S. (1996) Unilateral LTP triggers bilateral in- and Mobley, W.C. (2003) NGF signaling in sensory neurons:
creases in hippocampal neurotrophin and trk receptor evidence that early endosomes carry NGF retrograde signals.
mRNA expression in behaving rats: evidence for interhem- Neuron, 39(1): 69–84.
ispheric communication. J. Comp. Neurol., 368(3): 371–382. Do, V.H., Martinez, C.O., Martinez Jr., J.L. and Derrick, B.E.
Bramham, C.R. and Srebro, B. (1989) Synaptic plasticity in the (2002) Long-term potentiation in direct perforant path pro-
hippocampus is modulated by behavioral state. Brain Res., jections to the hippocampal CA3 region in vivo. J. Ne-
493: 74–86. urophysiol., 87(2): 669–678.
Bramham, C.R. and Wells, D.G. (2007) Dendritic mRNA: Dragoi, G., Harris, K.D. and Buzsaki, G. (2003) Place repre-
transport, translation, and function. Nat. Rev. Neurosci. (in sentation within hippocampal networks is modified by long-
press). term potentiation. Neuron, 39(5): 843–853.
Browne, G.J. and Proud, C.G. (2002) Regulation of peptide- Drake, C.T., Milner, T.A. and Patterson, S.L. (1999) Ultra-
chain elongation in mammalian cells. Eur. J. Biochem., structural localization of full-length trkB immunoreactivity
269(22): 5360–5368. in rat hippocampus suggests multiple roles in modulating
Bruel-Jungerman, E., Davis, S., Rampon, C. and Laroche, S. activity-dependent synaptic plasticity. J. Neurosci., 19(18):
(2006) Long-term potentiation enhances neurogenesis in the 8009–8026.
adult dentate gyrus. J. Neurosci., 26(22): 5888–5893. Dyer, J.R., Michel, S., Lee, W., Castellucci, V.F., Wayne, N.L.
Cammalleri, M., Lutjens, R., Berton, F., King, A.R., Simpson, and Sossin, W.S. (2003) An activity-dependent switch to cap-
C., Francesconi, W. and Sanna, P.P. (2003) Time-restricted independent translation triggered by eIF4E dephosphorylat-
role for dendritic activation of the mTOR-p70S6K pathway ion. Nat. Neurosci., 6(3): 219–220.
in the induction of late-phase long-term potentiation in the Errington, M.L., Galley, P.T. and Bliss, T.V. (2003) Long-term
CA1. Proc. Natl. Acad. Sci. U.S.A., 100(24): 14368–14373. potentiation in the dentate gyrus of the anaesthetized rat
Campenot, R.B. and MacInnis, B.L. (2004) Retrograde trans- is accompanied by an increase in extracellular gluta-
port of neurotrophins: fact and function. J. Neurobiol., 58(2): mate: real-time measurements using a novel dialysis elec-
217–229. trode. Philos. Trans. R. Soc. Lond. B Biol. Sci., 358(1432):
Carroll, M., Warren, O., Fan, X. and Sossin, W.S. (2004) 5-HT 675–687.
stimulates eEF2 dephosphorylation in a rapamycin-sensitive Feig, S. and Lipton, P. (1993) Pairing the cholinergic agonist
manner in Aplysia neuritis. J. Neurochem., 90(6): 1464–1476. carbachol with patterned Schaffer collateral stimulation in-
Casadio, A., Martin, K.C., Giustetto, M., Zhu, H., Chen, M., itiates protein synthesis in hippocampal CA1 pyramidal cell
Bartsch, D., Bailey, C.H. and Kandel, E.R. (1999) A tran- dendrites via a muscarinic, NMDA-dependent mechanism. J.
sient, neuron-wide form of CREB-mediated long-term facil- Neurosci., 13(3): 1010–1021.
itation can be stabilized at specific synapses by local protein Figurov, A., Pozzo Miller, L.D., Olafsson, P., Wang, T. and
synthesis. Cell, 99(2): 221–237. Lu, B. (1996) Regulation of synaptic responses to high-fre-
Castren, E. (2004) Neurotrophic effects of antidepressant drugs. quency stimulation and LTP by neurotrophins in the hippo-
Curr. Opin. Pharmacol., 4(1): 58–64. campus. Nature, 381(6584): 706–709.
467

Frey, U., Frey, S., Schollmeier, F. and Krug, M. (1996) Influ- potentiation and the consolidation of long-term memory. J.
ence of actinomycin D, a RNA synthesis inhibitor, on long- Neurosci., 20(11): 3993–4001.
term potentiation in rat hippocampal neurons in vivo and in Guzowski, J.F., Mcnaughton, B.L., Barnes, C.A. and Worley,
vitro. J. Physiol. Lond., 490(Pt 3): 703–711. P.F. (1999) Environment-specific expression of the immedi-
Frey, U., Matthies, H. and Reymann, K.G. (1991) The effect of ate-early gene Arc in hippocampal neuronal ensembles. Nat.
dopaminergic D1 receptor blockade during tetanization on Neurosci., 2(12): 1120–1124.
the expression of long-term potentiation in the rat CA1 re- Hanse, E. and Gustafsson, B. (1992) Long-term potentiation
gion in vitro. Neurosci. Lett., 129(1): 111–114. and field EPSPs in the lateral and medial perforant paths in
Fukazawa, Y., Saitoh, Y., Ozawa, F., Ohta, Y., Mizuno, K. the dentate gyrus in vitro: a comparison. Eur. J. Neurosci.,
and Inokuchi, K. (2003) Hippocampal LTP is accompanied 4(11): 1191–1201.
by enhanced F-actin content within the dendritic spine that is Harley, C.W., Walling, S.G. and Brown, R.A.M. (2005) The
essential for late LTP maintenance in vivo. Neuron, 38(3): three faces of norepinephrine: plasticity at the perforant path-
447–460. dentate gyrus synapse. In: Stanton P., Bramham C. and
Geinisman, Y. (2000) Structural synaptic modifications associ- Scharfman H. (Eds.), Synaptic plasticity and transsynaptic
ated with hippocampal LTP and behavioral learning. Cereb. signaling. Spring Science+Business Media, Berlin.
Cortex, 10(10): 952–962. Harris, K.M., Fiala, J.C. and Ostroff, L. (2003) Structural
Gelinas, J.N. and Nguyen, P.V. (2005) Beta-adrenergic receptor changes at dendritic spine synapses during long-term potent-
activation facilitates induction of a protein synthesis-depend- iation. Philos. Trans. R. Soc. Lond. B Biol. Sci., 358(1432):
ent late phase of long-term potentiation. J. Neurosci., 25(13): 745–748.
3294–3303. Havik, B., Rokke, H., Bardsen, K., Davanger, S. and Bram-
Ginty, D.D. and Segal, R.A. (2002) Retrograde neurotrophin ham, C.R. (2003) Bursts of high-frequency stimulation trig-
signaling: Trk-ing along the axon. Curr. Opin. Neurobiol., ger rapid delivery of pre-existing alpha-CaMKII mRNA to
12(3): 268–274. synapses: a mechanism in dendritic protein synthesis during
Gooney, M. and Lynch, M.A. (2001) Long-term potentiation in long-term potentiation in adult awake rats. Eur. J. Neurosci.,
the dentate gyrus of the rat hippocampus is accompanied by 17(12): 2679–2689.
brain-derived neurotrophic factor-induced activation of Husi, H., Ward, M.A., Choudhary, J.S., Blackstock, W.P. and
TrkB. J. Neurochem., 77(5): 1198–1207. Grant, S.G. (2000) Proteomic analysis of NMDA receptor-
Gooney, M., Messaoudi, E., Maher, F.O., Bramham, C.R. and adhesion protein signaling complexes. Nat. Neurosci., 3(7):
Lynch, M.A. (2004) BDNF-induced LTP in dentate gyrus is 661–669.
impaired with age: analysis of changes in cell signaling events. Inamura, N., Nawa, H. and Takei, N. (2005) Enhancement of
Neurobiol. Aging, 25(10): 1323–1331. translation elongation in neurons by brain-derived neurotro-
Gooney, M., Shaw, K., Kelly, A., O’Mara, S.M. and Lynch, phic factor: implications for mammalian target of rapamycin
M.A. (2002) Long-term potentiation and spatial learning are signaling. J. Neurochem., 95(5): 1438–1445.
associated with increased phosphorylation of TrkB and ex- Jacobsen, J.S., Wu, C.C., Redwine, J.M., Comery, T.A., Arias,
tracellular signal-regulated kinase (ERK) in the dentate R., Bowlby, M., Martone, R., Morrison, J.H., Pangalos,
gyrus: evidence for a role for brain-derived neurotrophic M.N., Reinhart, P.H. and Bloom, F.E. (2006) Early-onset
factor. Behav. Neurosci., 116(3): 455–463. behavioral and synaptic deficits in a mouse model of Al-
Gould, E. and Gross, C.G. (2002) Neurogenesis in adult mam- zheimer’s disease. Proc. Natl. Acad. Sci. U.S.A., 103(13):
mals: some progress and problems. J. Neurosci., 22(3): 5161–5166.
619–623. Jones, M.W., Errington, M.L., French, P.J., Fine, A., Bliss,
Graves, L., Pack, A. and Abel, T. (2001) Sleep and memory: a T.V., Garel, S., Charnay, P., Bozon, B., Laroche, S. and
molecular perspective. Trends Neurosci., 24(4): 237–243. Davis, S. (2001) A requirement for the immediate early gene
Gronli, J., Bramham, C.R., Murison, R., Kanhema, T., Fiske, Zif268 in the expression of late LTP and long-term memories.
E., Bjorvatn, B., Sorensen, E., Ursin, R. and Portas, C.M. Nat. Neurosci., 4(3): 289–296.
(2006) Chronic mild stress inhibits BDNF expression and Kanai, Y., Dohmae, N. and Hirokawa, N. (2004) Kinesin
CREB activation in the dentate gyrus but not in the hippo- transports RNA: isolation and characterization of an RNA-
campus proper. Pharmacol. Biochem. Behav., 85(4): transporting granule. Neuron, 43(4): 513–525.
842–849. Kang, H. and Schuman, E.M. (1996) A requirement for local
Gureviciene, I., Ikonen, S., Gurevicius, K., Sarkaki, A., protein synthesis in neurotrophin-induced hippocampal
van Groen, T., Pussinen, R., Ylinen, A. and Tanila, H. synaptic plasticity. Science, 273(5280): 1402–1406.
(2004) Normal induction but accelerated decay of LTP Kang, H., Welcher, A.A., Shelton, D. and Schuman, E.M.
in APP+PS1 transgenic mice. Neurobiol. Dis., 15(2): (1997) Neurotrophins and time: different roles for TrkB
188–195. signaling in hippocampal long-term potentiation. Neuron,
Guzowski, J.F., Lyford, G.L., Stevenson, G.D., Houston, F.P., 19(3): 653–664.
McGaugh, J.L., Worley, P.F. and Barnes, C.A. (2000) Inhi- Kanhema, T., Dagestad, G., Panja, D., Tiron, A., Messaoudi,
bition of activity-dependent arc protein expression in the rat E., Havik, B., Ying, S.W., Nairn, A.C., Sonenberg, N. and
hippocampus impairs the maintenance of long-term Bramham, C.R. (2006) Dual regulation of translation
468

initiation and peptide chain elongation during BDNF-in- Lin, Y.W., Yang, H.W., Wang, H.J., Gong, C.L., Chiu, T.H.
duced LTP in vivo: evidence for compartment-specific trans- and Min, M.Y. (2006) Spike-timing-dependent plasticity at
lation control. J. Neurochem., 19: 1328–1337. resting and conditioned lateral perforant path synapses on
Kelleher III, R.J., Govindarajan, A., Jung, H.Y., Kang, H. and granule cells in the dentate gyrus: different roles of N-methyl-
Tonegawa, S. (2004) Translational control by MAPK signa- D-aspartate and group I metabotropic glutamate receptors.
ling in long-term synaptic plasticity and memory. Cell, Eur. J. Neurosci., 23(9): 2362–2374.
116(3): 467–479. Ling, D.S., Benardo, L.S. and Sacktor, T.C. (2006) Protein
Kelly, A., Mullany, P.M. and Lynch, M.A. (2000) Protein syn- kinase Mzeta enhances excitatory synaptic transmission by
thesis in entorhinal cortex and long-term potentiation in increasing the number of active postsynaptic AMPA recep-
dentate gyrus. Hippocampus, 10(4): 431–437. tors. Hippocampus, 16(5): 443–452.
Kelly, M.T., Yao, Y., Sondhi, R. and Sacktor, T.C. (2007) Ling, D.S., Benardo, L.S., Serrano, P.A., Blace, N., Kelly,
Actin polymerization regulates the synthesis of PKMzeta in M.T., Crary, J.F. and Sacktor, T.C. (2002) Protein kinase
LTP. Neuropharmacology, 52(1): 41–45. Mzeta is necessary and sufficient for LTP maintenance. Nat.
Kesner, R.P., Lee, I. and Gilbert, P. (2004) A behavioral as- Neurosci., 5(4): 295–296.
sessment of hippocampal function based on a subregional Link, W., Konietzko, U., Kauselmann, G., Krug, M., Schwa-
analysis. Rev. Neurosci., 15(5): 333–351. nke, B., Frey, U. and Kuhl, D. (1995) Somatodendritic ex-
Klann, E. and Dever, T.E. (2004) Biochemical mechanisms for pression of an immediate early gene is regulated by synaptic
translational regulation in synaptic plasticity. Nat. Rev. Ne- activity. Proc. Natl. Acad. Sci. U.S.A., 92(12): 5734–5738.
urosci., 5(12): 931–942. Lisman, J.E. and Zhabotinsky, A.M. (2001) A model of synap-
Kokaia, M., Asztely, F., Olofsdotter, K., Sindreu, C.B., Kull- tic memory: a CaMKII/PP1 switch that potentiates trans-
mann, D.M. and Lindvall, O. (1998) Endogenous neurotro- mission by organizing an AMPA receptor anchoring
phin-3 regulates short-term plasticity at lateral perforant assembly. Neuron, 31(2): 191–201.
path-granule cell synapses. J. Neurosci., 18(21): 8730–8739. Lyford, G.L., Yamagata, K., Kaufmann, W.E., Barnes, C.A.,
Kossel, A.H., Cambridge, S.B., Wagner, U. and Bonhoeffer, T. Sanders, L.K., Copeland, N.G., Gilbert, D.J., Jenkins, N.A.,
(2001) A caged Ab reveals an immediate/instructive effect of Lanahan, A.A. and Worley, P.F. (1995) Arc, a growth factor
BDNF during hippocampal synaptic potentiation. Proc. and activity-regulated gene, encodes a novel cytoskeleton-
Natl. Acad. Sci. U.S.A., 98(25): 14702–14707. associated protein that is enriched in neuronal dendrites.
Kosub, K.A., Do, V.H. and Derrick, B.E. (2005) NMDA re- Neuron, 14(2): 433–445.
ceptor antagonists block heterosynaptic long-term depression Madison, D.V. and Nicoll, R.A. (1988) Enkephalin hyperpo-
(LTD) but not long-term potentiation (LTP) in the CA3 re- larizes interneurones in the rat hippocampus. J. Physiol., 398:
gion following lateral perforant path stimulation. Neurosci. 123–130.
Lett., 374(1): 29–34. Matsuzaki, M., Honkura, N., Ellis-Davies, G.C. and Kasai, H.
Krichevsky, A.M. and Kosik, K.S. (2001) Neuronal RNA (2004) Structural basis of long-term potentiation in single
granules: a link between RNA localization and stimulation- dendritic spines. Nature, 429(6993): 761–766.
dependent translation. Neuron, 32(4): 683–696. Meng, Y., Zhang, Y., Tregoubov, V., Janus, C., Cruz, L.,
Kuipers, S.D. and Bramham, C.R. (2006) Brain-derived ne- Jackson, M., Lu, W.Y., MacDonald, J.F., Wang, J.Y., Falls,
urotrophic factor mechanisms and function in adult synaptic D.L. and Jia, Z. (2002) Abnormal spine morphology and
plasticity: new insights and implications for therapy. Curr. enhanced LTP in LIMK-1 knockout mice. Neuron, 35(1):
Opin. Drug Discov. Devel., 9(5): 580–586. 121–133.
Kulla, A. and Manahan-Vaughan, D. (2002) Modulation by Messaoudi, E., Bardsen, K., Srebro, B. and Bramham, C.R.
serotonin 5-HT(4) receptors of long-term potentiation and (1998) Acute intrahippocampal infusion of BDNF induces
depotentiation in the dentate gyrus of freely moving rats. lasting potentiation of synaptic transmission in the rat dent-
Cereb. Cortex, 12(2): 150–162. ate gyrus. J. Neurophysiol., 79(1): 496–499.
Lee, I., Hunsaker, M.R. and Kesner, R.P. (2005a) The role of Messaoudi, E., Kanhema, T., Soule, J., Tiron, A., Dagyte, G.,
hippocampal subregions in detecting spatial novelty. Behav. da Silva, B. and Bramham, C.R. (2007) Sustained Arc
Neurosci., 119(1): 145–153. synthesis controls LTP consolidation through regulation
Lee, J.L., Everitt, B.J. and Thomas, K.L. (2004) Independent of local actin polymerization in the dentate gyrus in vivo.
cellular processes for hippocampal memory consolidation Submitted.
and reconsolidation. Science., 304(5672): 839–843. Messaoudi, E., Ying, S.W., Kanhema, T., Croll, S.D. and
Lee, P.R., Cohen, J.E., Becker, K.G. and Fields, R.D. (2005b) Bramham, C.R. (2002) BDNF triggers transcription-depend-
Gene expression in the conversion of early-phase to late- ent, late phase LTP in vivo. J. Neurosci., 22: 7453–7461.
phase long-term potentiation. Ann. N.Y. Acad. Sci., 1048: Mezey, S., Doyere, V., De Souza, I., Harrison, E., Cambon, K.,
259–271. Kendal, C.E., Davies, H., Laroche, S. and Stewart, M.G.
Li, L., Carter, J., Gao, X., Whitehead, J. and Tourtellotte, (2004) Long-term synaptic morphometry changes after in-
W.G. (2005) The neuroplasticity-associated arc gene is a di- duction of long-term potentiation and long-term depression
rect transcriptional target of early growth response (Egr) in the dentate gyrus of awake rats are not simply mirror
transcription factors. Mol. Cell Biol., 25(23): 10286–10300. phenomena. Eur. J. Neurosci., 19(8): 2310–2318.
469

Milner, T.A. and Drake, C.T. (2001) Ultrastructural evidence Plath, N., Ohana, O., Dammermann, B., Errington, M.L.,
for presynaptic mu opioid receptor modulation of synaptic Schmitz, D., Gross, C., Mao, X., Engelsberg, A., Mahlke, C.,
plasticity in NMDA-receptor-containing dendrites in the Welzl, H., Kobalz, U., Stawrakakis, A., Fernandez, E.,
dentate gyrus. Brain Res. Bull., 54(2): 131–140. Waltereit, R., Bick-Sander, A., Therstappen, E., Cooke, S.F.,
Min, M.Y., Asztely, F., Kokaia, M. and Kullmann, D.M. Blanquet, V., Wurst, W., Salmen, B., Bosl, M.R., Lipp, H.P.,
(1998) Long-term potentiation and dual-component quantal Grant, S.G., Bliss, T.V., Wolfer, D.P. and Kuhl, D. (2006)
signaling in the dentate gyrus. Proc. Natl. Acad. Sci. U.S.A., Arc/Arg3.1 is essential for the consolidation of synaptic
95(8): 4702–4707. plasticity and memories. Neuron, 52(3): 437–444.
Moga, D.E., Calhoun, M.E., Chowdhury, A., Worley, P., Ramirez-Amaya, V., Vazdarjanova, A., Mikhael, D., Rosi, S.,
Morrison, J.H. and Shapiro, M.L. (2004) Activity-regulated Worley, P.F. and Barnes, C.A. (2005) Spatial exploration-
cytoskeletal-associated protein is localized to recently acti- induced Arc mRNA and protein expression: evidence for se-
vated excitatory synapses. Neuroscience, 125(1): 7–11. lective, network-specific reactivation. J. Neurosci., 25(7):
Monteggia, L.M., Barrot, M., Powell, C.M., Berton, O., Ga- 1761–1768.
lanis, V., Gemelli, T., Meuth, S., Nagy, A., Greene, R.W. and Richter, J.D. and Sonenberg, N. (2005) Regulation of cap-de-
Nestler, E.J. (2004) Essential role of brain-derived neurotro- pendent translation by eIF4E inhibitory proteins. Nature,
phic factor in adult hippocampal function. Proc. Natl. Acad. 433(7025): 477–480.
Sci. U.S.A., 101(29): 10827–10832. Rodriguez, J.J., Davies, H.A., Silva, A.T., De Souza, I.E.,
Nagy, V., Bozdagi, O., Matynia, A., Balcerzyk, M., Okulski, P., Peddie, C.J., Colyer, F.M., Lancashire, C.L., Fine, A., Err-
Dzwonek, J., Costa, R.M., Silva, A.J., Kaczmarek, L. and ington, M.L., Bliss, T.V. and Stewart, M.G. (2005) Long-
Huntley, G.W. (2006) Matrix metalloproteinase-9 is required term potentiation in the rat dentate gyrus is associated with
for hippocampal late-phase long-term potentiation and mem- enhanced Arc/Arg3.1 protein expression in spines, dendrites
ory. J. Neurosci., 26(7): 1923–1934. and glia. Eur. J. Neurosci., 21(9): 2384–2396.
Nestler, E.J., Barrot, M., DiLeone, R.J., Eisch, A.J., Gold, S.J. Rosenblum, K., Futter, M., Voss, K., Erent, M., Skehel, P.A.,
and Monteggia, L.M. (2002) Neurobiology of depression. French, P., Obosi, L., Jones, M.W. and Bliss, T.V. (2002) The
Neuron, 34(1): 13–25. role of extracellular regulated kinases I/II in late-phase long-
Nguyen, P.V. and Kandel, E.R. (1996) A macromolecular syn- term potentiation. J. Neurosci., 22(13): 5432–5441.
thesis-dependent late phase of long-term potentiation requir- Routtenberg, A. and Rekart, J.L. (2005) Post-translational
ing cAMP in the medial perforant pathway of rat protein modification as the substrate for long-lasting mem-
hippocampal slices. J. Neurosci., 16(10): 3189–3198. ory. Trends Neurosci., 28(1): 12–19.
Oertner, T.G. and Matus, A. (2005) Calcium regulation of actin Rowan, M.J., Klyubin, I., Cullen, W.K. and Anwyl, R. (2003)
dynamics in dendritic spines. Cell Calcium, 37(5): 477–482. Synaptic plasticity in animal models of early Alzheimer’s
Okamoto, K., Nagai, T., Miyawaki, A. and Hayashi, Y. (2004) disease. Philos. Trans. R. Soc. Lond. B Biol. Sci., 358(1432):
Rapid and persistent modulation of actin dynamics regulates 821–828.
postsynaptic reorganization underlying bidirectional plastic- Sairanen, M., Lucas, G., Ernfors, P., Castren, M. and Castren,
ity. Nat. Neurosci., 7(10): 1104–1112. E. (2005) Brain-derived neurotrophic factor and antidepres-
Ostroff, L.E., Fiala, J.C., Allwardt, B. and Harris, K.M. (2002) sant drugs have different but coordinated effects on neuronal
Polyribosomes redistribute from dendritic shafts into spines turnover, proliferation, and survival in the adult dentate
with enlarged synapses during LTP in developing rat hippo- gyrus. J. Neurosci., 25(5): 1089–1094.
campal slices. Neuron, 35(3): 535–545. Sanes, J.R. and Lichtman, J.W. (1999) Can molecules explain
Palop, J.J., Chin, J., Bien-Ly, N., Massaro, C., Yeung, B.Z., long-term potentiation? Nat. Neurosci., 2(7): 597–604.
Yu, G.Q. and Mucke, L. (2005) Vulnerability of dentate Santi, S., Cappello, S., Riccio, M., Bergami, M., Aicardi, G.,
granule cells to disruption of arc expression in human am- Schenk, U., Matteoli, M. and Canossa, M. (2005)
yloid precursor protein transgenic mice. J. Neurosci., 25(42): Hippocampal neurons recycle BDNF for activity-dependent
9686–9693. secretion and LTP maintenance. EMBO J., 25(18):
Pastalkova, E., Serrano, P., Pinkhasova, D., Wallace, E., Fen- 4372–4380.
ton, A.A. and Sacktor, T.C. (2006) Storage of spatial infor- Scharfman, H., Goodman, J., Macleod, A., Phani, S.,
mation by the maintenance mechanism of LTP. Science, Antonelli, C. and Croll, S. (2005) Increased neurogenesis
313(5790): 1141–1144. and the ectopic granule cells after intrahippocampal BDNF
Patterson, S.L., Grover, L.M., Schwartzkroin, P.A. and Both- infusion in adult rats. Exp. Neurol., 192(2): 348–356.
well, M. (1992) Neurotrophin expression in rat hippocampal Scheetz, A.J., Nairn, A.C. and Constantine-Paton, M. (2000)
slices: a stimulus paradigm inducing LTP in CA1 evokes in- NMDA receptor-mediated control of protein synthesis at
creases in BDNF and NT-3 mRNAs. Neuron, 9(6): developing synapses. Nat. Neurosci., 3(3): 211–216.
1081–1088. Schratt, G.M., Nigh, E.A., Chen, W.G., Hu, L. and Greenberg,
Pinkstaff, J.K., Chappell, S.A., Mauro, V.P., Edelman, G.M. M.E. (2004) BDNF regulates the translation of a select group
and Krushel, L.A. (2001) Internal initiation of translation of of mRNAs by a mammalian target of rapamycin-phospha-
five dendritically localized neuronal mRNAs. Proc. Natl. tidylinositol 3-kinase-dependent pathway during neuronal
Acad. Sci. U.S.A., 98(5): 2770–2775. development. J. Neurosci., 24(33): 7366–7377.
470

Schratt, G.M., Tuebing, F., Nigh, E.A., Kane, C.G., Sabatini, A.S. and Abraham, W.C. (1999) A double dissociation
M.E., Kiebler, M. and Greenberg, M.E. (2006) A brain- within the hippocampus of dopamine D1/D5 receptor and
specific microRNA regulates dendritic spine development. beta-adrenergic receptor contributions to the persistence of
Nature, 439(7074): 283–289. long-term potentiation. Neuroscience, 92(2): 485–497.
Serrano, P., Yao, Y. and Sacktor, T.C. (2005) Persistent Takei, N., Inamura, N., Kawamura, M., Namba, H., Hara, K.,
phosphorylation by protein kinase Mzeta maintains Yonezawa, K. and Nawa, H. (2004) Brain-derived neurotro-
late-phase long-term potentiation. J. Neurosci., 25(8): phic factor induces mammalian target of rapamycin-depend-
1979–1984. ent local activation of translation machinery and protein
Shirayama, Y., Chen, A.C., Nakagawa, S., Russell, D.S. and synthesis in neuronal dendrites. J. Neurosci., 24(44):
Duman, R.S. (2002) Brain-derived neurotrophic factor pro- 9760–9769.
duces antidepressant effects in behavioral models of depres- Takei, N., Kawamura, M., Hara, K., Yonezawa, K. and Nawa,
sion. J. Neurosci., 22(8): 3251–3261. H. (2001) Brain-derived neurotrophic factor enhances neu-
Shimazu, K., Zhao, M., Sakata, K., Akbarian, S., Bates, B., ronal translation by activating multiple initiation processes:
Jaenisch, R. and Lu, B. (2006) NT-3 facilitates hippocampal comparison with the effects of insulin. J. Biol. Chem.,
plasticity and learning and memory by regulating neurogen- 276(46): 42818–42825.
esis. Learn. Mem., 13(3): 307–315. Tang, S.J., Reis, G., Kang, H., Gingras, A.C., Sonenberg, N.
Smart, F.M., Edelman, G.M. and Vanderklish, P.W. (2003) and Schuman, E.M. (2002) A rapamycin-sensitive signaling
BDNF induces translocation of initiation factor 4E to pathway contributes to long-term synaptic plasticity in the
mRNA granules: evidence for a role of synaptic microfila- hippocampus. Proc. Natl. Acad. Sci. U.S.A., 99(1): 467–472.
ments and integrins. Proc. Natl. Acad. Sci. U.S.A., 100(24): Tao, X., Finkbeiner, S., Arnold, D.B., Shaywitz, A.J. and
14403–14408. Greenberg, M.E. (1998) Ca2+ influx regulates BDNF tran-
Smith, W.B., Starck, S.R., Roberts, R.W. and Schuman, E.M. scription by a CREB family transcription factor-dependent
(2005) Dopaminergic stimulation of local protein synthe- mechanism. Neuron, 20(4): 709–726.
sis enhances surface expression of GluR1 and synaptic Tashiro, A., Sandler, V.M., Toni, N., Zhao, C. and Gage, F.H.
transmission in hippocampal neurons. Neuron, 45(5): (2006) NMDA-receptor-mediated, cell-specific integration of
765–779. new neurons in adult dentate gyrus. Nature, 442(7105):
Soule, J., Messaoudi, E. and Bramham, C.R. (2006) Brain- 929–933.
derived neurotrophic factor and control of synaptic consol- Tsankova, N.M., Berton, O., Renthal, W., Kumar, A., Neve,
idation in the adult brain. Biochem. Soc. Trans., 34(Pt 4): R.L. and Nestler, E.J. (2006) Sustained hippocampal chro-
600–604. matin regulation in a mouse model of depression and anti-
Stanton, P.K. and Sarvey, J.M. (1985a) Blockade of nor- depressant action. Nat. Neurosci., 9(4): 519–525.
epinephrine-induced long-lasting potentiation in the hippo- Vazdarjanova, A., Ramirez-Amaya, V., Insel, N., Plummer,
campal dentate gyrus by an inhibitor of protein synthesis. T.K., Rosi, S., Chowdhury, S., Mikhael, D., Worley, P.F.,
Brain Res., 361(1–2): 276–283. Guzowski, J.F. and Barnes, C.A. (2006) Spatial exploration
Stanton, P.K. and Sarvey, J.M. (1985b) Depletion of induces ARC, a plasticity-related immediate-early gene, only
norepinephrine, but not serotonin, reduces long- term po- in calcium/calmodulin-dependent protein kinase II-positive
tentiation in the dentate gyrus of rat hippocampal slices. principal excitatory and inhibitory neurons of the rat fore-
J. Neurosci., 5(8): 2169–2176. brain. J. Comp. Neurol., 498(3): 317–329.
Steward, O., Wallace, C.S., Lyford, G.L. and Worley, P.F. Villarreal, D.M., Do, V., Haddad, E. and Derrick, B.E. (2002)
(1998) Synaptic activation causes the mRNA for the IEG Arc NMDA receptor antagonists sustain LTP and spatial mem-
to localize selectively near activated postsynaptic sites on ory: active processes mediate LTP decay. Nat. Neurosci.,
dendrites. Neuron, 21(4): 741–751. 5(1): 48–52.
Steward, O. and Worley, P.F. (2001) A cellular mechanism Vo, N., Klein, M.F., Varlamova, O., Keller, D.M., Yamamoto,
for targeting newly synthesized mRNAs to synaptic sites T., Goodman, R.H. and Impey, S. (2005) A cAMP-response
on dendrites. Proc. Natl. Acad. Sci. U.S.A., 98(13): element binding protein-induced microRNA regulates neu-
7062–7068. ronal morphogenesis. Proc. Natl. Acad. Sci. U.S.A., 102(45):
Straube, T. and Frey, J.U. (2003) Involvement of beta-ad- 16426–16431.
renergic receptors in protein synthesis-dependent late long- Walling, S.G. and Harley, C.W. (2004) Locus ceruleus activa-
term potentiation (LTP) in the dentate gyrus of freely moving tion initiates delayed synaptic potentiation of perforant path
rats: the critical role of the LTP induction strength. Neuro- input to the dentate gyrus in awake rats: a novel beta-ad-
science, 119(2): 473–479. renergic- and protein synthesis-dependent mammalian plas-
Straube, T., Korz, V., Balschun, D. and Frey, J.U. (2003) Re- ticity mechanism. J. Neurosci., 24(3): 598–604.
quirement of beta-adrenergic receptor activation and protein Wang, D.C., Chen, S.S., Lee, Y.C. and Chen, T.J. (2006) Am-
synthesis for LTP-reinforcement by novelty in rat dentate yloid-beta at sublethal level impairs BDNF-induced arc ex-
gyrus. J. Physiol., 552(Pt 3): 953–960. pression in cortical neurons. Neurosci. Lett., 398(1–2): 78–82.
Swanson-Park, J.L., Coussens, C.M., Mason-Parker, S.E., Weeks, A.C., Ivanco, T.L., LeBoutillier, J.C., Racine, R.J. and
Raymond, C.R., Hargreaves, E.L., Dragunow, M., Cohen, Petit, T.L. (2001) Sequential changes in the synaptic
471

structural profile following long-term potentiation in the rat in synaptoneurosomes. Proc. Natl. Acad. Sci. U.S.A., 99(4):
dentate gyrus: III. Long-term maintenance phase. Synapse, 2368–2373.
40(1): 74–84. Ying, S.W., Futter, M., Rosenblum, K., Webber, M.J., Hunt,
Wibrand, K., Messaoudi, E., Havik, B., Steenslid, V., Løvlie, S.P., Bliss, T.V. and Bramham, C.R. (2002) Brain-derived
R., Steen, V.M. and Bramham, C.R. (2006) Identification of neurotrophic factor induces long-term potentiation in intact
genes co-upregulated with Arc during BDNF-induced long- adult hippocampus: requirement for ERK activation coupled
term potentiation in the adult rat dentate gyrus in vivo. Eur. to CREB and upregulation of Arc synthesis. J. Neurosci.,
J. Neurosci., 23: 1501–1511. 22(5): 1532–1540.
Wu, K., Xu, J.L., Suen, P.C., Levine, E., Huang, Y.Y., Mount, Young, J.Z., Isiegas, C., Abel, T. and Nguyen, P.V. (2006)
H.T., Lin, S.Y. and Black, I.B. (1996) Functional trkB ne- Metaplasticity of the late-phase of long-term potentiation: a
urotrophin receptors are intrinsic components of the adult critical role for protein kinase A in synaptic tagging. Eur. J.
brain postsynaptic density. Brain Res. Mol. Brain Res., Neurosci., 23(7): 1784–1794.
43(1–2): 286–290. Zhou, Q., Xiao, M. and Nicoll, R.A. (2001) Contribution of
Xie, C.W. and Lewis, D.V. (1991) Opioid-mediated facilitation cytoskeleton to the internalization of AMPA receptors. Proc.
of long-term potentiation at the lateral perforant path-dent- Natl. Acad. Sci. U.S.A., 98(3): 1261–1266.
ate granule cell synapse. J. Pharmacol. Exp. Ther., 256(1): Zito, K., Knott, G., Shepherd, G.M., Shenolikar, S. and
289–296. Svoboda, K. (2004) Induction of spine growth and synapse
Yin, Y., Edelman, G.M. and Vanderklish, P.W. (2002) The formation by regulation of the spine actin cytoskeleton.
brain-derived neurotrophic factor enhances synthesis of Arc Neuron, 44(2): 321–334.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 26

Comparison of cellular mechanisms of long-term


depression of synaptic strength at perforant
path–granule cell and Schaffer collateral–CA1
synapses

Beatrice Pöschel and Patric K. Stanton

Department of Cell Biology and Anatomy, New York Medical College, Valhalla, NY 10595, USA

Abstract: This chapter compares the cellular mechanisms that have been implicated in the induction and
expression of long-term depression (LTD) at Schaffer collateral–CA1 synapses to perforant path-dentate
gyrus (DG) synapses. In general, Schaffer collateral LTD and long-term potentiation (LTP) both appear to
be a complex combination of many alterations in synaptic transmission that occur at both presynaptic and
postsynaptic sites, while at perforant path synapses, most evidence has focused on postsynaptic long-term
alterations. Within the DG, the medial perforant path is far more studied than lateral perforant path
synapses, where most evidence relates to the induction of heterosynaptic LTD at lateral perforant path
synapses when LTP is induced in the medial perforant path. Of course, there remain many other classes of
synapses in the DG where synaptic plasticity, including LTD, have been largely neglected. It is clear that a
better understanding of the range of DG loci where long-lasting activity-dependent plasticity, both LTD and
LTP, are expressed will be essential to improve our understanding of the cognitive roles of such DG plasticity.

Keywords: CA1; dentate gyrus; hippocampus; long-term depression; perforant path; plasticity; Schaffer
collateral

Hippocampal long-term depression (LTD) Kobayashi et al., 1996), as well as many other
synapses and regions. The cellular mechanisms for
LTD is defined as a long lasting (hours to weeks), induction and expression of LTD can vary
usually synapse specific, decrease of the strength of depending on the stimulation protocol, brain sub-
synaptic transmission. It appears to be the coun- region, and stage of development (Domenici et al.,
terpart of long-term potentiation (LTP), a long- 1998; Kemp et al., 2000; Malenka and Bear, 2004).
lasting increase in synaptic strength. LTD can be The majority of studies elucidating the cellular
induced at synapses in the hippocampal subre- mechanisms of hippocampal LTD have been per-
gions CA1, CA3, and the dentate gyrus (DG) formed in area CA1. Studies of LTD in area CA1
(Dudek and Bear, 1992; O’Mara et al., 1995a; will be reviewed to develop a model of how LTD
can be induced and expressed, to be used as
Corresponding author. Tel.: +001-914-594-4883; a framework to compare LTD mechanisms in CA1
Fax: +001-914-594-4653; E-mail: patric_stanton@nymc.edu with those in the DG.

DOI: 10.1016/S0079-6123(07)63026-X 473


474

LTD in area CA1 (Schaffer collateral– CA1 stimulated between the bursts, then the test input
synapses) expresses LTD. In a closely-related form of LTD
called spike-timing LTD, repeated application of
LTD can be divided into two phases: induction single presynaptic stimuli that closely follow single
and expression. The induction phase involves cel- postsynaptic action potentials (APs) elicits LTD,
lular biochemical events which are triggered by while reversing the pairing (i.e., pre-AP) elicits LTP
particular patterns of synaptic activity within a (Magee and Johnston, 1997; Markram et al., 1997).
timeframe of minutes, and which initiate longer- Induction of LTD by pharmacological agents al-
lasting processes leading to the expression of LTD. lows for selective targeting of particular receptors
that activate different induction pathways. Thus,
particular signal cascades can be investigated with
Induction regard to their contribution to the induction of po-
tentially separable forms of LTD that might never-
LTD can be induced by different means, including theless share similar expression mechanisms. In area
both electrical stimulation and pharmacological CA1, LTD can be induced by agonist activation of
manipulations. While it is often mistakenly as- a number of transmitter receptors (metabotropic
sumed that these forms of LTD are similar mech- glutamate receptors (mGluRs): Manahan-Vaughan
anistically within and across brain regions, evidence and Reymann, 1995; N-methyl-D-aspartate recep-
is substantial that multiple forms of LTD with dis- tors (NMDAR): Lee et al., 1998; Kamal et al., 1999)
tinct induction and expression mechanisms coexist or activation/inactivation of intracellular messen-
within single synapses, and that the profiles of gers (i.e., simultaneous activation of protein kinase
expression of these forms can differ greatly between G (PKG) and inhibition of protein kinase A (PKA):
synapses. Santschi et al., 1999; Bailey et al., 2003).
Most commonly, LTD is induced by a prolonged The intracellular induction mechanisms of LTD
(10–15 min) train of low-frequency (1–2 Hz) stimu- vary depending on the brain region, developmental
lation (LFS; Dudek and Bear, 1992). This form of stage and triggering mechanism (Malenka and
LTD is called homosynaptic because only the stim- Bear, 2004). Different forms of LTD can coexist
ulated pathway exhibits a long-term change in in one region, as in area CA1 (Kemp and Bashir,
synaptic strength. Heterosynaptic LTD involves 1997; Manahan-Vaughan, 1997; Oliet et al., 1997;
interactions between separate synapses or groups Nicoll et al., 1998), as well as other brain regions.
of synapses that converge on the same neuron. The main forms of LTD at Schaffer collateral–CA1
In CA1, heterosynaptic LTD can be induced in a synapses are NMDAR dependent and mGluR-
non-stimulated pathway by the delivery of LTP- dependent LTD, which will be discussed below.
inducing stimuli to a parallel pathway that con- After LTD-inducing synaptic stimulation,
verges on the same neuron (Lynch et al., 1977). One NMDAR- and mGluR-dependent LTD are in-
computationally interesting form of heterosynaptic duced simultaneously in most cases. They can be
LTD is associative LTD (Stanton and Sejnowski, separated by selective pharmacological blockade of
1989). For associative forms of synaptic plasticity, NMDARs or mGluRs, respectively, or selectively
the relative timing of activity in two (or more) input induced by agonist application.
pathways is a key determinant of the direction of
change in synaptic strength, i.e., whether LTP or
LTD is induced. During induction of associative NMDAR/mGluR-dependent LTD
LTD, the activity of a test input is negatively cor-
related in time with activity of a conditioning train During initial investigations of LTD in CA1, LTD
of bursts applied to a separate pathway, which was elicited via a protocol of LFS trains (Dudek
leads to LTD in the test input. That is, if a series and Bear, 1992). It turned out that this stimulation
of brief high-frequency bursts of stimuli are applied induced a mixture of NMDAR- and mGluR-
to the conditioning input, and the test input is dependent LTDs. Thus, studies of the intracellular
475

Fig. 1. Cascades involved with NMDAR/mGluR-dependent LTD at Schaffer collateral–CA1 synapses. AA: arachidonic acid, AC:
adenylyl cyclase, Ca: Ca2+, cADPR: cyclic adenosine diphosphate ribose, CaM: calmodulin, CaMKII: calmodulin-dependent protein
kinase II, cAMP: cyclic adenosine monophosphate, cGMP: cyclic guanosine monophosphate, CN: calcineurin, DAG: diacyl glycerol,
HPETE: hydroperoxyeicosatetraenoic acid, I1: inhibitor 1, IP3: inositol-1,4,5-trisphosphate, IEG: immediate early genes, MAPK:
mitogen-activated protein kinase, mGluR: metabotropic glutamate receptor, NO: nitric oxide, NOS: NO-synthetase, P: phosphate,
PIP2: phosphatidylinositol-3,4-bisphosphate, PLA2: phospholipase A2, PKA: protein kinase A, PKC: protein kinase C, PKG: protein
kinase G, PLC: phospholipase C, Post: postsynapse, PP1: protein phosphatase 1, Pre: presynapse, PS: protein synthesis, VGCC:
voltage-gated Ca2+ channels.

mechanisms of stimulus-induced LTD did not dis- glutamate to activate the receptor allows them to
tinguish between NMDAR- and mGluR-depend- function as coincidence detectors of presynaptic
ent pathways. Later it was found that selective transmitter release and postsynaptic depolarization
activation of NMDARs and mGluRs can induce (Cotman et al., 1988), and because they gate the
LTD independently, and that NMDAR- and influx of Ca2+, they are the key factors in estab-
mGluR-dependent LTD pathways could be stud- lishing associative long-term changes (both LTP
ied independently. These studies will be reviewed and LTD) of synaptic strength. It has been shown
below (Fig. 1). that application of NMDAR antagonists block the
induction of LTD by prolonged LFS (1 Hz/15 min;
Induction. The first step in the induction of LTD Dudek and Bear, 1992; Mulkey and Malenka,
is the activation of glutamate receptors and/or 1992).
voltage-gated Ca2+ channels (VGCCs). NMDARs Other studies have revealed that activation of
have been a primary focus of attention, because mGluRs (Oliet et al., 1997; Manahan-Vaughan,
their requirement for membrane depolarization to 1997) and VGCCs (Christie et al., 1995; Cummings
relieve Mg2+-dependent channel block plus et al., 1996; Oliet et al., 1997) can also be necessary
476

for the induction of LTD in CA1, and the depend- associated with an increase in PP1 activity lasting
ence on these receptors varies with the precise 35 min after LTD induction, as well as PP2A ac-
stimulus pattern and level of postsynaptic depo- tivity increases lasting 65 min (Thiels et al., 1998,
larization. Activation of NMDARs, mGluRs, or 2000). CN is a phosphatase that is dependent
VGCCs each lead to elevations in intracellular on Ca2+ levels and is activated by Ca2+/CaM
[Ca2+]. Ca2+ is a key ion that triggers a variety of (Klee et al., 1979).
kinases and phosphatases, some mediating intra- Amongst its effects, CN has been demonstrated
cellular events, which lead to both short- and long- to dephosphorylate synapsin I (Hosaka et al., 1999),
term changes in synaptic strength. During the resulting in a gradual, activity-dependent reduction
induction of LTD, the rise in [Ca2+] is lower than of neurotransmitter release, since vesicles are held in
that occurring during induction of LTP (Hansel et the reserve pool by dephoshorylated synapsin and
al., 1996; Yang et al., 1999), which is thought to are prevented from joining the readily releasa
lead to a preferred activation of phosphatases with ble vesicle pool (RRP) (Hilfiker et al., 1999). CN
lower Kd’s for activation than most Ca2+-depend- also regulates the internalization of both alpha-
ent kinases (Lisman, 1989). Besides phosphatases amino-3-hydroxy-5-methyl-4-isoxazoleproprionate
(Mulkey et al., 1993; Thiels et al., 1998), LTD has (AMPA) receptors (AMPARs) (Beattie et al., 2000;
been shown to also depend on multiple kinases. Lin et al., 2000) and NMDARs (Shi et al., 2000).
It has been reported that protein kinase C (PKC) CN has been reported to contribute to downregu-
(Stanton, 1995; Hrabetova and Sacktor, 1996), lation of Ca2+ signaling by reducing both Ca2+
PKG (Reyes-Harde et al., 1999a, b), PKA (Bran- influx and Ca2+-induced Ca2+ release from intra-
don et al., 1995), Ca2+/calmodulin-dependent pro- cellular stores via dephosphorylation and inactiva-
tein kinase II (CaMKII) (Stevens et al., 1994; tion of VGCCs (L-, N-, P/Q-, or R-type;
Stanton and Gage, 1996), and mitogen-activated Armstrong, 1989) Additionally, it has been shown
protein kinase (MAPK) (Thiels et al., 2002; that NMDA-induced reduction in dendritic spine
Gallagher et al., 2004) can all play roles in the in- density in cultured hippocampal neurons is blocked
duction of LTD in area CA1. In some cases, these by pre-treatment with CN inhibitors and that this is
roles appear to be played presynaptically, and in probably via preventing the CN-mediated promo-
others postsynaptically, reinforcing the notion of tion of actin depolymerization (Halpain et al., 1998).
multiple forms of LTD. CN asserts its effects through disinhibition of
PP1 (postsynaptic serine/threonine PP1), so that
Expression. Phosphatases and kinases target sev- PP1 is only indirectly regulated by [Ca2+]. When
eral effector molecules that could cause changes activated, CN dephosphorylates and inactivates
in synaptic strength via effects on a wide range of inhibitor-1 (Cohen, 1989; Mulkey et al., 1994),
protein mechanisms, such as receptor number and which is normally bound to PP1. Inactivation of
sensitivity, ion-channel open probability and con- inhibitor-1 releases and activates PP1. PP1 is also
ductance, cytoskeletal machinery, mRNA and pro- regulated by PKA, which phosphorylates and in-
tein synthesis (PS), and vesicle-release probability. activates inhibitor-1 and thus inhibits PP1 (Cohen,
(a) Phosphatase pathways 1989). PP1 regulates gene expression via the de-
Phosphatases whose activation has been sug- phosphorylation and inactivation of the transcrip-
gested to be involved in the induction of LTD are tion factor cyclic adenosine monophosphate
protein phosphatase 2A (PP2A), calcineurin (CN; (cAMP) responsive element-binding protein
also, PP2B), and PP1. Inhibition of these phospha- (CREB; Hagiwara et al., 1992; Bito et al., 1996),
tases with the PP1/PP2A inhibitors okadaic acid which in its active form, initiates the transcription
and calyculin A (Mulkey et al., 1993) and CN of genes that contain a cAMP response element
inhibitors FK506 and cyclosporin A (Mulkey et al., (CRE) within their promoter sequence (Brindle
1994) have all been demonstrated to block the and Montminy, 1992). Another substrate of PP1
induction of LTD. Additionally, it has been shown (and/or PP2A) is AMPARs (Lee et al., 2000)
that the induction of LTD in CA1 in vivo is which are dephosphorylated at the PKA site
477

(SER-845) (Lee et al., 1998, 2000), resulting in a mGluRs are divided into three groups: I, II, and
decrease in their open probability (Banke et al., III. Group I mGluRs are positively coupled to
2000). Like CN, PP1 has been implicated in activ- adenylate cyclase and the cAMP-PKA pathway,
ity-dependent endocytotic internalization of AM- and they also can activate PKC. mGluRs II and
PARs, since CN and PP1 antagonists each inhibit mGluRs III are negatively coupled to adenylate
NMDA- and insulin-induced endocytosis of AM- cyclase. NMDARs contribute to intracellular
PARs (Beattie et al., 2000; Ehlers, 2000; Lin et al., [Ca2+] elevation via Ca2+ influx through
2000). NMDAR channels, followed by Ca2+-induced
PP2A, as well as PP1, have both been reported Ca2+ release from ryanodine-sensitive stores
to dephosphorylate the autophosphorylation site (Alford et al., 1992). Group I mGluRs promote
Ser-657/660 and the trans-phosphorylation site the release of cytosolic Ca2+ from inositol-1,4,5-
Ser-641 on the catalytic domain of PKC-a/PKC- trisphosphate (IP3)-sensitive intracellular stores
bII (Keranen et al., 1995; Thiels et al., 2000). PP1 via positive coupling to IP3. It has been demon-
can also dephosphorylate the autophosphorylat- strated that induction of LFS-LTD requires Ca2+
ion site Ser-641 on the catalytic domain of PKC- release from presynaptic, but not postsynaptic,
bII (Keranen et al., 1995). ryanodine-sensitive stores, and Ca2+ release from
However, the ability of phosphatases to dephos- postsynaptic IP3-sensitive stores (Reyes and Stanton,
phorylate substrates in vitro or in vivo does not 1996). Intracellular Ca2+ can activate PKA via
confirm that this dephosphorylation occurs during Ca2+/CaM-sensitive adenylyl cyclase (AC) (Eliot
the induction of LTD. For instance, PP1 and et al., 1989) which synthesizes cAMP, CaMKII via
PP2A have been shown to dephosphorylate Ca2+/CaM (Colbran, 1992), PKG via the Ca2+/
Thr286 of CaMKII in vitro (Shields et al., 1985; CaM-NOS-NO-GC-cGMP pathway (Garbers,
Strack et al., 1997) but hippocampal LTD can still 1990; Bredt et al., 1991), and PKC by Ca2+ paired
be induced in CaMKIIT286A mice (Krezel et al., with diacylglycerol (DAG; Huang, 1989).
1999; Parsley et al., 2007), indicating that this de- During the induction of LTD, kinases are bidi-
phosphorylation event is not required for at least rectionally regulated, as are their effector path-
some forms of LTD. ways. While sometimes phosphorylated and
Dephosphorylation events which have been activated kinases are necessary to induce LTD,
linked to decreases in synaptic strength during in other cases kinases are dephosphorylated and
LTD in area CA1 are: AMPAR dephosphorylat- inactivated to promote LTD.
ion of SER-845 by CN and PP1/PP2A (Lee et al.,
2000, 2003), AMPAR internalization (Carroll Ca2+/phospholipid-dependent protein kinase
et al., 1999; Heynen et al., 2000), and PP1/PP2A- (PKC). Reports indicate that inhibition of PKC
mediated dephosphorylation of PKC at Ser-657/ may be required for LTD, since a selective inhibitor
660 (Thiels et al., 2000). CN dephosphorylation of of PKC, chelerythrine, induces a depression which
synapsin I (Hosaka et al., 1999) could, by reducing shows mutual occlusion with LFS-LTD (Hrabetova
transmitter release (Hilfiker et al., 1999), contrib- and Sacktor, 1996) suggesting similar induction
ute to presynaptic LTD. Additional substrates de- and/or expression mechanisms. Indeed, LTD is as-
phosphorylated by CN that might play important sociated with a decrease in PKC activity (Hrabetova
roles in the long-lasting remodeling of synaptic and Sacktor, 1996; Thiels et al., 2000). This decrease
structure are microtubule-associated protein 2 in PKC activity has been shown to be mediated by
(MAP2), protein tau and tubulin (Goto et al., either a proteolysis-mediated decrease in PKCg and
1985), though a requirement of these events for PKMz activity (Hrabetova and Sacktor, 1996) or by
LTD has yet to be confirmed. dephosphorylation of catalytically relevant auto-
(b) Kinase pathways phosphorylation sites on the C terminus of PKC
Many kinases can be activated by Ca2+ and (Ser-657 on PKCa; Ser-660 on PKCbII) by PPs
activated or inactivated by distinct mGluR and (Thiels et al., 2000). While the decrease in activity of
other G protein-coupled metabotropic pathways. PKCa, b, and g was only transient (35–65 min,
478

Thiels et al., 2000), downregulation of the autono- slices, but not by postsynaptic intracellular infu-
mously active PKMz (Schwartz, 1993) persisted up sion of KN-62 into single CA1 pyramidal neurons,
to 120 min post-LFS (Hrabetova and Sacktor, 2001) indicating that the CaMKII that may be necessary
making it a special candidate for the maintenance for LTD induction is presynaptically localized
phase of LTD. It has been demonstrated that LTD (Stanton and Gage, 1996).
is associated with a decrease in pre- as well
as postsynaptic PKC substrate phosphorylation cAMP-dependent protein kinase (PKA). There
(Ramakers et al., 2000). Since it is known that are contradictory studies concerning the role of
PKC activators such as phorbol esters enhance pre- PKA in the induction of LTD. Brandon et al.
synaptic transmission (Chaki et al., 1994), it seems (1995) reported that the inhibition of PKA strongly
plausible that inactivation of constitutively active inhibits LFS-induced LTD at mouse Schaffer col-
PKC might lead to depression of release. On the lateral–CA1 synapses (Brandon et al., 1995), while
postsynaptic side, PKC is known to target the studies in knockout mice have reported that LTD
extracellular signal-regulated kinase (ERK) signa- in field CA1 is dramatically reduced if either the
ling cascade (Roberson et al., 1999) and could thus PKA regulatory subunit RIb (Brandon et al., 1995)
affect transcription processes during LTD (see or the PKA catalytic subunit Cb1 (Qi et al., 1996)
below). are missing. Interestingly Cb1- and RIb-containing
On the other hand, there are also studies sug- PKA holoenzymes show a higher sensitivity to
gesting a requirement for activation of PKC during cAMP than Cb- and RIb-containing PKA holoen-
LTD. PKC is known to phosphorylate SER-880 of zymes (Cadd et al., 1990) and could be preferen-
the AMPAR subunit GluR2 (McDonald et al., tially activated during LTD induction when Ca2+
2001), GluR2 SER-880 phosphorylation is in- influx, and perhaps, associated increases in [cAMP]
creased after LTD induction (Kim et al., 2001), may be less than those needed to elicit LTP.
preventing the phosphorylation of SER-880 on A number of studies explicitly contradict the
GluR2 by point mutation inhibits the constitutive notion that PKA activation is needed for LTD.
synaptic incorporation of GluR2 homomers, as Santschi et al. (1999) reported that multiple syn-
well as LTD (Seidenman et al., 2003) and removal thetic PKA inhibitors enhanced the magnitude of
of GluR2 containing AMPARs from synapses is LTD at these same synapses in rat hippocampal
one potential expression mechanism for LTD slices, while Nicholls et al. (2006) found that LTD
(Seidenman et al., 2003). Thus, it is feasible that is enhanced in a transgenic mouse model where
PKC activates a pathway that includes the phos- adenylate cyclase is constitutively inhibited by
phorylation of SER-880 on GluR2, followed by Gi2a and PKA stimulation prevented. Studies by
the endocytosis of GluR2 containing AMPARs to Kameyama et al. (1998) similarly support a role
express LTD. This pathway could be implemented for inhibition of PKA in promoting LTD; they
by the PKC isoforms (a, bI, bII, g) that have been found that inhibition of postsynaptic PKA in-
reported to be transiently (o15 min) activated duced a depression that occluded LFS-LTD and
after induction of LTD (Hrabetova and Sacktor, that LFS-LTD was inhibited by prior activation of
2001). Overall, multiple roles for both presynaptic postsynaptic PKA. Furthermore, LTD is blocked
and postsynaptic isoforms of PKC make the func- by the selective mGluRII antagonist EGLU, and
tions of these enzymes in induction and mainte- by the A1 adenosine receptor blocker DPCPX, in
nance of LTD potentially quite complex. juvenile rats (Santschi et al., 2006). Since mGluR-
IIs are negatively coupled to cAMP, it is a rea-
Ca2+/calmodulin-dependent protein kinase (CaM- sonable hypothesis that they mediate LTD
KII). It has been reported that mice that lack the by decreasing [cAMP] and PKA activity. The con-
gene for alpha CaMKII show reduced LTD tradictory data suggesting a positive role for PKA
(Stevens et al., 1994). Furthermore, induction of activity in LTD might indicate either a develop-
LFS-LTD is blocked by extracellular application mental switch and/or dependence on genus, since
of the CaMKII inhibitor KN-62 to hippocampal Kameyama et al. (1998) and Santschi et al. (1999)
479

used juvenile rats (14–21 days), while Brandon ryanodine-sensitive stores, suggesting that this signal
et al. (1995) and Qi et al. (1996) studied adult mice cascade and Ca2+ release from ryanodine-sensitive
(4–6 weeks). Distinct PKA isoforms may also play stores mediated a component of LFS-LTD.
roles in different forms of synaptic plasticity, i.e., Since it is known that cGMP can enhance Ca2+-
LTP or LTD (Brandon et al., 1997). release from ryanodine-sensitive stores via activa-
tion of cADP-ribose (Galione et al., 1993; Me-
cGMP-dependent protein kinase (PKG). LTD at szaros et al., 1993; Lee, 1997), it has been proposed
Schaffer collateral–CA1 synapses can also be that the PKG-ADP ribosyl cyclase-cADP ribose
blocked by PKG inhibitors (Reyes-Harde et al., cascade links cGMP to Ca2+-release from ryano-
1999a, b). It is thought that PKG is activated pre- dine-sensitive stores during LTD (Reyes-Harde
synaptically during LTD induction via a cascade et al., 1999a). Reyes-Harde et al. (1999a) verified
that includes the retrograde messenger nitric oxide that cGMP does stimulate the synthesis of cADP
(NO), activation of guanylyl cyclase, production ribose in hippocampal slices and found that an-
of guanosine 30 ,50 -cyclic monophosphate (cGMP) tagonists of cADP ribose receptors in presynaptic
and subsequent PKG activation. Although double terminals block the induction of a portion of LTD.
knockout mice lacking both the neuronal and
endothelial forms of NO synthase (nNOS and Presynaptic vesicular release. LFS-LTD has been
eNOS) did not show significantly reduced LTD shown to be associated with a long-term reduction
(Son et al., 1996; 3 months) and LFS-LTD was not in neurotransmitter release (Stanton et al., 2001,
affected by the NOS inhibitor L-NG-nitroarginine 2003; Zhang et al., 2006). There is evidence that key
(L-NOArg) (Cummings et al., 1994), several other opposing mediators of LTD on the presynaptic site
studies have demonstrated a role for NO for LFS- are both PKA and PKG. In fact, it has been shown
LTD. Izumi and Zorumski (1993) found a block that a chemical LTD (cLTD) can be induced by
of LFS-LTD by the NO inhibitor L-NG-mon- simultaneously raising [cGMP] while inhibiting
omethylargine (L-NMMA) as well as L-NOArg. PKA (Santschi et al., 1999), and that this is asso-
Santschi et al. (1999) partially reduced the magni- ciated with a decrease in the evoked release of the
tude of LFS-LTD by application of the compet- fluorescent marker of presynaptic activity FM1–43
itive NOS inhibitor L-nitroarginine (L-NA). from isolated rat hippocampal presynaptic termi-
Furthermore, Stanton et al. (2003) also reported nals (Bailey et al., 2003). Furthermore, cLTD is
a partial block of LFS-LTD by L-NA, as well as occluded by LFS-LTD (Santschi et al., 1999) and
by the extracellular NO scavenger hemoglobin. suppresses LFS-evoked FM1–43 release (Stanton
Additional support for the NO-cGMP-PKG path- et al., 2001), demonstrating that it mimics the signal
way comes from studies of Gage et al. (1997) and transduction pathways necessary for LFS-LTD,
Reyes-Harde et al. (1999a, b); Gage et al. (1997) i.e., simultaneous increase of [cGMP] and inhibi-
reported that presynaptic NO-sensitive guanylyl tion of PKA, that lead to a presynaptic decrease in
cyclase (NOGC) is necessary for the induction neurotransmitter release.
of LTD, since an extracellularly applied NOGC It has also been shown that a selective PKG
inhibitor prevented the induction of LTD, while inhibitor completely prevents induction of LTD
selective intracellular blockade of postsynaptic of presynaptic FM1–43 release from the RRP
NOGC did not. Furthermore, they also showed (Stanton et al., 2003). Further upstream, activa-
that the membrane-permeable cGMP analog tion of NMDARs and group II mGluRs have both
8-pCPTcGMP and the NO donor S-nitroso-N- been indicated to play a role in the induction of
acetylpenicillamine (SNAP) potentiated the induc- presynaptic LFS-LTD. Inhibition of NMDARs
tion of LTD by a weak submaximal LFS (1 Hz/400 or mGluRs (group I and II, but not group III)
stimuli). Reyes-Harde et al. (1999b) demonstrated have been shown to block induction of LFS-LTD
that the potentiation of LTD by SNAP was me- (Reyes-Harde et al., 1999b; Stanton et al., 2003;
diated via the NOGC-cGMP-PKG cascade and Santschi et al., 2006). The induction of LFS-LTD
required release of Ca2+ from Ca2+-mediated of FM1–43 release from the RRP by a 1–2 Hz
480

train is largely blocked by the NMDAR antago- from juvenile (P10–15; Fitzpatrick and Baudry,
nist AP-5 (Stanton et al., 2003). The competitive 1994) and adult rats (Normandin et al., 1996), but a
NOS inhibitor L-NA completely prevented the more selective PLA2 inhibitor did not block induc-
LFS-induced RRP release and the extracellular tion of LTD in adults (Stanton, 1996), (2) inhibitors
NO scavenger hemoglobin partial blocked LFS- of lipoxygenase pathways of AA metabolism de-
induced RRP release (Stanton et al., 2003) sug- creased the magnitude of LTD in juvenile (P20–25;
gesting that NMDARs mediate presynaptic LTD Chabot et al., 1998) and adult rats (Normandin et
via the retrograde messenger NO. Zhang et al. al., 1996). The 12-1ipoxygenase pathway may have
(2006) found that group II, but not group I cellular effects compatible with LTD expression,
mGluRs, contribute to the induction of presynap- such as inhibition of CaMKII (Piomelli et al., 1989)
tic LTD by a different, paired-pulse stimulation or reduction of transmitter release (Freeman et al.,
protocol, indicating that, depending upon the par- 1991) and changes in synaptic transmission through
ticular stimulus protocol, either NMDARs or the alteration of the affinity states of AMPARs
group I mGluRs can supply the necessary postsy- (Chabot et al., 1998). In detail, it was shown that
naptic [Ca2+], while any presynaptic receptors, 12-lipoxygenase products attenuate the potassium-
such as group II mGluRs, that inhibit adenylate induced release of endogenous glutamate from
cyclase can supply the necessary suppression of hippocampal mossy fiber synaptosomes (Freeman
[cAMP] in the presynaptic terminal. et al., 1991) and low concentrations of PLA2 caused
a decrease in AMPA binding due to a change in
Extracellular signal-regulated kinase (ERK). ERK/ receptor affinity that was blocked by lipoxygenase
MAPK cascade has also been implicated as a key inhibitors (Chabot et al., 1998). The PLA2-induced
signal transduction mechanism for coupling neuronal decrease in AMPA binding seems to be dependent
cell surface receptors to plasticity-induced transcrip- on protein phosphatase dephosphorylation, as the
tion. It has been demonstrated that ERK activation PP inhibitor okadaic acid resulted in a significant
is necessary for the induction of LTD, since an ERK reduction in the PLA2-induced decrease in AMPA
inhibitor blocks induction of LTD (Thiels et al., (Chabot et al., 1998). As with many biochemical
2002). Multiple second messengers, such as cAMP, cascades, the involvement/requirement for PLA2
PKA, [Ca2+], and DAG, can all control ERK signa- activation in LTD may prove to be stimulus
ling via the small G proteins Ras and Rap1 (Grewal protocol-specific.
et al., 1999) and thus link both NMDARs and
mGluRs to this pathway. During induction of LTD, Maintenance. It is suggested that long-lasting
ERK stimulates specific transcription pathways that expression of LTD of synaptic strength requires
differ from LTP; in LTD, ERK activation does not the regulated expression of proteins, some of them
result in the increased phosphorylation of the tran- with low resting concentrations. It has been dem-
scription factor CREB, but rather an increased phos- onstrated that LTD, like LTP (Stanton and
phorylation of the transcription factor Elk-1 (Thiels Sarvey, 1984), requires ongoing PS for stable
et al., 2002). expression. In slice cultures of 8–9 days old rats,
the induction of LFS-LTD has been reported to be
Phospholipase A2. It has been suggested that dur- inhibited by the PS inhibitor anisomycin, as well as
ing LTD Ca2+ activates Ca2+-sensitive phospholi- the transcriptional inhibitor actinomycin D (but
pase A2 (PLA2) (Clark et al., 1995) besides see Huber et al., 2000; Kauderer and Kandel,
phosphatases and kinases. The activation of PLA2 2000). While anisomycin blocks the late phase of
leads to arachidonic acid (AA) and lipoxygenase LTD when applied prior to an LFS, it does not
metabolite production. Findings which suggested affect the maintenance of LTD when applied
involvement of the 12-1ipoxygenase pathway in 30 min after LFS, suggesting that PS is only nec-
LTD are: (1) A non-selective PLA2 inhibitor, essary during the induction and early phases of
bromophenacylbromide, significantly reduced the LTD to synthesize proteins necessary for mainte-
extent of Schaffer collateral–CA1 LTD in slices nance of persistent depression.
481

LFS-LTD in adult rats (7–8 weeks) seems to be in juvenile rats (Kameyama et al., 1998), suggesting
only dependent on PS, but not transcription, as its that PP2B dephosphorylates other substrates, but
induction was blocked by anisomycin (in vivo: not AMPARs, during NMDA-LTD. Interestingly,
Manahan-Vaughan et al., 2000; in slices: Sajikumar the role of phosphatases for NMDA-LTD seems to
and Frey, 2003) but not actinomycin D (Manahan- be developmentally regulated. While inhibitors of
Vaughan et al., 2000). The blockade of LTD by PP1/PP2A do not affect NMDA-LTD in juvenile
anisomycin, applied before LFS, becomes effective rats (P21–30: Kameyama et al., 1998), they have
at approximately 5 h after LFS, suggesting that, been reported to block NMDA-LTD in adult rats
while existing protein levels can support alterations (6 months: Kamal et al., 1999). Furthermore, inhi-
in synaptic strength lasting many hours, the main- bition of PP2B completely blocks NMDA-LTD in
tenance of longer-lasting synaptic plasticity, like adult rats (6 months: Kamal et al., 1999), but only
long-term memory, requires ongoing PS. partially blocks LTD in juvenile rats (P21–30:
Kameyama et al., 1998) and is without effect on
NMDA-LTD in even younger rats (P14; Kamal
NMDAR-dependent LTD et al., 1999).
Brief application of NMDA (20–50 mM for
NMDAR-dependent LTD can be induced in area 1–2 min) with a concentration and application
CA1 via application of NMDA (Lee et al., 1998; duration that is similar to those that induce
Kamal et al., 1999), or LFS-stimulation under NMDA-LTD (3 min, 20 mM), causes a rapid, nearly
pharmacological mGluR blockade (Oliet et al., complete internalization of surface AMPARs in
1997; Zhang et al., 2006) (Fig. 1). hippocampal cultures (2–3 weeks in vitro; Beattie
Brief application of NMDA (3 min, 20 mM) to et al., 2000; Ehlers, 2000), suggesting that
hippocampal slices of young (P21–35: Lee et al., NMDAR-LTD, like LFS-LTD, may be expressed
1998; P14: Kamal et al., 1999) or adult rats by AMPAR endocytosis. Indeed, Nosyreva and
(6 months; Kamal et al., 1999) induces NMDAR- Huber (2005) confirmed that NMDA-induced LTD
dependent LTD blocked by the NMDAR antago- (3 min, 20 mM NMDA) in CA1 of neonatal hippo-
nist 2-amino-5-phosphonopentanoic acid (AP5). campal slices was associated with a decrease in sur-
NMDA-LTD and LFS-LTD have been shown to face expression of GluR1 and GluR2/3 at 10 min
occlude one another (Lee et al., 1998; Kamal et al., and 60 min, respectively. Like LFS-LTD, the
1999), suggesting shared induction and/or expres- NMDA-induced AMPAR internalization was
sion mechanisms. Indeed, in juvenile rats, NMDA- dependent on Ca2+ influx, being almost completely
LTD, like LFS-LTD (Lee et al., 2000), is associated inhibited by removing extracellular Ca2+ (Beattie
with dephosphorylation of the GluR1 subunit of et al., 2000) and by the Ca2+ chelator BAPTA-AM
AMPARs at the PKA-site (SER-845; Lee et al., (Ehlers, 2000). Like LFS-LTD, NMDA-induced
1998). Additionally, as for LFS-LTD (Kameyama AMPAR internalization was also dependent on
et al., 1998), inhibition of PKA seems to play a role CN, but not PP2A, since it was inhibited by the CN
for NMDA-LTD in juvenile rats, since both blocker FK-506, but not by low concentrations of
NMDA-LTD and AMPAR dephosphorylation okadaic acid (10 nM) that selectively inhibits PP2A
are inhibited by prior activation of postsynaptic (Beattie et al., 2000; Ehlers, 2000).
PKA (Kameyama et al., 1998). Nevertheless, in ju- There are contradictory findings concerning the
venile rats, NMDA-LTD and the NMDA-induced role of PP1 in AMPAR endocytosis. While Beattie
dephosphorylation of AMPARs cannot be blocked et al. (2000) could not inhibit NMDA-induced
by high concentrations of inhibitors of PP1 and AMPAR internalization with the PP1/PP2A inhib-
PP2A (Kameyama et al., 1998), in marked contrast itors calyculin A or okadaic acid (1 mM), Ehlers
to earlier findings for LFS-LTD (Mulkey et al., (2000) reported an almost complete inhibition of
1993; Lee et al., 2000). Even though PP2B inhibitors NMDA-induced AMPAR internalization by these
did not affect NMDA-induced dephosphorylation inhibitors, as well as the PP1-selective inhibitor
of AMPARs, they partially inhibited NMDA-LTD tautomycin. While the reason for these divergent
482

findings is unclear, there were slightly different in- they demonstrated that NOS activity is required
duction protocols used; Beattie et al. (2000) pre- for this presynaptic NMDAR-dependent LTD, as
pared dissociated cell cultures from areas CA1/ they could block it with the NOS inhibitor L-NA,
CA3 from P0 rat pups, while Ehlers (2000) pre- suggesting NO as a retrograde transmitter that es-
pared hippocampal mixed cell cultures from 1 to 2 tablishes presynaptic NMDAR-dependent LTD.
days old rats, with both maintained 2–3 weeks in Finally, it has been shown that T-type VGCCs
vitro before commencing experiments. Beattie et al. and activation of PKC do not appear to play a role
(2000) induced AMPAR internalization by appli- in NMDAR-dependent LTD (as opposed to
cation of 50 mM NMDA for 1 min in the presence mGluR-dependent LTD (Oliet et al., 1997), as
of both AMPAR and mGluR antagonists, while the T-type VGCC channel blocker Ni2+ and the
Ehlers (2000) applied 20 mM NMDA alone for PKC inhibitory peptide, PKC19–36, did not affect
1–2 min, possibly suggesting an additional role for the induction of NMDAR-dependent LTD (Oliet
ambient activation of mGluRs in the activation of et al., 1997).
phosphatases that lead to AMPAR internalization.
Interestingly, the findings of Ehlers (2000) seem
mGluR-dependent LTD
to support the hypothesis that AMPAR dephos-
phorylation is a step toward AMPAR internaliza-
It is now clear that there exist multiple forms of
tion. They observed that application of NMDA,
mGluR-dependent LTD, depending on develop-
which induces AMPAR endocytosis, additionally
mental stage and mGluR subtype involved. The
induces the rapid dephosphorylation of GluR1 at
mGluRs are a family of heterotrimeric guanosine
the SER-845 PKA site. Both events are temporally
triphosphate-binding protein (G protein)-coupled
correlated, as NMDA-induced AMPAR internali-
receptors (Sugiyama et al., 1987; Tanabe et al.,
zation peaked shortly after maximal dephos-
1992). They are divided into three groups, mGluR
phorylation of GluR1. Additionally, internalized
I, mGluR II, and mGluR III, which are made up
GluR1 subunits showed no phosphorylation at the
of eight subtypes. This categorization is based on
PKA site SER-845, but were still phosphorylated at
their amino acid sequence homology, agonist
the CaMKII site SER-831. Both AMPAR end-
selectivity, and second messenger coupling (Conn
ocytosis and dephosphorylation of GluR1 at the
and Pin, 1997).
SER-845 PKA site, were prevented by Ca2+-free
medium, FK-506, high concentrations of okadaic
acid (1 mM) or tautomycin, but not prevented by mGluR-dependent LTD in neonatal rats. In neo-
low concentrations of okadaic acid. Ehlers (2000) natal rats electrical stimulation (5 Hz/3 min) induces
suggested that NMDAR activity results in Ca2+ both NMDAR-independent and mGluR-dependent
influx and activation of a PP cascade (including LTD (P3–7: Bolshakov and Siegelbaum, 1994).
PP2B and PP1) that triggers AMPAR dephos- Metabotropic GluR-dependent LTD has been
phorylation followed by endocytosis. reported to require activation of group I mGluRs
There is evidence that NMDAR-dependent (but group II mGluR blockade was not tested; Oliet
LTD is expressed not only postsynaptically as et al., 1997/P6–7; Nosyreva and Huber, 2005/
AMPAR dephosphorylation and endocytosis (Lee P8–15). mGluR activation is only necessary for in-
et al., 1998; Beattie et al., 2000; Ehlers, 2000; duction of LTD in neonatal rats, since maintenance
Zhang et al., 2006), but presynaptically as a long- is not inhibited by the mGluRI inhibitor (+)-a-
term reduction in neurotransmitter release (Zhang methyl-4-carboxyphenylglycine (MCPG; Bolshakov
et al., 2006). Zhang et al. (2006) showed, by eval- and Siegelbaum, 1994) (Fig. 2).
uating changes in paired-pulse ratio (PPR), two- mGluR-dependent LTD, which is not affected by
photon imaging of FM1–43 release and variance- NMDAR blockade (Li et al., 2002), can be also
mean analyses, that NMDA-LTD is associated induced by 1 Hz stimulation in neonatal rats (P6–8;
with reduced transmitter release probability in Li et al., 2002; P4–10; Feinmark et al., 2003).
slices from juvenile rats (P14–19). Additionally, mGluR-LTD induced by 1 Hz in neonatal rats has
483

Fig. 2. Cascades involved with mGluR-dependent LTD in neonatal rats at Schaffer collateral–CA1 synaspes. AA: arachidonic acid,
Ca: Ca2+, HPETE: hydroperoxyeicosatetraenoic acid, IP3: inositol-1,4,5-trisphosphate, MAPK: mitogen-activated protein kinase,
mGluR: metabotropic glutamate receptor, PIP2: phosphatidylinositol-3,4-bisphosphate, PLA2: phospholipase A2, PLC: phospholipase
C, Post: postsynapse, VGCC: voltage-gated Ca2+ channels.

been claimed to be dependent on group II mGluRs as postsynaptic mechanisms like PS or changes in


only (as it was not affected by an mGluRI blocker; the surface expression of AMPARs play no role
Li et al., 2002) but other groups detected a blockade in group I mGluR-LTD (Nosyreva and Huber,
of LTD after mGluR5 blockade (group II mGluR 2005). Instead, group I mGluR-LTD in neonatal
blockade was not tested; Feinmark et al., 2003). rats appears to be expressed presynaptically as a
Induction of group I mGluR-LTD is prevented long-term decrease in transmitter release, since the
by the Ca2+ chelator EGTA and by the blocker of expression of LTD at this age is accompanied by
L-type VGCCs nitrendipine in slices from neonatal changes in PPR (Nosyreva and Huber, 2005),
(P3–8) rats (Bolshakov and Siegelbaum, 1994), coefficient of variation of EPSCs (Bolshakov and
suggesting that a combination of Ca2+ influx via Siegelbaum, 1994), a decrease in the frequency but
these channels and release from intracellular stores not amplitude, of miniature EPSCs (Bolshakov and
is needed for this form of LTD. Elevation of Siegelbaum, 1994), and a decreased rate of release
postsynaptic [Ca2+] and mGluR activation (with of the fluorescent marker FM1–43 (N-(3-triethyl-
the specific mGluR agonist ACPD) is sufficient to ammoniumpropyl)-4-(4-(dibutylamino) styryl) pyr-
induce LTD (Bolshakov and Siegelbaum, 1995), idinium dibromide; Zakharenko et al., 2002).
indicating a postsynaptic site of induction. Never- Since this form of LTD is induced postsynaptic-
theless, studies indicate that group I mGluR-LTD ally but expressed presynaptically, a retrograde mes-
in neonatal rats is not expressed postsynaptically, senger is needed to transfer the signal from the
484

postsynaptic to the presynaptic site (Bolshakov and juvenile rats by LFS stimulation (5 Hz/3 min;
Siegelbaum, 1994). AA has been suggested to be the P11-35; Oliet et al., 1997; Nicoll et al., 1998) and
retrograde transmitter involved in neonatal mGluR- by paired-pulse LFS (PP-LFS; paired pulses at
LTD (P4–7; Bolshakov and Siegelbaum, 1995), be- 50 ms interstimulus intervals, applied at a 1 Hz
cause AA application can induce LTD (Bolshakov frequency for 15 min; P21–30; Huber et al., 2000)
and Siegelbaum, 1995), an inhibitor of PLA2 that in the presence of the NMDAR antagonist
initiates the AA signaling cascade by liberating AP5. Additionally, chemical activation of group
AA from membrane phospholipids, blocks LTD I mGluRs with (S)-3,5-dihydroxyphenylglycine
(Bolshakov and Siegelbaum, 1995) and metabolites (DHPG) can induce an LTD in juvenile rats (Hub-
of AA have been shown to mediate mGluR-LTD in er et al., 2001) which is independent of NMDAR
neonatal rats (Feinmark et al., 2003). Feinmark activation (Huang and Hsu, 2006). LFS-induced
et al. (2003) showed that 12(S)-hydroperoxyeicosa- group I mGluR-dependent and DHPG-LTD show
5Z,8Z,10E,14Z-tetraenoic acid (12-S-HPETE), a 12- mutual occlusion, which suggests similar induction/
lipoxygenase metabolite of AA, is actively recruited expression mechanisms. Group I mGluR-dependent
during the induction of mGluR-LTD, since (1) a LTD in juvenile rats seems to be mediated by group
mouse in which the leukocyte-type 12-lipoxygenase I mGluRs, since it is blocked by the selective group
(the neuronal isoform) was deleted through homol- 1 antagonist AIDA ([CRS]-1-aminoindan-1,5-di-
ogous recombination was deficient in mGluR-LTD, carboxylic acid) (Oliet et al., 1997), but not by the
(2) pharmacological inhibition of 12-lipoxygenase selective group II mGluR antagonist EGLU (2S-a-
also blocked induction of mGluR-LTD, and (3) di- ethylglutamic acid), or the selective group III
rect application of 12-S-HPETE to hippocampal mGluR antagonist MAP4 ((S)-2-amino-2-methyl-
slices induced a LTD of synaptic transmission that 4-phosphonobutanic acid) (Huang and Hsu, 2006).
mimicked and occluded mGluR-LTD induced by The responsible group I mGluR subtype appears to
synaptic stimulation. be mGluR5, since group I mGluR-LTD is not
Group I mGluR-LTD has also been reported to affected by the mGluR1 selective antagonist 4CPG
require the postsynaptic activation of p38 MAPK in ((S)-4-carboxyphenylglycine) (Oliet et al., 1997;
neonatal rats, since introduction of the p38 MAPK Huang and Hsu, 2006), is blocked by the mGluR5
inhibitor SB203580 into a postsynaptic CA1 py- selective antagonist MPEP (2-methyl-6-(phenyl-
ramidal neuron greatly diminished the induction of ethynyl)pyridine) (Huang et al., 2004; Huang and
this form of LTD, while the p42/44 inhibitor PD Hsu, 2006) and is absent in homozygote mGluR5
98059 had no effect (P4–11; Bolshakov et al., 2000). knockout mice (Huber et al., 2001). Established
Additionally, it has been demonstrated that dial- DHPG-LTD is completely reversed by application
yzing CA1 neurons with the active, dually phos- of the mGluR antagonists MCPG or LY341495 and
phorylated p38 MAPK induces a progressive LTD is reestablished once the antagonists are
depression in the amplitude of the EPSC which washed out, suggesting that the induction of
occludes LTD in response to subsequent 5 Hz elec- DHPG-LTD involves either an alteration in func-
trical stimulation. P38 MAPK provides a link to the tional state of the receptors or the switching on of a
AA pathway since P38 MAPK can activate cyto- tonically active mGluR receptor (Huang and Hsu,
solic PLA2 (cPLA2) by phosphorylation (Borsch- 2006). In contrast to the requirement for persistent
Haubold et al., 1997). cPLA2 is also regulated by mGluRI activation, the Gq-mediated downstream
[Ca2+]; Ca2+ promotes binding of the enzyme to pathways of group I mGluRs do not seem to be
phospholipid substrates by causing the translocation necessary for the maintenance of mGluRI-LTD;
of cPLA2 to nuclear and other cellular membranes neither the phospholipase C (PLC) blocker U73122,
through a Ca2+-dependent lipid-binding motif the PKC inhibitor Bis-1 nor the depletion of intra-
(Clark et al., 1995). cellular Ca2+ pools by thapsigargin or cyclopiazonic
acid have any effect on pre-established DHPG-LTD
mGluR-dependent LTD in juvenile rats. Group I (Reyes and Stanton, 1996; Huang and Hsu, 2006).
mGluR-dependent LTD has been induced in Instead, a downstream protein tyrosine phosphatase
485

(PTP) signaling cascade seems to be required for the of group I mGluR-LTD in juvenile rats is
expression of DHPG-LTD (see below; Huang and prevented by blockers of T-type VGCCs, but not
Hsu, 2006) (Fig. 3). L-type VGCCs (Oliet et al., 1997; Nicoll et al.,
The induction of group I mGluR-LTD requires 1998), suggesting a developmental switch. Addi-
postsynaptic activation of PKC (Oliet et al., 1997; tionally, it has been shown that group I mGluR-
Nicoll et al., 1998; Huang and Hsu, 2006), but is not LTD in juvenile rats is prevented by inhibitors of
affected by PP1/2A inhibitors (P11–35: Oliet et al., mRNA translation but not transcription (Huber
1997; Nicoll et al., 1998; P21–28: Huang and Hsu, et al., 2000), indicating that new PS from existing
2006), contrasting with NMDAR-LTD, which does mRNA species is a component of mGluR-LTD by
appear to involve postsynaptic activation of PP2B this developmental stage.
(P21–30: Kameyama et al., 1998). Like NMDAR- There are contradictory findings of the role of
LTD, induction of group I mGluR-LTD relies on p38 MARK for the induction of group I mGluR-
increases in postsynaptic [Ca2+] (Oliet et al., 1997; LTD; one study reported that group I mGluR-
Nicoll et al., 1998; but see Fitzjohn et al., 2001/ LTD was not affected by p38 MARK inhibitors,
P12–18). but was blocked by p42/44 MARK and ERK
In contrast to group I mGluR-LTD in neonatal inhibitors instead (Gallagher et al., 2004). Another
rats (Bolshakov and Siegelbaum, 1994), induction study found no effect of p42/44 MARK inhibition

Fig. 3. Cascades involved with mGluR-dependent LTD in juvenile rats at Schaffer collateral–CA1 synapses. Ca: Ca2+, DAG: diacyl
glycerol, IP3: inositol-1,4,5-trisphosphate, MAPK: mitogen-activated protein kinase, mGluR: metabotropic glutamate receptor, P:
phosphate, PIP2: phosphatidylinositol-3,4-bisphosphate, PKC: protein kinase C, PLC: phospholipase C, Post: postsynapse, PTP:
protein tyrosine phosphatase, VGCC: voltage-gated Ca2+ channels.
486

on group I mGluR-LTD, but a potent effect of (Huang et al., 2004; Nosyreva and Huber, 2005;
p38 MARK inhibition (Huang et al., 2004). Huang and Hsu, 2006). During expression of
Huang et al. (2004) further showed, with specific group I mGluR-LTD, the decrease in AMPAR
inhibitors, that during group I mGluR-LTD, p38 surface expression has been shown to be regulated
MARK is activated via a specific cascade that in- by p38 MARK. p38 MARK accelerates a rapid
cludes activation of mGluR5, release of G protein loss of surface AMPARs by stimulating the activ-
bg-subunits, which, in turn, promotes the ex- ity of guanyl nucleotide dissociation inhibitor
change of GDP with GTP of the small GTPase (GDI), extracting Rab5 from endosomal mem-
Rap1, causing Rap1 to activate a sequential kinase branes and forming a GDI-Rab5 complex (Cavalli
cascades that includes MAPK kinase 3/6 and p38 et al., 2001; Huang et al., 2004). This complex
MAPK. is required for the sequestration of the ligand-
There is evidence for both pre- and postsynaptic receptor-complex into clathrin-coated pits
expression of group I mGluR-LTD in juvenile rats. (McLauchlan et al., 1998). Huang and Hsu
A presynaptic site of expression is suggested by (2006) showed that a transient activation of p38
studies showing that expression of LTD is accom- MARK is sufficient to stimulate AMPAR end-
panied by a change in PPR (Fitzjohn et al., 2001; ocytosis during group I mGluR-dependent LTD,
Nosyreva and Huber, 2005), increase in EPSC fail- since maintenance of previously-induced DHPG-
ure rate (Fitzjohn et al., 2001), a change in the LTD is not affected by the selective p38 MAPK
coefficient of variation of EPSCs (Fitzjohn et al., inhibitor SB203580, and the increase in phosphor-
2001), a decrease in the frequency but not the am- ylation of p38 MAPK elicited by DHPG is tran-
plitude, of miniature EPSCs (Fitzjohn et al., 2001), sient. In contrast to p38 MAPK, postsynaptic
and a decrease in the frequency, but not in ampli- PTPs appear to be necessary for the continuing
tude, of Sr2+ asynchronously-evoked quantal events expression of group I mGluR-dependent LTD,
(which only occurs at the subset of synapses that are because DHPG-LTD is reversed by postsynaptic
stimulated; Oliet et al., 1997). But these data are loading of pyramidal neurons with a PTP inhibitor
challenged by conflicting findings from Zhang et al. (Huang and Hsu, 2006). Interestingly, it has been
(2006) which showed a lack of presynaptic changes demonstrated that levels of tyrosine phosphorylat-
for DHPG-LTD and pharmacologically isolated ion of GluR2 AMPAR subunits (phosphorylated
group I mGluR-LTD, measured using changes in under basal conditions (Ahmadian et al., 2004)),
PPR, two-photon imaging of FM1–43 release, and but not GluR1 subunits, are dramatically de-
variance-mean analysis. This study did, however, creased during DHPG-LTD, suggesting that AM-
indicate that coactivation of group II mGluRs could PARs are targets of PTPs during maintenance of
recruit presynaptic changes in release. group I mGluR-dependent LTD.
There is evidence that group I mGluR-LTD is Another chemically induced form of LTD at
expressed by postsynaptic NMDAR and AMPAR Schaffer collateral–CA1 synapses in hippocampal
internalization, since group I mGluR activation slices from juvenile rats (14–24 days old) involves
induces a PS-dependent loss of postsynaptic AM- the activation of group II mGluRs (Santschi et al.,
PARs and NMDARs in cultured hippocampal 2006; Zhang et al., 2006). While the selective
neurons (Snyder et al., 2001), and group I mGluR mGluR II agonist DCGIV [(2S,20 R, 30 R)-2-(20 ,30 -
activation induces an LTD in cultured slices which dicarboxycyclopropyl)glycine] alone only induces
is dependent on AMPAR endocytosis (Xiao et al., a short-term depression which fully reverses upon
2001). In acute hippocampal slices from juvenile drug washout, a combination of DCGIV applica-
rats, group I mGluR activation produces a de- tion and elevation of endogenous intracellular
crease in AMPAR surface expression (GluR1 and [cGMP] by application of the type V phosphodi-
GluR2/3) which was completely blocked by the esterase inhibitor zaprinast, is sufficient to induce
broad-range mGluR antagonist LY341495, while LTD (Santschi et al., 2006). It is probable that
the late phase (60 min after LTD induction) was mGluRs II promote LTD by inhibiting adenylate
blocked by the translational inhibitor anisomyin cyclase, because pairing application of zaprinast
487

with either activation of A1 adenosine receptors, CaMK activity may tend to inhibit DHPG-
which also inhibit adenylate cyclase, or with direct induced LTD. Another possibility is that DHPG
pharmacologic inhibition of PKA activity, are simultaneously induces KN-62 insensitive LTD
each sufficient to elicit LTD (Santschi et al., 2006). and a smaller KN-62-sensitive LTP; blockade of
LTD elicited by any of these methods of pairing the latter leading to the apparent enhancement
elevated [cGMP] with reduced [cAMP], is induced of the former.
and expressed presynaptically: selective postsynap- DHPG-LTD is also facilitated by inhibitors of
tic inhibition of PKA does not enhance LTD, PP1 and PP2A, but not affected by inhibition of
while postsynaptic inhibition of PKG does not PP2B (Schnabel et al., 2001), which indicates that
block any of these chemical forms of LTD (San- some form of phosphorylation of targets of PP1
tschi et al., 1999). Consistent with this conclusion, and PP2A is important to DHPG-induced LTD
it has been shown that coapplication of zaprinast and that this phosphorylation is actively opposed
with a synthetic PKA inhibitor (H-89) significantly by dephosphorylation by PP1 or PP2A.
and persistently decreases evoked presynaptic ve- DHPG-induced LTD can be reversed by the
sicular release of the fluorescent marker FM1–43 non-selective mGluR antagonist MCPG (Palmer
from Schaffer collateral terminals in hippocam- et al., 1997; Fitzjohn et al., 1998, 1999; Watabe
pal slices (Stanton et al., 2001, 2003), and from et al., 2002), and by the group I/II mGluR antag-
isolated rat hippocampal presynaptic terminals onist LY341495 (Fitzjohn et al., 1998; Watabe
(Bailey et al., 2003). et al., 2002), but LTD is reestablished following
washout of either antagonist. One explanation for
mGluR-dependent LTD in adult rats. (a) DHPG- the transient effect of the mGluR antagonists
induced mGluRI-dependent LTD could be that endogenous glutamate tonically ac-
The group I mGluR agonist DHPG also induces tivates mGluRIs. A second possibility is that mGlu
LTD in slices from adult rats (8–10 weeks old), un- receptor agonists and mGlu receptor antagonists
der conditions of increased excitability (in Mg2+- switch the mode of activity of mGluRI (i.e., act as
free medium; Palmer et al., 1997). Interestingly, this agonists and inverse agonists) with respect to
form of LTD does not occlude stimulus-induced downstream events, i.e., they activate opposing
(2 Hz/10 min) LTD (as opposed to the studies of signaling pathways. MCPG-induced reversal of
Huber and colleagues, see above). DHPG-induced LTD does not involve activation
Based on studies in immature rats, it was of CaMKII, PKA, or release of Ca2+ from intra-
expected that DHPG induces LTD via the activa- cellular stores (Schnabel et al., 1999a, b), but does
tion of group I mGluR-coupled intracellular signal involve the activation of PP1/PP2A (Schnabel et
cascades like PKC and PKA activation, and release al., 2001). Thus, DHPG-induced LTD seems to
of Ca2+ from intracellular stores. Surprisingly, depend on the net activity of some yet to be de-
DHPG-induced LTD has been reported not to be termined protein kinase(s) and PP1/PP2A. Acti-
affected by PKC or PKA inhibitors, or by depleting vation of mGluRs by DHPG switches the
intracellular Ca2+ stores (Schnabel et al., 1999a, receptor-effector cascade into a mode that pro-
2001). Instead, induction, but not expression, of motes phosphorylation. Application of mGluRI
DHPG-LTD is prevented by inhibitors of PTPs antagonists can temporarily reverse the phosphor-
(Moult et al., 2002), but the molecular targets of ylation mode via transient activation of a cascade
tyrosine dephosphorylation remain to be identified. that involves dephosphorylation via PP1/PP2A.
DHPG-LTD is enhanced by treatment with While DHPG-LTD is induced postsynaptically
KN-62, an inhibitor of Ca2+/CaM-dependent pro- (group I mGluRs are only located postsynaptical-
tein kinases (CaMKs) (including CaMKII; Schna- ly; Conn and Pin, 1997), it has been suggested to
bel et al., 1999b), which indicates that some form be expressed presynaptically as inhibition of volt-
of dephosphorylation of proteins which are phos- age-sensitive Ca2+ channels and K+ channels
phorylated by CaMKs may be important to (Watabe et al., 2002; Tan et al., 2003) at Schaffer
DHPG-induced LTD. It could be that basal collateral–CA1 synapses. A presynaptic locus of
488

expression is supported by studies reporting that: voltage levels driven by many other synaptic con-
(1) DHPG induces a persistent depression of tacts, it appears that long-term changes are driven
NMDA, as well as AMPA, receptor-mediated by pairing of local spine [Ca2+] and postsynaptic
components of EPSCs (Watabe et al., 2002), (2) depolarization within a certain range. Postsynap-
the expression of LTD is accompanied by a change tically, increases in [Ca2+] that preferentially ac-
in PPR (Faas et al., 2002; Rouach and Nicoll, tivate PPs and, perhaps independently, AMPAR
2003; Tan et al., 2003); and (3) a lack of change in internalization appear to elicit postsynaptic LTD.
postsynaptic sensitivity to AMPA in CA1, meas- Presynaptically, the postsynaptic Ca2+-dependent
ured by focally elicited AMPAR-currents evoked activation of NOS generates NO which diffuses
by rapid uncaging of glutamate, following DHPG- locally to proximal presynaptic terminals, where it
LTD (Rammes et al., 2003). The depression must pair with activation of G protein-coupled
induced by DHPG is dependent on postsynaptic, receptors that inhibit adenylate cyclase to drive
GTP-dependent signaling pathways, as it is a cyclic GMP/PKG/cyclic ADPribose and Ca2+-
blocked after replacement of GTP with guano- dependent presynaptic cascade that causes LTD of
sine-50 -O-(2-thiodiphosphate) (GDPgS; Watabe et vesicular transmitter release from the rapidly-
al., 2002). recycling vesicle pool. NMDAR-dependent LTD
(b) Stimulus-induced mGluR-dependent LTD appears to involve both presynaptic and postsy-
Stimulus-induced (2 Hz, 900 pulses during naptic alterations, while group I mGluR-depend-
GABAA-blockade) LTD in adult rats (7 weeks ent LTD is expressed purely postsynaptically, and
old) shows different mechanisms than DHPG- the activation of group II mGluRs recruits presy-
induced LTD (Otani and Connor, 1998). LTD de- naptic cascades that may or may not be shared
pends on mGluR, but not NMDAR, activation in with NMDAR-LTD. Thus, excitatory LTD in
adult rats, as it is blocked by the mGluR antag- field CA1 is a complex of multiple alterations
onist MCPG, but not the NMDAR antagonist across the synapse. The extensive information now
AP5. An increase in postsynaptic [Ca2+] is needed available about relatively rapid biochemical events
to induce this form of LTD, since postsynaptic during the induction of LTD is not, as of yet,
infusion of the Ca2+ chelator BAPTA blocked in- matched by sufficient data about whether and how
duction of LTD. A transient (20–30 s) increase in these changes lead to precise stimulation or sup-
intracellular [Ca2+] could be measured during the pression of particular genes, or of retraction or
2 Hz stimulation. This [Ca2+] increase was the re- even complete elimination of synaptic contacts.
sult of voltage-gated influx, rather than intracel-
lular Ca2+ release triggered by mGluR activation,
since it was blocked by hyperpolarization of the LTD in the dentate gyrus
neuronal soma to 110 mV, but not by MCPG.
Injection of the PKC inhibitor peptide PKC19–36 Induction
into the postsynaptic neuron completely blocked
LTD, verifying that PKC activation, probably via LTD in the DG can also be induced by a variety of
mGluRI, is necessary for the induction of this stimulus protocols. The most common induction
form of LTD. protocol is LFS (1–2 Hz) of the medial perforant
path (MPP) to induce homosynaptic LTD (in
Summary: CA1 LTD. Current data on LTD at vitro: O’Mara et al., 1995a; in vivo: Abraham,
Schaffer collateral–CA1 synapses is becoming sub- 1996; Trommer et al., 1996; Pöschel and Man-
stantial and reveals a synapse where distinct pre- ahan-Vaughan, 2007). Heterosynaptic LTD has
and postsynaptic cascades cause long-term altera- also been demonstrated at both medial and lateral
tions on both sides of the synapse. Because only perforant path (LPP) synapses by high-frequency
the postsynaptic dendritic spine would seem to be stimulation (HFS) of the other synaptic input,
privy to Hebbian information about both presy- (Abraham and Goddard, 1983; Doyere et al.,
naptic glutamate release and global postsynaptic 1997). Associative LTD can also be induced in the
489

lateral perforant path, where it seems to share ex- causes the release of cytosolic Ca2+ from intra-
pression mechanisms with heterosynaptic LTD cellular stores (Conn and Pin, 1997).
(Christie et al., 1995). Christie et al. (1995) induced The mGluRI agonist DHPG induces an mGluRI-
primed-associative LTD of the lateral path by al- dependent LTD at medial perforant path synapses
ternating high-frequency conditioning bursts to which does not require NMDAR activation (Camo-
the medial path and single shocks to the lateral deca et al., 1999). However, this LTD seems to occur
path at 100 ms intervals, provided these were ad- only in juvenile/young adult animals (up to 5 weeks
ministered 10 min after a homosynaptic priming of age in rats; Camodeca et al., 1999; Wang et al.,
stimulation of the lateral path (5 Hz, 80 pulses). 2007). There exist conflicting results concerning the
Lin et al. (2006) elicited LTD at lateral perforant involvement of PKC activation in DHPG-LTD.
path synapses by pairing antidromically-evoked While an early study reported a strong inhibition
postsynaptic APs (aAPs) with synaptic excitatory (but not complete block) of DHPG-LTD by the
postsynaptic potentials (fEPSPs), providing that PKC inhibitor Bisindolylmaleimide I (Bis I, 1 mM,
the fEPSP occurred rapidly after the aAP. Agonist inhibits a, b, g, e at this concentration; Camodeca
activation of mGluRs can also induce multiple et al., 1999), a later study by the same group did not
forms of LTD in the DG (O’Mara et al., 1995a; find an effect of Bis I on DHPG-LTD using identical
Camodeca et al., 1999; Huang et al., 1999a, b, c; protocols (Rush et al., 2002). Thus, it remains to be
Klausnitzer et al., 2004; Pöschel and Manahan- determined whether PKC activity is necessary for the
Vaughan, 2005). induction of DHPG-LTD. Furthermore, it has been
reported that DHPG-LTD is dependent on the ac-
tivation of p38 MAPK (Rush et al., 2002), as well as
non-receptor tyrosine kinases (Camodeca et al.,
mGluR agonist-induced LTD in the medial 1999), since the p38 MAPK inhibitor SB203580
perforant path (4-(4-fluorophenyl)-2-(4-methylsulfonylphenyl)-5-(4-
pyridyl) imidazol) and non-receptor tyrosine kinase
Metabotropic GluR activation is reported to be nec- inhibitor lavendustin A each strongly inhibited the
essary for the induction of LFS-LTD at medial per- induction of DHPG-LTD. Additional support for a
forant path-DG synapses (O’Mara et al., 1995a; role for tyrosine kinases and p38 MAPK in DHPG-
Huang et al., 1997; Camodeca et al., 1999; LTD is the finding that activation of MAPK path-
Klausnitzer et al., 2004; Pöschel et al., 2005). The ways (including p38 MAPK) by G protein-coupled
mGluR family of heterotrimeric guanosine tripho- receptors requires activation of the Src family of
sphate-binding protein (G protein)-coupled recep- non-receptor tyrosine kinases (Luttrell et al., 1999).
tors (Sugiyama et al., 1987; Tanabe et al., 1992) are MAPKs phosphorylate transcription and translation
divided into three groups: mGluR I, II, and III. factors, thereby regulating gene expression and PS
Despite their different intracellular coupling, agonist (Thomas and Huganir, 2004; Kelleher et al., 2004).
activation of all three groups has surprisingly been Consistent with this is a study showing that DHPG-
shown to result in the induction of LTD (Huang LTD is inhibited by PS blockade with either an-
et al., 1999a, b, c; Camodeca et al., 1999; Naie and isomycin or emetine (Naie and Manahan-Vaughan,
Manahan-Vaughan, 2005; Pöschel and Manahan- 2005). It seems that DHPG-LTD is expressed only
Vaughan, 2005) (Fig. 4) postsynaptically, as it is not accompanied by any
Group I mGluRs include the receptor subtypes change in paired-pulse depression (Camodeca et al.,
mGluR1 and mGluR5 (and their splice variants). 1999).
Group I mGluRs are coupled to a pathway that Some mGluRI subtypes (mGluR1a, mGluR5a
involves intracellular Ca2+ release from intracel- and mGluR5b) stimulate cAMP formation by po-
lular stores and PKC activation. They produce Gq- tentiating AC activation triggered by neurotrans-
mediated activation of PLC which hydrolyzes in- mitters that act on GS-coupled receptors (Conn
ositol-bisphosphate to IP3 and DAG. DAG, with and Pin, 1997; Moldrich et al., 2002). For instance,
Ca2+, coactivates protein kinase C, while IP3 in cultured striatal neurons, the ability of group I
490

PKC?

mGluRI
tyrosine kinase

p38 MAPK
mRNA- mRNA-
transcription translation

mGluRII ?
+ + +

PKC
2+
PKA
T-VGCC Ca
Post

Fig. 4. Cascades involved with mGluR-induced LTD at medial perforant path-DG synapses. Ca: Ca2+, MAPK: mitogen-activated
protein kinase, mGluR: metabotropic glutamate receptor, PKA: protein kinase A, PKC: protein kinase C, Post: postsynapse, VGCC:
voltage-gated Ca2+ channels.

mGluRs to potentiate cAMP formation induced potentiation is dependent on the coactivation of


by D1 dopamine receptor agonists correlates with group II mGluRs plus Gs-coupled receptors. The
their ability to generate IP3, and this potentiation potentiation of cAMP accumulation seems to be
requires PKC activity (Paolillo et al., 1998). Nev- mediated by synergistic interaction between group
ertheless, mGluRI potentiation of the AC-cAMP- I and group II mGluRs, or group II mGluRs and
PKA pathway does not seem to play a necessary b-adrenoceptors, which induces endogenous adeno-
role in inducing DHPG-LTD, because the PKA sine release (Moldrich et al., 2002). Adenosine sub-
inhibitor H-89 does not affect/prevent induction of sequently activates both A2A receptors, which are
DHPG-LTD (Camodeca et al., 1999). positively coupled to cAMP, and A1 receptors cou-
Group II mGluRs consist of mGluR2 and pled negatively to adenylate cyclase.
mGluR3, which inhibit forskolin and receptor- Metabotropic GluRII-dependent LTD is induced
induced increases in [cAMP]. A Gi-type of G pro- in the DG by selective agonist activation of
tein is probably involved in this coupling (Conn and mGluRIIs with (2S,20 R,30 R)-2-(20 30 -dicarboxycy-
Pin, 1997). It has also been reported that activation clopropyl) glycine (DCG-IV) or (1S,2S,5R,6S)-2-
of group II mGluRs potentiates increases in [cAMP] aminobicyclo[3.1.0]hexane-2-6-dicarboxylic acid
triggered by neurotransmitters that act on Gs- (LY354740): Huang et al., 1999a, b, c; (S)-4-car-
coupled receptors in adult hippocampal slices (Conn boxy-3-hydroxyphenylglycine (4C3HPG): Pöschel
and Pin, 1997; Moldrich et al., 2002), i.e., this and Manahan-Vaughan, 2005), and also by
491

activation of the mGluRII subtype mGluR3 (with Group III mGluRs comprise the subtypes
N-acetylaspartylglutamate (NAAG): Huang et al., mGlu4, 6, 7, and 8 (and their splice variants) and
1999b; Pöschel and Manahan-Vaughan, 2005). Per- are linked to the inhibition of Gs-activated ACs
fusion with mGluRII agonists in vitro induces a (Conn and Pin, 1997). The induction pathways of
rapid depression of MPP-evoked EPSPs which is Group III mGluR-induced LTD have not been
maintained in the presence of the agonist but, fol- investigated, but it has been shown that Group III
lowing washout, is partially reversed to a depression mGluR-induced LTD is expressed independently
level which is maintained throughout the experiment of PS (Naie and Manahan-Vaughan, 2005).
(LTD was induced lasting at least 1 h following
washout; Huang et al., 1999a, b). Intracerebroven-
tricular injection of mGluRII agonists in vivo in- Electrically-induced LTD at medial PP-DG
duces a rapid (mGluRII agonist) or slowly synapses
developing depression (mGluR3 agonist) of MPP-
evoked EPSPs which is maintained throughout the Although an early study reported that the non-
experiment (up to 25 h after LFS; Pöschel and competitive NMDAR open-channel blockers keta-
Manahan-Vaughan, 2005). The rapidly evoked mine and phencyclidine were able to block the in-
reversible depression of the EPSP by mGluRII ago- duction of medial perforant path LTD (Desmond
nists in vitro was associated with a change in paired- et al., 1991), follow-up studies, using the specific
pulse depression which was not maintained during NMDAR antagonist AP5, have not confirmed the
LTD following washout of the agonist (Huang et dependence of LFS-LTD on NMDAR activation
al., 1999a, b), indicating that the reversible depres- at medial PP-DG synapses (O’Mara et al., 1995a;
sion may be expressed via a presynaptic reduction in in vitro: Trommer et al., 1996; Wang et al., 1997;
the probability of transmitter release, but that LTD in vivo: Pöschel and Manahan-Vaughan, 2005,
is induced either by a presynaptic reduction in the 2007). Nevertheless, postsynaptic intracellular
number of active release sites, and/or postsynaptic [Ca2+] elevations are a necessary step in the in-
changes. duction of LFS-LTD at medial PP-DG synapses,
While Huang et al. (1999a) reported a partial and can be provided either via Ca2+ influx into
block of mGluRII-induced LTD by the NMDAR dentate granule cells through low voltage-acti-
antagonist AP5, Pöschel and Manahan-Vaughan vated Ca2+ channels (T-type), or Ca2+ release
(2005) could not confirm a dependence of from intracellular stores (O’Mara et al., 1995b;
mGluR3-induced LTD on any NMDAR activa- Wang et al., 1997), or both. The intracellular
tion. This may mean that NMDAR blockade stores involved include both group I mGluR-acti-
affects only mGluR2-dependent LTD. In vitro, vated IP3-receptor-dependent and ryanodine-
mGluRII-induced LTD has been shown to depend receptor-sensitive postsynaptic stores (O’Mara
on depolarization, since the cessation of test stim- et al., 1995b; Wang et al., 1997). Ryanodine-re-
ulation during mGluRII agonist application pre- ceptor-sensitive Ca2+ stores are a target of the
vented the induction of LTD (Huang et al., NO-cGMP pathway via a PKG-dependent activa-
1999a). Furthermore, mGluRII-induced LTD, tion of a cyclase/hydrolase that generates cyclic
but not the reversible initial depression, in vitro ADPribose, a messenger that either enhances or
requires the activation of VGCCs, and of PKA or causes release from RyR-sensitive stores which
PKC (Huang et al., 1999a, c). The downstream mediate Ca2+-dependent Ca2+ release (CDCR).
pathways of PKA and PKC have yet to be inves- The NO-cGMP pathway has been shown to play a
tigated, but PS does not seem to be required for role in the induction of LFS-LTD at PP-DG
the induction of mGluRII-induced LTD, since it is synapses, as inhibition of NOS, guanylyl cyclase
not blocked in vivo by the translation-inhibitor and postsynaptic PKG all prevent the induction of
anisomycin or the transcription-inhibitor act- LFS-LTD (Wu et al., 1997, 1998). Furthermore,
inomycin D (Pöschel and Manahan-Vaughan, cGMP-dependent LTD induced by the type V
2005). phosphodiesterase inhibitor zaprinast shows
492

mutual occlusion with LFS-LTD, and zaprinast- et al., 1999), and with mGluRII-induced LTD
induced LTD is dependent on mGluR, but not (Huang et al., 1999a), which suggests common in-
NMDAR, activation (Wu et al., 1998). Since duction and/or maintenance of intracellular path-
postsynaptic inhibition of PKG and postsynaptic ways shared between a compound LFS-LTD and
blockade of ryanodine receptors were reportedly both mGluR I and II-dependent LTDs. Induction
sufficient to strongly inhibit LFS-LTD in the DG, of LFS-LTD fully blocks the subsequent induction
the NO-cGMP pathway seems to be induced of either mGluRI-induced LTD or mGluRII-
postsynaptically in this region, with NO acting induced LTD, whereas the induction of either
not as retrograde messenger as in CA1, but as an mGluRI-induced LTD or mGluRII-induced LTD
intercellular postsynaptic messenger in the DG only partially inhibits the subsequent induction of
(Fig. 5). LFS-LTD, suggesting that mGluRIs and mGluRIIs
LFS-LTD in the DG appears to be dependent on contribute with parallel/independent intracellular
all subtypes of mGluRs at medial PP-DG synapses. pathways to LFS-LTD. Furthermore, LFS-LTD is
It has been reported that LFS-LTD shows mutual blocked by antagonists of mGluRIs ((AIDA):
occlusion with mGluRI-induced LTD (Camodeca Camodeca et al., 1999), mGluRIIs (2S,1S0 ,2S0 -2-

Fig. 5. Cascades involved with LFS-induced LTD at medial perforant path-DG synapses. AC: adenylyl cyclase, Ca: Ca2+, cADPR:
cyclic adenosine diphosphate ribose, CaM: calmodulin, cGMP: cyclic guanosine monophosphate, DAG: diacyl glycerol, IP3: inositol-
1,4,5-trisphosphate, MAPK: mitogen-activated protein kinase, mGluR: metabotropic glutamate receptor, NO: nitric oxide, NOS: NO-
synthetase, PIP2: phosphatidylinositol-3,4-bisphosphate, PKA: protein kinase A, PKC: protein kinase C, PKG: protein kinase G,
PLC: phospholipase C, Post: postsynapse, VGCC: voltage-gated Ca2+ channels.
493

methyl-2-(20 -carboxycyclopropyl)glycine (MCCG): AMPARs are completely removed postsynaptical-


Huang et al., 1997; NAAG: Pöschel et al., 2005), ly). This interpretation is supported by the findings
and mGluRIIIs ((R.S)-r-cyclopropyl-4-phosphono- of Camodeca et al. (1999) and Zhang et al. (2006),
phenylglycine (CPPG): Klausnitzer et al., 2004). who found that DHPG-LTD in the DG was not
Thus, all mGluR subgroups seem to contribute to accompanied by a change in paired-pulse depres-
the induction of LFS-LTD, despite their different sion (Camodeca et al., 1999) and group I mGluRs
intracellular signaling pathways. Kinases which are did not contribute to the presynaptic expression of
part of different mGluR intracellular pathways, LFS-LTD in field CA1 (Zhang et al., 2006).
such as PKA (Huang et al., 1999a, c), PKC (Wang
et al., 1998; Huang et al., 1999a, c) and MAPK
(Murray and O’Connor, 2003), have all been Heterosynaptic LTD at lateral perforant path
reported to be necessary for the induction of LFS- synapses
LTD. Application of a PKC activator can also
directly induce a depression which occludes subse- Induction of heterosynaptic LTD in the lateral
quent induction of LFS-LTD (Wang et al., 1998) perforant path of the DG (by HFS stimulation of
supporting a role for PKC in mediating some por- the medial PP) has been shown in vitro (Christie
tion of LFS-LTD at MPP-DG synapses. The mon- and Abraham, 1992b) and in vivo (Abraham et al.,
omeric G protein Ras activates certain MAPK 2006) to depend on activation of both NMDARs
pathways (Thomas and Huganir, 2004) and the Ras and L-type VGCCs. Furthermore, immediate early
inhibitor manumycin A has also been shown to gene expression has been associated with the per-
significantly attenuate LFS-LTD (Murray and sistence of heterosynaptic LTD at both medial and
O’Connor, 2004). lateral perforant path synapses on dentate granule
One stimulation protocol (step depolarization of cells, i.e., the greatest immediate early gene ex-
a DG granule neuron from a holding potential pression occurred following a stimulus protocol
of 70 mV to 50 mV for 1.1 s applied during which consistently gave the longest-lasting LTD
HFS medial perforant path stimulus trains in the (Abraham et al., 1994).
presence of an NMDAR blocker) can also induce Induction of associative LTD in the lateral per-
LTD which depends only on mGluRI activation forant path, which is induced after priming of the
(Wu et al., 2001). This form of LTD required lateral path, does not depend on NMDAR activa-
membrane depolarization, was mediated by Ca2+ tion (Christie and Abraham, 1992a). Nevertheless,
influx via L-type Ca2+ channels plus Ca2+ release heterosynaptic LTD and prime-associative LTD of
from ryanodine-receptor-sensitive intracellular the lateral path seem to share a common expression
Ca2+ stores, and required activation of both mechanism, since they occlude one another (Christie
MAPK and PKC (Wu et al., 2001, 2004, Wang et al., 1995). It remains to be determined what non-
et al., 2007). Wu et al. (2001) detected decreases in NMDARs are involved in either.
potency and increases in failure rate after induc- The one study of spike-timing-dependent LTD
tion of mGluRI-dependent LTD. Although these at lateral perforant path synapses reported a
changes can be attributed to either presynaptic or dependence on NMDAR activation, and block-
postsynaptic changes, Wu et al. (2001) argue in ade by inhibitors of protein kinase C and PP2B
favor of a postsynaptic site of expression, stating (Lin et al., 2006).
that a decrease in potency could also be caused by Morphological changes associated with LTD
a decrease in the probability of release if more than have been so far investigated in only two studies;
one synapse was being activated and the increase Mezey et al. (2004), who reported that heterosy-
in failure rate, according to the silent synapse hy- naptic LTD of the medial path of the DG results in
pothesis of Liao et al. (1995), could also be caused an input-specific increase in axodendritic synapse
by a postsynaptic change associated with a failure density, whereas LTP at the lateral path was asso-
of detection of released transmitter as active ciated with an increase in perforated axospinous
synapses are converted into silent synapses (i.e., synapses in the potentiated area, and Zhou et al.
494

(2004), who reported an NMDAR, CN, and cofilin- References


dependent shrinkage of dendritic spines following
induction of LTD at Schaffer collateral–CA1 Abraham, W.C. (1996) Induction of heterosynaptic and ho-
mosynaptic LTD in hippocampal sub-regions in vivo. J.
synapses. Thus, the range of morphological altera-
Physiol. Paris., 90: 305–306.
tions that may be associated with LTD, their func- Abraham, W.C., Christie, B.R., Logan, B., Lawlor, P. and
tional effects on transmission, and the time course Dragunow, M. (1994) Immediate early gene expression as-
of such changes during the progression of LTD, are sociated with the persistence of heterosynaptic long-term de-
still unknown. pression in the hippocampus. Proc. Natl. Acad. Sci. U.S.A.,
91: 10049–10053.
Abraham, W.C. and Goddard, G.V. (1983) Assymetric rela-
Summary: DG LTD tionships between homosynaptic long-term potentiation
and heterosynaptic long-term depression. Nature, 305:
717–719.
This chapter makes it clear that significantly less is Abraham, W.C., Maxon-Parker, S.E., Irvine, G.I., Logan, B.
known about LTD in the DG compared to field and Gill, A.I. (2006) Induction and activity-dependent re-
CA1, or even the hippocampus proper in general. versal of persistent LTP and LTD in lateral perforant path
The majority of this evidence points to postsynap- synapses in vivo. Neurobiol. Learn. Mem., 86: 82–90.
Ahmadian, G., Ju, W., Liu, L., Wyszynski, M., Lee, S.H.,
tic alterations in MPP-DG LTD, and suggests that
Dunah, A.W., Taghibiglou, C., Wang, Y., Lu, J., Wong,
an NO/cGMP/PKG cascade produces long-term T.P., Sheng, M. and Wang, Y.T. (2004) Tyrosine phosphor-
postsynaptic LTD here, in contrast to CA1. In this ylation of GluR2 is required for insulin-stimulated AMPA
respect, MPP-LTD may have more in common receptor endocytosis and LTD. EMBO J., 23: 1040–1050.
with LTD in the striatum, than in field CA1. Alford, S., Frenguelli, B.G. and Collingridge, G.L. (1992)
However, there are enough studies suggesting Characterisation of Ca2+ signals induced in hippocampal
CA1 neurones by the synaptic activation of NMDA recep-
long-term presynaptic alterations in mGluR3- tors. J. Physiol. Lond., 469: 693–716.
dependent LTD (Pöschel et al., 2005), and in the Armstrong, D.L. (1989) Calcium channels regulation by cal-
actions of brain-derived neurotrophic factor cineurin, a Ca2+-activated phosphatase in mammalian brain.
(BDNF) that promote/elicit LTP (Gooney et al., Trends Neurosci., 12: 117–122.
2002), to leave the question of presynaptic long- Bailey, C.P., Trejos, J.A., Schanne, F.A. and Stanton, P.K.
(2003) Pairing elevation of [cyclic GMP] with inhibition of
term plasticity at MPP-DG synapses open to fur- PKA produces long-term depression of glutamate release
ther study. Likewise, the relative lack of much from isolated rat hippocampal presynaptic terminals. Eur. J.
knowledge about LPP-DG synapse LTD (or LTP) Neurosci., 17: 903–908.
makes this a fertile area of future study. Most re- Banke, T.G., Bowie, D., Lee, H., Huganir, R.L., Schousboe, A.
and Traynelis, S.F. (2000) Control of GluR1 AMPA receptor
ports have examined heterosynaptic LTD of LPP-
function by cAMP-dependent protein kinase. J. Neurosci.,
DG synapses, which are likely to involve postsy- 20: 89–102.
naptic alterations that depend on the spread of Beattie, E.C., Carroll, R.C., Yu, X., Morishita, W., Yasuda, H.,
postsynaptic depolarization and activation of volt- von Zastrow, M. and Malenka, R.C. (2000) Regulation of
age-dependent Ca2+ channels, and may differ sub- AMPA receptor endocytosis by a signaling mechanism
stantially from homosynaptic LPP LTD. Even shared with LTD. Nat. Neurosci., 3: 1291–1300.
Bito, H., Deisseroth, K. and Tsien, R.W. (1996) CREB phos-
more so than in field CA1, the connections be- phorylation and dephosphorylation: a Ca2+- and stimulus
tween biochemical cascades necessary for the duration-dependent switch for hippocampal gene expression.
induction of LTD, and eventual synaptic remode- Cell, 87: 1203–1214.
ling, are unexplored territory in the DG. Bolshakov, V.Y., Carboni, L., Cobb, M.H., Siegelbaum, S.A.
and Belardetti, F. (2000) Dual MAP kinase pathways medi-
ate opposing forms of long-term plasticity at CA3-CA1
Acknowledgments synapse. Nat. Neurosci., 3: 1107–1112.
Bolshakov, V.Y. and Siegelbaum, S.A. (1994) Postsynaptic in-
duction and presynaptic expression of hippocampal long-
This work was supported by National Institutes of term depression. Science, 264: 1148–1152.
Health Grant R01-NS44421 (to P.K.S.) and Na- Bolshakov, V.Y. and Siegelbaum, S.A. (1995) Hippocampal
tional Institute of Drug Abuse Grant R01-HD45754 long-term depression: arachidonic acid as a potential retro-
(to Dr. Peter Mundel). grade messenger. Neuropharmacology, 34: 1581–1587.
495

Borsch-Haubold, A.G., Kramer, R.M. and Watson, S.P. (1997) Cohen, P. (1989) The structure and regulation of protein
Phosphorylation and activation of cytosolic phospholipase phosphatases. Annu. Rev. Biochem., 58: 453–508.
A2 by 38-kDa mitogen-activated protein kinase in collagen- Colbran, R.J. (1992) Regulation and role of brain calcium/cal-
stimulated human platelets. Eur. J. Biochem., 245: 751–759. modulin-dependent protein kinase II. Neurochem. Int., 21:
Brandon, E.P., Idzerda, R.L. and McKnight, G.S. (1997) PKA 469–497.
isoforms, neural pathways, and behaviour: making the con- Conn, P.J. and Pin, J.P. (1997) Pharmacology and functions of
nection. Curr. Opin. Neurobiol., 7: 397–403. metabotropic glutamate receptors. Annu. Rev. Pharmacol.
Brandon, E.P., Zhuo, M., Huang, Y.Y., Qi, M., Gerhold, K.A., Toxicol., 37: 205–237.
Burton, K.A., Kandel, E.R., McKnight, G.S. and Idzerda, Cotman, C.W., Monaghan, D.T. and Ganong, A.H. (1988)
R.L. (1995) Hippocampal long-term depression and depot- Excitatory amino acid neurotransmission: NMDA receptors
entiation are defective in mice carrying a targeted disruption and Hebb-type synaptic plasticity. Ann. Rev. Neurosci., 11:
of the gene encoding the RI beta subunit of cAMP-dependent 61–80.
protein kinase. Proc. Natl. Acad. Sci. U.S.A., 92: 8851–8855. Cummings, J.A., Mulkey, R.M., Nicoll, R.A. and Malenka,
Bredt, D.S., Hwang, P.M., Glatt, C.E., Lowenstein, C., Reed, R.C. (1996) Ca2+ signaling requirements for long-term de-
R.R. and Snyder, S.H. (1991) Cloned and expressed nitric pression in the hippocampus. Neuron, 16: 825–833.
oxide synthase structurally resembles cytochrome P-450 red- Cummings, J.A., Nicola, S.M. and Malenka, R.C. (1994) In-
uctase. Nature, 351: 714–718. duction in the rat hippocampus of long-term potentiation
Brindle, P.K. and Montminy, M.R. (1992) The CREB family of (LTP) and long-term depression (LTD) in the presence of a
transcription activators. Curr. Opin. Genet. Dev., 2: 199–204. nitric oxide synthase inhibitor. Neurosci. Lett., 176: 110–114.
Cadd, G.G., Uhler, M.D. and McKnight, G.S. (1990) Holoen- Desmond, N.L., Colbert, C.M., Zhang, D.X. and Levy, W.B.
zymes of cAMP-dependent protein kinase containing the (1991) NMDA receptor antagonists block the induction of
neural form of type I regulatory subunit have an increased long-term depression in the hippocampal dentate gyrus of the
sensitivity to cyclic nucleotides. J. Biol. Chem., 265: anesthetized rat. Brain Res., 552: 93–98.
19502–19506. Domenici, M.R., Berretta, N. and Cherubini, E. (1998) Two
Camodeca, N., Breakwell, N.A., Rowan, M.J. and Anwyl, R. distinct forms of long-term depression coexist at the mossy
(1999) Induction of LTD by activation of group I mGluR in fiber-CA3 synapse in the hippocampus during development.
the dentate gyrus in vitro. Neuropharmacology, 38: Proc. Natl. Acad. Sci. U.S.A., 95: 8310–8315.
1597–1606. Doyere, V., Srebro, B. and Laroche, S. (1997) Heterosynaptic
Carroll, R.C., Lissin, D.V., von Zastrow, M., Nicoll, R.A. and LTD and depotentiation in the medial perforant path of the
Malenka, R.C. (1999) Rapid redistribution of glutamate re- dentate gyrus in the freely moving rat. J. Neurophysiol., 77:
ceptors contributes to long-term depression in hippocampal 571–578.
cultures. Nat. Neurosci., 2: 454–460. Dudek, S.M. and Bear, M.F. (1992) Homosynaptic long-term
Cavalli, V., Vilbois, F., Corti, M., Marcote, M.J., Tamura, K., depression in area CA1 of hippocampus and effects of N-
Karin, M., Arkinstall, S. and Gruenberg, J. (2001) The stress- methyl-D-aspartate receptor blockade. Proc. Natl. Acad. Sci.
induced MAP kinase p38 regulates endocytic trafficking via U.S.A., 89: 4363–4367.
the GDI:Rab5 complex. Mol. Cell, 7: 421–432. Ehlers, M.D. (2000) Reinsertion or degradation of AMPA re-
Chabot, C., Gagne, J., Giguere, C., Bernard, J., Baudry, M. ceptors determined by activity-dependent endocytic sorting.
and Massicotte, G. (1998) Bidirectional modulation of Neuron, 28: 511–525.
AMPA receptor properties by exogenous phospholipase A2 Eliot, L.S., Dudai, Y., Kandel, E.R. and Abrams, T.W. (1989)
in the hippocampus. Hippocampus, 8: 299–309. Ca2+/calmodulin sensitivity may be common to all forms of
Chaki, S., Muramatsu, M. and Otomo, S. (1994) Involvement neural adenylate cyclase. Proc. Natl. Acad. Sci. U.S.A., 86:
of protein kinase C activation in regulation of acetylcholine 9564–9568.
release from rat hippocampal slices by minaprine. Neuroc- Faas, G.C., Adwanikar, H., Gereau IV, R.W. and Saggau, P.
hem. Int., 24: 37–41. (2002) Modulation of presynaptic calcium transients by me-
Christie, B.R. and Abraham, W.C. (1992a) Priming of associ- tabotropic glutamate receptor activation: a differential role in
ative long-term depression in the dentate gyrus by theta fre- acute depression of synaptic transmission and long-term de-
quency synaptic activity. Neuron, 9: 79–84. pression. J. Neurosci., 22: 6885–6890.
Christie, B.R. and Abraham, W.C. (1992b) NMDA-dependent Feinmark, S.J., Begum, R., Tsvetkov, E., Goussakov, I., Funk,
heterosynaptic long-term depression in the dentate gyrus of C.D., Siegelbaum, S.A. and Bolshakov, V.Y. (2003) 12-lip-
anaesthetized rats. Synapse, 10: 1–6. oxygenase metabolites of arachidonic acid mediate metabo-
Christie, B.R., Stellwagen, D. and Abraham, W.C. (1995) Ev- tropic glutamate receptor-dependent long-term depression at
idence for common expression mechanisms underlying hete- hippocampal CA3-CA1 synapses. J. Neurosci., 23:
rosynaptic and associative long-term depression in the 11427–11435.
dentate gyrus. J. Neurophysiol., 74: 1244–1247. Fitzjohn, S.M., Bortolotto, Z.A., Palmer, M.J., Doherty, A.J.,
Clark, J.D., Schievella, A.R., Nalefski, E.A. and Lin, L.L. Ornstein, P.L., Schoepp, D.D., Kingston, A.E., Lodge, D.
(1995) Cytosolic phospholipase A2. J. Lipid. Mediat. Cell and Collingridge, G.L. (1998) The potent mGlu receptor an-
Signal, 12: 83–117. tagonist LY341495 identifies roles for both cloned and novel
496

mGlu receptors in hippocampal synaptic plasticity. Neuro- induce LTP and LTD in neocortical pyramidal cells. J.
pharmacology, 37: 1445–1458. Physiol. Paris, 90: 317–319.
Fitzjohn, S.M., Kingston, A.E., Lodge, D. and Collingridge, Heynen, A.J., Quinlan, E.M., Bae, D.C. and Bear, M.F. (2000)
G.L. (1999) DHPG-induced LTD in area CA1 of juvenile rat Bidirectional, activity-dependent regulation of glutamate re-
hippocampus; characterization and sensitivity to novel ceptors in the adult hippocampus in vivo. Neuron, 28:
mGluR antagonists. Neuropharmacology, 38: 1577–1583. 527–536.
Fitzjohn, S.M., Palmer, M.J., May, J.E., Neeson, A., Morris, Hilfiker, S., Pieribone, V.A., Czernik, A.J., Kao, H.T., Augus-
S.A. and Collingridge, G.L. (2001) A characterisation of tine, G.J. and Greengard, P. (1999) Synapsins as regulators
long-term depression induced by metabotropic glutamate re- of neurotransmitter release. Philos. Trans. R. Soc. Lond. B
ceptor activation in the rat hippocampus in vitro. J. Physiol., Biol. Sci., 354: 269–279.
537: 421–430. Hosaka, M., Hammer, R.E. and Sudhof, T.C. (1999) A
Fitzpatrick, J.S. and Baudry, M. (1994) Blockade of long-term phospho-switch controls the dynamic association of synap-
depression in neonatal hippocampal slices by a phospholi- sins with synaptic vesicles. Neuron, 24: 377–387.
pase A2 inhibitor. Dev. Brain Res., 78: 81–86. Hrabetova, S. and Sacktor, T.C. (1996) Bidirectional regulation
Freeman, E.J., Damron, D.S., Terrian, D.M. and Dorman, of protein kinase M zeta in the maintenance of long-term
R.V. (1991) 12-Lipoxygenase products attenuate the gluta- potentiation and long-term depression. J. Neurosci., 16:
mate release and calcium accumulation evoked by depolari- 5324–5333.
zation of hippocampal mossy fiber nerve endings. J. Hrabetova, S. and Sacktor, T.C. (2001) Transient transloca-
Neurochem., 56: 1079–1082. tion of conventional protein kinase C isoforms and per-
Gage, A.T., Reyes, M. and Stanton, P.K. (1997) Nitric-oxide- sistent downregulation of atypical protein kinase Mzeta in
guanylyl-cyclase-dependent and independent components of long-term depression. Brain Res. Mol. Brain Res., 95:
multiple forms of long-term synaptic depression. Hippocam- 146–152.
pus, 7: 286–295. Huang, C.C. and Hsu, K.S. (2006) Sustained activation of me-
Galione, A., White, A., Willmott, N., Turner, M., Potter, tabotropic glutamate receptor 5 and protein tyrosine
B.V.L. and Watson, S.P. (1993) cGMP mobilizes intracellu- phosphatases mediate the expression of (S)-3,5-di-
lar Ca21 in sea urchin eggs by stimulating cyclic ADP-ribose hydroxyphenylglycine-induced long-term depression in the
synthesis. Nature, 365: 456–459. hippocampal CA1 region. J. Neurochem., 96: 179–194.
Gallagher, S.M., Daly, C.A., Bear, M.F. and Huber, K.M. Huang, C.C., You, J.L., Wu, M.Y. and Hsu, K.S. (2004) Rap1-
(2004) Extracellular signal-regulated protein kinase activa- induced p38 mitogen-activated protein kinase activation
tion is required for metabotropic glutamate receptor-depend- facilitates AMPA receptor trafficking via the GDI.Rab5
ent long-term depression in hippocampal area CA1. J. complex. Potential role in (S)-3,5-dihydroxyphenylglycene-
Neurosci., 24: 4859–4864. induced long term depression. Biol. Chem., 279:
Garbers, D.L. (1990) The guanylyl cyclase receptor family. New 12286–12292.
Biol., 2: 499–504. Huang, K.P. (1989) The mechanism of protein kinase C acti-
Gooney, M., Shaw, K., Kelly, A., O’Mara, S.M. and Lynch, vation. Trends Neurosci., 12: 425–432.
M.A. (2002) Long-term potentiation and spatial learning are Huang, L., Killbride, J., Rowan, M.J. and Anwyl, R. (1999a)
associated with increased phosphorylation of TrkB and ex- Activation of mGluRII induces LTD via activation of pro-
tracellular signal-regulated kinase (ERK) in the dentate tein kinase A and protein kinase C in the dentate gyrus of the
gyrus: evidence for a role for brain-derived neurotrophic hippocampus in vitro. Neuropharmacology, 38: 73–83.
factor. Behav. Neurosci., 116: 455–463. Huang, L.Q., Rowan, M.J. and Anwyl, R. (1997) mGluR II
Goto, S., Yamamoto, H., Fukunaga, K., Iwasa, T., Matsuk- agonist inhibition of LTP induction, and mGluR II antag-
ado, Y. and Miyamoto, E. (1985) Dephosphorylation of mi- onist inhibition of LTD induction, in the dentate gyrus in
crotubule-associated protein 2, tau factor, and tubulin by vitro. Neuroreport, 8: 687–693.
calcineurin. J. Neurochem., 45: 276–283. Huang, L., Rowan, M.J. and Anwyl, R. (1999b) Induction
Grewal, S.S., York, R.D. and Stork, P.J. (1999) Extracellular- of long-lasting depression by (+)-alpha-methyl-4-car-
signal-regulated kinase signalling in neurons. Curr. Opin. boxyphenylglycine and other group II mGlu receptor lig-
Neurobiol., 9: 544–553. ands in the dentate gyrus of the hippocampus in vitro. Eur. J.
Hagiwara, M., Alberts, A., Brindle, P., Meinkoth, J., Fe- Pharmacol., 366: 151–158.
ramisco, J., Deng, T., Karin, M., Shenolikar, S. and Mont- Huang, L.Q., Rowan, M.J. and Anwyl, R. (1999c) Role of
miny, M. (1992) Transcriptional attenuation following protein kinases A and C in the induction of mGluR depend-
cAMP induction requires PP-1-mediated dephosphorylation ent long-term depression in the medial perforant path of the
of CREB. Cell, 70: 105–113. rat dentate gyrus in vitro. Neurosci. Lett., 274: 71–74.
Halpain, S., Hipolito, A. and Saffer, L. (1998) Regulation of F- Huber, K.M., Kayser, M.S. and Bear, M.F. (2000) Role for
actin stability in dendritic spines by glutamate receptors and rapid dendritic protein synthesis in hippocampal mGluR-
calcineurin. J. Neurosci., 18: 9835–9844. dependent long-term depression. Science, 288: 1254–1257.
Hansel, C., Artola, A. and Singer, W. (1996) Different thresh- Huber, K.M., Roder, J.C. and Bear, M.F. (2001) Chemical
old levels of postsynaptic [Ca2+]i have to be reached to induction of mGluR5- and protein synthesis-dependent
497

long-term depression in hippocampal area CA1. J. Ne- subunit is required for synaptic plasticity and retention of
urophysiol., 86: 321–325. spatial memory. Cell, 112: 631–643.
Izumi, Y. and Zorumski, C.F. (1993) Nitric oxide and long- Li, S.T., Kato, K., Tomizawa, K., Matsushita, M., Moriwaki,
term synaptic depression in the rat hippocampus. Neurore- A., Matsui, H. and Mikoshiba, K. (2002) Calcineurin plays
port, 4: 1131–1134. different roles in group II metabotropic glutamate receptor-
Kamal, A., Ramakers, G.M., Urban, I.J., De Graan, P.N. and and NMDA receptor-dependent long-term depression. J.
Gispen, W.H. (1999) Chemical LTD in the CA1 field of the Neurosci., 22: 5034–5041.
hippocampus from young and mature rats. Eur. J. Neurosci., Liao, D., Hessler, N.A. and Malinow, R. (1995) Activation of
11: 3512–3516. postsynaptically silent synapses during pairing-induced LTP
Kameyama, K., Lee, H.K., Bear, M.F. and Huganir, R.L. in CA1 region of hippocampal slice. Nature, 375(6530):
(1998) Involvement of a postsynaptic protein kinase A subst- 400–404.
rate in the expression of homosynaptic long-term depression. Lin, J.W., Ju, W., Foster, K., Lee, S.H., Ahmadian, G.,
Neuron, 21: 1163–1175. Wyszynski, M., Wang, Y.T. and Sheng, M. (2000) Distinct
Kauderer, B.S. and Kandel, E.R. (2000) Capture of a protein molecular mechanisms and divergent endocytotic pathways
synthesis-dependent component of long-term depression. of AMPA receptor internalization. Nat. Neurosci., 3:
Proc. Natl. Acad. Sci. U.S.A., 97: 13342–13347. 1282–1290.
Kelleher III, R.J., Govindarajan, A., Jung, H.Y., Kang, H. and Lin, Y.W., Yang, H.W., Wang, H.J., Gong, C.L., Chiu, T.H.
Tonegawa, S. (2004) Translational control by MAPK signa- and Min, M.Y. (2006) Spike-timing-dependent plasticity at
ling in long-term synaptic plasticity and memory. Cell, 116: resting and conditioned lateral perforant path synapses on
467–479. granule cells in the dentate gyrus: different roles of N-methyl-
Kemp, N., McQueen, J., Faulkes, S. and Bashir, Z.I. (2000) D-aspartate and group I metabotropic glutamate receptors.
Different forms of LTD in the CA1 region of the hippocam- Eur. J. Neurosci., 23: 2362–2374.
pus: role of age and stimulus protocol. Eur. J. Neurosci., 12: Lisman, J. (1989) A mechanism for the Hebb and the anti-Hebb
360–366. processes underlying learning and memory. Proc. Natl. Acad.
Keranen, L.M., Dutil, E.M. and Newton, A.C. (1995) Protein Sci. U.S.A., 86: 9574–9578.
kinase C is regulated in vivo by three functionally distinct Luttrell, L.M., Ferguson, S.S., Daaka, Y., Miller, W.E., Ma-
phosphorylations. Curr. Biol., 5: 1394–1403. udsley, S., Della Rocca, G.J., Kawakatsu, H., Owada, K.,
Kim, C.H., Chung, H.J., Lee, H.K. and Huganir, R.L. (2001) Luttrell, D.K., Caron, M.G. and Lefkowitz, R.J. (1999) Beta-
Interaction of the AMPA receptor subunit GluR2/3 with arrestin-dependent formation of beta2 adrenergic receptor-
PDZ domains regulates hippocampal long-term depression. Src protein kinase complexes. Science, 283: 655–661.
Proc. Natl. Acad. Sci. U.S.A., 98: 11725–11730. Lynch, G.S., Dunwiddie, T. and Gribkoff, V. (1977) Heterosy-
Klausnitzer, J., Kulla, A. and Manahan-Vaughan, D. (2004) naptic depression: a postsynaptic correlate of long-term po-
Role of the group III metabotropic glutamate receptor in tentiation. Nature, 266: 737–739.
LTP, depotentiation and LTD in dentate gyrus of freely Magee, J.C. and Johnston, D. (1997) A synaptically controlled,
moving rats. Neuropharmacology, 46: 160–170. associative signal for Hebbian plasticity in hippocampal neu-
Klee, C.B., Crouch, T.H. and Krinks, M.H. (1979) Calcineurin: rons. Science, 275: 209–213.
a calcium and calmodulin-binding protein of the nervous Malenka, R.C. and Bear, M.F. (2004) LTP and LTD: an em-
system. Proc. Natl. Acad. Sci. U.S.A., 76: 6270–6273. barrassment of riches. Neuron, 44: 5–21.
Kobayashi, M., Manabe, T. and Takahashi, T. (1996) Presy- Manahan-Vaughan, D. (1997) Group 1 and 2 metabotropic
naptic long-term depression at the hippocampal mossy fiber- glutamate receptors play differential roles in hippocampal
CA3 synapse. Science, 273: 648–650. long-term depression and long-term potentiation in freely
Krezel, W., Giese, K.P., Silva, A.J. and Chapman, P.F. (1999) moving rats. J. Neurosci., 17: 3303–3311.
Long-term depression is unimpaired in hippocampus of adult Manahan-Vaughan, D. and Reymann, K.G. (1995) 1S,3R-
alpha CaMKIIT286A mice. Soc. Neurosci. Abst., 25: 987. ACPD dose-dependently induces a slow-onset potentiation in
Lee, H.C. (1997) Mechanisms of calcium signaling by cyclic the rat hippocampal CA1 region in vivo. Neuropharm., 34:
ADP-ribose and NAADP. Physiol. Rev., 77: 1133–1164. 1103–1105.
Lee, H.K., Barbarosie, M., Kameyama, K., Bear, M.F. and Manahan-Vaughan, D., Kulla, A. and Frey, J.U. (2000) Re-
Huganir, R.L. (2000) Regulation of distinct AMPA receptor quirement of translation but not transcription for the main-
phosphorylation sites during bidirectional synaptic plasticity. tenance of long-term depression in the CA1 region of freely
Nature, 405: 955–959. moving rats. J. Neurosci., 20: 8572–8576.
Lee, H.K., Kameyama, K., Huganir, R.L. and Bear, M.F. Markram, H., Lubke, J., Frotscher, M. and Sakmann, B. (1997)
(1998) NMDA induces long-term synaptic depression and Regulation of synaptic efficacy by coincidence of postsynap-
dephosphorylation of the GluR1 subunit of AMPA receptors tic APs and EPSPs. Science, 275: 213–215.
in hippocampus. Neuron, 21: 1151–1162. McDonald, B.J., Chung, H.J. and Huganir, R.L. (2001) Iden-
Lee, H.K., Takamiya, K., Han, J.S., Man, H., Kim, C.H., tification of protein kinase C phosphorylation sites within the
Rumbaugh, G., Yu, S., Ding, L., He, C. and Petralia, R.S. AMPA receptor GluR2 subunit. Neuropharmacology, 41:
(2003) Phosphorylation of the AMPA receptor GluR1 672–679.
498

McLauchlan, H., Newell, J., Morrice, N., Osborne, A., West, Oliet, S.H., Malenka, R.C. and Nicoll, R.A. (1997) Two distinct
M. and Smythe, E. (1998) A novel role for Rab5-GDI in forms of long-term depression coexist in CA1 hippocampal
ligand sequestration into clathrin-coated pits. Curr. Biol., 8: pyramidal cells. Neuron, 18: 969–982.
34–45. O’Mara, S.M., Rowan, M.J. and Anwyl, R. (1995a) Metabo-
Meszaros, L.G., Bak, J. and Chu, A. (1993) Cyclic ADP-ribose tropic glutamate receptor-induced homosynaptic long-term
as an endogenous regulator of the non-skeletal type ryano- depression and depotentiation in the dentate gyrus of the rat
dine receptor Ca2+ channel. Nature, 364: 76–79. hippocampus in vitro. Neuropharmacology, 34: 983–989.
Mezey, S., Doyere, V., De Souza, I., Harrison, E., Cambon, K., O’Mara, S.M., Rowan, M.J. and Anwyl, R. (1995b) Dantrolene
Kendal, C.E., Davies, H., Laroche, S. and Stewart, M.G. inhibits long-term depression and depotentiation of synaptic
(2004) Long-term synaptic morphometry changes after in- transmission in the rat dentate gyrus. Neuroscience, 68:
duction of long-term potentiation and long-term depression 621–624.
in the dentate gyrus of awake rats are not simply mirror Otani, S. and Connor, J.A. (1998) Requirement of rapid Ca2+
phenomena. Eur. J. Neurosci., 19: 2310–2318. entry and synaptic activation of metabotropic glutamate re-
Moult, P.R., Schnabel, R., Kilpatrick, I.C., Bashir, Z.I. and ceptors for the induction of long-term depression in adult rat
Collingridge, G.L. (2002) Tyrosine dephosphorylation un- hippocampus. J. Physiol., 511: 761–770.
derlies DHPG-induced LTD. Neuropharmacology, 43: Palmer, M.J., Irving, A.J., Seabrook, G.R., Jane, D.E. and
175–180. Collingridge, G.L. (1997) The group I mGlu receptor agonist
Mulkey, R.M., Endo, S., Shenolikar, S. and Malenka, R.C. DHPG induces a novel form of LTD in the CA1 region of the
(1994) Involvement of a calcineurin/inhibitor-1 phosphatase hippocampus. Neuropharmacology, 36: 1517–1532.
cascade in hippocampal long-term depression. Nature, 369: Paolillo, M., Montecucco, A., Zanassi, P. and Schinelli, S.
486–488. (1998) Potentiation of dopamine-induced cAMP formation
Mulkey, R.M., Herron, C.E. and Malenka, R.C. (1993) An by group I metabotropic glutamate receptors via protein
essential role for protein phosphatases in hippocampal long- kinase C in cultured striatal neurons. Eur. J. Neurosci., 10:
term depression. Science, 261: 1051–1055. 1937–1945.
Mulkey, R.M. and Malenka, R.C. (1992) Mechanisms under- Parsley, S.L., Pilgram, S.M., Soto, F., Giese, K.P. and
lying induction of homosynaptic long-term depression in area Edwards, F.A. (2007) Enriching the environment of alpha-
CA1 of the hippocampus. Neuron, 9: 967–975. CaMKIIT286A mutant mice reveals that LTD occurs in
Murray, H.J. and O’Connor, J.J. (2003) A role for COX-2 and memory processing but must be subsequently reversed by
p38 mitogen activated protein kinase in long-term depression LTP. Learn Mem., 14: 75–83.
in the rat dentate gyrus in vitro. Neuropharmacology, 44: Piomelli, D., Wang, J.K., Sihra, T.S., Nairn, A.C., Czernik,
374–380. A.J. and Greengard, P. (1989) Inhibition of Ca2+/calmod-
Murray, H.J. and O’Connor, J.J. (2004) A role for monomeric ulin-dependent protein kinase 11 by arachidonic acid and its
G-proteins in synaptic plasticity in the rat dentate gyrus in metabolites. Proc. Natl. Acad. Sci. U.S.A., 86: 8550–8554.
vitro. Brain Res., 1000: 85–91. Pöschel, B. and Manahan-Vaughan, D. (2005) Group II
Naie, K. and Manahan-Vaughan, D. (2005) Investigations of mGluR-induced long term depression in the dentate gyrus
the protein synthesis dependency of mGluR-induced long- in vivo is NMDA receptor-independent and does not require
term depression in the dentate gyrus of freely moving rats. protein synthesis. Neuropharmacology, 49: 1–12.
Neuropharmacology, 49: 35–44. Pöschel, B. and Manahan-Vaughan, D. (2007) Persistent
Nicholls, R.E., Zhang, X.L., Bailey, C.P., Conklin, B.R., Kan- (424 h) long-term depression in the dentate gyrus of freely
del, E.R. and Stanton, P.K. (2006) mGluR2 acts through moving rats is not dependent on activation of NMDA re-
inhibitory Galpha subunits to regulate transmission and ceptors, L-type voltage-gated calcium channels or protein
long-term plasticity at hippocampal mossy fiber-CA3 synthesis. Neuropharmacology, 52: 46–54.
synapses. Proc. Natl. Acad. Sci. U.S.A., 103: 6380–6385. Pöschel, B., Wroblewska, B., Heinemann, U. and Manahan-
Nicoll, R.A., Oliet, S.H. and Malenka, R.C. (1998) NMDA Vaughan, D. (2005) The metabotropic glutamate receptor
receptor-dependent and metabotropic glutamate receptor- mGluR3 is critically required for hippocampal long-term de-
dependent forms of long-term depression coexist in CA1 pression and modulates long-term potentiation in the dentate
hippocampal pyramidal cells. Neurobiol. Learn. Mem., 70: gyrus of freely moving rats. Cereb. Cortex, 15: 1414–1423.
62–72. Qi, M., Zhuo, M., Skalhegg, B.S., Brandon, E.P., Kandel, E.R.,
Normandin, M., Gagne, J., Bernard, J., Lie, R., Miceli, D., McKnight, G.S. and Idzerda, R.L. (1996) Impaired hippo-
Baudry, M. and Massicotte, G. (1996) Involvement of the 12- campal plasticity in mice lacking the Cbeta1 catalytic subunit
lipoxygenase pathway of arachidonic acid metabolism in ho- of cAMP-dependent protein kinase. Proc. Natl. Acad. Sci.
mosynaptic long term depression of the rat hippocampus. U.S.A., 93: 1571–1576.
Brain Res., 730: 40–46. Ramakers, G.M., Heinen, K., Gispen, W.H. and de Graan,
Nosyreva, E.D. and Huber, K.M. (2005) Developmental switch P.N. (2000) Long term depression in the CA1 field is as-
in synaptic mechanisms of hippocampal metabotropic gluta- sociated with a transient decrease in pre- and postsynaptic
mate receptor-dependent long-term depression. J. Neurosci., PKC substrate phosphorylation. J. Biol. Chem., 275:
25: 2992–3001. 28682–28687.
499

Rammes, G., Palmer, M., Eder, M., Dodt, H.U., Zieglgans- Seidenman, K.J., Steinberg, J.P., Huganir, R. and Malinow,
berger, W. and Collingridge, G.L. (2003) Activation of mGlu R. (2003) Glutamate receptor subunit 2 Serine 880 phos-
receptors induces LTD without affecting postsynaptic sensi- phorylation modulates synaptic transmission and mediates
tivity of CA1 neurons in rat hippocampal slices. J. Physiol., plasticity in CA1 pyramidal cells. J. Neurosci., 23:
546: 455–460. 9220–9228.
Reyes, M. and Stanton, P.K. (1996) Induction of hippocampal Shi, J., Townsend, M. and Constantine-Paton, M. (2000) Ac-
long-term depression requires release of Ca2+ from separate tivity-dependent induction of tonic calcineurin activity me-
presynaptic and postsynaptic intracellular stores. J. Neuro- diates a rapid developmental downregulation of NMDA
sci., 16: 5951–5960. receptor currents. Neuron, 28: 103–114.
Reyes-Harde, M., Empson, R., Potter, B.V., Galione, A. and Shields, S.M., Ingebritsen, T.S. and Kelly, P.T. (1985) Identi-
Stanton, P.K. (1999a) Evidence of a role for cyclic ADP- fication of protein phosphatase 1 in synaptic junctions: de-
ribose in long-term synaptic depression in hippocampus. phosphorylation of endogenous calmodulin-dependent
Proc. Natl. Acad. Sci. U.S.A., 96: 4061–4066. kinase II and synapse-enriched phosphoproteins. J. Neuro-
Reyes-Harde, M., Potter, B.V., Galione, A. and Stanton, P.K. sci., 5: 3414–3422.
(1999b) Induction of hippocampal LTD requires nitric-oxide- Snyder, E.M., Philpot, B.D., Huber, K.M., Dong, X., Fallon,
stimulated PKG activity and Ca2+ release from cyclic ADP- J.R. and Bear, M.F. (2001) Internalization of ionotropic
ribose-sensitive stores. J. Neurophysiol., 82: 1569–1576. glutamate receptors in response to mGluR activation. Nat.
Roberson, E.D., English, J.D., Adams, J.P., Selcher, J.C., Ko- Neurosci., 4: 1079–1085.
ndratick, C. and Sweatt, J.D. (1999) The mitogen-activated Son, H., Hawkins, R.D., Martin, K., Kiebler, M., Huang, P.L.,
protein kinase cascade couples PKA and PKC to cAMP re- Fishman, M.C. and Kandel, E.R. (1996) Long-term potent-
sponse element binding protein phosphorylation in area CA1 iation is reduced in mice that are doubly mutant in end-
of hippocampus. J. Neurosci., 19: 4337–4348. othelial and neuronal nitric oxide synthase. Cell, 87:
Rouach, N. and Nicoll, R.A. (2003) Endocannabinoids con- 1015–1023.
tribute to short-term but not long-term mGluR-induced de- Stanton, P.K. (1995) Transient protein kinase C activation
pression in the hippocampus. Eur. J. Neurosci., 18: primes long-term depression and suppresses long-term po-
1017–1020. tentiation of synaptic transmission in hippocampus. Proc.
Rush, A.M., Wu, J., Rowan, M.J. and Anwy, R. (2002) Group Natl. Acad. Sci. U.S.A., 92: 1724–1728.
I metabotropic glutamate receptor (mGluR)-dependent long- Stanton, P.K. (1996) Phospholipase A2 activation is not re-
term depression mediated via p38 mitogen-activated protein quired for long-term synaptic depression. Eur. J. Pharmacol.,
kinase is inhibited by previous high-frequency stimulation 273: R7–R9.
and activation of mGluRs and protein kinase C in the rat Stanton, P.K. and Gage, A.T. (1996) Distinct synaptic loci of
dentate gyrus in vitro. J. Neurosci., 22: 6121–6128. Ca2+/calmodulin-dependent protein kinase II necessary for
Sajikumar, S. and Frey, J.U. (2003) Anisomycin inhibits the long-term potentiation and depression. J. Neurophysiol., 76:
late maintenance of long-term depression in rat hippocampal 2097–2101.
slices in vitro. Neurosci. Lett., 338: 147–150. Stanton, P.K., Heinemann, U. and Muller, W. (2001) FM1-43
Santschi, L., Reyes-Harde, M. and Stanton, P.K. (1999) Chem- imaging reveals cGMP-dependent long-term depression of
ically induced, activity-independent LTD elicited by simul- presynaptic transmitter release. J. Neurosci., 21: RC167.
taneous activation of PKG and inhibition of PKA. J. Stanton, P.K. and Sarvey, J.M. (1984) Blockade of long-term
Neurophysiol., 82: 1577–1589. potentiation in rat hippocampal CA1 region by inhibitors of
Santschi, L.A., Zhang, X.L. and Stanton, P.K. (2006) Activa- protein synthesis. J. Neurosci., 4: 3080–3088.
tion of receptors negatively coupled to adenylate cyclase is Stanton, P.K. and Sejnowski, T.J. (1989) Associative long-term
required for induction of long-term synaptic depression at depression in the hippocampus induced by hebbian covari-
Schaffer collateral-CA1 synapses. J. Neurobiol., 66: 205–219. ance. Nature, 339: 215–218.
Schnabel, R., Kilpatrick, I.C. and Collingridge, G.L. (1999a) Stanton, P.K., Winterer, J., Bailey, C.P., Kyrozis, A., Raginov,
An investigation into signal transduction mechanisms in- I., Laube, G., Veh, R.W., Nguyen, C.Q. and Muller, W.
volved in DHPG-induced LTD in the CA1 region of the (2003) Long-term depression of presynaptic release from the
hippocampus. Neuropharmacology, 38: 1585–1596. readily releasable vesicle pool induced by NMDA receptor-
Schnabel, R., Kilpatrick, I.C. and Collingridge, G.L. (2001) dependent retrograde nitric oxide. J. Neurosci., 23:
Protein phosphatase inhibitors facilitate DHPG-induced 5936–5944.
LTD in the CA1 region of the hippocampus. Br. J. Pharma- Stevens, C.F., Tonegawa, S. and Wang, Y. (1994) The role of
col., 132: 1095–1101. calcium-calmodulin kinase II in three forms of synaptic plas-
Schnabel, R., Palmer, M.J., Kilpatrick, I.C. and Collingridge, ticity. Curr. Biol., 4: 687–693.
G.L. (1999b) A CaMKII inhibitor, KN-62, facilitates Strack, S., Barban, M.A., Wadzinski, B.E. and Colbran, R.J.
DHPG-induced LTD in the CA1 region of the hippocam- (1997) Differential inactivation of postsynaptic density-asso-
pus. Neuropharmacology, 38: 605–608. ciated and soluble Ca2+/calmodulin-dependent protein kin-
Schwartz, J.H. (1993) Cognitive kinases. Proc. Natl. Acad. Sci. ase II by protein phosphatases 1 and 2A. J. Neurochem., 68:
U.S.A., 90: 8310–8313. 2119–2128.
500

Sugiyama, H., Ito, I. and Hirono, C. (1987) A new type of gyrus of the rat hippocampus in vitro. J. Physiol., 513:
glutamate receptor linked to inositol phospholipid metabo- 467–475.
lism. Nature, 325: 531–533. Watabe, A.M., Carlisle, H.J. and O’Dell, T.J. (2002) Postsy-
Tan, Y., Hori, N. and Carpenter, D.O. (2003) The mechanism naptic induction and presynaptic expression of group 1
of presynaptic long-term depression mediated by group I mGluR-dependent LTD in the hippocampal CA1 region. J.
metabotropic glutamate receptors. Cell Mol. Neurobiol., 23: Neurophysiol., 87: 1395–1403.
187–203. Wu, J., Rowan, M.J. and Anwyl, R. (2004) Synaptically stim-
Tanabe, Y., Masu, M., Ishii, T., Shigemoto, R. and Nakanishi, ulated induction of group I metabotropic glutamate receptor-
S. (1992) A family of metabotropic glutamate receptors. dependent long-term depression and depotentiation is inhib-
Neuron, 8: 169–179. ited by prior activation of metabotropic glutamate receptors
Thiels, E., Kanterewicz, B.I., Knapp, L.T., Barrionuevo, G. and protein kinase C. Neuroscience, 123: 507–514.
and Klann, E. (2000) Protein phosphatase-mediated regula- Wu, J., Rush, A., Rowan, M.J. and Anwyl, R. (2001) NMDA
tion of protein kinase C during long-term depression in the receptor- and metabotropic glutamate receptor-dependent
adult hippocampus in vivo. J. Neurosci., 20: 7199–7207. synaptic plasticity induced by high frequency stimulation in
Thiels, E., Kanterewicz, B.I., Norman, E.D., Trzaskos, J.M. the rat dentate gyrus in vitro. J. Physiol., 533: 745–755.
and Klann, E. (2002) Long-term depression in the adult hip- Wu, J., Wang, Y., Rowan, M.J. and Anwyl, R. (1997) Evidence
pocampus in vivo involves activation of extracellular signal- for involvement of the neuronal isoform of nitric oxide synt-
regulated kinase and phosphorylation of Elk-1. J. Neurosci., hase during induction of long-term potentiation and long-
22: 2054–2062. term depression in the rat dentate gyrus in vitro. Neurosci-
Thiels, E., Norman, E.D., Barrionuevo, C. and Klann, E. ence, 78: 393–398.
(1998) Transient and persistent increases in protein phos- Wu, J., Wang, Y., Rowan, M.J. and Anwyl, R. (1998) Evidence
phatase activity during long-term depression in the adult for involvement of the cGMP-protein kinase G signaling
hippocampus in vivo. Neuroscience, 86: 1023–1029. system in the induction of long-term depression, but not
Thomas, G.M. and Huganir, R.L. (2004) MAPK cascade sig- long-term potentiation, in the dentate gyrus in vitro. J.
nalling and synaptic plasticity. Nat. Rev. Neurosci., 5: Neurosci., 18: 3589–3596.
173–183. Xiao, M.Y., Zhou, Q. and Nicoll, R.A. (2001) Metabotropic
Trommer, B.L., Liu, Y.B. and Pasternak, J.F. (1996) Long- glutamate receptor activation causes a rapid redistribution of
term depression at the medial perforant path-granule cell AMPA receptors. Neuropharmacology, 41: 664–671.
synapse in developing rat dentate gyrus. Brain Res. Dev. Yang, S.N., Tang, Y.G. and Zucker, R.S. (1999) Selective in-
Brain Res., 96: 97–108. duction of LTP and LTD by postsynaptic [Ca2]i elevation. J.
Wang, Q., Chang, L., Rowan, M.J. and Anwyl, R. (2007) Neurophysiol., 81: 781–787.
Developmental dependence, the role of the kinases p38 Zakharenko, S.S., Zablow, L. and Siegelbaum, S.A. (2002)
MAPK and PKC, and the involvement of tumor necrosis Altered presynaptic vesicle release and cycling during
factor-R1 in the induction of mGlu-5 LTD in the dentate mGluR-dependent LTD. Neuron, 35: 1099–1110.
gyrus. Neuroscience, 144: 110–118. Zhang, X.L., Zhou, Z.Y., Winterer, J., Muller, W. and Stanton,
Wang, Y., Rowan, M.J. and Anwyl, R. (1997) Induction P.K. (2006) NMDA-dependent, but not group I metabo-
of LTD in the dentate gyrus in vitro is NMDA receptor tropic glutamate receptor-dependent, long-term depression at
independent, but dependent on Ca2+ influx via low-voltage- Schaffer collateral-CA1 synapses is associated with long-term
activated Ca2+ channels and release of Ca2+ from intracel- reduction of release from the rapidly recycling presynaptic
lular stores. J. Neurophysiol., 77: 812–825. vesicle pool. J. Neurosci., 26: 10270–10280.
Wang, Y., Wu, J., Rowan, M.J. and Anywl, R. (1998) Role of Zhou, Q., Homma, K.J. and Poo, M.M. (2004) Shrinkage
protein kinase C in the induction of homosynaptic long-term of dendritic spines associated with long-term depression of
depression by brief low frequency stimulation in the dentate hippocampal synapses. Neuron, 44: 749–757.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 27

Structural reorganization of the dentate gyrus


following entorhinal denervation: species differences
between rat and mouse

Thomas Deller, Domenico Del Turco, Angelika Rappert and Ingo Bechmann

Institute of Clinical Neuroanatomy, J.W. Goethe-University, Theodor-Stern-Kai 7, D-60590 Frankfurt/Main, Germany

Abstract: Deafferentation of the dentate gyrus by unilateral entorhinal cortex lesion or unilateral perforant
pathway transection is a classical model to study the response of the central nervous system (CNS) to
denervation. This model has been extensively characterized in the rat to clarify mechanisms underlying
denervation-induced gliosis, transneuronal degeneration of denervated neurons, and collateral sprouting of
surviving axons. As a result, candidate molecules have been identified which could regulate these changes,
but a causal link between these molecules and the postlesional changes has not yet been demonstrated. To
this end, mutant mice are currently studied by many groups. A tacit assumption is that data from the rat
can be generalized to the mouse, and fundamental species differences in hippocampal architecture and the
fiber systems involved in sprouting are often ignored. In this review, we will (1) provide an overview of some
of the basics and technical aspects of the entorhinal denervation model, (2) identify anatomical species
differences between rats and mice and will point out their relevance for the axonal reorganization process,
(3) describe glial and local inflammatory changes, (4) consider transneuronal changes of denervated dentate
neurons and the potential role of reactive glia in this context, and (5) summarize the differences in
the reorganization of the dentate gyrus between the two species. Finally, we will discuss the use of the
entorhinal denervation model in mutant mice.

Keywords: entorhinal cortex lesion; perforant pathway transection; sprouting; transneuronal degeneration;
glia; inflammation; regeneration

Introduction removal of the cellular debris, surviving non-


lesioned axons sprout and reinnervate the dener-
Injuries to the central nervous system (CNS) cause vated target cells, and denervated neurons remodel
local damage at the lesion site as well as denerva- their dendritic arbor in response to denervation
tion of connected brain regions. In response to this and reinnervation. In the past, much effort has
denervation, complex cellular changes occur in the been devoted to understanding these changes at
denervated brain area (see Steward, 1994b, for re- the molecular level. The rationale for this is that
view): Glial cells react and participate in the the denervation of brain areas is a very common
and detrimental side effect of all kinds of brain
Corresponding author. Tel.: +49-69-6301-6361; lesions (Steward, 1994b). It can occur following
Fax: +49-69-6301-6425; E-mail: T.Deller@em.uni-frankfurt.de traumatic brain damage, ischemia, inflammation,

DOI: 10.1016/S0079-6123(07)63027-1 501


502

and neurodegeneration and may contribute sig- often been interpreted solely on the basis of data
nificantly to the severity of the clinical symptoms obtained in rats. It is assumed that the reorgan-
of these and other brain diseases. It is one of the ization of the mouse dentate gyrus is more or less
motivations of the scientists working in the field similar to that of the rat dentate gyrus following
today that understanding the reorganization proc- entorhinal deafferentation. However, this assump-
esses in denervated brain areas and learning how tion may not be true, given the important biolog-
to influence them in a beneficial fashion may lead ical differences between these two species and even
to new treatment strategies for brain diseases. between different mouse strains.
One of the classical model systems used to study Therefore, it may be prudent to keep data from
the reorganization of denervated brain regions is the two species separate. In this review, we will
the reorganization of the rat dentate gyrus follow- compare the reorganization of the dentate gyrus
ing entorhinal denervation (Lynch and Cotman, following entorhinal denervation in rats and mice.
1975; Cotman and Nadler, 1978; Gall and Lynch, We will first give an overview over some of the
1980; Gall et al., 1986; Steward, 1991; Deller and fundamentals and technical aspects of the model
Frotscher, 1997; Frotscher et al., 1997; Collazos- system. Second, we will identify species differences
Castro and Nieto-Sampedro, 2001; Deller et al., between the non-lesioned dentate gyrus of rats and
2001; Ramirez, 2001; Savaskan and Nitsch, 2001). mice and will point out their relevance for the ax-
This model is well established, and structural, onal reorganization process. Third, glial changes
functional, and molecular changes have been stud- occurring in the dentate gyrus of both species after
ied in detail over the years. A fairly large number denervation will be described and considered as a
of candidate molecules have been identified that local inflammatory response of the brain. Fourth,
could regulate the reorganization of the dentate transneuronal changes of denervated dentate neu-
gyrus, including growth-associated molecules, rons will be considered, and the potential role of
neurotrophic factors, cell adhesion molecules, reactive glia in this context will be reviewed. Fifth,
extracellular matrix molecules, and a variety differences in the reorganization of the dentate
of cytokines (see Deller and Frotscher, 1997; gyrus between the two species will be summarized
Collazos-Castro and Nieto-Sampedro, 2001; and, finally, the use of the entorhinal lesion model
Deller et al., 2001; Savaskan and Nitsch, in mutant mice will be discussed.
2001, for review). In the majority of studies,
however, correlations between molecular changes
and the reorganization process were reported, so Entorhinal denervation in rats and mice: technical
it remains to be seen that these candidate mole- aspects
cules are causally linked to the postlesional re-
organization process. An attractive experimental The entorhinal denervation model was established
strategy to address these issues of causality is to in the 1970s, when it was first shown that unilat-
use genetically engineered mice and to analyze eral removal of entorhinal afferents to the rat
the reorganization of the dentate gyrus following dentate gyrus results in degenerative and regener-
entorhinal denervation in these mutants (Fagan ative changes in the so-called ‘‘entorhinal’’ zone of
et al., 1998; Finsen et al., 1999; Jensen et al., the molecular layer, i.e., the outer two-thirds of the
1999; Jensen et al., 2000b; Del Turco et al., 2001, molecular layer (Lynch et al., 1972; Cotman and
2003; White et al., 2001a, b; Kadish and van Nadler, 1978). This model is considered particu-
Groen, 2003; Rappert et al., 2004). larly useful, because of the relatively simple and
Although the entorhinal denervation model has highly laminated cyto- and fiber architecture of the
now frequently been used in mice, the reorganiza- dentate gyrus, which makes it possible to distin-
tion of the denervated dentate gyrus has only been guish between denervated and non-denervated lay-
partially characterized in this species. In fact, in ers and enables researchers to differentiate
the absence of data from the mouse, results ob- between the various cellular changes that occur
tained after entorhinal denervation in mice have within a denervated layer or in its vicinity (Cotman
503

and Nadler, 1978; Steward, 1991; Deller et al., preferred method to de-entorhinate the mouse
2001). dentate gyrus (Fig. 1; Finsen et al., 1999; Jensen
A confusing element of the entorhinal denerva- et al., 1999, 2000b; Drojdahl et al., 2002, 2004;
tion model in rats is the fact that the method to Ying et al., 2002; Del Turco et al., 2003; Kovac
produce the lesion has changed over the years. In et al., 2004; Rappert et al., 2004; Nielsen et al.,
the beginning, entorhinal cortex lesions (ECL) 2006; Pedersen et al., 2006). Chemical lesions of
were performed using electrolytic damage to the the entorhinal cortex itself were less frequently
entorhinal cortex itself, thereby removing the cells performed (Kadish and van Groen, 2002, 2003;
of origin of the perforant pathway and denervat- Kadish et al., 2002). Thus, the lesioning techniques
ing the dentate gyrus (Lynch et al., 1972; Steward differ between rats and mice; in mice the mechan-
et al., 1974; Zimmer and Hjorth-Simonsen, 1975; ical perforant pathway transection is the most
Rose et al., 1976; Gall et al., 1979a, b; Poirier common, whereas electrolytic lesions are more
et al., 1990; Parent et al., 1993; Guthrie et al., 1995, typical in rats.
1997; Jucker et al., 1996). Later, other methods Finally, the unilateral ECL/perforant pathway
were also used, for example, mechanical or elec- transection model should also be distinguished from
trolytic transection of the perforant pathway several related models that have also been employed
(Benowitz et al., 1990; Nitsch and Frotscher, to study reactive changes in the denervated dentate
1992; Nitsch et al., 1992; Beck et al., 1993; Gwag gyrus in the past. Among these models are bilateral
et al., 1994; Deller et al., 1995a) or chemical lesions ECL (Lynch et al., 1976; Hardman et al., 1997;
(Ueki et al., 1996). Often, these lesioning tech- Faillace et al., 2006), a combination of ECL
niques are regarded as more or less equivalent, with fimbria-fornix-transection (Beck et al., 1993;
since all of them result in a robust denervation of Peterson, 1994), and fimbria-fornix-transection with-
the dentate gyrus. However, they may differ in out ECL (Hjorth-Simonsen and Jeune, 1972; Lynch
their acute and chronic effects on the dentate et al., 1974). All of these models remove additional or
gyrus: whereas electrolytic lesions strongly stimu- neurochemically different projection systems to the
late the perforant pathway and may have an dentate gyrus, resulting in an altered reorganization
epileptogenic effect, non-electrolytic lesions are process compared to unilateral ECL/perforant path-
less likely to do so (Dasheiff and McNamara, way transection.
1982; Campbell et al., 1984; Kelley and Steward,
1996a). Since a variety of molecules and genes are
regulated by activity, gene expression — at least The non-denervated dentate gyrus: neuroanatomical
during the initial phase of the reorganization proc- species differences
ess — could be influenced by the lesioning tech-
nique (Kelley and Steward, 1996b, 1998; Steward The normal anatomy of the dentate gyrus is the basis
et al., 1997). Moreover, as the extent of necrotic for a meaningful interpretation of experimentally
and apoptotic degeneration certainly varies be- induced changes in this brain region. In the rat, the
tween the different methods, it is likely that there neuroanatomy of the dentate gyrus has been thor-
are also fundamental differences with regard to the oughly investigated during the last 50 years. The
induced inflammatory and immune responses. In cellular architecture of neuronal and non-neuronal
addition to such differences relating to the lesion- cells, the connectivity, and the basic neurochemistry
ing techniques, differences in rat strains were of the dentate gyrus were studied using anatomical
rarely considered. In summary, data from the rat techniques for cellular labeling, retro- and antero-
ECL/perforant pathway transection model are grade tracing, and immunohistochemistry at the
plentiful, yet some caution is warranted when re- light- and electron-microscopic level. These neuro-
sults from two labs using different lesioning tech- anatomical facts have been reviewed in numerous
niques or different rat strains are compared. papers and book chapters (e.g., Frotscher, 1988;
In mice, mechanical transection of the perforant Amaral and Witter, 1995) as well as in other chap-
pathway using wire and plate knifes has been the ters of this book. In contrast, the mouse dentate
504

Fig. 1. The mouse dentate gyrus after entorhinal denervation. (A, B) Fink-Heimer degeneration stain in a control mouse hippocampus
(A) and 5 days after entorhinal denervation (B). Note darkly stained degeneration product in the denervated zone of the molecular
layer (B, arrowheads). (C, D) Fluoro-Jade C degeneration stain of the contralateral (C) and ipsilateral (D) mouse dentate gyrus four
days after denervation. Note increased reactivity throughout the denervated portion of the molecular layer (D). (E, F) Acetylcho-
linesterase (AChE) staining of the control mouse dentate gyrus (E) and 6 weeks after entorhinal denervation (F). A dense AChE-
positive fiber band is present in the outer molecular layer (OML) and middle molecular layer (MML; arrowheads). EC, entorhinal
cortex; CA1, CA3, Hippocampal subfields; GCL, granule cell layer; IML, inner molecular layer; ML; molecular layer; H, hilus. Scale
bars: A, B: 300 mm; C–F: 100 mm.
505

gyrus has not been as thoroughly studied. Although associational (C/A) projection to the inner molecu-
data exist on its general cellular and fiber architec- lar layer. Although this projection is formed by
ture (e.g., Stanfield and Cowan, 1979a, b; Stanfield mossy cells in both species, these cells contain
et al., 1979; Soriano et al., 1994; Supèr and Soriano, calretinin in mice but not in rats (Liu et al., 1996;
1994; Deller et al., 1999a, b, 2002; Drakew et al., Blasco-Ibanez and Freund, 1997; Fujise et al., 1998).
2002; Gebhardt et al., 2002; Del Turco et al., 2004), In addition, the C/A projection, in mice covers only
data on the connectivity of identified cells are scarce. one-fifth of the molecular layer in contrast to one-
Only a handful of tracing studies were performed in third in rats, is more diffusely organized and
mice, in which tracing was combined with electron the border of the C/A fiber plexus towards the
microscopy to demonstrate synaptic connections of entorhinal zone of the dentate gyrus is indistinct
identified neurons (Blasco-Ibanez and Freund, 1997; (Figs. 2a, b), suggesting that the border between the
Deller et al., 1999b; Del Turco et al., 2003; Otal two layers is not as strict as in the rat (Del Turco
et al., 2006). Similarly, the topographical organiza- et al., 2003). A third and striking anatomical
tion of the afferent projection systems, the cells of difference relates to the crossed entorhinodentate
origin of the various pathways, and the neurochem- projection. This projection system seems to be non-
istry of the target cells have not been studied in de- existent in mice (van Groen et al., 2002, 2003),
tail. Strain differences have rarely been considered. whereas it is readily detectable in rats (Steward
Thus, neuroanatomical data on the non-denervated et al., 1974; Steward, 1976a; Deller et al., 1996a).
mouse dentate gyrus are often incomplete and lack Thus, the entorhinodentate pathway and at least
the depth and precision of the rat data. two other fiber systems are different in rats and
Nevertheless, on the basis of the available neuro- mice, which could influence the pattern of axonal
anatomical data several anatomical differences are reorganization in the denervated dentate gyrus.
evident, which are likely to result in significant
differences in the reorganization of the dentate gyrus
following entorhinal denervation. One relevant Reorganization of axons and fiber systems in the
difference concerns the ipsilateral entorhinal projec- denervated dentate gyrus
tion to the dentate gyrus, i. e., the projection system
that is lost as a consequence of ECL/perforant The reorganization and collateral sprouting of
pathway transection. This projection has been stud- non-lesioned axons in the denervated dentate
ied in rats (e.g. Blackstad, 1958; Amaral and Witter, gyrus is probably the most prominent and best-
1989; Deller, 1998) and mice (Deller et al., 1999a; known aspect of the entorhinal denervation model
van Groen et al., 2002, 2003) using anterograde in rats. It was also the first to be described (Lynch
tracing and almost identical patterns of termination et al., 1972). When discussing this aspect of the
were observed: In both species, the majority of entorhinal denervation model, we should distin-
entorhinodentate fibers terminates in a laminar guish between changes at the level of overall
fashion in the outer portion of the molecular number of synapses and terminals, and changes
layers of the dentate gyrus and even individual observed at the level of identified fiber systems.
axons have a very similar morphology. However,
the portion of the molecular layer covered by ento-
rhinal fibers appears to be greater in the mouse. In Changes in synapse and bouton densities following
this species, entorhinal fibers cover approximately entorhinal denervation in rat and mouse
four-fifth of the molecular layer (Deller et al., 1999a;
van Groen et al., 2002, 2003) compared to two- In rats, degenerative and regenerative changes of
thirds in rats (Blackstad, 1958; Amaral and Witter, synapses and terminals have been studied in detail.
1989; Deller, 1998). Thus, entorhinal denervation Using electron microscopy, synapse counts follow-
in mice will result in a more extensive denervation ing ECL show 80–90% of all synapses are lost in the
of granule cell dendrites than in rats. A second denervated zone of the dentate gyrus, followed by
relevant difference concerns the commissural and reactive changes that replace up to 60–80% of all
506
507

lost synapses by 4 weeks postlesion (Matthews et al., revealed an increase in staining in the denervated
1976a, b; Steward and Vinsant, 1983). In a recent dentate gyrus compared to the contralateral side by
stereological study, these data were largely con- 4 weeks survival time (van Groen et al., 2002;
firmed, although synapse recovery observed with the Kadish and van Groen, 2002, 2003).
more sensitive stereological tools was somewhat less
than reported before (Marrone et al., 2004b). To
visualize total changes in presynaptic terminals, Sprouting of identified fiber systems following
immunohistochemical markers for presynaptic bou- entorhinal denervation
tons such as synaptophysin have also been em-
ployed in rats (Jucker et al., 1996; Miwa and Ueki, The recovery of synapse densities following ento-
1996; Stone et al., 1998; Kadish and van Groen, rhinal denervation in rats and mice raises the
2003). Although changes in synaptophysin-positive question which fibers contribute to this reinnerva-
boutons following entorhinal denervation seem to tion and whether this process is, in fact, restorative
correlate with the quantitative electron microscopic in nature. As axons from the ipsilateral entorhinal
data, a systematic comparison of the two techniques cortex are completely removed from the dentate
has not yet been performed, and it remains to be gyrus and these fibers cannot regenerate in the
seen whether or not changes in synaptophysin- adult brain, reinnervation apparently occurs from
positive puncta accurately reflect changes at the non-lesioned axons. The new connections thus
ultrastructural level. differ anatomically as well as functionally from the
For the denervated mouse dentate gyrus, virtu- ones removed by entorhinal denervation.
ally no quantitative electron-microscopic data are In rats, the sprouting fiber systems were iden-
available. Thus, the extent of both synapse loss and tified using histochemical as well as retrograde
reactive synaptogenesis are unknown. Entorhinal and anterograde tracing methods. Four fiber sys-
denervation has, however, been shown qualitatively tems, which are already present in the non-
in mice using silver impregnation of degenerating esioned dentate gyrus, are believed to contribute
fibers and terminals (Fig. 1; Steward, 1992; Del to the reinnervation process (see Cotman
Turco et al., 2003). In addition, stereological esti- and Nadler, 1978; Steward, 1994b; Deller and
mates of the total volume of the dentate gyrus fol- Frotscher, 1997, for review): (i) glutamatergic
lowing entorhinal denervation indicate that the C/A fibers to the inner molecular layer (mossy cell
dentate gyrus shrinks to 60% of its original size, axons), (ii) GABAergic C/A fibers to the outer
suggesting that considerable degenerative changes molecular layer (iii) glutamatergic cholinergic
occur following entorhinal denervation (Phinney septohippocampal fibers, and (iv) crossed ento-
et al., 2004). Similarly, entorhinal reinnervation has rhinodentate fibers. In the following paragraphs,
to some extent been demonstrated qualitatively. we will consider the contribution of each of these
Synaptophysin-immunostaining was used to dem- fiber systems to the reinnervation process and will
onstrate changes in presynaptic terminals and point out potentially relevant species differences.

Fig. 2. Sprouting of commissural/associational fibers in the mouse dentate gyrus following entorhinal denervation. (A, B) Dentate
gyrus of a control mouse shown at low (A) and high magnification (B). The commissural projections to the inner molecular layer (IML)
and the middle and outer molecular layer (OML; arrow in B) were labeled using Phaseolus vulgaris-leucoagglutinin (PHAL)-tracing.
(C, D) Dentate gyrus of a mouse 2 months after entorhinal denervation shown at low (C) and high (D) magnification. Compared with
the control, the commissural fiber plexus has increased in width and PHAL-labeled commissural fibers are regularly observed in the
MML (arrows). (E, F) Dentate gyrus of a control mouse stained for calretinin shown at low (E) and high (F) magnification. Calretinin-
immunostaining labels mossy cells in the hilus (H) and interneurons in the molecular layer. Calretinin-immunoreactive mossy cell
axons are abundant in the IML. (G, H) Dentate gyrus of a mouse 2 months following entorhinal denervation shown at low (G) and
high (H) magnification. Note that compared to the control, the density of calretinin-immunoreactive fibers has increased throughout
the molecular layer and that the calretinin-positive fiber plexus in the IML has expanded in width (H). CA3, hippocampal subfield;
GCL, granule cell layer; H, hilus; OML, outer molecular layer; IML, inner molecular layer; MML, middle molecular layer. Scale bars:
A, C, E, G: 100 mm; B, D, F, H: 25 mm.
508

Reorganization of commissural/associational mossy cells in the hilus (Blasco-Ibanez and Freund,


cell axons after entorhinal denervation 1997). In mice, these cells contain the calcium-
binding protein calretinin (Figs. 2e, f), which
As mentioned above, the C/A projection to the facilitates the identification of their axons in
inner molecular layer of the dentate gyrus in the the inner molecular layer (Liu et al., 1996;
rat arises from glutamate-immunoreactive mossy Blasco-Ibanez and Freund, 1997; Fujise et al.,
cells (Soriano and Frotscher, 1994). These neurons 1998; Deller et al., 1999a; Del Turco et al., 2003).
are located in the hilus and project both contra- The contribution of this projection to the reor-
laterally (commissural fibers) and ipsilaterally ganization of the mouse dentate gyrus following
(associational fibers). Their axons terminate in entorhinal denervation was first studied by Shi and
a characteristic laminar pattern in the inner one- Stanfield (1996), who used horseradish peroxidase
third of the molecular layer, i.e., complementary to tracing and observed an ingrowth of C/A fibers
the entorhinodentate fiber projection to the outer into the denervated molecular layer (Shi and Stan-
two-thirds of the molecular layer (Amaral and field, 1996). We confirmed and extended these data
Witter, 1995). The contribution of this projection using anterograde Phaseolus vulgaris-leucoagglutinin
to the reorganization of the rat dentate gyrus is (PHAL)-tracing in combination with immuno-
age-dependent (Gall and Lynch, 1978, 1980, 1981; staining for calretinin at the light as well as the
Stäubli et al., 1984; Gall et al., 1986): In young electron-microscopic level (Del Turco et al., 2003).
rats, injured within the first two-postnatal weeks, In this study we showed that, in contrast to the
extensive sprouting of C/A axon collaterals into adult rat, many mossy cell axons leave the main
the deafferented adjacent outer molecular layer C/A fiber plexus in the inner molecular layer in
can be observed and some of these fibers extend response to entorhinal denervation, invade the
all the way to the hippocampal fissure. The degree denervated adjacent zone (Figs. 2c, d, g, h), and
of this reactive growth response diminishes with form new synapses outside their normal terminal
maturation, and is almost completely lost in the zone. This pattern is striking in its distinction from
adult rat brain (Gall and Lynch, 1981; Deller et al., the layer-specific sprouting in adult rats and is
1996c). In addition, a 30–40 mm outward expan- rather reminiscent of the sprouting seen in imma-
sion of the entire C/A fiber plexus was also ture rats (Gall and Lynch, 1981).
observed (Cotman and Nadler, 1978; Gall and In addition, an expansion of the entire C/A
Lynch, 1981). Although the nature of this termination zone was also observed in mice
phenomenon was a matter of debate for many (Del Turco et al., 2003; Phinney et al., 2004).
years (see Deller and Frotscher, 1997, for review), Phinney et al. (2004) studied this expansion using
available data support the interpretation that the three-dimensional stereological tools and com-
expansion of the C/A plexus is not the result pared the expansion of the inner molecular layer
of translaminar sprouting but rather caused by in single sections with changes of the total volume
reorganization and shrinkage processes occurring of the inner molecular layer following denervation.
within the non-denervated inner molecular layer Although this study confirmed a significant
itself (Hoff et al., 1981; Caceres and Steward, 1983; increase in the width of the C/A termination zone
Frotscher et al., 1997; Marrone et al., 2004a, using single sections, it did not detect an increase
b; Deller et al., 2006). We conclude, therefore, that in its total volume. In fact, shrinkage of the dent-
(1) C/A mossy cell axons do not leave their ate gyrus in all three dimensions, especially the
home territory to invade the denervated zone compression of the dentate gyrus along the septo-
in adult rats, and (2) axonal remodeling occurs temporal axis, seemed to underlie the expansion.
not only in the denervated zone but also within The authors concluded that the expansion of the
the adjacent non-denervated inner molecular inner molecular layer does not indicate an increase
layer. in the size of this lamina, and therefore should not
In adult C57BL/6 mice, the C/A projection to be used as an indicator of C/A sprouting in
the inner molecular layer also arises from mossy C57BL/6 mice.
509

Reorganization of GABAergic commissural/ Following entorhinal denervation in rats,


associational axons after entorhinal denervation changes in the density of AChE-fibers in the den-
ervated zone were observed and interpreted as a
The GABAergic C/A projection in rats that ter- sign of cholinergic sprouting (Lynch et al., 1972;
minates in the outer molecular layer of the dentate Nadler et al., 1977a, b; Zimmer et al., 1986). These
gyrus (Amaral and Witter, 1989; Deller et al., observations were later corroborated by antero-
1995a, b, 1996b; Zappone and Sloviter, 2001) also grade tracing studies of septohippocampal fibers
exists in mice (Deller et al., 1999a; Gebhardt et al., (Nadler and Evenson, 1982; Stanfield and Cowan,
2002; Del Turco et al., 2004). 1982; Nyakas et al., 1988). In spite of these data,
In rats, the reaction of this projection was studied however, doubts remained. It was pointed out that
at the single-fiber level using anterograde tracing, changes in AChE-staining could also be the result
electron microscopy, and GABA-postembedding of denervation-induced changes in gene expression
(Deller et al., 1995a). This study demonstrated that (Chen et al., 1983; McKeon et al., 1989) or tissue
GABAergic commissural axons sprout new collat- shrinkage (Storm-Mathisen, 1974). In addition,
erals within the denervated zone and form synapses biochemical studies did not reveal any evidence for
there. We concluded from this observation that cholinergic sprouting (Aubert et al., 1994), and a
(1) GABAergic fibers can sprout and reinnervate the non-stereological quantitative analysis of choline
dentate gyrus in response to the loss of glutamatergic acetyltransferase (ChAT)-positive fibers did not
afferents, and (2) that C/A fiber sprouting in the confirm the AChE-data (Henderson et al., 1998).
adult rat is layer specific. Thus, both the C/A fibers Thus, it was questioned whether AChE-histo-
to the outer molecular layer as well as C/A fiber to chemistry is an appropriate technique to visualize
the inner molecular layer sprout within their home cholinergic sprouting in the denervated rat dentate
territories in response to entorhinal denervation gyrus (Aubert et al., 1994) and whether cholinergic
in adult rats (Deller et al., 1995a, 1996c; Deller and sprouting occurs at all (Storm-Mathisen, 1974).
Frotscher, 1997; Frotscher et al., 1997). In mice, The first of these issues could be resolved by using
a systematic analysis of the GABAergic C/A pro- the immunotoxin 192 IgG-saporin, which selec-
jection to the molecular layer following entorhinal tively destroys cholinergic neurons in the basal
denervation has not yet been reported. forebrain (Wiley et al., 1991; Book et al., 1992;
Heckers et al., 1994). Following entorhinal dener-
Reorganization of acetylcholinesterase (AChE): vation, rats were treated with 192 IgG-saporin and
positive fibers after entorhinal denervation the AChE-positive fiber band could be completely
abolished (Naumann et al., 1997). Thus, AChE-
A third projection system contributing to the reor- histochemistry is appropriate to visualize septo-
ganization of the dentate gyrus following entorhinal hippocampal fibers selectively in the denervated
denervation is the cholinergic septohippocampal dentate gyrus. However, the second of these issues
projection. In non-lesioned rats, this projection arises yet awaits clarification. Since it will be essential to
from cholinergic neurons located in the diagonal rule out shrinkage as a cause, cholinergic sprout-
band of Broca (Nyakas et al., 1987; Jakab and ing should be reinvestigated in rats using assump-
Leranth, 1995). In contrast to the C/A and ent- tion-free stereological methodology.
orhinal fibers, this projection system terminates in all In non-lesioned mice, the septohippocampal pro-
layers of the rat and mouse dentate gyrus (Amaral jection appears to be generally similar to the one
and Witter, 1995). To visualize the cholinergic septo- observed in rats (Linke et al., 1994). Similarly, in
hippocampal fibers in the rat dentate gyrus, acetyl- denervated mice, an increase in the density of
cholinesterase (AChE) histochemistry has been AChE-positive fibers is observed in the denervated
widely used, since it has been shown to be suffi- outer molecular layer (Steward, 1992, 1994a; Shi
ciently specific in this brain area (Eckenstein and and Stanfield, 1996; Kadish and van Groen, 2002;
Sofroniew, 1983; Levey et al., 1983, 1984; Naumann Phinney et al., 2004). Analogous to the situation in
et al., 1997). rats, this increase in AChE-fiber density (Figs. 1e, f)
510

was interpreted as a sign of cholinergic sprouting 1976a; Cotman et al., 1977). Morphological anal-
(Steward, 1992, 1994a; Shi and Stanfield, 1996; ysis demonstrated robust sprouting of the crossed
Kadish and van Groen, 2002). To verify this finding, entorhinodentate projection: It increases its fiber
and to account for tissue shrinkage that occurs in and terminal density by sixfold and of its synapse
the denervated dentate gyrus, a three-dimensional density by approximately 100-fold. The reorgani-
stereological analysis of total AChE-fiber length was zation of single fibers was also revealed using
performed (Phinney et al., 2004). Although this anterograde PHAL-tracing (Deller et al., 1996a).
study confirmed a significant increase in the density Whereas crossed entorhinodentate axons showed
of AChE-fibers in single sections, an increase in the only few branches and formed almost exclusively
total length of AChE fibers could not be detected. en passant boutons in non-lesioned animals,
Stereological measurements of the volume of the crossed entorhinodentate fibers displayed high
non-lesioned and denervated dentate gyrus revealed densities of small axonal extensions and tangle-
that shrinkage of the dentate gyrus in all three di- like formations resembling axonal and synaptic
mensions, rather than growth of fibers was respon- clusters in denervated rats. In keeping with these
sible for the increase in AChE-fiber density. The morphological data, electrophysiological studies
authors concluded that the increase in AChE fiber showed that the functional characteristics of the
density in the denervated outer molecular layer crossed entorhinodentate projection become sim-
should not be used as an indicator of lesion-induced ilar to the ipsilateral projection following dener-
cholinergic sprouting in mice. In summary, there is vation (Steward et al., 1973; White et al., 1976;
at present no direct evidence for sprouting of septo- Harris et al., 1978; Reeves and Steward, 1986,
hippocampal fibers following entorhinal denerva- 1988). Finally, behavioral experiments confirmed
tion in mice. that sprouting of crossed entorhinodentate fibers
can indeed ameliorate some, but not all, of the
behavioral deficits observed after unilateral ento-
Reorganization of crossed entorhinodentate fibers rhinal denervation (Loesche and Steward, 1977;
after entorhinal denervation Scheff and Cotman, 1977; Ramirez and Stein,
1984; Schenk and Morris, 1985). In summary,
A fourth projection system, contributing to the these data indicate that (1) crossed entorhinodent-
reorganization of the denervated dentate gyrus in ate sprouting contributes significantly to the
rats, is the crossed entorhinodentate (temporo- reinnervation of the rat dentate gyrus after ento-
dentate) pathway. Its cells of origin, trajectory, rhinal denervation, and (2) this sprouting of a
and its termination pattern have been well char- homologous fiber system is probably functionally
acterized using retrograde and anterograde tracing restorative.
techniques (Steward et al., 1974; Goldowitz et al., In mice, the crossed entorhinodentate pathway
1975; Steward, 1976b, 1980; Steward and Scoville, seems to be essentially non-existent (van Groen
1976; Steward and Vinsant, 1978; Amaral and et al., 2002, 2003), and it is unlikely that the very
Witter, 1995; Deller et al., 1996a). These studies few fibers observed in the contralateral dentate
revealed that the crossed entorhinodentate projec- gyrus of mice could contribute significantly to the
tion is weaker but anatomically homologous to the reinnervation of the dentate gyrus following den-
ipsilateral pathway. From a functional point of ervation. However, it should also be pointed out
view this homology makes the crossed entorhino- that this has not yet been systematically analyzed
dentate pathway particularly interesting, because and that it is unclear to what extent sprouting can
its sprouting could contribute to a restoration be elicited from the few crossed entorhinodentate
of function following entorhinal denervation. Ac- fibers observed in mice. In summary, we conclude
cordingly, morphological, electrophysiological, that the entorhinodentate projection is almost ab-
and behavioral changes caused by the reorganiza- sent in mice and that there is presently no evidence
tion of crossed entorhinodentate fibers were stud- for crossed entorhinodentate fiber sprouting in this
ied in detail (Steward et al., 1976, 1988; Steward, species.
511

Glial changes following entorhinal denervation regulatory functions have also been proposed, and
it was suggested that microglia could initiate a
In recent years, the role of non-neuronal cells in sequence of events, which could eventually induce
the reorganization of the dentate gyrus after ento- or regulate axon sprouting (Gage et al., 1988;
rhinal denervation has received considerable Fagan and Gage, 1990; Eyupoglu et al., 2003). In
attention. The reaction of resident glia as well as short, it was suggested that entorhinal terminal
the recruitment of blood cells into the zone of degeneration activates microglia. These activated
denervation have been studied, and a role of these cells phagocytose degenerating terminals and
cells in the reorganization process has been pro- release interleukin-1, thereby activating the
posed; these non-neuronal cells are not only im- astroglia. Activated astrocytes, in turn, secrete
portant for the removal of cellular debris in the neurotrophic factors, which induce sprouting of
zone of denervation, but they may also regulate axons carrying specific receptors (Gage et al.,
axonal growth and transneuronal degeneration 1988; Eyupoglu et al., 2003).
processes. Thus, non-neuronal cells may play In mice, the morphologic transformation of mi-
a central role in the reorganization of the dentate croglia is fairly similar to the changes observed in
gyrus following denervation (see below). rats in that an accumulation of reactive microglia
occurs within the denervated parts of the molec-
ular layer (Fig. 3a) as early as 24 h after entorhinal
Microglia denervation (Schauwecker and Steward, 1997;
Jensen et al., 1999; Wirenfeldt et al., 2003;
In rats, microglial cells react very rapidly to ento- Rappert et al., 2004; Pedersen et al., 2006). At
rhinal denervation by proliferation, and by trans- this time, the microglia have transformed into
formation from a ramified to an ‘‘activated’’ what is regarded as an ‘‘amoeboid’’ phenotype,
phenotype (Gall et al., 1979b; Fagan and Gage, and the relative depletion of the adjacent non-
1990, 1994; Gehrmann et al., 1991; Jensen et al., denervated inner molecular layer suggests that
1994; Hailer et al., 1997). Using histological and they migrate into the zone of axonal degeneration.
immunohistochemical techniques, microglial reac- In fact, we have demonstrated that the microglial
tions were visible as early as 24 h postlesion, were accumulation is severely impaired in mice deficient
strongest at approximately 3 days postlesion, and for CXCR3, a chemokine receptor crucial for the
returned to normal levels during the first 2 weeks attraction of microglia (Rappert et al., 2002, 2004).
after denervation. Electron microscopy and Mini The other known factor contributing to the mi-
Ruby-tracing demonstrated that reactive micro- croglial accumulation is proliferation, which has
glial cells phagocytose axonal debris and myelin been demonstrated in mice using modern stereo-
sheaths (Gehrmann et al., 1991; Jensen et al., 1994; logical techniques (Wirenfeldt et al., 2003; Ladeby
Bechmann and Nitsch, 1997). Interestingly, mi- et al., 2005a, b). The molecular regulation of the
croglial activation was also observed, though to microglial response following entorhinal denerva-
a much lesser extent, in adjacent non-denervated tion has been studied in greater detail in mice
areas, suggesting that microglial cells are actively compared to rats. Besides CXCR3 (Rappert et al.,
recruited to the denervated zone (Gall et al., 2004), the peripheral benzodiazepine binding site
1979b; Jensen et al., 1994). Molecules that could ligand PK11195 (Pedersen et al., 2006), and inter-
regulate this highly specific recruitment are feron-gamma (Jensen et al., 2000a), interleukin
integrin adhesion molecules, such as leukocyte 1beta and the transforming growth factor beta1
function antigen-1 (LFA-1), very late antigen-4 (Eyupoglu et al., 2003) have been implicated. In
(VLA-4), and the ligand for LFA-1, intercellular addition to the role of microglia as phagocytes,
adhesion molecule-1 (ICAM-1), which are all however, regulatory functions were also proposed
expressed by microglial cell in the zone of dener- and, recently, experimental evidence for a link be-
vation (Hailer et al., 1997). In addition to the tween microglia activation and dendritic atrophy
role of microglia as debris-removing phagocytes, of GABAergic parvalbumin-positive basket cells
512

Fig. 3. Microglial cells play a role in transneuronal degeneration of parvalbumin-positive dendrites. (A) Overview of the entorhinal-
hippocampal area of a control mouse after denervation of the middle molecular layer (MML) of the dentate gyrus. Mac-1 positive
microglia accumulate in the zone of anterograde axonal degeneration. The insert shows the morphology of the activated microglial
cells. This microglial response is less pronounced in CXCR3 / (KO) mice compared to wild-type (WT) controls. (B–E) If less
microglial cells are recruited, denervated dendrites are preserved. The loss of parvalbumin-positive dendrites in the dentate gyrus is
evident in WT mice (B, D) and clearly diminished in CXCR3 / animals (C, E). The boxes in (B) and (C) indicate the areas shown at
higher magnification in (D) and (E), respectively. Scale bars: A: 300 mm; inset: 50 mm; B, C: 100 mm; D, E: 25 mm.
513

in the denervated dentate gyrus was shown participate in the removal of degenerating termi-
(Rappert et al., 2004; see also below). nals and spines (Bechmann and Nitsch, 1997;
Deller et al., 1997). Increases in GFAP messenger
RNA (mRNA) slightly preceded changes in GFAP
Blood-derived cells
and were observed as early as 12 h postlesion
(Steward et al., 1990, 1993). Interestingly, GFAP
Besides migration and proliferation of resident
mRNA changes depended on postlesional seizure
microglia, a third mechanism contributing to the
activity, demonstrating that changes in GFAP
accumulation of microglia/macrophages within
expression in the early postlesional time period
the zone of denervation is the recruitment of
may be caused by changes in activity rather than
blood-derived monocytic cells. At least in the
denervation (Kelley and Steward, 1996b). In addi-
mouse, these cells are able to invade the denervat-
tion, it was shown that reactive astrocytes change
ed hippocampus, where they transform into mi-
the molecular composition of the denervated
croglia-like elements within 72 h after entorhinal
molecular layer (Gall et al., 1979b; Gage et al.,
denervation (Bechmann et al., 2005; Ladeby et al.,
1988; Steward, 1991; Deller et al., 2000, 2001):
2005b).
Astrocyte-derived neurotrophic factors (Kawaja
As all antibodies used for microglia-detection
and Gage, 1991; Yoshida and Gage, 1991;
also bind to monocytes/macrophages, the invasion
Gomez-Pinilla et al., 1992; Guthrie et al., 1995,
of hematopoietic cells as a source of ‘‘new’’ mi-
1997; Lee et al., 1997), cell-adhesion molecules
croglia have been difficult to study. Fagan and
(Styren et al., 1994, 1995; Jucker et al., 1996), and
Gage (1994) analyzed rats using i.v. injection of
a variety of extracellular matrix molecules (Deller
DiI-labeled low density lipoprotein (LDL) parti-
et al., 2000) were studied and it was suggested that
cles, but could not find any evidence for the infil-
these astrocyte-derived molecules could regulate
tration of monocytes. In their preparations, T cells
and pattern the sprouting response.
were also not visible (Fagan and Gage, 1994).
In mice, less information is available. Astroglial
Later, using the R73 antibody, we demonstrated
changes were studied using GFAP-immunohisto-
the presence of CD4 a/b in the middle molecular
chemistry, revealing astroglial migration and an
layer after entorhinal denervation in the same rat
increased expression of GFAP-immunoreactivity
strain (Bechmann et al., 2001a), but the question
throughout the denervated zone (Steward and
whether monocytes are also recruited has only
Trimmer, 1997; Del Turco et al., 2003). Changes
been addressed in mice, where CCL-2 has been
in GFAP mRNA were measured using northern
identified as a major chemoattractant molecule
blot (Steward and Trimmer, 1997; Steward et al.,
(Babcock et al., 2003). As described below, recent
1997) and at the single-cell level using laser micro-
evidence links chemokine-mediated recruitment of
dissection in combination with quantitative real-
microglia to transsynaptic dendritic changes in the
time-polymerase chain reaction (qRT-PCR)
dentate gyrus.
(Burbach et al., 2004a, b). These postlesional
changes in GFAP mRNA depend, as in the rat,
Astroglia on activity and protein synthesis (Steward, 1994a;
Steward et al., 1997). In contrast to rats, however,
In rats, astroglial activation after entorhinal dener- studies analyzing changes in the expression of
vation was observed at slightly later time points astrocyte-derived molecules following entorhinal
than the activation of microglia and was charac- denervation are rare (Poulsen et al., 2000; Mayer
terized by proliferation, upregulation of glial fibril- et al., 2005; Schafer et al., 2005).
lary acidic protein (GFAP), and migration into the In summary, morphological evidence supports
denervated zone of the dentate gyrus (Rose et al., a phagocytic role of astrocytes in the denervated
1976; Gall et al., 1979b; Gage et al., 1988; Fagan dentate gyrus of the adult rat and correlative data
and Gage, 1990; Steward et al., 1993). Within the implicate astrocytes in the patterning of the
denervated zone, astrocytes were shown to sprouting response. In mice, the cellular response
514

of astrocytes appears to be similar to the one 2006), suggesting that sprouting axons could be
observed in rats. However, information about myelinated by newly formed oligodendroglial cells.
molecular changes induced by reactive astrocytes In summary, oligodendrocytes react to ento-
is scarce and the role of astrocytes in this species rhinal denervation in rats as well as mice. Their
remains to be elucidated. surface molecules may be involved in the regula-
tion of sprouting, and newly formed oligodendro-
cytes may contribute to the myelination of
sprouting axons.
Oligodendroglia and NG2-positive cells
Dendritic changes following entorhinal denervation
It is now well established that oligodendrocytes
and some of their surface molecules are potent in-
Denervated neurons in the CNS remodel their
hibitors of axonal growth (Kapfhammer and
dendritic field. This reorganization process
Schwab, 1992; Schwab et al., 1993). Since these
involves spines and dendrites, and may, if the den-
inhibitory molecules could influence sprouting,
ervation is extensive, lead to the death of the
they were studied following entorhinal denerva-
denervated target cell (Steward, 1994b; Kovac
tion. In addition, the entorhinal denervation
et al., 2004). In the rat, dentate gyrus granule cells
model was used to study the myelination of newly
(Nafstad, 1967; Steward and Scoville, 1976;
formed (sprouting) axons in the normal adult
Amaral and Witter, 1995) and parvalbumin-
CNS. Changes in putative oligodendroglial cell
positive GABAergic interneurons (Kosaka et al.,
precursors, NG2-cells (Dawson et al., 2000, 2003)
1987; Zipp et al., 1989; Nitsch et al., 1990) are the
were also analyzed.
targets of entorhinal afferents. These fibers im-
In rats, the first evidence for a reaction of
pinge on their distal dendrites and form between
oligodendroglial cells to entorhinal denervation
80 and 90% of all synapses on their distal dendritic
was provided by Gall et al. (1979b), who observed
field (Matthews et al., 1976a; Steward and Vinsant,
signs of oligodendroglial cell proliferation. Later,
1983). In non-lesioned C57BL/6 mice, the ana-
changes in myelination and Nogo-expression were
tomical situation appears to be similar (Deller,
analyzed (Meier et al., 2003, 2004). These studies
1998; Gebhardt et al., 2002; Del Turco et al.,
revealed correlations between the axonal sprouting
2004). Thus, ECL/perforant pathway transection
response and the time course of demyelination,
will denervate both cell types in both species.
suggesting that oligodendroglial-derived inhibitory
molecules could regulate sprouting. Consistent
with the results of Gall et al. (1979a, b), we re- Transneuronal changes in granule cells and
cently showed that oligodendroglial precursor parvalbumin-positive neurons
cells, NG2-positive cells, also react to entorhinal
denervation in rats (Dehn et al., 2006). In rats, transneuronal changes in granule cell mor-
In mice, an analysis of the dynamics of oligoden- phology after entorhinal denervation have been
drocytes following entorhinal denervation was per- studied using Golgi-impregnated (Parnavelas et al.,
formed by Drojdahl and colleagues (2004). In 1974; Caceres and Steward, 1983) and intracellularly
addition, changes in oligodendroglial-derived mole- labeled (Diekmann et al., 1996) granule cells. These
cules, such as changes in the expression of myelin studies agree that the dendritic arbor of granule cells
basic protein (Jensen et al., 2000b; Drojdahl is remodeled following denervation, and that gran-
et al., 2004) and myelin-associated glycoprotein ule cells lose spines and dendrites during the first
(Mingorance et al., 2005), were reported. Similar to 2 weeks after entorhinal denervation. By 30 days
the observations in rats, proliferation of NG2-pos- postlesion, there was a recovery of the dendritic ar-
itive oligodendroglial precursors and the formation bor detected by the Golgi method (Caceres and
of new oligodendrocytes was demonstrated follow- Steward, 1983), whereas permanent transneuronal
ing entorhinal denervation in mice (Nielsen et al., changes of granule cell dendrites were reported
515

using the intracellularly labeled granule cells cells and parvalbumin-positive neurons following
(Diekmann et al., 1996). entorhinal denervation? Although several hypoth-
Dendritic remodeling of parvalbumin-positive eses have been proposed, including excitotoxicity,
GABAergic neurons was also studied following elevation of intracellular calcium levels, and lack
entorhinal denervation (Nitsch and Frotscher, of growth factors (Steward, 1991; Nitsch and
1991, 1992, 1993; Nitsch et al., 1992). Similar to Frotscher, 1992; Sattler et al., 1998; Arundine and
granule cells, the distal dendrites of parvalbumin- Tymianski, 2003; Kovac et al., 2004), the role of
positive neurons demonstrated degenerative these mechanisms and molecules has not yet been
changes (Nitsch and Frotscher, 1991, 1992, clarified. As far as granule cells are concerned, no
1993). These degenerative changes persisted, with studies to date have explained the extensive
only a slight recovery by 2 months postlesion transneuronal degeneration of the denervated
(Nitsch and Frotscher, 1991). granule cell arbor. In the case of the degeneration
In mice, morphological changes of granule cells of parvalbumin-positive neurons, Nitsch and
have not yet been studied at the level of single Frotscher (1992) showed that the application of
cells, and at present only indirect — and contro- the glutamate receptor antagonist MK801 prior to
versial — data are available. Following ibotenic entorhinal denervation could prevent the atrophy
acid injections into the entorhinal cortex, Kadish of parvalbumin-immunoreactive dendrites (Nitsch
and van Groen (2003) did not observe atrophy of and Frotscher, 1992). They suggested that an
the denervated molecular layer of the mouse dent- NMDA-receptor mediated influx of calcium could
ate gyrus, and suggested that granule cell dendritic cause the degeneration of parvalbumin-positive
atrophy is limited in this species. In contrast, oth- dendrites. In mice, degeneration of parvalbumin-
ers (Phinney et al., 2004) who used mechanical positive dendrites was recently investigated
transection and measured the width and the vol- (Rappert et al., 2004) and microglia and chemo-
ume of the molecular layer using stereological kines were implicated in the regulation of trans-
methods demonstrated a>40% reduction in vol- neuronal degeneration. Since this study proposed
ume 45 days postlesion, suggesting substantial a novel mechanism for denervation-induced trans-
granule cell atrophy. A detailed temporal analysis neuronal degeneration a more detailed review of
on the postlesional morphological changes of sin- this study and the potential role of chemokines
gle granule cells is needed to clarify this issue, and appears to be warranted.
its potential differences from the rat.
In contrast to granule cells, morphological
changes of parvalbumin-positive neurons have Transneuronal degeneration of parvalbumin-positive
been studied in more detail in mouse (Rappert dendrites: a role for the chemokine receptor CXCR3
et al., 2004). A temporal analysis demonstrated
changes in parvalbumin-positive dendrites that Chemokines are a diverse family of pro-inflamma-
were quite similar to rats. Parvalbumin-positive tory cytokines. These small homologous peptides
dendrites were rapidly lost in the denervated mo- (8–14 kDa) (Rollins, 1997; Baggiolini, 1998) play
lecular layer of C57BL/6 mice and no recovery was an important role in recruiting inflammatory cells
observed up to 31 days postlesion (Fig. 3). How- into tissues in response to infection and inflam-
ever, a detailed morphological analysis of identi- mation (Karpus and Ransohoff, 1998; Mackay,
fied parvalbumin-positive cells was not performed. 2001). Additionally, these molecules are involved
in the pathogenesis of many important neuroin-
flammatory diseases ranging from multiple sclero-
Transneuronal changes: insights into cellular sis and stroke to HIV encephalopathy (Luster,
mechanisms 1998; Asensio and Campbell, 1999).
For many chemokine receptors in the adult
What are the potential mechanisms involved in the CNS, the corresponding ligand can only be de-
transneuronal dendritic degeneration of granule tected under pathological conditions, for example,
516

in the vicinity of an inflammatory lesion (Asensio CXCL10 mRNA has been found in the lesioned
and Campbell, 1999; Bacon and Harrison, 2000) hippocampus of the mouse (Babcock et al., 2003)
suggesting that chemokine-receptors and their lig- and the protein could be localized on neurons by
ands are primarily involved in the regulation of our group (Rappert et al., 2004). In mice deficient
pathophysiological processes. In line with this in- for CXCR3, the accumulation of microglia in
terpretation, several studies have shown that zones of denervation was dramatically reduced,
chemokines recruit leukocytes into inflamed brain demonstrating a crucial role for CXCR3 in mi-
tissue (Fife et al., 2000; Izikson et al., 2000; Siebert croglial recruitment and suggesting CXCL10 as
et al., 2000; Huang et al., 2001; Babcock and the signaling partner (Rappert et al., 2004).
Owens, 2003). Messenger RNA for CCL5, CCL2, The importance of CXCR3 is underscored by
CXCL10, CCL3, CCL4 and CXCL2 were specifi- the preservation of parvalbumin-positive dendrites
cally induced in the lesion-reactive hippocampus after entorhinal denervation in CXCR3-deficient
and at the site of axonal transection by 24 h after mice (Rappert et al., 2004). The absence of the
axotomy, while other chemokines like XCL1, dendritic loss in the CXCR3 knockout animals is
CCL11 and CCL1 were not detected (Babcock not simply delayed, since a difference in the den-
et al., 2003). Interestingly, CCL2 was upregulated dritic length compared to wild-type animals was
in the denervated hippocampus already at 3 h after still evident 31 days after entorhinal denervation.
axotomy, while an increase in CCL5 was first de- Although a reduction of dendrites developed, there
tected after 24–48 h. The different time course of were significantly more dendrites in the KO in
induction of these two chemokines suggests that comparison to the wild-type animal (Figs 3b–e).
chemokines could have different functions follow- This is consistent with previous observations that
ing ECL. In fact, as pointed out above, analyses the elimination of microglia in rat entorhinohip-
of chemokine receptor-deficient mice indicated pocampal slice cultures diminishes lesion-induced
that the CCR2 ligand CCL2 (MCP-1) is critical dendritic loss (Eyupoglu et al., 2003). Although
for leukocyte infiltration, whereas CCR5 ligands there is no mechanistic explanation for this phe-
such as CCL5 (RANTES) are not (Babcock et al., nomenon, these data argue against the concept
2003). that ‘‘retraction’’ is dependent only on the loss of
What could be the role of the chemokines and a specific input, and demonstrates a causal link
their receptors in the reorganization of the dentate between microglial response and the structural
gyrus following ECL? Since chemokine receptors outcome of postlesional changes. Whether the mi-
are expressed in neurons and glia, and chemokines croglia actively eliminate denervated dendritic seg-
are mostly produced by glial cells, it has been sug- ments or provide signals inducing their retraction
gested that they are involved in glia-to-glia, glia- remains to be seen.
to-neuron, and neuron-to-glia interactions
(Hesselgesser and Horuk, 1999; Ambrosini and
Aloisi, 2004; Adler et al., 2005). In the context of Summary
ECL the chemokine receptor CXCR3 and its lig-
ands CXCL9, CXL10, CXCL11 and CCL21 are In the present review, we have summarized and
especially interesting. For example, CCL21 is spe- compared the available data on the structural
cifically expressed by damaged neurons (Biber reorganization of the dentate gyrus following
et al., 2001; de Jong et al., 2005) and not by any entorhinal denervation in rats and mice. We con-
glial cell type (Biber et al., 2001; Alt et al., 2002; clude on the basis of this comparison, that in mice
Columba-Cabezas et al., 2003). Since CCL21 in- neither the normal anatomy of the dentate gyrus
duces intracellular calcium signals and chemotaxis and nor many aspects concerning its structural re-
of cultured microglia (Biber et al., 2001; Rappert organization following denervation have yet been
et al., 2002), a potential role of CCL21 in neuron- described in depth. The available data, however,
microglia communication has been proposed — already point to important species differences in
but has to be proven in vivo. As a second ligand, both aspects, anatomy and reorganization. In the
517

following paragraphs, we will discuss these species Fernaud-Espinosa, 2000), and it could be demon-
differences in the context of sprouting and inflam- strated that several growth inhibitory extracellular
mation and will make suggestions for the use of matrix molecules such as tenascin-C (Deller et al.,
the entorhinal denervation model in mutant mice. 1997), DSD-1-proteoglycan (Deller et al., 1997),
neurocan (Haas et al., 1999), brevican (Thon et al.,
2000), and NG2 (Dehn et al., 2006) are upregu-
How can the species difference in the sprouting lated within the denervated zone of the molecular
response be explained? layer after entorhinal denervation. Because these
molecules form a sharp border towards the adja-
Our comparison of sprouting following entorhinal cent non-denervated zones, we suggested that
denervation in rats and mice revealed major spe- these molecules delineate boundaries of axonal
cies differences. Some of these differences may growth following entorhinal lesion (Deller et al.,
readily be explained on the basis of differences in 2000, 2001). This hypothesis, although certainly
normal anatomy. For example, if crossed entorhi- attractive, still awaits experimental verification.
nodentate fibers do not exist in mice, sprouting of In mice, the situation is different and C/A fibers
this fiber projection is not to be expected. Other to the inner molecular layer sprout across laminar
species differences, for example the response of the boundaries. One explanation could be that in this
C/A fibers to the inner molecular layer are more species, the expression pattern, the time course
difficult to explain. This projection system exists in of expression, or the order of assembly of growth
rats as well as mice and only subtle anatomical inhibitory extracellular matrix molecules could
differences are observed between the two species be different. So far, only few of the extracellular
(Del Turco et al., 2003). Nevertheless, in adult rats matrix molecules mentioned above have been
these C/A fibers stay in their home territory investigated: Brevican (Mayer et al., 2005) and
whereas they sprout across the laminar bounda- NG2 (Nielsen et al., 2006) were analyzed and den-
ries in mice. In the absence of major anatomical ervation-induced changes were reported which
differences, an explanation for this phenomenon were similar to those observed for these molecules
should be sought on the molecular level. in rats. However, subtle differences with regard to
In recent years, we have studied molecules which the time course of expression and the cellular dis-
could regulate the layer-specific sprouting of C/A tribution of the two molecules were also observed.
fibers to the inner molecular layer in rats (Frotsc- Whether or not these differences — or differences
her et al., 1997; Deller et al., 2000, 2001). Whereas in the expression of other extracellular matrix
glia-derived neurotrophic factors and cytokines molecules — could result in an absent or less
are probably involved in the induction of new effective growth barrier between the non-dener-
axonal collaterals (Gomez-Pinilla et al., 1992; vated inner molecular layer and the denervated
Guthrie et al., 1995, 1997; Fagan et al., 1997; zone in mice remains to be seen.
Lee et al., 1997; Woods et al., 1998, 1999), other Another explanation could be that mossy cells in
molecules may be better suited for the guidance of mice and rats differ in their ability to grow, i.e., in
newly formed fibers. A group of molecules which their intrinsic growth competence (Benowitz and
could regulate a layer-specific growth response by Routtenberg, 1997; Caroni, 1997, 1998). It is well
forming a growth barrier between the non-dener- established that the growth competence of neurons
vated inner and the denervated middle- and outer depends on the expression of growth-associated
molecular layer are extracellular matrix molecules proteins, such as growth-associated protein 43
synthesized by reactive glia (Rose et al., 1976; or cortical cytoskeleton-associated protein 23
Gage et al., 1988). These molecules are thought to (Aigner et al., 1995; Benowitz and Routtenberg,
form boundaries for growing axons during devel- 1997; Caroni et al., 1997; Frey et al., 2000; Laux
opment (Faissner and Steindler, 1995; Höke and et al., 2000). These molecules are strongly expressed
Silver, 1996; Pearlman and Sheppard, 1996; during brain development and are downregulated
Margolis and Margolis, 1997; Bovolenta and postnatally (Caroni, 1997). Intriguingly, the ability
518

of rat mossy cells to sprout across laminar bound- reorganization, but as yet an impact on the sprout-
aries (Lynch et al., 1973; Gall and Lynch, 1980, ing response has not been proven. However, in both
1981; Gall et al., 1986) and the downregulation of rats and mice, interfering with microglial recruit-
growth-associated proteins (Caroni, 1997) occurs ment impacts on the loss of denervated dendrites
during the same developmental time window, sug- (Eyupoglu et al., 2003; Rappert et al., 2004).
gesting that the ability of mossy cell axons to sprout Phagocytosis of distal segments by activated micro-
into the denervated dentate gyrus indeed depends on glia is an attractive explanation for this observation,
their intrinsic growth competence. In contrast, in which is currently tested in our laboratory.
mice, in which translaminar mossy cell sprouting The immune system of rodents is capable of
is found in the adult, mossy cell growth competence mounting strong immune responses to myelin-
is relatively strong. Recent observations in trans- associated epitopes, but entorhinal denervation
genic mice overexpressing growth-associated pro- does not induce evident autoimmune demyelina-
teins indicate that the intrinsic growth-competence tion. This could be attributed to a state of immune
of mossy cells affects the strength and extent of the ignorance, as Fagan and Gage (1994) did not find
translaminar commissural sprouting response in infiltration of haematogenous cells into zones of
mice following entorhinal denervation (Del Turco axonal degeneration after entorhinal denervation
et al., 2001). in the rat. In the same strain, we detected infil-
Finally, competitive interactions of sprouting trating CD4 T cells interacting with MHC-II pos-
C/A mossy cell axons with other sprouting fiber itive microglia. These myelin-phagocytosing cells
systems could also play a role (Cotman and exhibited the phenotype of immature antigen-pre-
Nadler, 1978; Nadler et al., 1980; Steward, senting cells which is likely to cause T cell anergy
1994b). In rats, crossed entorhinodentate fibers (Bechmann et al., 2001a). Moreover, astrocytes
sprout within the denervated molecular layer expressed the death-ligand CD95L (FasL) which
(Steward et al., 1974; Deller et al., 1996a). Since causes apoptosis of activated T cells (Bechmann
recent data indicate that C/A fibers are repelled by et al., 1999, 2000, 2002). Thus, immune tolerance
entorhinal fibers (Borrell et al., 1999; Deller et al., seems to be actively maintained and, in fact, we
1999a), sprouting of this pathway would block have shown that autoimmunity to myelin is di-
translaminar ingrowth of C/A fibers into the den- minished following entorhinal denervation in
ervated zone in rats. In contrast, C/A fibers could mice. Strikingly, axonal antigens are found in cer-
grow into the denervated mouse fascia dentata vical lymph nodes after entorhinal denervation
because virtually all entorhinal fibers are removed and their appearance is accompanied by T cell
from the mouse dentate gyrus after unilateral ent- apoptosis (Mutlu et al., 2007). Monocytic cells are
orhinal denervation. If this hypothesis were true, known to pass the vascular wall to reside in peri-
then the absence of the crossed entorhinodentate vascular spaces in rats and mice under normal
pathway in mice would suffice to explain the major conditions (Bechmann et al., 2001b, c), but
species differences between rats and mice with re- they cross the glia limitans and transform into
gard to the axonal reorganization of the dentate microglia after entorhinal denervation in mice
gyrus following denervation. (Bechmann et al., 2005). CCL2 has been identified
as the key chemokine driving this recruitment
(Babcock et al., 2003), while it is unclear which
Inflammatory changes following denervation and signals adopt the invading macrophages into ram-
the immune response to entorhinal denervation ified microglia. One of the key questions is whether
such ‘‘microglia’’ transform into dendritic cells
The morphologic transformation of astrocytes and capable of leaving the neuropil to present antigens
microglia and their recruitment into the zones of in lymphoid organs. In fact, after entorhinal den-
axonal degeneration are similar in mice and rats. As ervation in mice, T cells specifically target one
described above, much has been speculated with re- cervical lymph node via the cribroid plate and the
gard to the contribution of glia to postlesional nasal mucosa (Goldmann et al., 2006).
519

It is also noteworthy that permeability of the usually the rat, were used to interpret data obtained
blood-brain barrier for IgG has been demon- in the other, usually the mouse. Since this assump-
strated after entorhinal denervation in the rat tion was never proven, we have compiled and com-
a decade ago (Jensen et al., 1997), but the role of pared the data available on structural changes
antibodies in this context and their impact on following entorhinal denervation in both species in
postlesional reorganization is completely unclear. this review. This comparative analysis revealed ma-
It is now clear that entorhinal denervation in jor differences between the entorhinal model system
mice and rats involves a systemic immune in rats and mice. We conclude, that a non-critical
response, which impacts on the structural out- transfer of data from one species to the other is not
come at least at the levels of phagocytosis of warranted.
growth-inhibiting debris. T cells have also been Given the difficulty in generalizing data from
shown to secrete neurotrophins (Kerschensteiner the rat to mouse, all observations made in rats
et al., 1999), but entorhinal denervation-studies in need first to be verified in the context of the mouse
mice bearing myelin-specific T cell receptors have model before mutants should be analyzed. In ad-
not yet been performed. It is evident from count- dition, strain differences should also be considered
less immunological studies that there are signifi- and it will probably be necessary to reinvestigate
cant species and strain differences in regard to the certain aspects specifically in the background
strength and consequence of immune responses to strain of a given mutant. Only then, will it be pos-
brain antigens (Gold et al., 2006). In regard to the sible to gain meaningful and fundamental biolog-
neurological outcome, they can be protective or ical insights from the entorhinal denervation
detrimental. The same type of lesion can evoke model in mice. Although certainly demanding,
overall protective or detrimental immune re- such an approach may yield additional scientific
sponses solely depending on strain subjected to benefits, since the investigation of species and
the damage (Kipnis et al., 2001). Thus, beyond strain differences in hippocampal connectivity and
species differences, strain differences must also be plasticity may open up new strategies for a com-
expected when evaluating the role of immune cells parative, yet functionally oriented analysis. This
in postlesional plasticity. way, a methodologically more careful approach
The intrinsic immune suppression of the brain would be richly awarded and of considerable value
may have evolved to cope with situations bearing to the scientific community.
high evolutionary pressure such as infection, where it
is better for the individual to tolerate certain infec-
tious agents than to eliminate all infected neurons. Acknowledgments
Under conditions such as axonal injuries, which cer-
tainly provided low pressure, it may be beneficial for This review is dedicated to Michael Frotscher,
the neurological outcome to therapeutically foster M.D., who is one of the pioneers of hippocampal
immune responses (Bechmann, 2005). This can be research and the entorhinal denervation model,
tested in mouse models of entorhinal denervation. and who contributed to many of the studies
reviewed here. This study was supported by the
Deutsche Forschungsgemeinschaft (DE 551/8-1,
Conclusions for the use of the entorhinal denervation SFB 507 B16), German Israeli Foundation (T.D.),
model in rats and mice and Gisela Stadelmann-Stiftung (D.D.T.)

In summary, ECL/perforant pathway transections


have been used to denervate the dentate gyrus of rats
References
and mice in order to study the reorganization of the
brain following injury. Under the assumption that Adler, M.W., Geller, E.B., Chen, X. and Rogers, T.J. (2005)
the reorganization of the dentate gyrus is similar in Viewing chemokines as a third major system of communica-
the two related rodent species, data obtained in one, tion in the brain. AAPS J., 7: E865–E870.
520

Aigner, L., Arber, S., Kapfhammer, J.P., Laux, T., Schneider, programmed cell death during the course of anterograde de-
C., Botteri, F., Brenner, H.R. and Caroni, P. (1995) Over- generation. Glia, 32: 25–41.
expression of the neural growth-associated protein GAP-43 Bechmann, I., Mor, G., Nilsen, J., Eliza, M., Nitsch, R. and
induces nerve sprouting in the adult nervous system of Naftolin, F. (1999) FasL (CD95L, Apo1L) is expressed in the
transgenic mice. Cell, 83: 269–278. normal rat and human brain: evidence for the existence of an
Alt, C., Laschinger, M. and Engelhardt, B. (2002) Functional immunological brain barrier. Glia, 27: 62–74.
expression of the lymphoid chemokines CCL19 (ELC) and Bechmann, I. and Nitsch, R. (1997) Astrocytes and microglial cells
CCL 21 (SLC) at the blood-brain barrier suggests their incorporate degenerating fibers following entorhinal lesion: a
involvement in G-protein-dependent lymphocyte recruit- light, confocal, and electron microscopical study using a
ment into the central nervous system during experimental phagocytosis-dependent labeling technique. Glia, 20: 145–154.
autoimmune encephalomyelitis. Eur. J. Immunol., 32: Bechmann, I., Peter, S., Beyer, M., Gimsa, U. and Nitsch, R.
2133–2144. (2001a) Presence of B7-2 (CD86) and lack of B7-1 (CD80) on
Amaral, D.G. and Witter, M.P. (1989) The three dimensional myelin phagocytosing MHC-II-positive rat microglia is as-
organization of the hippocampal formation: a review of sociated with nondestructive immunity in vivo. FASEB J.,
anatomical data. Neuroscience, 31: 571–591. 15: 1086–1088.
Amaral, D.G. and Witter, M.P. (1995) The hippocampal for- Bechmann, I., Priller, J., Kovac, A., Bontert, M., Wehner, T.,
mation. In: Paxinos G. (Ed.), The Rat Nervous System Klett, F.F., Bohsung, J., Stuschke, M., Dirnagl, U. and
(2nd ed.). Academic Press, San Diego, CA, pp. 443–494. Nitsch, R. (2001b) Immune surveillance of mouse brain peri-
Ambrosini, E. and Aloisi, F. (2004) Chemokines and glial cells: vascular spaces by blood-borne macrophages. Eur. J. Ne-
a complex network in the central nervous system. Neuroc- urosci., 14: 1651–1658.
hem. Res., 29: 1017–1038. Bechmann, I., Steiner, B., Gimsa, U., Mor, G., Wolf, S., Beyer,
Arundine, M. and Tymianski, M. (2003) Molecular mecha- M., Nitsch, R. and Zipp, F. (2002) Astrocyte-induced T cell
nisms of calcium-dependent neurodegeneration in excitotox- elimination is CD95 ligand dependent. J. Neuroimmunol.,
icity. Cell Calcium, 34: 325–337. 132: 60–65.
Asensio, V.C. and Campbell, I.L. (1999) Chemokines in the Beck, K.D., Lamballe, F., Klein, R., Barbacid, M., Schau-
CNS: plurifunctional mediators in diverse states. Trends wecker, P.E., McNeill, T.H., Finch, C.E., Hefti, F. and Day,
Neurosci., 22: 504–512. J.R. (1993) Induction of noncatalytic TrkB neurotrophin re-
Aubert, I., Poirier, J., Gauthier, S. and Quirion, R. (1994) ceptors during axonal sprouting in the adult hippocampus.
Multiple cholinergic markers are unexpectedly not altered in J. Neurosci., 13: 4001–4014.
the rat dentate gyrus following entorhinal cortex lesions. Benowitz, L.I., Rodriguez, W.R. and Neve, R.L. (1990) The
Neuroscience, 14: 2476–2484. pattern of GAP-43 immunostaining changes in the rat hip-
Babcock, A. and Owens, T. (2003) Chemokines in experimental pocampal formation during reactive synaptogenesis. Mol.
autoimmune encephalomyelitis and multiple sclerosis. Adv. Brain Res., 8: 17–23.
Exp. Med. Biol., 520: 120–132. Benowitz, L.I. and Routtenberg, A. (1997) GAP-43: an intrinsic
Babcock, A.A., Kuziel, W.A., Rivest, S. and Owens, T. (2003) determinant of neuronal development and plasticity. Trends
Chemokine expression by glial cells directs leukocytes to sites Neurosci., 20: 84–91.
of axonal injury in the CNS. J. Neurosci., 23: 7922–7930. Biber, K., Sauter, A., Brouwer, N., Copray, S.C. and Boddeke,
Bacon, K.B. and Harrison, J.K. (2000) Chemokines and their H.W. (2001) Ischemia-induced neuronal expression of the
receptors in neurobiology: perspectives in physiology and microglia attracting chemokine secondary lymphoid-tissue
homeostasis. J. Neuroimmunol., 104: 92–97. chemokine (SLC). Glia, 34: 121–133.
Baggiolini, M. (1998) Chemokines and leukocyte traffic. Blackstad, T.W. (1958) On the termination of some afferents to
Nature, 392: 565–568. the hippocampus and fascia dentata: an experimental study
Bechmann, I. (2005) Failed central nervous system regenera- in the rat. Acta Anat. (Basel), 35: 202–214.
tion: a downside of immune privilege? Neuromolecular Med., Blasco-Ibanez, J.M. and Freund, T.F. (1997) Distribution,
7: 217–228. ultrastructure and connectivity of calretinin-immunoreactive
Bechmann, I., Goldmann, J., Kovac, A.D., Kwidzinski, E., mossy cells of the mouse dentate gyrus. Hippocampus,
Simburger, E., Naftolin, F., Dirnagl, U., Nitsch, R. and 7: 307–320.
Priller, J. (2005) Circulating monocytic cells infiltrate layers Book, A.A., Wiley, R.G. and Schweitzer, J.B. (1992) Specificity
of anterograde axonal degeneration where they transform of 192 IgG-saporin for NGF receptor- positive cholinergic
into microglia. FASEB J., 19: 647–649. basal forebrain neurons in the rat. Brain Res., 590: 350–355.
Bechmann, I., Kwidzinski, E., Kovac, A.D., Simburger, E., Borrell, V., Ruiz, M., del Rio, J.A. and Soriano, E. (1999)
Horvath, T., Gimsa, U., Dirnagl, U., Priller, J. and Nitsch, Development of commissural connections in the hippocam-
R. (2001c) Turnover of rat brain perivascular cells. Exp. Ne- pus of reeler mice: evidence of an inhibitory influence of
urol., 168: 242–249. Cajal-Retzius cells. Exp. Neurol., 156: 268–282.
Bechmann, I., Lossau, S., Steiner, B., Mor, G., Gimsa, U. and Bovolenta, P. and Fernaud-Espinosa, I. (2000) Nervous system
Nitsch, R. (2000) Reactive astrocytes upregulate Fas (CD95) proteoglycans as modulators of neurite outgrowth. Prog.
and Fas ligand (CD95L) expression but do not undergo Neurobiol., 61: 113–132.
521

Burbach, G.J., Dehn, D., Del Turco, D., Staufenbiel, M. and Dehn, D., Burbach, G.J., Schafer, R. and Deller, T. (2006)
Deller, T. (2004a) Laser microdissection reveals regional and NG2 upregulation in the denervated rat fascia dentata fol-
cellular differences in GFAP mRNA upregulation following lowing unilateral entorhinal cortex lesion. Glia, 53: 491–500.
brain injury, axonal denervation, and amyloid plaque dep- Del Turco, D., Gebhardt, C., Burbach, G.J., Pleasure, S.J.,
osition. Glia, 48: 76–84. Lowenstein, D.H. and Deller, T. (2004) Laminar organiza-
Burbach, G.J., Dehn, D., Nagel, B., Del Turco, D. and Deller, tion of the mouse dentate gyrus: insights from BETA2/Neuro
T. (2004b) Laser microdissection of immunolabeled as- D mutant mice. J. Comp. Neurol., 477: 81–95.
trocytes allows quantification of astrocytic gene expression. Del Turco, D., Gebhardt, C., Woods, A.G., Kapfhammer, J.P.,
J. Neurosci. Methods, 138: 141–148. Naumann, T., Frotscher, M., Caroni, P. and Deller, T. (2001)
Caceres, A. and Steward, O. (1983) Dendritic reorganization in Overexpression of the growth associated protein CAP-23 re-
the denervated dentate gyrus of the rat following entorhinal sults in translaminar commissural sprouting in the hippo-
cortical lesions: a Golgi and electron microscopic analysis. J. campus after entorhinal cortex lesion in adult mice. Soc.
Comp. Neurol., 136: 287–295. Neurosci. Abstr., 27: 698.12.
Campbell, K.A., Bank, B. and Milgram, N.W. (1984) Del Turco, D., Woods, A.G., Gebhardt, C., Phinney, A.L.,
Epileptogenic effects of electrolytic lesions in the hippocam- Jucker, M., Frotscher, M. and Deller, T. (2003) Comparison
pus: role of iron deposition. Exp. Neurol., 86: 506–514. of commissural sprouting in the mouse and rat fascia dentata
Caroni, P. (1997) Intrinsic neuronal determinants that promote after entorhinal cortex lesion. Hippocampus, 13: 685–699.
axonal sprouting and elongation. Bio Essays, 19: 767–775. Deller, T. (1998) The anatomical organization of the rat fascia
Caroni, P. (1998) Driving the growth cone. Science, 281: dentate — new aspects of laminar organization as revealed by
1465–1466. anterograde tracing with Phaseolus vulgaris-leucoagglutinin.
Caroni, P., Aigner, L. and Schneider, C. (1997) Intrinsic neu- Anat. Embryol., 197: 89–103.
ronal determinants locally regulate extrasynaptic and synap- Deller, T., Bas, O.C., Vlachos, A., Merten, T., Del Turco, D.,
tic growth at the adult neuromuscular junction. J. Cell Biol., Dehn, D., Mundel, P. and Frotscher, M. (2006) Plasticity of
136: 679–692. synaptopodin and the spine apparatus organelle in the rat
Chen, L.L., Van Hoesen, G.E., Barnes, C.L. and West, J.R. fascia dentata following entorhinal cortex lesion. J. Comp.
(1983) Enhanced acetylcholinesterase staining in the hippo- Neurol., 499: 471–484.
campal perforant pathway zone after combined lesions of the Deller, T., Drakew, A. and Frotscher, M. (1999a) Different
septum and entorhinal cortex. Brain Res., 272: 354–359. primary target cells are important for fiber lamination in the
Collazos-Castro, J.E. and Nieto-Sampedro, M. (2001) Devel- fascia dentata: a lesson from reeler mutant mice. Exp. Ne-
opmental and reactive growth of dentate gyrus afferents: urol., 156: 239–253.
Cellular and molecular interactions. Restor. Neurol. Neuro- Deller, T., Drakew, A., Heimrich, B., Förster, E., Tielsch, A.
sci., 19: 169–187. and Frotscher, M. (1999b) The hippocampus of the reeler
Columba-Cabezas, S., Serafini, B., Ambrosini, E. and Aloisi, F. mutant mouse: Fiber segregation in area CA1 depends on the
(2003) Lymphoid chemokines CCL19 and CCL21 are ex- position of the postsynaptic target cells. Exp. Neurol., 156:
pressed in the central nervous system during experimental 254–267.
autoimmune encephalomyelitis: implications for the mainte- Deller, T. and Frotscher, M. (1997) Lesion-induced plasticity of
nance of chronic neuroinflammation. Brain Pathol., 13: central neurons: sprouting of single fibers in the rat hippo-
38–51. campus after unilateral entorhinal lesion. Prog. Neurobiol.,
Cotman, C.W., Gentry, C. and Steward, O. (1977) Synaptic 53: 687–727.
replacement in the dentate gyrus after unilateral entorhinal Deller, T., Frotscher, M. and Nitsch, R. (1995a) Morphological
lesion: electron microscopic analysis of the extent of replace- evidence for the sprouting of inhibitory commissural fibers in
ment of synapses by the remaining entorhinal cortex. J. Ne- response to the lesion of the excitatory entorhinal input to the
urocytol., 6: 455–464. rat dentate gyrus. J. Neurosci., 15: 6868–6878.
Cotman, C.W. and Nadler, J.V. (1978) Reactive synaptogenesis Deller, T., Frotscher, M. and Nitsch, R. (1996a) Sprouting of
in the hippocampus. In: Cotman C.W. (Ed.), Neuronal Plas- crossed entorhinodentate fibers after a unilateral entorhinal
ticity. Raven Press, New York, pp. 227–271. lesion: anterograde tracing of fiber reorganization with
Dasheiff, R.M. and McNamara, J.O. (1982) Electrolytic Phaseolus vulgaris-leucoagglutinin (PHAL). J. Comp. Ne-
entorhinal cortex lesions cause seizures. Brain Res., 231: urol., 365: 42–55.
444–450. Deller, T., Haas, C.A., Deissenrieder, K., Del Turco, D., Coulin,
Dawson, M.R., Levine, J.M. and Reynolds, R. (2000) NG2- C., Gebhardt, C., Drakew, A., Schwarz, K., Mundel, P. and
expressing cells in the central nervous system: are they Frotscher, M. (2002) Laminar distribution of synaptopodin in
oligodendroglial progenitors? J. Neurosci. Res., 61: normal and reeler mouse brain depends on the position of
471–479. spine-bearing neurons. J. Comp. Neurol., 453: 33–44.
Dawson, M.R., Polito, A., Levine, J.M. and Reynolds, R. Deller, T., Haas, C.A. and Frotscher, M. (2000) Reorganization
(2003) NG2-expressing glial progenitor cells: an abundant of the rat fascia dentata after a unilateral entorhinal cortex
and widespread population of cycling cells in the adult rat lesion: Role of the extracellular matrix. Ann. N.Y. Acad. Sci.,
CNS. Mol. Cell Neurosci., 24: 476–488. 911: 207–220.
522

Deller, T., Haas, C.A. and Frotscher, M. (2001) Sprouting in Evidence for normal aging of the septo-hippocampal
the hippocampus after entorhinal cortex lesion is layer-spe- cholinergic system in apoE ( / ) mice but impaired clear-
cific but not translaminar: which molecules may be involved? ance of axonal degeneration products following injury.
Restor. Neurol. Neurosci., 19: 159–167. Exp. Neurol., 151: 314–325.
Deller, T., Haas, C.A., Naumann, T., Joester, A., Faissner, A. Fagan, A.M., Suhr, S.T., Lucidi-Phillipi, C.A., Peterson, D.A.,
and Frotscher, M. (1997) Upregulation of astrocyte-derived Holtzman, D.M. and Gage, F.H. (1997) Endogenous FGF-2
tenascin-c correlates with neurite outgrowth in the rat dentate is important for cholinergic sprouting in the denervated hip-
gyrus after unilateral entorhinal cortex lesion. Neuroscience, pocampus. J. Neurosci., 17: 2499–2511.
81: 829–846. Faillace, M.P., Zwiller, J., Di Scala, G. and Bernabeu, R. (2006)
Deller, T., Nitsch, R. and Frotscher, M. (1995b) Phaseolus vul- Odor increases [3H]phorbol dibutyrate binding to protein
garis-leucoagglutinin (PHAL) tracing of commissural fibers kinase C in olfactory structures of rat brain. Effect of
to the rat dentate gyrus: Evidence for a previously unknown entorhinal cortex lesion. Brain Res., 1068: 16–22.
commissural projection to the outer molecular layer. J. Faissner, A. and Steindler, D. (1995) Boundaries and inhibitory
Comp. Neurol., 352: 55–68. molecules in developing neural tissues. Glia, 13: 233–254.
Deller, T., Nitsch, R. and Frotscher, M. (1996b) Heterogeneity Fife, B.T., Huffnagle, G.B., Kuziel, W.A. and Karpus, W.J.
of the commissural projection to the rat dentate gyrus: a (2000) CC chemokine receptor 2 is critical for induction of
Phaseolus vulgaris-leucoagglutinin tracing study. Neurosci- experimental autoimmune encephalomyelitis. J. Exp. Med.,
ence, 75: 111–121. 192: 899–905.
Deller, T., Nitsch, R. and Frotscher, M. (1996c) Layer-specific Finsen, B., Jensen, M.B., Lomholt, N.D., Hegelund, I.V., Po-
sprouting of commissural fibers to the rat fascia dentata after ulsen, F.R. and Owens, T. (1999) Axotomy-induced glial re-
unilateral entorhinal cortex lesion: a Phaseolus vulgaris-le- actions in normal and cytokine transgenic mice. Adv. Exp.
ucoagglutinin tracing study. Neuroscience, 71: 651–660. Med. Biol., 468: 157–171.
Diekmann, S., Ohm, T.G. and Nitsch, R. (1996) Long-lasting Frey, D., Laux, T., Xu, L., Schneider, C. and Caroni, P. (2000)
transneuronal changes in rat dentate granule cell dendrites Shared and unique roles of CAP23 and GAP43 in actin reg-
after entorhinal cortex lesion-a combined intracellular injec- ulation, neurite outgrowth, and anatomical plasticity. J. Cell
tion and electron microscopy study. Brain Pathol., 6: Biol., 149: 1443–1454.
205–214. Frotscher, M. (1988) Neuronal elements in the hippocampus
Drakew, A., Deller, T., Heimrich, B., Gebhardt, C., Del Turco, and their synaptic connections. Adv. Anat. Embryol., 111:
D., Tielsch, A., Forster, E., Herz, J. and Frotscher, M. (2002) 2–17.
Dentate granule cells in reeler mutants and VLDLR and Frotscher, M., Heimrich, B. and Deller, T. (1997) Sprouting in
ApoER2 knockout mice. Exp. Neurol., 176: 12–24. the hippocampus is layer-specific. Trends Neurosci., 20:
Drojdahl, N., Fenger, C., Nielsen, H.H., Owens, T. and Finsen, 218–223.
B. (2004) Dynamics of oligodendrocyte responses to antero- Fujise, N., Liu, Y., Hori, N. and Kosaka, T. (1998) Distribu-
grade axonal (Wallerian) and terminal degeneration in nor- tion of calretinin immunoreactivity in the mouse dentate
mal and TNF-transgenic mice. J. Neurosci. Res., 75: gyrus: II. Mossy cells, with special reference to their dorso-
203–217. ventral difference in calretinin immunoreactivity. Neurosci-
Drojdahl, N., Hegelund, I.V., Poulsen, F.R., Wree, A. and ence, 82: 181–200.
Finsen, B. (2002) Perforant path lesioning induces sprouting Gage, F.H., Olejniczak, P. and Armstrong, D.M. (1988) As-
of CA3-associated fibre systems in mouse hippocampal for- trocytes are important for sprouting in the septohippocampal
mation. Exp. Brain Res., 144: 79–87. circuit. Exp. Neurol., 102: 2–13.
Eckenstein, F. and Sofroniew, M.V. (1983) Identification of cen- Gall, C., Ivy, G. and Lynch, G. (1986) Neuroanatomical plas-
tral cholinergic neurons containing both choline acetytransf- ticity. Its role in organizing and reorganizing the central
erase and acetylcholinesterase and of central neurons containing nervous system. Hum. Growth, 2: 411–436.
only acetylcholinesterase. J. Neurosci., 3: 2286–2291. Gall, C. and Lynch, G. (1978) Rapid axon sprouting in the
Eyupoglu, I.Y., Bechmann, I. and Nitsch, R. (2003) Modifica- neonatal rat hippocampus. Brain Res., 153: 357–362.
tion of microglia function protects from lesion-induced neu- Gall, C. and Lynch, G. (1981) Fiber architecture of the dentate
ronal alterations and promotes sprouting in the gyrus following ablation of the entorhinal cortex in rats of
hippocampus. FASEB J., 17: 1110–1111. different ages: evidence for two forms of axon sprouting in
Fagan, A.M. and Gage, F.H. (1990) Cholinergic sprouting in the immature brain. Neuroscience, 6: 903–910.
the hippocampus: a proposed role for il-1. Exp. Neurol., 110: Gall, C. and Lynch, G.S. (1980) The regulation of fiber growth
105–120. and synaptogenesis in the developing hippocampus. Curr.
Fagan, A.M. and Gage, F.H. (1994) Mechanismus of sprouting Top. Dev. Biol., 1: 159–180.
in the adult central nervous system: cellular responses in ar- Gall, C., McWilliams, R. and Lynch, G. (1979a) The effect of
eas of terminal degeneration and reinnervation in the rat collateral sprouting on the density of innervation of normal
hippocampus. Neuroscience, 58: 705–725. target sites: implications for theories on the regulation of the
Fagan, A.M., Murphy, B.A., Patel, S.N., Kilbridge, J.F., size of the developing synaptic domains. Brain Res., 175:
Mobley, W.C., Bu, G. and Holtzman, D.M. (1998) 37–47.
523

Gall, C., Rose, G. and Lynch, G. (1979b) Proliferative and Hardman, R., Evans, D.J., Fellows, L., Hayes, B., Rupniak,
migratory activity of glial cells in the partially deafferented H.T., Barnes, J.C. and Higgins, G.A. (1997) Evidence for
hippocampus. J. Comp. Neurol., 183: 539–550. recovery of spatial learning following entorhinal cortex le-
Gebhardt, C., Del Turco, D., Drakew, A., Tielsch, A., Herz, J., sions in mice. Brain Res., 758: 187–200.
Frotscher, M. and Deller, T. (2002) Abnormal positioning of Harris, E.W., Lasher, S.S. and Steward, O. (1978) Habitutation
granule cells alters afferent fiber distribution in the mouse like decrements in transmission along the normal and lesion-
fascia dentata: morphologic evidence from reeler, apolipo- induced temporodentate pathways in the rat. Brain Res., 151:
protein E receptor 2-, and very low density lipoprotein 623–631.
receptor knockout mice. J. Comp. Neurol., 445: 278–292. Heckers, S., Ohtake, T., Wiley, R.G., Lappi, D.A., Geula, C.
Gehrmann, J., Schoen, S.W. and Kreutzberg, G.W. (1991) and Mesulam, M.M. (1994) Complete and selective choli-
Lesion of the rat entorhinal cortex leads to a rapid microglial nergic denervation of rat neocortex and hippocampus but not
reaction in the dentate gyrus. Acta Neuropathol., 82: amygdala by an immunotoxin against the p75 NGF receptor.
442–455. J. Neurosci., 14: 1271–1289.
Gold, R., Linington, C. and Lassmann, H. (2006) Understand- Henderson, Z., Harrison, P.S., Jagger, E. and Beeby, J.H.
ing pathogenesis and therapy of multiple sclerosis via animal (1998) Density of choline acetyltransferase-immunreactive
models: 70 years of merits and culprits in experimental terminals in the rat dentate gyrus after entorhinal cortex le-
autoimmune encephalomyelitis research. Brain, 129: sions: a quantitative light microscope study. Exp. Neurol.,
1953–1971. 152: 50–63.
Goldmann, J., Kwidzinski, E., Brandt, C., Mahlo, J., Richter, Hesselgesser, J. and Horuk, R. (1999) Chemokine and chemo-
D. and Bechmann, I. (2006) T cells traffic from brain to kine receptor expression in the central nervous system. J.
cervical lymph nodes via the cribroid plate and the nasal Neurovirol., 5: 13–26.
mucosa. J. Leukoc. Biol., 80: 797–801. Hjorth-Simonsen, A. and Jeune, B. (1972) Origin and termina-
Goldowitz, D., White, W.F., Steward, O., Cotman, C.W. and tion of the hippocampal perforant path in the rat studied by
Lynch, G. (1975) Anatomical evidence for a projection from silver impregnation. J. Comp. Neurol., 144: 215–232.
the entorhinal cortex to the contralateral dentate gyrus of the Hoff, S.F., Scheff, S.W., Kwan, A.Y. and Cotman, C.W. (1981)
rat. Exp. Neurol., 47: 433–441. A new type of lesion-induced synaptogenesis: I. Synaptic
Gomez-Pinilla, F., Lee, J.W. and Cotman, C.W. (1992) Basic turnover in non-denervated zones of the dentate gyrus in
FGF in adult rat brain: cellular distribution and response to young adult rats. Brain Res., 222: 1–13.
entorhinal lesion and fimbria-fornix transection. J. Neurosci., Höke, A. and Silver, J. (1996) Proteoglycans and other repul-
12: 345–355. sive molecules in glial boundaries during development and
van Groen, T., Kadish, I. and Wyss, J.M. (2002) Species differ- regeneration of the nervous system. Prog. Brain Res., 108:
ences in the projections from the entorhinal cortex to the 149–163.
hippocampus. Brain Res. Bull., 57: 553–556. Huang, D.R., Wang, J., Kivisakk, P., Rollins, B.J. and Ran-
van Groen, T., Miettinen, P. and Kadish, I. (2003) The ento- sohoff, R.M. (2001) Absence of monocyte chemoattractant
rhinal cortex of the mouse: organization of the projection to protein 1 in mice leads to decreased local macrophage
the hippocampal formation. Hippocampus, 13: 133–149. recruitment and antigen-specific T helper cell type 1 immune
Guthrie, K.M., Nguyen, T. and Gall, C.M. (1995) Insulin-like response in experimental autoimmune encephalomyelitis. J.
growth factor-1 mRNA is increased in deafferented hippo- Exp. Med., 193: 713–726.
campus: spatiotemporal correspondence of a trophic event Izikson, L., Klein, R.S., Charo, I.F., Weiner, H.L. and Luster,
with axon sprouting. J. Comp. Neurol., 352: 147–160. A.D. (2000) Resistance to experimental autoimmune encep-
Guthrie, K.M., Woods, A.G., Nguyen, T. and Gall, C.M. halomyelitis in mice lacking the CC chemokine receptor
(1997) Astroglial ciliary neurotrophic factor mRNA expres- (CCR)2. J. Exp. Med., 192: 1075–1080.
sion is increased in fields of axonal sprouting in deafferented Jakab, R.L. and Leranth, C. (1995) Septum. In: Paxinos G.
hippocampus. J. Comp. Neurol., 386: 137–148. (Ed.), The Rat Nervous System (2nd ed.). Academic Press,
Gwag, B.J., Sessler, F.M., Kimmerer, K. and Springer, J.E. San Diego, CA, pp. 405–442.
(1994) Neurotrophic factor mRNA expression in the dentate Jensen, M.B., Finsen, B. and Zimmer, J. (1997) Morphological
gyrus is increased following angular bundle transection. and immunophenotypic microglial changes in the denervated
Brain Res., 647: 23–29. fascia dentata of adult rats: correlation with blood-brain
Haas, C.A., Rauch, U., Thon, N., Merten, T. and Deller, T. barrier damage and astroglial reactions. Exp. Neurol., 143:
(1999) Entorhinal cortex lesion in adult rats induce the 103–116.
expression of the neuronal chondroitin sulfate proteoglycan Jensen, M.B., Gonzàlez, B., Castellano, B. and Zimmer, J.
neurocan in reactive astrocytes. J. Neurosci., 19: (1994) Microglial and astroglial reactions to anterograde ax-
9953–9963. onal degeneration: a histochemical and immuncytochemical
Hailer, N.P., Bechmann, I., Heizmann, S. and Nitsch, R. (1997) study of the adult rat fascia dentata after entorhinal perfo-
Adhesion molecule expression on phagocytic microglial cells rant path lesions. Exp. Brain Res., 98: 245–260.
following anterograde degeneration of perforant path axons. Jensen, M.B., Hegelund, I.V., Lomholt, N.D., Finsen, B. and
Hippocampus, 7: 341–349. Owens, T. (2000a) IFNgamma enhances microglial reactions
524

to hippocampal axonal degeneration. J. Neurosci., 20: H., Wekerle, H. and Hohlfeld, R. (1999) Activated human T
3612–3621. cells, B cells, and monocytes produce brain-derived neurotro-
Jensen, M.B., Hegelund, I.V., Poulsen, F.R., Owens, T., phic factor in vitro and in inflammatory brain lesions: a ne-
Zimmer, J. and Finsen, B. (1999) Microglial reactivity cor- uroprotective role of inflammation? J. Exp. Med., 189:
relates to the density and the myelination of the anterograd- 865–870.
ely degenerating axons and terminals following perforant Kipnis, J., Yoles, E., Schori, H., Hauben, E., Shaked, I. and
path denervation of the mouse fascia dentata. Neuroscience, Schwartz, M. (2001) Neuronal survival after CNS insult is
93: 507–518. determined by a genetically encoded autoimmune response. J.
Jensen, M.B., Poulsen, F.R. and Finsen, B. (2000b) Axonal Neurosci., 21: 4564–4571.
sprouting regulates myelin basic protein gene expression in Kosaka, T., Katsumaru, H., Hama, K., Wu, J.Y. and Heiz-
denervated mouse hippocampus. Int. J. Dev. Biol., 18: mann, C.W. (1987) GABAergic neurons containing CA2+ -
221–235. binding protein parvalbumin in the rat hippocampus and
de Jong, E.K., Dijkstra, I.M., Hensens, M., Brouwer, N., van dentate gyrus. Brain Res., 419: 119–130.
Amerongen, M., Liem, R.S., Boddeke, H.W. and Biber, K. Kovac, A.D., Kwidzinski, E., Heimrich, B., Bittigau, P., Deller,
(2005) Vesicle-mediated transport and release of CCL21 in T., Nitsch, R. and Bechmann, I. (2004) Entorhinal cortex
endangered neurons: a possible explanation for microglia lesion in the mouse induces transsynaptic death of perforant
activation remote from a primary lesion. J. Neurosci., 25: path target neurons. Brain Pathol., 14: 249–257.
7548–7557. Ladeby, R., Wirenfeldt, M., Dalmau, I., Gregersen, R.,
Jucker, M., D‘Amato, F., Mondadori, C., Mohajeri, H., Ma- Garcia-Ovejero, D., Babcock, A., Owens, T. and Finsen, B.
gyar, J., Bartsch, U. and Schachner, M. (1996) Expression of (2005a) Proliferating resident microglia express the stem
the neural adhesion molecule L1 in the deafferented dentate cell antigen CD34 in response to acute neural injury. Glia,
gyrus. Neuroscience, 75: 703–715. 50: 121–131.
Kadish, I. and van Groen, T. (2002) Low levels of estrogen Ladeby, R., Wirenfeldt, M., Garcia-Ovejero, D., Fenger, C.,
significantly diminish axonal sprouting after entorhinal cor- Dissing-Olesen, L., Dalmau, I. and Finsen, B. (2005b) Mi-
tex lesions in the mouse. J. Neurosci., 22: 4095–4102. croglial cell population dynamics in the injured adult central
Kadish, I. and van Groen, T. (2003) Differences in lesion- nervous system. Brain Res. Brain Res. Rev., 48: 196–206.
induced hippocampal plasticity between mice and rats. Laux, T., Fukami, K., Thelen, M., Golub, T., Frey, D. and
Neuroscience, 11: 499–509. Caroni, P. (2000) GAP43, MARCKS, and CAP23 modulate
Kadish, I., Pradier, L. and van Groen, T. (2002) Transgenic PI(4,5)P(2) at plasmalemmal rafts, and regulate cell cortex
mice expressing the human presenilin 1 gene demonstrate actin dynamics through a common mechanism. J. Cell Biol.,
enhanced hippocampal reorganization following entorhinal 149: 1455–1472.
cortex lesion. Brain Res. Bull., 57: 587–594. Lee, M.Y., Deller, T., Kirsch, M., Frotscher, M. and Hofmann,
Kapfhammer, J.P. and Schwab, M.E. (1992) Modulators of H.D. (1997) Differential regulation of CNTF and CNTF
neuronal migration and neurite growth. Curr. Opin. Cell receptor alpha expression in astrocytes and neurons of the
Biol., 4: 863–868. fascia dentata following entorhinal cortex lesion. J. Neuro-
Karpus, W.J. and Ransohoff, R.M. (1998) Chemokine regula- sci., 17: 1137–1146.
tion of experimental autoimmune encephalomyelitis: tempo- Levey, A.I., Wainer, B.H., Mufson, E.J. and Mesulam, M.M.
ral and spatial expression patterns govern disease (1983) Co-localization of acetylcholinesterase and choline
pathogenesis. J. Immunol., 161: 2667–2671. acetyltransferase in the rat cerebrum. Neuroscience, 9: 9–22.
Kawaja, M.D. and Gage, F.H. (1991) Reactive astrocytes are Levey, A.I., Wainer, B.H., Rye, D.B., Mufson, E.J. and Me-
substrates for the growth of adult CNS axons in the presence sulam, M.M. (1984) Choline acetyltransferase-immunoreac-
of elevated levels of nerve growth factor. Neuron, 7: tive neurons intrinsic to rodent cortex and distinction from
1019–1030. acetylcholinesterase-positive neurons. Neuroscience, 13:
Kelley, M.S. and Steward, O. (1996a) The process of reinner- 341–353.
vation in the dentate gyrus of adult rats: physiological events Linke, R., Schwegler, H. and Boldyreva, M. (1994) Cholinergic
at the time of the lesion and during the early postlesion and GABAergic septo-hippocampal projection neurons in
period. Exp. Neurol., 139: 73–82. mice: a retrograde tracing study combined with double
Kelley, M.S. and Steward, O. (1996b) The role of postlesion immunocytochemistry for choline acetyltransferase and parv-
seizures and spreading depression in the upregulation of glial albumin. Brain Res., 653: 73–80.
fibrillary acidic protein mRNA after entorhinal cortex Liu, Y., Fujise, N. and Kosaka, T. (1996) Distribution of ca-
lesions. Exp. Neurol., 139: 83–94. lretinin immunoreactivity in the mouse dentate gyrus. I.
Kelley, M.S. and Steward, O. (1998) Injury-induced physiolog- General description. Exp. Brain Res., 108: 389–403.
ical events that may modulate gene expression in neurons and Loesche, J. and Steward, O. (1977) Behavioral correlates of
glia. Rev. Neurosci., 8: 147–177. denervation and reinnervation of the hippocampal formation
Kerschensteiner, M., Gallmeier, E., Behrens, L., Leal, V.V., of the rat: recovery of alternation performance following
Misgeld, T., Klinkert, W.E., Kolbeck, R., Hoppe, E., Or- unilateral entorhinal cortex lesions. Brain Res. Bull., 2:
opeza-Wekerle, R.L., Bartke, I., Stadelmann, C., Lassmann, 31–39.
525

Luster, A.D. (1998) Chemokines–chemotactic cytokines that Mingorance, A., Fontana, X., Soriano, E. and del Rio, J.A.
mediate inflammation. N. Engl. J. Med., 338: 436–445. (2005) Overexpression of myelin-associated glycoprotein af-
Lynch, G. and Cotman, C. (1975) The hippocampus as a model ter axotomy of the perforant pathway. Mol. Cell Neurosci.,
for studying anatomical plasticity in the adult brain. In: Is- 29: 471–483.
aacson R.L. and Pribram K.H. (Eds.), The Hippocampus, Miwa, C. and Ueki, A. (1996) Effects of entorhinal cortex le-
Vol. 1. Plenum Press, New York, pp. 123–154. sion on learning behavior and on hippocampus in the rat.
Lynch, G., Gall, C., Rose, G. and Cotman, C.W. (1976) Psychiatry Clin. Neurosci., 50: 223–230.
Changes in the distribution of the dentate gyrus associational Mutlu, L., Brandt, C., Kwidzinski, E., Sawitzki, B., Gimsa, U.,
system following unilateral or bilateral entorhinal lesion in Mahlo, J., Aktas, O., Nitsch, R., van Zwam, M., Laman, J.
the adult rat. Brain Res., 110: 57–71. and Bechmann, I. (2007) Tolerogenic effect of fiber tract in-
Lynch, G., Matthews, D.A., Mosko, S., Parks, T. and Cotman, jury: Peripheral apoptosis of T cells and reduction of EAE
C.W. (1972) Induced acetylcholinesterase-rich layer in rat severity following entorhinal cortex lesion. Exp. Brain Res.,
dentate gyrus following entorhinal lesions. Brain Res., 42: 178: 542–553.
311–318. Nadler, J.V., Cotman, C., Paoletti, C. and Lynch, G.S. (1977a)
Lynch, G., Stanfield, B. and Cotman, C.W. (1973) Develop- Histochemical evidence of altered development of cholinergic
mental differences in postlesion axonal growth in the hippo- fibers in the rat dentate gyrus following lesions. II. Effects of
campus. Brain Res., 59: 155–168. partial entorhinal and simultaneous multiple lesions. J.
Lynch, G.S., Stanfield, B., Parks, T. and Cotman, C.W. (1974) Comp. Neurol., 171: 589–604.
Evidence for selective post-lesion axonal growth in the dent- Nadler, J.V., Cotman, C.W. and Lynch, G.S. (1977b) Histo-
ate gyrus of the rat. Brain Res., 69: 1–11. chemical evidence of altered development of cholinergic fib-
Mackay, C.R. (2001) Chemokines: immunology’s high impact ers in the rat dentate gyrus following lesions. I. Time course
factors. Nat. Immunol., 2: 95–101. after complete unilateral entorhinal lesion at various ages. J.
Margolis, R.U. and Margolis, R.K. (1997) Chondroitin sulfate Comp. Neurol., 171: 561–587.
proteoglycans as mediators of axon growth and pathfinding. Nadler, J.V. and Evenson, D.A. (1982) Autoradiographic ev-
Cell Tissue Res., 290: 343–348. idence that septohippocampal fibers reinnervate fascia den-
Marrone, D.F., LeBoutillier, J.C. and Petit, T.L. (2004a) tata denervated by entorhinal lesion during development.
Changes in synaptic ultrastructure during reactive syna- Anat. Embryol., 165: 113–123.
ptogenesis in the rat dentate gyrus. Brain Res., 1005: Nadler, J.V., Perry, B.W. and Cotman, C.W. (1980) Interaction
124–136. with CA4-derived fibers accounts for distribution of septo-
Marrone, D.F., LeBoutillier, J.C. and Petit, T.L. (2004b) Com- hippocampal fibers in rat fascia denata after entorhinal le-
parative analyses of synaptic densities during reactive syna- sion. Exp. Neurol., 68: 185–194.
ptogenesis in the rat dentate gyrus. Brain Res., 996: 19–30. Nafstad, P.H.J. (1967) An electron microscope study on the
Matthews, D.A., Cotman, C.W. and Lynch, G. (1976a) An termination of the perforant path fibres in the hippocampus
electron microscopic study of lesion-induced synaptogenesis and fascia denata. Z. Zellforsch. Mikrosk. Anat., 76:
in the dentate gyrus of the adult rat. I. Magnitude and time 532–542.
course of degeneration. Brain Res., 115: 1–21. Naumann, T., Deller, T., Bender, R. and Frotscher, M. (1997)
Matthews, D.A., Cotman, C.W. and Lynch, G. (1976b) An 192 IgG-saporin-induced loss of cholinergic neurons in
electron microscopic study of lesion-induced synaptogenesis the septum abolishes cholinergic sprouting after unilat-
in the dentate gyrus of the adult rat. II. Reappearance of eral entorhinal lesion in the rat. Eur. J. Neurosci., 9:
morphologically normal synaptic contacts. Brain Res., 115: 1304–1313.
23–41. Nielsen, H.H., Ladeby, R., Drojdahl, N., Peterson, A.C. and
Mayer, J., Hamel, M.G. and Gottschall, P.E. (2005) Evidence Finsen, B. (2006) Axonal degeneration stimulates the forma-
for proteolytic cleavage of brevican by the ADAMTSs in the tion of NG2+ cells and oligodendrocytes in the mouse. Glia,
dentate gyrus after excitotoxic lesion of the mouse entorhinal 54: 105–115.
cortex. BMC Neurosci., 6: 52. Nitsch, R., Bader, S. and Frotscher, M. (1992) Reorganization
McKeon, R.J., Vietje, B.P. and Wells, J. (1989) Increase in of input synapses of parvalbumin-containing neurons in the
acetylcholinesterase in the molecular layer of the denate rat fascia dentata following entorhinal lesion. Neurosci.
gyrus in the absence of septal inputs following selective gran- Lett., 135: 33–36.
ule cell lesions. Brain Res., 503: 317–321. Nitsch, R. and Frotscher, M. (1991) Maintenance of peripheral
Meier, S., Brauer, A.U., Heimrich, B., Nitsch, R. and Sa- dendrites of GABAergic neurons requires specific input.
vaskan, N.E. (2004) Myelination in the hippocampus during Brain Res., 554: 304–307.
development and following lesion. Cell Mol. Life Sci., 61: Nitsch, R. and Frotscher, M. (1992) Reduction of posttrau-
1082–1094. matic transneuronal ‘‘early gene’’ activation and dendritic
Meier, S., Brauer, A.U., Heimrich, B., Schwab, M.E., Nitsch, atrophy by the N-methyl-D-aspartate receptor antagonist
R. and Savaskan, N.E. (2003) Molecular analysis of Nogo MK-801. Proc. Natl. Acad. Sci. U.S.A., 89: 5197–5200.
expression in the hippocampus during development and fol- Nitsch, R. and Frotscher, M. (1993) Transneuronal changes in
lowing lesion and seizure. FASEB J., 17: 1153–1155. dendrites of GABAergic parvalbumin-containing neurons of
526

the rat fascia dentata following entorhinal lesion. Hippo- R. and Kettenmann, H. (2004) CXCR3-dependent microglial
campus, 3: 481–490. recruitment is essential for dendrite loss after brain lesion. J.
Nitsch, R., Soriano, E. and Frotscher, M. (1990) The parval- Neurosci., 24: 8500–8509.
bumin-containing nonpyramidal neurons in the rat hippo- Rappert, A., Biber, K., Nolte, C., Lipp, M., Schubel, A., Lu, B.,
campus. Anat. Embryol., 181: 413–425. Gerard, N.P., Gerard, C., Boddeke, H.W. and Kettenmann,
Nyakas, C., Luiten, P.G.M., Balkan, B. and Spencer, D.G.J. H. (2002) Secondary lymphoid tissue chemokine (CCL21)
(1988) Changes in septohippocampal projections after lateral activates CXCR3 to trigger a Cl current and chemotaxis in
entorhinal or combined entorhinal-raphè lesions as studied murine microglia. J. Immunol., 168: 3221–3226.
by anterograde tracing methods. Brain Res. Bull., 21: Reeves, T.M. and Steward, O. (1986) Emergence of the capacity
285–293. for LTP during reinnervation of the dentate gyrus: evidence
Nyakas, C., Luiten, P.G.M., Spencer, D.G. and Traber, J. that abnormally shaped spines can mediate LTP. Exp. Brain
(1987) Detailed projection patterns of septal and diagonal Res., 65: 167–175.
band efferents to the hippocampus in the rat with emphasis Reeves, T.M. and Steward, O. (1988) Changes in the firing
on innervation of CA1 and dentate gyrus. Brain Res. Bull., properties of neurons in the dentate gyrus with denervation
18: 533–545. and reinnervation: implications for behavioral recovery. Exp.
Otal, R., Burgaya, F., Frisen, J., Soriano, E. and Martinez, A. Neurol., 102: 37–49.
(2006) Ephrin-A5 modulates the topographic mapping and Rollins, B.J. (1997) Chemokines. Blood, 90: 909–928.
connectivity of commissural axons in murine hippocampus. Rose, G., Lynch, G. and Cotman, C.W. (1976) Hypertrophy
Neuroscience, 141: 109–121. and redistribution of astrocytes in the deafferented dentate
Parent, A., Dea, D., Quirion, R. and Poirier, J. (1993) gyrus. Brain Res. Bull., 1: 87–92.
[3H]phorbol ester binding sites and neuronal plasticity in Sattler, R., Charlton, M.P., Hafner, M. and Tymianski, M.
the hippocampus following entorhinal cortex lesions. Brain (1998) Distinct influx pathways, not calcium load, determine
Res., 607: 23–32. neuronal vulnerability to calcium neurotoxicity. J. Neuroc-
Parnavelas, J.G., Lynch, G., Brecha, N., Cotman, C.W. and hem., 71: 2349–2364.
Globus, A. (1974) Spine loss and regrowth in hippocampus Savaskan, N.E. and Nitsch, R. (2001) Molecules involved in
following deafferentation. Nature, 248: 71–73. reactive sprouting in the hippocampus. Rev. Neurosci., 12:
Pearlman, A.L. and Sheppard, A.M. (1996) Extracellular matrix 195–215.
in early cortical development. Prog. Brain Res., 108: 117–134. Schafer, M., Brauer, A.U., Savaskan, N.E., Rathjen, F.G. and
Pedersen, M.D., Minuzzi, L., Wirenfeldt, M., Meldgaard, M., Brummendorf, T. (2005) Neurotractin/kilon promotes neu-
Slidsborg, C., Cumming, P. and Finsen, B. (2006) Up-reg- rite outgrowth and is expressed on reactive astrocytes after
ulation of PK11195 binding in areas of axonal degeneration entorhinal cortex lesion. Mol. Cell Neurosci., 29: 580–590.
coincides with early microglial activation in mouse brain. Schauwecker, P.E. and Steward, O. (1997) Genetic influences
Eur. J. Neurosci., 24: 991–1000. on cellular reactions to brain injury: activation of microglia
Peterson, G.M. (1994) Sprouting of central noradrenergic in denervated neuropil in mice carrying a mutation (WLDs)
fibers in the dentate gyrus following combined lesions of its that causes delayed wallerian degeneration. J. Comp. Ne-
entorhinal and septal afferents. Hippocampus, 4: 635–648. urol., 380: 82–94.
Phinney, A.L., Calhoun, M.E., Woods, A.G., Deller, T. and Scheff, S.W. and Cotman, C.W. (1977) Recovery of spontane-
Jucker, M. (2004) Stereological analysis of the reorganization ous alternation following lesions of the entorhinal cortex in
of the dentate gyrus following entorhinal cortex lesion in adult rats: possible correlation to axon sprouting. Behav.
mice. Eur. J. Neurosci., 19: 1731–1740. Bio., 21: 286–293.
Poirier, J., May, P.C., Osterburg, H.H., Geddes, J., Cotman, Schenk, F. and Morris, R.G.M. (1985) Dissociation between
C.W. and Finch, C.E. (1990) Selective alterations of RNA in components of spatial memory in rats after recovery from the
rat hippocampus after entorhinal cortex lesioning. Proc. effects of retrohippocampal lesions. Exp. Brain Res., 58: 11–28.
Natl. Acad. Sci. U.S.A., 87: 303–307. Schwab, M.E., Kapfhammer, J.P. and Bandtlow, C.E. (1993)
Poulsen, F.R., Lagord, C., Courty, J., Pedersen, E.B., Bar- Inhibitors of neurite growth. Annu. Rev. Neurosci., 16:
ritault, D. and Finsen, B. (2000) Increased synthesis of he- 565–595.
parin affin regulatory peptide in the perforant path lesioned Shi, B. and Stanfield, B.B. (1996) Differential sprouting re-
mouse hippocampal formation. Exp. Brain Res., 135: sponses in axonal fiber systems in the dentate gyrus following
319–330. lesions of the perforant path in WLDs mutant mice. Brain
Ramirez, J.J. (2001) The role of axonal sprouting in functional Res., 740: 89–101.
reorganization after CNS injury: lessons from the hippocam- Siebert, H., Sachse, A., Kuziel, W.A., Maeda, N. and Bruck,
pal formation. Restor. Neurol. Neurosci., 19: 237–262. W. (2000) The chemokine receptor CCR2 is involved in ma-
Ramirez, J.J. and Stein, D.G. (1984) Sparing and recovery of crophage recruitment to the injured peripheral nervous sys-
spatial alternation performance after entorhinal cortex le- tem. J. Neuroimmunol., 110: 177–185.
sions in rats. Behav. Brain Res., 13: 53–61. Soriano, E. and Frotscher, M. (1994) Mossy cells of the rat
Rappert, A., Bechmann, I., Pivneva, T., Mahlo, J., Biber, K., fascia dentata are glutamate-immunoreactive. Hippocampus,
Nolte, C., Kovac, A.D., Gerard, C., Boddeke, H.W., Nitsch, 4: 65–70.
527

Soriano, E., del Rio, J.A., Martinez, A. and Super, H. (1994) Steward, O., Cotman, C.W. and Lynch, G.S. (1976) A quan-
Organization of the embryonic and early postnatal murine titative autoradiographic and electrophysiological study of
hippocampus. I. Immunocytochemical characterization of the reinnervation of the dentate gyrus by the contralateral
neuronal populations in the subplate and marginal zone. J. entorhinal cortex following ipsilateral lesions. Brain Res.,
Comp. Neurol., 342: 571–595. 114: 181–200.
Stanfield, B.B., Caviness, V.S.J. and Cowan, W.M. (1979) The Steward, O., Kelley, M.S. and Schauwecker, P.E. (1997) Signals
organization of certain afferents to the hippocampus and that regulate astroglial gene expression: induction of GFAP
dentate gyrus in normal and reeler mice. J. Comp. Neurol., mRNA following seizures or injury is blocked by protein
185: 461–483. synthesis inhibitors. Exp. Neurol., 148: 100–109.
Stanfield, B.B. and Cowan, W.M. (1979a) The development of Steward, O., Kelley, M.S. and Torre, E.R. (1993) The process
the hippocampus and dentate gyrus in normal and reeler of reinnervation in the dentate gyrus of adult rats: temporal
mice. J. Comp. Neurol., 185: 423–460. relationship between changes in the level of glial fibrillary
Stanfield, B.B. and Cowan, W.M. (1979b) The morphology of acidic protein (GFAP ) and GFAP mRNA in reactive as-
the hippocampus and dentate gyrus in normal and reeler trocytes. Exp. Neurol., 124: 167–183.
mice. J. Comp. Neurol., 185: 393–422. Steward, O. and Scoville, S.A. (1976) Cells of origin of
Stanfield, B.B. and Cowan, W.M. (1982) The sprouting of entorhinal cortical afferents to the hippocampus and fascia
septal afferents to the dentate gyrus after lesions of the dentata of the rat. J. Comp. Neurol., 169: 347–370.
entorhinal cortex in adult rats. Brain Res., 232: 162–170. Steward, O., Torre, E.R., Phillips, L. and Trimmer, P.A. (1990)
Stäubli, U., Gall, C. and Lynch, G. (1984) The distribution of The process of reinnervation in the dentate gyrus of adult
the commissural-associational afferents of the dentate gyrus rats: time course of increases in mRNA for glial fibrillary
after perforant path lesions in one-day-old rats. Brain Res., acidic protein. J. Neurosci., 10: 2373–2384.
292: 156–159. Steward, O. and Trimmer, P.A. (1997) Genetic influences on
Steward, O. (1976a) Reinnervation of the dentate gyrus by ho- cellular reactions to CNS injury: the reactive response of as-
mologous afferents following entorhinal cortical lesion in trocytes in denervated neuropil regions in mice carrying a
adult rats. Science, 194: 426–428. mutation (Wld(S)) that causes delayed Wallerian degenera-
Steward, O. (1976b) Topographic organization of the projec- tion. J. Comp. Neurol., 380: 70–81.
tions from the entorhinal area to the hippocampal formation Steward, O. and Vinsant, S.L. (1978) Identification of the cells
of the rat. J. Comp. Neurol., 167: 285–314. of origin of a central pathway which sprouts following lesions
Steward, O. (1980) Trajectory of contralateral entorhinal axons in mature rats. Brain Res., 147: 223–243.
which reinnervate the fascia dentata of the rat following Steward, O. and Vinsant, S.L. (1983) The process of reinner-
ipsilateral entorhinal lesions. Brain Res., 183: 277–289. vation in the dentate gyrus of the adult rat: a quantitative
Steward, O. (1991) Synapse replacement on cortical neurons electron microscopic analysis of terminal proliferation
following denervation. Cereb. Cortex, 9: 81–132. and reactive synaptogenesis. J. Comp. Neurol., 214: 370–386.
Steward, O. (1992) Signals that induce sprouting in the central Steward, O., Vinsant, S.L. and Davis, L. (1988) The process of
nervous system: sprouting is delayed in a strain of mouse reinnervation in the dentate gyrus of adlut rats: an ultra-
exhibiting delayed axonal degeneration. Exp. Neurol., 118: structural study of changes in presynaptic terminals as a re-
340–351. sult of sprouting. J. Comp. Neurol., 267: 203–210.
Steward, O. (1994a) Cholinergic sprouting is blocked by re- Stone, D.J., Rozovsky, I., Morgan, T.E., Anderson, C.P. and
peated induction of electroconvulsive seizures, a manipula- Finch, C.E. (1998) Increased synaptic sprouting in response to
tion that induces a persistent reactive state in astrocytes. Exp. estrogen via an apolipoprotein E-dependent mechanism: im-
Neurol., 129: 103–111. plications for Alzheimer’s disease. J. Neurosci., 18: 3180–3185.
Steward, O. (1994b) Reorganization of neuronal circuitry fol- Storm-Mathisen, J. (1974) Choline acetyltransferase and ace-
lowing central nervous system trauma: naturally occurring tylcholinesterase in fascia dentata following lesion of the
processes and opportunities for therapeutic intervention. In: entorhinal afferents. Brain Res., 80: 181–197.
Salzman S.K. and Faden A.I. (Eds.), The Neurobiology of Styren, S.D., Lagenaur, C.F., Miller, P.D. and DeKosky, S.T.
Central Nervous System Trauma. Oxford University Press, (1994) Rapid expression and transport of embryonic N-CAM
New York, pp. 266–287. in dentate gyrus following entorhinal cortex lesion: ultra-
Steward, O., Cotman, C.W. and Lynch, G. (1974) Growth of a structural analysis. J. Comp. Neurol., 349: 486–492.
new fiber projection in the brain of adult rats: reinnervation Styren, S.D., Miller, P.D., Lagenaur, C.F. and DeKosky, S.T.
of the dentate gyrus by the contralateral entorhinal cortex (1995) Alternate strategies in lesion-induced reactive syna-
following ipsilateral entorhinal lesion. Exp. Brain Res., 20: ptogenesis: differential expression of L1 in two populations
45–66. of sprouting axons. Exp. Neurol., 131: 165–173.
Steward, O., Cotman, C.W. and Lynch, G.S. (1973) Re-estab- Supèr, H. and Soriano, E. (1994) The organization of the em-
lishment of electrophysiologically functional entorhinal cor- bryonic and early postnatal murine hippocampus. II. Devel-
tical input to the dentate gyrus deafferented by ipsilateral opment of entorhinal, commissural, and septal connections
entorhinal lesions: innervation by the contralateral cortex. studied with the lipophilic tracer dil. J. Comp. Neurol., 344:
Exp. Brain Res., 18: 396–414. 101–120.
528

Thon, N., Haas, C.A., Rauch, U., Merten, T., Fässler, R., are severely attenuated in middle aged and aged rats. Ne-
Frotscher, M. and Deller, T. (2000) The chondroitin sulphate uroscience, 83: 663–668.
proteoglycan brevican is upregulated by astrocytes after Woods, A.G., Poulsen, F.R. and Gall, C.M. (1999) Dexa-
entorhinal cortex lesions in adult rats. Eur. J. Neurosci., 12: methasone specifically suppresses microglial trophic
2547–2558. responses to hippocampal deafferentation. Neuroscience,
Ueki, A., Miwa, C., Oohara, K. and Miyoshi, K. (1996) His- 91: 1277–1289.
tological evidence for cholinergic alteration in the hippo- Ying, G.X., Huang, C., Jiang, Z.H., Liu, X., Jing, N.H. and
campus following entorhinal cortex lesion. J. Neurol. Sci., Zhou, C.F. (2002) Up-regulation of cystatin C expression in
142: 7–11. the murine hippocampus following perforant path transec-
White, F., Nicoll, J.A. and Horsburgh, K. (2001a) Alterations tions. Neuroscience, 112: 289–298.
in ApoE and ApoJ in relation to degeneration and regener- Yoshida, K. and Gage, F.H. (1991) Fibroblast growth factors
ation in a mouse model of entorhinal cortex lesion. Exp. stimulate nerve growth factor synthesis and secretion by
Neurol., 169: 307–318. astrocytes. Brain Res., 538: 118–126.
White, F., Nicoll, J.A., Roses, A.D. and Horsburgh, K. (2001b) Zappone, C.A. and Sloviter, R.S. (2001) Commissurally pro-
Impaired neuronal plasticity in transgenic mice expressing jecting inhibitory interneurons of the rat hippocampal dent-
human apolipoprotein E4 compared to E3 in a model of ate gyrus: a colocalization study of neuronal markers and the
entorhinal cortex lesion. Neurobiol. Dis., 8: 611–625. retrograde tracer Fluoro-Gold. J. Comp. Neurol., 441:
White, W.F., Goldowitz, D., Lynch, G. and Cotman, C.W. 324–344.
(1976) Electrophysiological analysis of the projection from Zimmer, J. and Hjorth-Simonsen, A. (1975) Crossed pathways
the contralateral entorhinal cortex to the dentate gyrus in from the entorhinal area to the fascia dentata. II. Provokable
normal rats. Brain Res., 114: 201–209. in rats. J. Comp. Neurol., 161: 71–102.
Wiley, R.G., Oeltmann, T.N. and Lappi, D.A. (1991) Zimmer, J., Laurberg, S. and Sunde, N. (1986) Non-cholinergic
Immunolesioning: selective destruction of neurons using afferents determine the distribution of the cholinergic septo-
immunotoxin to rat NGF receptor. Brain Res., 562: 149–153. hippocampal projection: a study of the AChE staining pat-
Wirenfeldt, M., Dalmau, I. and Finsen, B. (2003) Estimation of tern in the rat fascia dentata and hippocampus after lesions,
absolute microglial cell numbers in mouse fascia dentata us- x-irradiation, and intracerebral grafting. Exp. Brain Res., 64:
ing unbiased and efficient stereological cell counting princi- 158–168.
ples. Glia, 44: 129–139. Zipp, F., Nitsch, R., Soriano, E. and Frotscher, M. (1989)
Woods, A.G., Guthrie, K.M., Kurlawalla, M.A. and Gall, Entorhinal fibers form synaptic contacts on parvalbumin-
C.M. (1998) Deafferentation-induced increases in hippocam- immunoreactive neurons in the rat fascia dentata. Brain Res.,
pal insulin-like growth factor-1 messenger RNA expression 495: 161–166.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 28

Adult neurogenesis in the intact and epileptic


dentate gyrus

Jack M. Parent

Department of Neurology, University of Michigan Medical Center, 109 Zina Pitcher Place, 5021 BSRB,
Ann Arbor, MI 48109-2200, USA

Abstract: Neurogenesis persists throughout life in the adult mammalian dentate gyrus. Adult-born dentate
granule cells integrate into existing hippocampal circuitry and may provide network plasticity necessary for
certain forms of hippocampus-dependent learning and memory. Neural stem cells and neurogenesis in the
adult dentate gyrus are regulated by a variety of environmental, physiological, and molecular factors. These
include aging, stress, exercise, neurovascular components of the stem cell niche, growth factors, neuro-
transmitters, and hormones. Seizure activity also influences dentate granule cell neurogenesis. Production
of adult-born neurons increases in rodent models of temporal lobe epilepsy, and both newborn and
pre-existing granule neurons contribute to aberrant axonal reorganization in the epileptic hippocampus.
Prolonged seizures also disrupt the migration of dentate granule cell progenitors and lead to hilar-ectopic
granule cells. The ectopic granule neurons appear to integrate abnormally and contribute to network
hyperexcitability. Similar findings of granule cell layer dispersion and ectopic granule neurons in
human TLE suggest that aberrant neurogenesis contributes to epileptogenesis or learning and memory
disturbances in this epilepsy syndrome.

Keywords: neurogenesis; stem cell; epileptogenesis; temporal lobe epilepsy; migration; dentate granule cell

Dentate granule cell neurogenesis persists (DGCs). Persistent DGC neurogenesis occurs in
throughout life all mammalian species examined to date, including
human and nonhuman primates (Eriksson et al.,
New neurons are generated throughout life in the 1998; Gould et al., 1998; Kornack and Rakic,
mammalian dentate gyrus (Altman and Das, 1965; 1999). Although the majority of DGCs in the rat
Kaplan and Hinds, 1977; Cameron et al., 1993; are produced near the end of the first postnatal
Kuhn et al., 1996). This process, termed neuro- week, new DGCs continue to be generated at a
genesis, begins with neural stem or progenitor cells lower rate throughout adulthood and into senes-
that give rise to transit amplifying precursors to cence (Schlessinger et al., 1975; Kuhn et al., 1996)
expand the numbers of progeny. In the dentate (Fig. 1). The human dentate gyrus, for example,
gyrus, these progeny form dentate granule cells generates neurons well into the seventh decade of
life (Eriksson et al., 1998; Kempermann et al.,
2004). Estimates suggest that approximately 9000
Corresponding author. Tel.: +1 734 763 3776; DGCs are added daily to the dentate gyrus of
Fax: +1 734 763 7686; E-mail: parent@umich.edu young adult rats, and 6% of the granule cell

DOI: 10.1016/S0079-6123(07)63028-3 529


530

Fig. 1. Neonatal and adult dentate granule cell neurogenesis. (A, B) Bromodeoxyuridine (BrdU) immunostaining of coronal sections
through the dentate gyrus of 10 day (A) and 10 week (B) old rats shows proliferating granule cell progenitors concentrated at the inner
part of the granule cell layer at the hilar (h) border. Dentate granule cell neurogenesis peaks at the end of the 1st postnatal week (A) but
persists into adulthood (B). BrdU was injected 6 days earlier for both. (C) Retroviral nuclear localization signal-b-galactosidase
reporter labeling of proliferating cell clusters in the dentate subgranular zone (arrows) of an adult rat. Inset shows the right cluster at
higher magnification. Retrovirus was injected 3 days earlier. (D) Doublecortin immunolabeling of immature neurons (arrows) in the
adult rat, with cell bodies close to the SGZ but processes throughout the granule cell and molecular layer gcl, granule cell layers.

layer is thought to consist of recently born neurons Integration and function of adult-born DGCs
(Cameron and McKay, 2001).
Dentate gyrus neural stem cells reside in the Increasing attention has focused on the potential
subgranular zone (SGZ) at the border of the hilus for adult-generated DGCs to integrate into exist-
and DGC layer. The stem cells are a subtype of ing neural circuitry. Combined retrograde tracer
radial glia-like astrocyte that expresses glial fibril- and mitotic labeling studies using tritiated thymi-
lary acidic protein (GFAP) and nestin (Seri et al., dine or its analogue, bromodeoxyuridine (BrdU),
2001; Filippov et al., 2003). They proliferate and in adult rodent indicate that mossy fibers of
give rise to clusters of transit amplifying precur- newborn DGCs project to appropriate targets in
sors that divide further to generate neuroblasts hippocampal area CA3 (Stanfield and Trice, 1988;
(Fig. 1). The SGZ precursors and neuroblasts are Hastings and Gould, 1999; Markakis and Gage,
characterized by a sequence of marker expression 1999). Acute hippocampal slice recordings from
as they mature (reviewed in Kempermann et al., rodents in which adult-generated DGCs were labe-
2004). The progenitors express doublecortin led with retroviral reporters or by transgenic
(DCX) and polysialylated neural cell adhesion manipulations show that the cells integrate into
molecule (PSA-NCAM) during cell division and as hippocampal circuitry and acquire some electro-
early postmitotic neurons (Fig. 1D). With further physiological characteristics of mature DGCs
differentiation, postmitotic neurons continue to (Song et al., 2002b; van Praag et al., 2002; Wadi-
expresss DCX and PSA-NCAM, but also begin to che et al., 2005; Ge et al., 2006). The newborn
express calretinin and Prox1; subsequently, they neurons appear to receive gamma-aminobutyric
differentiate further into mature granule neurons acid (GABA)ergic synaptic inputs at 1 week after
that are immunoreactive for Prox1 and calbindin. birth and glutamatergic inputs within 2 weeks.
531

Adult-generated DGCs initially show depolarizing entire spectrum of biological functions involving
responses to GABA (Tozuka et al., 2005; Wadiche adult DGC neurogenesis awaits more specific
et al., 2005; Ge et al., 2006), similar to immature means of manipulating DGC progenitors in the
neurons during early development. Unlike those in adult.
the neonate, however, adult-generated DGCs take
much longer (perhaps 3–6 months) to obtain a full
dendritic arborization pattern (van Praag et al., Regulation of adult DGC neurogenesis
2002). Compared to neighboring mature DGCs,
immature DGCs in the adult have a higher input A host of physiological states and molecular
resistance and a subthreshold calcium conductance factors appear to regulate adult hippocampal
that may enhance their excitability. In vitro elect- neurogenesis (see Table 1). Neuronal production
rophysiological studies also show a decreased declines markedly with age (Seki and Arai, 1993;
threshold for long-term potentiation (LTP) or Kuhn et al., 1996), probably due to reduced DGC
long-term depression, as well as higher-amplitude precursor proliferation and altered differentiation
LTP, in newborn DGCs compared to their mature or delayed maturation of their progeny (Kemper-
counterparts (Wang et al., 2000; Schmidt-Hieber mann et al., 1998; Rao et al., 2005). The mecha-
et al., 2004). nisms likely reflect active suppression of
The precise function of DGCs generated in neurogenesis with age because removal of cortico-
adulthood is unknown. Most of the experimental steroids by adrenalectomy or the infusion of spe-
evidence to date, however, supports a role for cific growth factors significantly restores DGC
DGC neurogenesis in learning and memory func- production in aged rats (Cameron and McKay,
tion (see Doetsch and Hen, 2005 for review). Stim- 1999; Lichtenwalner et al., 2001; Jin et al., 2003).
ulation of adult DGC neurogenesis with In addition to age, acute or chronic stress de-
behavioral interventions such as exercise or envi- creases DGC neurogenesis in young adult rodents
ronmental enrichment is associated with better and primates (Mirescu and Gould, 2006).
performance on certain hippocampal learning Adult neurogenesis is stimulated by environ-
tasks (Bruel-Jungerman et al., 2005; van Praag mental enrichment, which increases the survival
et al., 2005; Meshi et al., 2006), although the link and neuronal differentiation of DGCs generated in
between increased neurogeneis and enrichment- early adulthood through senescence (Kemper-
enhanced performance has recently been called mann et al., 1997, 1998, 2002; Nilsson et al.,
into question (Meshi et al., 2006). The degree of 1999). Exercise also enhances adult neurogenesis,
adult neurogenesis in some inbred mouse strains but this effect relates to increased precursor pro-
also correlates positively with learning ability liferation rather than survival (van Praag et al.,
(Kempermann and Gage, 2002). More direct 1999). Growth factors may play a significant role
evidence that neurogenesis supports hippocam- in mediating these effects, as blocking brain up-
pal-dependent learning and memory derives take of insulin-like growth factor-1 (IGF-1) or
from experiments in which the depletion of vascular endothelial growth factor (VEGF) pre-
adult-generated neurons disrupts certain forms of vents the increase in DGC production induced by
learning (Shors et al., 2001, 2002; Winocur et al., enrichment or exercise (Trejo et al., 2001; Fabel
2006). These data are supported by the findings et al., 2003; Cao et al., 2004). Administration of
described above that suggest LTP is more easily these growth factors also increases adult neuro-
elicited and of higher amplitude in immature, genesis in the rodent dentate gyrus (Aberg et al.,
adult-born DGCs (Wang et al., 2000; Schmidt- 2000; Jin et al., 2002; Cao et al., 2004; Schanzer
Hieber et al., 2004). Recent animal studies also et al., 2004), but it is unclear whether this effect
provide evidence that DGC neurogenesis in the results from increased progenitor cell proliferation
adult is necessary for the positive effects of anti- (Jin et al., 2002) or survival (Schanzer et al., 2004).
depressant medication (Santarelli et al., 2003). Other mediators of adult hippocampal neuro-
Despite these accumulating data, knowledge of the genesis include hormones and neurotransmitters.
532

Table 1. Modulation of adult hippocampal neurogenesis

Variable Change in Mechanism Change in Proliferation Neuronal Survival


neurogenesis neurogenesis differentiation

Physiologic state
Aging Decreased Proliferation/ Decreased m m
neuronal
differentiation
Stress Decreased Proliferation Decreased m
Exercise Increased Proliferation/ Increased m m? m?
?survival/
?differentiation
Environmental Increased Survival Increased m
enrichment
Growth factors
bFGF Increased (early Proliferation Increased m
postnatal) (postnatal)
BDNF Increased Differentiation/ Increased m m
survival
VEGF Increased Proliferation/ Increased m m
survival
IGF-1 Increased Proliferation/ Increased k m
survival/
?differentiation
Neurotransmitters/neuromodulators
Glutamate Decreased Proliferation Decreased m
GABA Increased Differentiation/ Increased m? m
?proliferation
Serotonin Increased Proliferation Increased m
Nitrous oxide Decreased Decreased Decreased k m
proliferation
(promotes
differentiation)
Transcription factors
Wnt Increased Proliferation/ Increased m m?
?differentiation
Sonic Increased Proliferation Increased m
hedgehog
Sox2 Increased ?Proliferation/ m? m?
?differentiation

Adrenalectomy and corticosteroid replacement receptor antagonists or lesion of DGC afferents


studies have shown that circulating stress hor- exerts the opposite effect (Cameron et al., 1995).
mones suppress the rate of DGC production in Serotonin (5-hydroxytryptamine) also influences
adulthood (reviewed in Mirescu and Gould, 2006). adult neurogenesis. Chronic treatment with anti-
In contrast, cell proliferation in the rat dentate depressant drugs that act as serotonin reuptake
gyrus increases during proestrus, and the effect is inhibitors increases cell proliferation in the rodent
abolished by ovariectomy (Tanapat et al., 1999). dentate gyrus (Malberg et al., 2000), and blockade
Glutamatergic inputs also appear to influence of serotonin synthesis or administration of a
dentate gyrus neurogenesis in the adult rodent. serotonin neurotoxin decreases new cell produc-
Activation of N-methyl-D-aspartate (NMDA) re- tion (Brezun and Daszuta, 1999). In terms of other
ceptors attenuates cell proliferation, and suppress- neuromodulators, the number of newly generated
ing glutamatergic neurotransmission with NMDA DGCs is negatively regulated by nitric oxide,
533

which reduces DGC precursor proliferation but endogenous repair after brain injury or neurode-
promotes neuronal differentiation (Packer et al., generation. Various brain insults in the adult,
2003). Neuronal addition thus continues in adult- including mechanical lesions, hypoglycemia,
hood under the influence of a balance between fluid percussion injury, and stroke, stimulate pro-
multiple positive and negative regulators of ne- genitor cell proliferation in the hippocampal
urogenic activity. dentate gyrus (Gould and Tanapat, 1997; Liu
The local microenvironment in regions of on- et al., 1998; Jin et al., 2001; Suh et al., 2005).
going neurogenesis in the adult mammalian brain As in the intact brain, many newly generated cells
provides what is referred to as a ‘‘stem cell niche.’’ die, but the vast majority of those that survive
This niche governs neuronal production and in- differentiate into neurons, and typically less
cludes progenitors, astrocytes, microvasculature, than 20% of adult-born cells become glia or
and microglia (Wurmser et al., 2004). Astrocytes remain undifferentiated. Similar findings are seen
appear to influence all phases of adult neurogen- after seizure-induced injury in temporal lobe
esis, including neural precursor cell proliferation, epilepsy (TLE) models, which were the first brain
migration, and differentiation (Lim and Alvarez- injury models used to demonstrate altered hippo-
Buylla, 1999; Song et al., 2002a). Studies of ne- campal neurogenesis (Parent et al., 1997). Studies
urogenesis in the adult rodent dentate gyrus and of adult rodent models of limbic epileptogenesis
subventricular zone (SVZ) implicate radial glia- or acute seizures indicate that prolonged
like astrocytes as the neural stem cells from which seizures potently stimulate DGC neurogenesis
neuroblasts derive (Doetsch et al., 1999; Seri et al., (Bengzon et al., 1997; Gray and Sundstrom,
2001; Garcia et al., 2004; Merkle et al., 2004). 1998; Parent et al., 1997, 1998; Scott et al.,
A vascular presence in the stem cell niche likely 1998). In the adult rodent kainate and pilocar-
modulates neurogenesis as precursors in the dent- pine models of temporal lobe epilepsy,
ate gyrus and SVZ reside in close proximity to the chemoconvulsant-induced status epilepticus (SE)
microvasculature (Palmer et al., 2000; Shen et al., increases dentate gyrus cell proliferation approx-
2004), and their proliferative activity appears to be imately 5–10-fold after a latent period of at least
tightly linked with angiogenesis. In the adult, sup- several days (Parent et al., 1997; Gray and
pression of hippocampal neurogenesis by irradia- Sundstrom, 1998) (Fig. 2A and B). Approxi-
tion may reflect disruption of local angiogenesis mately 80–90% of the newly generated cells
(Monje et al., 2002), and infusion of VEGF into differentiate into DGCs.
the adult brain increases the production both of DGC neurogenesis is also enhanced by kindling-
endothelial cells and DGCs (Jin et al., 2002). In induced epileptogenesis. Electrical kindling of the
contrast to astrocytes and endothelial cells, recent amygdala (Parent et al., 1998; Scott et al., 1998),
studies suggest that reactive microglia disrupt hippocampus (Bengzon et al., 1997) or perforant
neurogenesis in the adult dentate gyrus. Micro- path (Nakagawa et al., 2000) acutely increases
glial reaction and inflammation induced by brain dentate gyrus cell proliferation and neurogenesis.
irradiation, seizures, or lipopolysaccharide treat- Similar neurogenic effects occur after acute
ment decreases the survival of adult-generated seizures induced by intermittent perforant path
DGCs, and suppressing microglial activation with or hippocampal stimulation in adult rats (Bengzon
minocycline or indomethacin treatment partially et al., 1997; Parent et al., 1997). Even single,
restores neurogenesis (Ekdahl et al., 2003; Monje discrete seizure-like afterdischarges induced by
et al., 2003). hippocampal stimulation lead to increased num-
bers of newly differentiated DGCs (Bengzon et al.,
1997). Although more severe seizures enhance
Seizure-induced hippocampal neurogenesis neurogenesis to a greater extent, the survival of
the newborn DGCs may decrease with increased
The persistence of neural stem cells in the adult seizure severity (Mohapel et al., 2004). This effect
mammalian forebrain raises the possibility for may relate to the degree of microglial activation as
534

Fig. 2. Increased cell proliferation and altered dentate gyrus neuroblast migration after SE. (A, B) BrdU incorporation (arrows) in
adult rat dentate gyrus 7 days after saline (A) or pilocarpine (B) treatment followed by a 2 day post-BrdU survival. Subgranular zone
BrdU labeling increases markedly after SE (B) compared to the control (A). (C, D) Doublecortin (DCX) immunoreactivity in the
dentate gyrus of a control (Con; C) and an adult rat 14 days after pilocarpine-induced SE (14d; D). Note the increased hilar DCX
expression and chains of DCX-positive cells extending into the hilus (arrows in D) after SE. E, PSA-NCAM+ neuroblast chains (red,
arrows) alongside GFAP+ hilar astrocytes (green). Dashed lines in C–E denote the granule cell layer (gcl) — hilar (h) border.
(See Color Plate 28.2 in color plate section.)

it is reversed in part by minocycline treatment to project axons aberrantly in the epileptogenic


(Ekdahl et al., 2003). hippocampal formation.
An increasingly recognized abnormality of
DGC neurogenesis in experimental TLE is the
ectopic location of adult-born granule neurons.
Seizure-induced hippocampal neurogenesis: The DGC layer in human TLE is often abnormal
functional significance due to dispersion and the presence of ectopic
granule-like neurons in the hilus and inner molec-
In the pilocarpine model of TLE, aberrant mossy ular layer (Houser, 1990; Parent et al., 2006a).
fiber sprouting occurs in parallel with the gener- An early report of seizure-induced neurogenesis
ation of increased numbers of DGCs (Parent et al., described newly differentiating neurons with gran-
1997). Evidence suggests that developing axons ule cell morphology in the dentate hilus and inner
from newborn DGCs contribute to status epilep- molecular layer after pilocarpine-induced SE in
ticus (SE)-induced aberrant mossy fiber reorgan- adult rats (Parent et al., 1997). Hilar-ectopic
ization in both hippocampal area CA3 and the DGCs are found in several different experimental
dentate inner molecular layer (Parent et al., 1997). TLE models and may persist for many months
Eliminating DGC progenitors by brain irradiation (Parent et al., 1997, 2006a; Scharfman et al., 2000;
before and after pilocarpine treatment, however, Dashtipour et al., 2001). The cells resemble the
does not suppress mossy fiber sprouting (Parent granule-like neurons identified in surgical speci-
et al., 1999). Thus, seizure-induced injury appears mens from humans with temporal lobe epilepsy,
to induce both pre-existing and adult-born DGCs both in terms of their morphology and marker
535

Fig. 3. Schematic showing the effects of prolonged seizures on caudal subventricular zone (SVZ) and dentate gyrus progenitors. In the
intact adult rat (top), progenitors located in the infracallosal SVZ (purple) give rise to white matter oligodendrocytes (orange), while
those in the dentate gyrus (green) give rise to DGCs in the granule cell layer (gcl). After seizure-induced hilar and pyramidal cell layer
(pcl) injury (bottom; jagged arrows), progenitors migrate aberrantly from the caudal SVZ to form glia (a; purple) in the hippocampus
proper or from the dentate subgranular zone to form hilar-ectopic DGCs (b). (See Color Plate 28.3 in color plate section.)

expression (Scharfman et al., 2000; Dashtipour seizures exhibit a much higher percentage of per-
et al., 2001; Parent et al., 2006a). Studies of ne- sistent basal dendrites than is normally found
urogenesis in the pilocarpine TLE model indicate (Scharfman et al., 2000; Dashtipour et al., 2001).
that the hilar-ectopic DGCs migrate aberrantly Work by Dashtipour and colleagues suggests that
from the dentate subgranular proliferative zone to the basal dendrites of hilar-ectopic DGCs receive
the hilus (Parent et al., 2006a) (Fig. 3). increased excitatory input (Dashtipour et al.,
Using intracellular recordings in hippocampal 2001), suggesting that this structural plasticity
slices from epileptic adult rats, Scharfman and may be a mechanism for seizure generation or
colleagues (Scharfman et al., 2000) have shown propagation. Further evidence supporting the
that the hilar-ectopic granule-like cells are hyper- epileptogenic nature of hilar-ectopic DGCs comes
excitable and fire in abnormal bursts synchro- from the work of Jung and colleagues (Jung et al.,
nously with CA3 pyramidal cells. Expression of 2004). Their group used the antimitotic agent
the activation-induced immediate early gene c-fos AraC to inhibit DGC neurogenesis after pilocar-
by hilar-ectopic DGCs during spontaneous limbic pine treatment, and found that these rats devel-
seizures also suggests that they participate in ep- oped fewer and shorter spontaneous recurrent
ileptic circuitry (Scharfman et al., 2002). In addi- seizures than controls. Taken together, these data
tion, many putatively newborn DGCs located in suggest that hilar-ectopic DGCs integrate abnor-
the hilus or hilar aspect of the DGC layer after mally, are hyperexcitable, and thus may contribute
536

Brain insult in childhood (e.g.,


prolonged febrile seizure)

RECURRENT IMPAIRED
Altered developmental cues SEIZURES LEARNING/
MEMORY

Aberrant migration
of DGC progenitors Ectopically integrated DGCs

Fig. 4. Aberrant neurogenesis hypothesis of epileptogenesis and memory disturbance in mesial temporal lobe epilepsy. Note the
potential maladaptive positive feedback loop of recurrent seizures on aberrant neurogenesis.

to seizure generation or propagation. The presence and olfactory bulb of adult rodents (Gould and
of hilar-ectopic DGCs in hippocampi surgically McEwen, 1993; Biebl et al., 2000). Seizures also
resected to treat intractable epilepsy (Houser, increase the expression of molecules with the
1990; Parent et al., 2006a) raises the possibility potential to increase neurogenesis or gliogenesis
that abnormalities in dentate gyrus persistent ne- such as growth factors (Humpel et al., 1993) and
urogenesis contribute to epileptogenesis or cogni- neurotrophins (Isackson et al., 1991). Specific ne-
tive dysfunction in human TLE (Fig. 4). urotransmitters or neuromodulatory systems
that normally influence DGC neurogenesis (see
Table 1) also may be altered by seizure activity.
Seizure-induced neurogenesis: mechanisms In terms of potential mechanisms underlying
hilar- or molecular-layer ectopic DGC formation,
The mechanisms by which seizure activity stimu- molecular factors that influence neuronal migra-
lates neurogenesis or gliogenesis are unknown. tion during development are prime candidates.
Experiments in which proliferating cells were labe- The expression of one of these developmental fac-
led with BrdU prior to seizure induction have tors, reelin, persists in the hippocampal dentate
shown that epileptic activity stimulates dentate gyrus of adult humans and rodents, and has been
gyrus and caudal SVZ precursors that are actively implicated in DGC layer dispersion in human
proliferating prior to injury (Parent et al., 1999, TLE (Haas et al., 2002). The expression of reelin
2006b). Huttman and colleagues used a reporter decreases markedly in the dentate gyrus after
mouse in which green fluorescent protein selec- pilocarpine-induced SE (JMP, personal communi-
tively labeled nestin-expressing stem-like cells to cation), suggesting that loss of this migration
show that kainate-induced seizures specifically guidance factor may be responsible for the aber-
increased the proliferation of the radial glia-like rant migration of DGC progenitors during epile-
progenitors in dentate gyrus (Huttmann et al., ptogenesis. Another potential mechanism is loss of
2003). Seizures may act to increase neurogenesis GABA caused by SE-induced depletion of dentate
indirectly through activation of astrocytes, as as- interneurons, as this neurotransmitter influences
trocytes stimulate hippocampal neurogenesis via DGC differentiation (see Table 1) and GABA
wnt signaling and perhaps other mechanisms decreases neuroblast migration in the other adult
(Song et al., 2002a; Lie et al., 2005). Alternatively, neurogenic region, the SVZ-olfactory bulb path-
the SE-induced death of some mature DGCs may way (Liu et al., 2005). In an interesting study,
increase cell turnover in the dentate gyrus via other Scharfman and colleagues showed that hippocam-
mechanisms. Cell death is associated with subse- pal brain-derived neurotrophic factor (BDNF)
quent cell birth in a number of postnatal infusion in adult rat increased DGC neurogenesis
neurogenic systems, including the dentate gyrus and led to the appearance of ectopic DGCs
537

(Scharfman et al., 2005). This result raises the pos- Altman, J. and Das, G.D. (1965) Autoradiographic and histo-
sibility that BDNF is involved in seizure-induced, logical evidence of postnatal hippocampal neurogenesis in
aberrant DGC neurogenesis as well. rats. J. Comp. Neurol., 124: 319–335.
Bengzon, J., Kokaia, Z., Elmer, E., Nanobashvili, A., Kokaia,
The multiple steps of DGC neurogenesis are in- M. and Lindvall, O. (1997) Apoptosis and proliferation of
fluenced by a delicate balance of factors affecting dentate gyrus neurons after single and intermittent limbic
DGC progenitors in the adult. Improved under- seizures. Proc. Natl. Acad. Sci. U.S.A., 94: 10432–10437.
standing of the biological functions of adult-born Biebl, M., Cooper, C.M., Winkler, J. and Kuhn, H.G. (2000)
DGCs, their regulation, and how they are influ- Analysis of neurogenesis and programmed cell death reveals
a self-renewing capacity in the adult rat brain. Neurosci.
enced by brain pathology may provide important Lett., 291: 17–20.
insights into the pathophysiology of TLE, its as- Brezun, J.M. and Daszuta, A. (1999) Depletion in serotonin
sociated cognitive impairments and other brain decreases neurogenesis in the dentate gyrus and the subven-
disorders. Manipulation of adult neurogenesis for tricular zone of adult rats. Neuroscience, 89: 999–1002.
therapeutic purposes therefore is likely to involve Bruel-Jungerman, E., Laroche, S. and Rampon, C. (2005) New
neurons in the dentate gyrus are involved in the expression
suppression of aberrant integration in certain
of enhanced long-term memory following environmental
instances rather than simply stimulation of neuro- enrichment. Eur. J. Neurosci., 21: 513–521.
genesis for neuronal replacement. Cameron, H.A., McEwen, B.S. and Gould, E. (1995) Regula-
tion of adult neurogenesis by excitatory input and NMDA
receptor activation in the dentate gyrus. J. Neurosci., 15:
4687–4692.
Cameron, H.A. and McKay, R.D. (1999) Restoring production
Abbreviations
of hippocampal neurons in old age. Nat. Neurosci., 2:
894–897.
BDNF brain-derived neurotrophic Cameron, H.A. and McKay, R.D. (2001) Adult neurogenesis
factor produces a large pool of new granule cells in the dentate
BrdU bromodeoxyuridine gyrus. J. Comp. Neurol., 435: 406–417.
DCX doublecortin Cameron, H.A., Woolley, C.S., McEwen, B.S. and Gould, E.
(1993) Differentiation of newly born neurons and glia in the
DGC dentate granule cell dentate gyrus of the adult rat. Neuroscience, 56: 337–344.
GABA gamma-aminobutyric acid Cao, L., Jiao, X., Zuzga, D.S., Liu, Y., Fong, D.M., Young, D.
GFAP glial fibrillary acidic protein and During, M.J. (2004) VEGF links hippocampal activity
IGF-1 insulin-like growth factor-1 with neurogenesis, learning and memory. Nat. Genet., 36:
LTP long-term potentiation 827–835.
Dashtipour, K., Tran, P.H., Okazaki, M.M., Nadler, J.V. and
mTLE medial temporal lobe epi- Ribak, C.E. (2001) Ultrastructural features and synaptic
lepsy connections of hilar ectopic granule cells in the rat dentate
NMDA N-methyl-D-aspartate gyrus are different from those of granule cells in the granule
PSA-NCAM polysialylated neural cell ad- cell layer. Brain Res., 890: 261–271.
hesion molecule Doetsch, F., Caille, I., Lim, D.A., Garcia-Verdugo, J.M. and
Alvarez-Buylla, A. (1999) Subventricular zone astrocytes are
SE status epilepticus
neural stem cells in the adult mammalian brain. Cell, 97:
SGZ subgranular zone 703–716.
SVZ subventricular zone Doetsch, F. and Hen, R. (2005) Young and excitable: the func-
TLE temporal lobe epilepsy tion of new neurons in the adult mammalian brain. Curr.
VEGF vascular endothelial growth Opin. Neurobiol., 15: 121–128.
Ekdahl, C.T., Claasen, J.H., Bonde, S., Kokaia, Z. and
factor
Lindvall, O. (2003) Inflammation is detrimental for neuro-
genesis in adult brain. Proc. Natl. Acad. Sci. U.S.A., 100:
13632–13637.
References Eriksson, P.S., Perfilieva, E., Bjork-Eriksson, T., Alborn, A.M.,
Nordborg, C., Peterson, D.A. and Gage, F.H. (1998) Ne-
Aberg, M.A., Aberg, N.D., Hedbacker, H., Oscarsson, J. and urogenesis in the adult human hippocampus. Nat. Med., 4:
Eriksson, P.S. (2000) Peripheral infusion of IGF-I selectively 1313–1317.
induces neurogenesis in the adult rat hippocampus. J. Ne- Fabel, K., Fabel, K., Tam, B., Kaufer, D., Baiker, A., Simm-
urosci., 20: 2896–2903. ons, N., Kuo, C.J. and Palmer, T.D. (2003) VEGF is
538

necessary for exercise-induced adult hippocampal neurogen- Jin, K., Sun, Y., Xie, L., Batteur, S., Mao, X.O., Smelick, C.,
esis. Eur. J. Neurosci., 18: 2803–2812. Logvinova, A. and Greenberg, D.A. (2003) Neurogenesis
Filippov, V., Kronenberg, G., Pivneva, T., Reuter, K., Steiner, and aging: FGF-2 and HB-EGF restore neurogenesis in
B., Wang, L.P., Yamaguchi, M., Kettenmann, H. and Kem- hippocampus and subventricular zone of aged mice. Aging
permann, G. (2003) Subpopulation of nestin-expressing pro- Cell, 2: 175–183.
genitor cells in the adult murine hippocampus shows Jin, K., Zhu, Y., Sun, Y., Mao, X.O., Xie, L. and Greenberg,
electrophysiological and morphological characteristics of as- D.A. (2002) Vascular endothelial growth factor (VEGF)
trocytes. Mol. Cell Neurosci., 23: 373–382. stimulates neurogenesis in vitro and in vivo. Proc. Natl.
Garcia, A.D., Doan, N.B., Imura, T., Bush, T.G. and Acad. Sci. U.S.A., 99: 11946–11950.
Sofroniew, M.V. (2004) GFAP-expressing progenitors are Jung, K.H., Chu, K., Kim, M., Jeong, S.W., Song, Y.M.,
the principal source of constitutive neurogenesis in adult Lee, S.T., Kim, J.Y., Lee, S.K. and Roh, J.K. (2004)
mouse forebrain. Nat. Neurosci., 7: 1233–1241. Continuous cytosine-b-D-arabinofuranoside infusion re-
Ge, S., Goh, E.L., Sailor, K.A., Kitabatake, Y., Ming, G.L. duces ectopic granule cells in adult rat hippocampus with
and Song, H. (2006) GABA regulates synaptic integration of attenuation of spontaneous recurrent seizures following
newly generated neurons in the adult brain. Nature, 439: pilocarpine-induced status epilepticus. Eur. J. Neurosci.,
589–593. 19: 3219–3226.
Gould, E. and McEwen, B.S. (1993) Neuronal birth and death. Kaplan, M.S. and Hinds, J.W. (1977) Neurogenesis in the adult
Curr. Opin. Neurobiol., 3: 676–682. rat: electron microscopic analysis of light radioautographs.
Gould, E. and Tanapat, P. (1997) Lesion-induced proliferation Science, 197: 1092–1094.
of neuronal progenitors in the dentate gyrus of the adult rat. Kempermann, G. and Gage, F.H. (2002) Genetic determinants
Neuroscience, 80: 427–436. of adult hippocampal neurogenesis correlate with acquisition,
Gould, E., Tanapat, P., McEwen, B.S., Flugge, G. and but not probe trial performance, in the water maze task.
Fuchs, E. (1998) Proliferation of granule cell precursors in Eur. J. Neurosci., 16: 129–136.
the dentate gyrus of adult monkeys is diminished by stress. Kempermann, G., Gast, D. and Gage, F.H. (2002) Neuroplas-
Proc. Natl. Acad. Sci. U.S.A., 95: 3168–3171. ticity in old age: sustained fivefold induction of hippocampal
Gray, W.P. and Sundstrom, L.E. (1998) Kainic acid increases neurogenesis by long-term environmental enrichment. Ann.
the proliferation of granule cell progenitors in the dentate Neurol., 52: 135–143.
gyrus of the adult rat. Brain Res., 790: 52–59. Kempermann, G., Jessberger, S., Steiner, B. and Kronenberg,
Haas, C.A., Dudeck, O., Kirsch, M., Huszka, C., Kann, G., G. (2004) Milestones of neuronal development in the adult
Pollak, S., Zentner, J. and Frotscher, M. (2002) Role for hippocampus. Trends Neurosci., 27: 447–452.
reelin in the development of granule cell dispersion in tem- Kempermann, G., Kuhn, H.G. and Gage, F.H. (1997) More
poral lobe epilepsy. J. Neurosci., 22: 5797–5802. hippocampal neurons in adult mice living in an enriched en-
Hastings, N.B. and Gould, E. (1999) Rapid extension of axons vironment. Nature, 386: 493–495.
into the CA3 region by adult-generated granule cells. J. Kempermann, G., Kuhn, H.G. and Gage, F.H. (1998) Expe-
Comp. Neurol., 413: 146–154. rience-induced neurogenesis in the senescent dentate gyrus. J.
Houser, C.R. (1990) Granule cell dispersion in the dentate Neurosci., 18: 3206–3212.
gyrus of humans with temporal lobe epilepsy. Brain Res., Kornack, D.R. and Rakic, P. (1999) Continuation of neuro-
535: 195–204. genesis in the hippocampus of the adult macaque monkey.
Humpel, C., Lippoldt, A., Chadi, G., Ganten, D., Olson, L. and Proc. Natl. Acad. Sci. U.S.A., 96: 5768–5773.
Fuxe, K. (1993) Fast and widespread increase of basic fib- Kuhn, H.G., Dickinson-Anson, H. and Gage, F.H. (1996)
roblast growth factor messenger RNA and protein in the Neurogenesis in the dentate gyrus of the adult rat: age-related
forebrain after kainate-induced seizures. Neuroscience, 57: decrease of neuronal progenitor proliferation. J. Neurosci.,
913–922. 16: 2027–2033.
Huttmann, K., Sadgrove, M., Wallraff, A., Hinterkeuser, S., Lichtenwalner, R.J., Forbes, M.E., Bennett, S.A., Lynch, C.D.,
Kirchhoff, F., Steinhauser, C. and Gray, W.P. (2003) Seizures Sonntag, W.E. and Riddle, D.R. (2001) Intracerebroven-
preferentially stimulate proliferation of radial glia-like astro- tricular infusion of insulin-like growth factor-I ameliorates
cytes in the adult dentate gyrus: functional and immunocyto- the age-related decline in hippocampal neurogenesis. Neuro-
chemical analysis. Eur. J. Neurosci., 18: 2769–2778. science, 107: 603–613.
Isackson, P.J., Huntsman, M.M., Murray, K.D. and Gall, Lie, D.C., Colamarino, S.A., Song, H.J., Desire, L., Mira, H.,
C.M. (1991) BDNF mRNA expression is increased in adult Consiglio, A., Lein, E.S., Jessberger, S., Lansford, H.,
rat forebrain after limbic seizures: temporal patterns of in- Dearie, A.R. and Gage, F.H. (2005) Wnt signalling regu-
duction distinct from NGF. Neuron, 6: 937–948. lates adult hippocampal neurogenesis. Nature, 437:
Jin, K., Minami, M., Lan, J.Q., Mao, X.O., Batteur, S., 1370–1375.
Simon, R.P. and Greenberg, D.A. (2001) Neurogenesis Lim, D.A. and Alvarez-Buylla, A. (1999) Interaction between
in dentate subgranular zone and rostral subventricular zone astrocytes and adult subventricular zone precursors stimu-
after focal cerebral ischemia in the rat. Proc. Natl. Acad. Sci. lates neurogenesis. Proc. Natl. Acad. Sci. U.S.A., 96:
U.S.A., 98: 4710–4715. 7526–7531.
539

Liu, J., Solway, K., Messing, R.O. and Sharp, F.R. (1998) In- neurogenesis in experimental temporal lobe epilepsy. Ann.
creased neurogenesis in the dentate gyrus after transient glo- Neurol., 59: 81–91.
bal ischemia in gerbils. J. Neurosci., 18: 7768–7778. Parent, J.M., Janumpalli, S., McNamara, J.O. and Lowenstein,
Liu, X., Wang, Q., Haydar, T.F. and Bordey, A. (2005) Nonsy- D.H. (1998) Increased dentate granule cell neurogenesis fol-
naptic GABA signaling in postnatal subventricular zone lowing amygdala kindling in the adult rat. Neurosci. Lett.,
controls proliferation of GFAP-expressing progenitors. Nat. 247: 9–12.
Neurosci., 8: 1179–1187. Parent, J.M., Tada, E., Fike, J.R. and Lowenstein, D.H. (1999)
Malberg, J.E., Eisch, A.J., Nestler, E.J. and Duman, R.S. Inhibition of dentate granule cell neurogenesis with brain ir-
(2000) Chronic antidepressant treatment increases neurogen- radiation does not prevent seizure-induced mossy fiber
esis in adult rat hippocampus. J. Neurosci., 20: 9104–9110. synaptic reorganization in the rat. J. Neurosci., 19:
Markakis, E.A. and Gage, F.H. (1999) Adult-generated neu- 4508–4519.
rons in the dentate gyrus send axonal projections to field CA3 Parent, J.M., Yu, T.W., Leibowitz, R.T., Geschwind, D.H.,
and are surrounded by synaptic vesicles. J. Comp. Neurol., Sloviter, R.S. and Lowenstein, D.H. (1997) Dentate granule
406: 449–460. cell neurogenesis is increased by seizures and contributes to
Merkle, F.T., Tramontin, A.D., Garcia-Verdugo, J.M. and aberrant network reorganization in the adult rat hippocam-
Alvarez-Buylla, A. (2004) Radial glia give rise to adult neural pus. J. Neurosci., 17: 3727–3738.
stem cells in the subventricular zone. Proc. Natl. Acad. Sci. van Praag, H., Kempermann, G. and Gage, F.H. (1999) Run-
U.S.A., 101: 17528–17532. ning increases cell proliferation and neurogenesis in the adult
Meshi, D., Drew, M.R., Saxe, M., Ansorge, M.S., David, D., mouse dentate gyrus. Nat. Neurosci., 2: 266–270.
Santarelli, L., Malapani, C., Moore, H. and Hen, R. (2006) van Praag, H., Schinder, A.F., Christie, B.R., Toni, N., Palmer,
Hippocampal neurogenesis is not required for behavioral T.D. and Gage, F.H. (2002) Functional neurogenesis in the
effects of environmental enrichment. Nat. Neurosci., 9: adult hippocampus. Nature, 415: 1030–1034.
729–731. van Praag, H., Shubert, T., Zhao, C. and Gage, F.H. (2005)
Mirescu, C. and Gould, E. (2006) Stress and adult neurogenesis. Exercise enhances learning and hippocampal neurogenesis in
Hippocampus, 16: 233–238. aged mice. J. Neurosci., 25: 8680–8685.
Mohapel, P., Ekdahl, C.T. and Lindvall, O. (2004) Status ep- Rao, M.S., Hattiangady, B., Abdel-Rahman, A., Stanley, D.P.
ilepticus severity influences the long-term outcome of neuro- and Shetty, A.K. (2005) Newly born cells in the ageing dent-
genesis in the adult dentate gyrus. Neurobiol. Dis., 15: ate gyrus display normal migration, survival and neuronal
196–205. fate choice but endure retarded early maturation. Eur. J.
Monje, M.L., Mizumatsu, S., Fike, J.R. and Palmer, T.D. Neurosci., 21: 464–476.
(2002) Irradiation induces neural precursor-cell dysfunction. Santarelli, L., Saxe, M., Gross, C., Surget, A., Battaglia, F.,
Nat. Med., 8: 955–962. Dulawa, S., Weisstaub, N., Lee, J., Duman, R., Arancio, O.,
Monje, M.L., Toda, H. and Palmer, T.D. (2003) Inflammatory Belzung, C. and Hen, R. (2003) Requirement of hippocampal
blockade restores adult hippocampal neurogenesis. Science, neurogenesis for the behavioral effects of antidepressants.
302: 1760–1765. Science, 301: 805–809.
Nakagawa, E., Aimi, Y., Yasuhara, O., Tooyama, I., Shimada, Schanzer, A., Wachs, F.P., Wilhelm, D., Acker, T., Cooper-
M., McGeer, P.L. and Kimura, H. (2000) Enhancement of Kuhn, C., Beck, H., Winkler, J., Aigner, L., Plate, K.H. and
progenitor cell division in the dentate gyrus triggered by in- Kuhn, H.G. (2004) Direct stimulation of adult neural stem
itial limbic seizures in rat models of epilepsy. Epilepsia, 41: cells in vitro and neurogenesis in vivo by vascular endothelial
10–18. growth factor. Brain Pathol., 14: 237–248.
Nilsson, M., Perfilieva, E., Johansson, U., Orwar, O. and Er- Scharfman, H., Goodman, J., Macleod, A., Phani, S., Anton-
iksson, P.S. (1999) Enriched environment increases neuro- elli, C. and Croll, S. (2005) Increased neurogenesis and the
genesis in the adult rat dentate gyrus and improves spatial ectopic granule cells after intrahippocampal BDNF infusion
memory. J. Neurobiol., 39: 569–578. in adult rats. Exp. Neurol., 192: 348–356.
Packer, M.A., Stasiv, Y., Benraiss, A., Chmielnicki, E., Grinb- Scharfman, H.E., Goodman, J.H. and Sollas, A.L. (2000)
erg, A., Westphal, H., Goldman, S.A. and Enikolopov, G. Granule-like neurons at the hilar/CA3 border after status
(2003) Nitric oxide negatively regulates mammalian adult epilepticus and their synchrony with area CA3 pyramidal
neurogenesis. Proc. Natl. Acad. Sci. U.S.A., 100: 9566–9571. cells: functional implications of seizure-induced neurogenesis.
Palmer, T.D., Willhoite, A.R. and Gage, F.H. (2000) Vascular J. Neurosci., 20: 6144–6158.
niche for adult hippocampal neurogenesis. J. Comp. Neurol., Scharfman, H.E., Sollas, A.L. and Goodman, J.H. (2002)
425: 479–494. Spontaneous recurrent seizures after pilocarpine-induced sta-
Parent, J.M., von dem Bussche, N. and Lowenstein, D.H. tus epilepticus activate calbindin-immunoreactive hilar cells
(2006b) Prolonged seizures recruit caudal subventricular zone of the rat dentate gyrus. Neuroscience, 111: 71–81.
glial progenitors into the injured hippocampus. Hippocam- Schlessinger, A.R., Cowan, W.M. and Gottlieb, D.I. (1975) An
pus, 16: 321–328. autoradiographic study of the time of origin and the pattern
Parent, J.M., Elliott, R.C., Pleasure, S.J., Barbaro, N.M. and of granule cell migration in the dentate gyrus of the rat. J.
Lowenstein, D.H. (2006a) Aberrant seizure-induced Comp. Neurol., 159: 149–175.
540

Schmidt-Hieber, C., Jonas, P. and Bischofberger, J. (2004) Stanfield, B.B. and Trice, J.E. (1988) Evidence that granule cells
Enhanced synaptic plasticity in newly generated granule cells generated in the dentate gyrus of adult rats extend axonal
of the adult hippocampus. Nature, 429: 184–187. projections. Exp. Brain Res., 72: 399–406.
Scott, B.W., Wang, S., Burnham, W.M., De Boni, U. and Suh, S.W., Fan, Y., Hong, S.M., Liu, Z., Matsumori, Y., Wein-
Wojtowicz, J.M. (1998) Kindling-induced neurogenesis in the stein, P.R., Swanson, R.A. and Liu, J. (2005) Hypoglycemia
dentate gyrus of the rat. Neurosci. Lett., 248: 73–76. induces transient neurogenesis and subsequent progenitor cell
Seki, T. and Arai, Y. (1993) Highly polysialylated neural cell loss in the rat hippocampus. Diabetes, 54: 500–509.
adhesion molecule (NCAM-H) is expressed by newly gener- Tanapat, P., Hastings, N.B., Reeves, A.J. and Gould, E. (1999)
ated granule cells in the dentate gyrus of the adult rat. Estrogen stimulates a transient increase in the number of new
J. Neurosci., 13: 2351–2358. neurons in the dentate gyrus of the adult female rat. J. Ne-
Seri, B., Garcia-Verdugo, J.M., McEwen, B.S. and Alvarez- urosci., 19: 5792–5801.
Buylla, A. (2001) Astrocytes give rise to new neurons in Tozuka, Y., Fukuda, S., Namba, T., Seki, T. and Hisatsune, T.
the adult mammalian hippocampus. J. Neurosci., 21: (2005) GABAergic excitation promotes neuronal differenti-
7153–7160. ation in adult hippocampal progenitor cells. Neuron, 47:
Shen, Q., Goderie, S.K., Jin, L., Karanth, N., Sun, Y., 803–815.
Abramova, N., Vincent, P., Pumiglia, K. and Temple, S. Trejo, J.L., Carro, E. and Torres-Aleman, I. (2001) Circulating
(2004) Endothelial cells stimulate self-renewal and expand insulin-like growth factor I mediates exercise-induced in-
neurogenesis of neural stem cells. Science, 304: creases in the number of new neurons in the adult hippo-
1338–1340. campus. J. Neurosci., 21: 1628–1634.
Shors, T.J., Miesegaes, G., Beylin, A., Zhao, M., Rydel, T. and Wadiche, L.O., Bromberg, D.A., Bensen, A.L. and Westbrook,
Gould, E. (2001) Neurogenesis in the adult is involved in the G.L. (2005) GABAergic signaling to newborn neurons in
formation of trace memories. Nature, 410: 372–376. dentate gyrus. J. Neurophysiol., 94: 4528–4532.
Shors, T.J., Townsend, D.A., Zhao, M., Kozorovitskiy, Y. and Wang, S., Scott, B.W. and Wojtowicz, J.M. (2000) Heteroge-
Gould, E. (2002) Neurogenesis may relate to some but not all nous properties of dentate granule neurons in the adult rat. J.
types of hippocampal-dependent learning. Hippocampus, 12: Neurobiol., 42: 248–257.
578–584. Winocur, G., Wojtowicz, J.M., Sekeres, M., Snyder, J.S. and
Song, H., Stevens, C.F. and Gage, F.H. (2002a) Astroglia in- Wang, S. (2006) Inhibition of neurogenesis interferes with
duce neurogenesis from adult neural stem cells. Nature, 417: hippocampus-dependent memory function. Hippocampus,
39–44. 16: 296–304.
Song, H.J., Stevens, C.F. and Gage, F.H. (2002b) Neural stem Wurmser, A.E., Palmer, T.D. and Gage, F.H. (2004) Neuro-
cells from adult hippocampus develop essential properties of science: cellular interactions in the stem cell niche. Science,
functional CNS neurons. Nat. Neurosci., 5: 438–445. 304: 1253–1255.
Plate 28.2. Increased cell proliferation and altered dentate gyrus neuroblast migration after SE. (A, B) BrdU incorporation (arrows) in
adult rat dentate gyrus 7 days after saline (A) or pilocarpine (B) treatment followed by a 2 day post-BrdU survival. Subgranular zone
BrdU labeling increases markedly after SE (B) compared to the control (A). (C, D) Doublecortin (DCX) immunoreactivity in the
dentate gyrus of a control (Con; C) and an adult rat 14 days after pilocarpine-induced SE (14d; D). Note the increased hilar DCX
expression and chains of DCX-positive cells extending into the hilus (arrows in D) after SE. E, PSA-NCAM+ neuroblast chains (red,
arrows) alongside GFAP+ hilar astrocytes (green). Dashed lines in C–E denote the granule cell layer (gcl) — hilar (h) border. (For
B/W version, see page 534 in the volume.)

Plate 28.3. Schematic showing the effects of prolonged seizures on caudal subventricular zone (SVZ) and dentate gyrus progenitors. In
the intact adult rat (top), progenitors located in the infracallosal SVZ (purple) give rise to white matter oligodendrocytes (orange), while
those in the dentate gyrus (green) give rise to DGCs in the granule cell layer (gcl). After seizure-induced hilar and pyramidal cell layer
(pcl) injury (bottom; jagged arrows), progenitors migrate aberrantly from the caudal SVZ to form glia (a; purple) in the hippocampus
proper or from the dentate subgranular zone to form hilar-ectopic DGCs (b). (For B/W version, see page 535 in the volume.)
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 29

Unmasking recurrent excitation generated by mossy


fiber sprouting in the epileptic dentate gyrus: an
emergent property of a complex system

Thomas P. Sutula1, and F. Edward Dudek2

1
Department of Neurology H6/570 CSC, University of Wisconsin, 600 Highland Avenue, Madison, WI 53792, USA
2
Department of Physiology, University of Utah, Salt Lake City, UT, USA

Abstract: Seizure-induced sprouting of the mossy fiber pathway in the dentate gyrus has been observed
nearly universally in experimental models of limbic epilepsy and in the epileptic human hippocampus.
The observation of progressive mossy fiber sprouting induced by kindling demonstrated that even a few
repeated seizures are sufficient to alter synaptic connectivity and circuit organization. As it is now
recognized that seizures induce synaptic reorganization in hippocampal and cortical pathways, the im-
plications of seizure-induced synaptic reorganization for circuit properties and function have been subjects
of intense interest. Detailed anatomical characterization of the sprouted mossy fiber pathway has revealed
that the overwhelming majority of sprouted synapses in the inner molecular layer of the dentate gyrus form
recurrent excitatory connections, and are thus likely to contribute to recurrent excitation and potentially to
enhanced susceptibility to seizures. Nevertheless, difficulties in detecting functional abnormalities in circuits
reorganized by mossy fiber sprouting and the fact that some sprouted axons appear to form synapses with
inhibitory interneurons have been cited as evidence that sprouting may not contribute to seizure suscep-
tibility, but could form recurrent inhibitory circuits and be a compensatory response to prevent seizures.
Quantitative analysis of the synaptic connections of the sprouted mossy fiber pathway, assessment of the
functional features of sprouted circuitry using reliable physiological measures, and the perspective
of complex systems analysis of neural circuits strongly support the view that the functional effects of the
recurrent excitatory circuits formed by mossy fiber sprouting after seizures or injury emerge only con-
ditionally and intermittently, as observed with spontaneous seizures in human epilepsy. The recognition
that mossy fiber sprouting is induced after hippocampal injury and seizures and contributes conditionally
to emergence of recurrent excitation has provided a conceptual framework for understanding how injury
and seizure-induced circuit reorganization may contribute to paroxysmal network synchronization, epile-
ptogenesis, and the consequences of repeated seizures, and thus has had a major influence on understanding
of fundamental aspects of the epilepsies.

Keywords: hippocampus; seizures; epilepsy; synaptic reorganization; granule cells; plasticity

Corresponding author. Tel.: +1 608 263 5448; Fax:


+1 608 263 9405; E-mail: sutula@neurology.wisc.edu

DOI: 10.1016/S0079-6123(07)63029-5 541


542

Introduction Status epilepticus induced by chemoconvulsants or


electrical stimulation induces macroscopic damage
Axon sprouting and growth are important cellular in the hilus of the dentate gyrus, subfields of the
processes in the coordinated sequence of cell birth, hippocampus, and a variety of extrahippocampal
differentiation, migration, and neurite outgrowth regions, and is accompanied by reactive sprout-
that culminate in the formation of synapses, ing of the mossy fiber axons in the dentate gyrus
circuits, and networks in the developing nervous (Ben-Ari et al., 1979, 1980; Nadler et al., 1980b;
system. The earliest events in formation of neural Ben-Ari, 1985). Mossy fiber sprouting in these
circuits are genetically determined, but as devel- models is a consequence of both direct neurotoxic
opment proceeds, activity-dependent sprouting damage and excitotoxicity induced by status epi-
and axon growth play an important role in or- lepticus (Ben-Ari et al., 1980). The massive, mac-
ganizing and refining neural connections (Sur and roscopic damage associated with status epilepticus
Rubenstein, 2005). Electrical activity in immature in these models precluded definitely distinguishing
circuits promotes synapse formation and organizes whether axon sprouting and reorganization was a
synaptic connectivity into networks supporting consequence of direct injury, lesion-induced deaff-
adult function, behavior, and adaptation. The ob- erentation, seizures, or a combination of these
servation, nearly 20 years ago, that repeated brief processes.
seizures evoked by kindling induced sprouting
of the mossy fiber pathway in the dentate gyrus of
adult rats demonstrated that the activity-depend- Progressive mossy fiber sprouting induced by
ent processes of sprouting and axon growth re- repeated brief seizures
quired for circuit formation in the developing
nervous system also remodel and reorganize adult Axon sprouting was definitively identified as a
pathways after brief seizures (Sutula et al., 1988). consequence of the recurring brief seizures that
define epilepsy by the observation in kindled rats
that mossy fiber axons expanded their terminal
The association of axon sprouting with injury in field to the supragranular region of the dentate
neural circuits gyrus in the absence of overt damage or injury
(Sutula et al., 1988). Mossy fiber sprouting, as
Sprouting in the adult nervous system was first demonstrated by Timm histochemistry, develops
recognized in response to injury and damage. In after only a few brief seizures, progresses with re-
addition to the mossy fiber pathway, the septo- peated seizures, one is permanent (Sutula et al.,
hippocampal, associational, and crossed ent- 1988; Represa et al., 1989a; Cavazos et al., 1991,
orhinal pathways within the dentate gyrus and 1992). Further evidence that mossy fiber sprouting
hippocampal formation undergo sprouting and is routinely and reliably induced by seizures was
rearrangement of synaptic connectivity in response provided by studies indicating that sprouting is
to a variety of lesions and injuries (Steward et al., induced by seizures propagating into the hippo-
1973, 1974, 1976; Steward, 1976; Steward and campus from remote regions (Golarai et al., 1992).
Messenheimer, 1978; Nadler et al., 1980a, 1981; Mossy fiber sprouting has been observed in genetic
Staubli et al., 1984; Steward, 1992). The first sug- mouse models of epilepsy (Stanfield, 1989; Qiao
gestion of a link between seizures, sprouting, and and Noebels, 1993), and after seizures evoked by
circuit reorganization was the observation that maximal electroshock, flurothyl, pentylenetetrazol,
sprouted axons of the crossed entorhinal pathway intrahippocampal tetanus toxin, and by hype-
to the dentate gyrus gained access to hippocampal rthermia in immature rats (Golarai et al., 1992;
networks modified by kindling (Messenheimer et Holmes et al., 1998, 1999; Anderson et al., 1999;
al., 1979). Evidence of relationships between in- Gombos et al., 1999; Vaidya et al., 1999; Bender
jury, sprouting, and seizures was provided by et al., 2003). Mossy fiber sprouting has been ob-
studies of kainic acid-induced status epilepticus. served in primates after complex partial seizures
543

induced by alumina gel injection (Ribak et al., neural circuits (Cavazos and Sutula, 1990;
1998). Sprouting has been observed in CA3, CA1, Cavazos et al., 1991, 1994; Bengzon et al., 1997;
and neocortex (Represa and Ben-Ari, 1992; Salin Pretel et al., 1997; Zhang et al., 1998; Dalby
et al., 1995; Esclapez et al., 1999; Lehmann et al., et al., 1998; Haas et al., 2001; Kotloski et al.,
2001; Smith and Dudek, 2001; Cavazos et al., 2002). A single kindled seizure doubles the rate of
2004; Cross and Cavazos, 2007), and in the human apoptosis in the hilar neurons of the dentate gyrus
epileptic temporal lobe (Represa et al., 1989b; (Bengzon, 1997), and kindling induces apoptosis in
Sutula et al., 1989; Houser et al., 1990; El Bahh et CA1, subiculum, and neocortex (Pretel, 1997).
al., 1999). Axonal remodeling may be a continuing Cumulative neuronal loss has been detected after
process in chronic poorly controlled epilepsy as repeated evoked secondary generalized seizures
expression of growth associated proteins B-50 and evoked by kindling using stereological counting
polysialyated neural cell adhesion molecule (PSA- methods in the CA1, CA3 subfields of the hippo-
NCAM) are observed in resected human epileptic campus and the hilus of the dentate gyrus resem-
hippocampus (Mikkonen et al., 1998). Because bling classical human hippocampal sclerosis
mossy fiber sprouting is progressively induced by (Cavazos and Sutula, 1990; Cavazos et al., 1994;
repeated brief seizures, poorly controlled seizures Kotloski et al., 2002). As neuronal loss is progres-
in humans may induce progressive sprouting and sively induced by kindling and occurs in topo-
synaptic reorganization. graphically specific patterns in hippocampal and
limbic areas as a function of the site of seizure
initiation, mossy fiber sprouting after seizures in
Seizure-induced mossy fiber sprouting is the adult nervous system is closely associated with
accompanied by progressive neuronal loss seizure-induced neuronal loss, implying that even a
relatively small number of repeated seizures which
As sprouting is induced by status epilepticus with induce sprouting are potentially accompanied not
extensive injury and by direct damage to hippo- only by neuronal loss, but by a variety of cellular
campal pathways, and also by kindling which is alterations.
not accompanied by obvious overt damage, there
has been uncertainty about whether seizures can
induce sprouting in the absence of neuronal dam- Mossy fiber sprouting is associated with a variety of
age. Mossy fiber sprouting has been induced by seizure-induced cellular alterations
repetitive stimulation trains that evoke long-term
potentiation (Adams et al., 1997), and in associ- Repeated seizures not only induce mossy fiber
ation with spatial learning in behavioral tasks such sprouting and neuronal loss, but also a diverse
as the Morris water maze, which would not be range of cellular and molecular alterations that
expected to produce neuronal loss (Ramirez- may potentially modify functional properties
Amaya et al., 1999; Holahan et al., 2006). In the of neural circuits. The range of alterations in-
case of mossy fiber sprouting induced by seizures, duced in neurons and circuits by seizures spans
however, it is difficult to conclusively eliminate the molecular to network levels. Neurons are born and
possibility that mild neuronal loss may be occur- added to neural circuits in the dentate gyrus after
ring even with the most sensitive immunohisto- status epilepticus and only a few partial seizures
chemical and stereological techniques for detection induced by kindling stimulation (Parent et al.,
of apoptosis and neuronal loss. Studies from mul- 1998), integrate into local circuits, and become
tiple laboratories using techniques including unbi- activated during recurring seizures (Scharfman
ased estimates of neuronal numbers, TUNEL et al., 2002). Seizures alter dendritic branching
staining for apoptotic neurons, and silver impreg- and promote spine loss (Jiang M et al., 1998), and
nation for degenerating cells have demonstrated induce growth of basal dendrites by granule cells
that single and repeated brief seizures induce neu- of the dentate gyrus (Buckmaster and Dudek,
ronal death which contributes to reorganization of 1999; Ribak et al., 2000). Repeated brief seizures
544

evoked by kindling upregulate GFAP (glial fibril- deal of attention because it is easily detected by a
lary acidic protein) mRNA and protein levels in a variety of methods and is a prominent histological
time-dependent manner (Hansen et al., 1990; alteration, seizures also induce axon growth in the
Bonthius et al., 1995), cause glial cell hypertrophy hilus, promote infrapyramidal to suprapryamidal
and proliferation (Khurgel and Ivy, 1996; Stringer, (interblade) connections not observed in normal
1996). Kindled seizures modify vesicle release pro- rats, and expand the terminal field of the mossy
teins (Matveeva et al., 2006), alter extracellular fiber pathway in the inner molecular layer of the
matrix proteins (Niquet et al., 1995), and modify dentate gyrus along the septotemporal axis of the
motor cortex maps (van Rooyen et al., 2006). The hippocampus over distances as long as 700 mm
association of mossy fiber sprouting with a diverse (Sutula et al., 1998; Buckmaster and Dudek, 1999)
variety of activity-dependent seizure-induced al- (Fig. 1). As an example of the potential functional
terations in hippocampal pathways in both exper- effects of sprouting, reorganization along the
imental and human epilepsy has implications for septotemporal axis may have specific effects on
assessment of the functional effects of sprouting hippocampal dependent spatial memory, as right-
and these other associated alterations, and poses and left-specific place cells are organized in lamel-
significant experimental challenges for interpreta- lar patterns along this axis (Hampson et al., 1999).
tion of how and to what degree any one of these Afferent projections to the dentate gyrus are topo-
seizure-induced alterations contribute to func- graphically organized in the rat along the septo-
tional properties of reorganized neural circuits. temporal axis (Dolorfo and Amaral, 1998), and as
sprouting in the inner molecular layer varies along
this axis as a function of site of seizure induction,
Assessing the functional effects of mossy fiber
might disrupt the normal topographic organiza-
sprouting
tion of afferent inputs undergoing processing
in hippocampal circuitry, which would be expected
The functional effects of sprouting are unavoida-
to result in functional abnormalities including
bly confounded by numerous neuronal and glial
paradigm specific behavioral and cognitive
alterations. Despite this potential complexity, the
dysfunction.
pattern of alterations of the terminal field of the
reorganized sprouted pathway and the synaptic
targets of sprouted axons are important determi-
Functional implications of seizure-induced
nants of the functional effects of mossy fiber
alterations in synaptic targets of mossy fibers
sprouting. Detailed anatomic characterization of
the terminal field of the sprouted mossy pathway
Formation of synaptic contacts by sprouted axons
and quantitative morphological analysis of the
on principal cells or inhibitory interneurons deter-
synaptic targets of sprouted mossy fiber terminals
mines whether seizure-induced reorganization
have generated clear and unambiguous predictions
results in new excitatory or inhibitory neural
about the possible functional effects of the
circuits. There is considerable information about
sprouted pathway in hippocampal circuits.
the targets of mossy fiber axons in both normal
and reorganized circuitry after seizures. The
Functional implications of seizure-induced number of synapses formed on principal cells
alterations in the terminal field of the mossy fiber or interneurons determines whether the overall
pathway impact of the new circuits is likely to be net ex-
citation or inhibition. While both types of synaptic
Reorganization of terminal fields by sprouting contacts have been observed in both normal and
would be expected to have potentially significant epileptic hippocampus, the new synapses formed
functional consequences in neural networks. While by sprouted axons in the inner molecular layer
mossy fiber sprouting into the supragranular are predominantly on principal cells and form re-
region of the dentate gyrus has received a great current excitatory circuits (Franck et al., 1995;
545

Fig. 1. Anatomical features and alterations in terminal fields of sprouted mossy fibers in the reorganized dentate gyrus. (A) Sprouting
by a biocytin-filled mossy fiber axon of a granule cell from a rat that received kainic acid in the supragranular region and inner
molecular layer and growth of the axon plexus in the hilus with increased numbers of synaptic varicosities. In addition, growth of basal
dendrites is well documented, but not shown. (B) Sprouted axons not only extend into the supragranular and inner molecular layer, but
cross the hilus from the infrapyramidal blade and thereby increase connectivity between blades of the dentate gyrus. (C) Sprouted
mossy fiber axons establish connectivity in the supragranular layer along the septotemporal axis as demonstrated by a biocytin filled
granule cell in a saggital section extending from the septal to temporal pole of the hippocampus and dentate gyrus. (D) The sprouted
axons of the granule cell in (C) shown at higher magnification establish connectivity in the supragranular layer extending nearly 700 mm
along the septotemporal axis. Scale bars (in mm): A, B, C ¼ 50, D ¼ 500. Adapted with permission from Sutula et al., 1998.

Okazaki et al., 1995; Wenzel et al., 2000; Buck- et al., 1986; Lim et al., 1997; Zhang and Houser,
master et al., 2002; Cavazos et al., 2003) (Fig. 2). 1999; Gonzales et al., 2001). Although the ‘‘giant
Analysis of the synaptic targets of sprouted mossy mossy fiber terminal’’ is a characteristic distin-
fibers at the ultrastructural EM level has been guishing anatomical feature of mossy fiber axons
facilitated by the ability to identify mossy fiber in the normal dentate gyrus and hippocampus, the
terminals based on size, morphological features, more numerous terminals of mossy fibers of the
immunoreactivity to dynophin, and labeling by normal dentate gyrus are small and form asym-
Timm histochemistry (Zhang and Houser, 1999; metric excitatory contacts with mossy cells and
Buckmaster et al., 2002; Cavazos et al., 2003). The with inhibitory interneurons in the hilus (Acsady
typical mossy fiber in the normal dentate gyrus et al., 1998). About 98% of the small terminals in
forms about 8–10 giant terminals that have asym- the hilus are on inhibitory interneurons (Acsady
metric (excitatory) synaptic contacts with neurons et al., 1998). These observations imply that while
in the hilus and with pyramidal neurons in CA3 the net contribution of the mossy fiber pathway
(Amaral, 1979; Amaral and Dent, 1981; Claiborne to CA3 is excitatory, collaterals in the hilus may
546

Fig. 2. Examples of synaptic targets of sprouted mossy fibers in the reorganized dentate gyrus. Presynaptic terminals from sprouted
mossy fibers form axo-spinous, axo-dendritic, and axo-somatic synapses, and occasionally form synapses with interneurons. (A) A
presynaptic terminal of a sprouted mossy fiber axon labeled with silver grains from pre-embedding Timm methods in the inner
molecular layer of the dentate gyrus of a kindled rat that experienced 50 class V seizures forms a synapse with a spine head (s)
connected to a dendrite (d) in the inner molecular layer. The same terminal also makes contact with another dendrite in cross section.
(B) A presynaptic terminal labeled by dynorphin-A immunoreactivity from a sprouted mossy fiber axon forms an asymmetrical
synapse with a dendritic spine in the inner molecular layer of a rat that experienced 50 class V kindled seizures. The spine appears
perforated and contains a spine apparatus. (C) A dendrite (d) in the inner molecular layer of the dentate gyrus is extensively contacted
by synaptic terminals from sprouted mossy fiber axons labeled by preembedding Timm technique in a kindled rat that experienced 50
class V seizures. (D) A dentate basket cell with prominent nuclear infolding receives numerous abnormally large presynaptic terminals
(shown in inset) exhibiting the morphological characteristics of mossy fiber boutons (cytoplasm with tightly packed vesicles, mi-
tochondria, and dense core vesicles), which form asymmetrical synapses with the dendrite. (E) A sprouted mossy fiber presynaptic
terminal (arrow) labeled by dynorphin-A is apposed to the plasma membrane of the cell bodies of granule cells (gc) in the granule cell
layer of the dentate gyrus of a rat that experienced 50 class V kindled seizures. (F) The sprouted synaptic terminal in (E) shown at
higher magnification forms an asymmetric postsynaptic density on the granule cell. Scale bars (in mm): A, B, C, E, F ¼ 1; D ¼ 5,
inset ¼ 1.25. Adapted with permission from Cavazos et al. (2003, panels A, B, C, E, F), and Sloviter et al. (2006; panel D).

predominantly contribute to local recurrent inhib- features of the normal mossy fiber pathway indi-
itory circuits. In addition, mossy fibers in the nor- cate that functional contributions of asymmetric
mal dentate gyrus can form dense innervation excitatory terminals of mossy fibers to net excita-
patterns along interneurons extending dendrites tion and inhibition may vary systematically in
into the molecular layer (Ribak and Peterson, different regions of hippocampal circuitry, even
1991). These regionally specific anatomical from the hilus to CA3.
547

It should not be surprising that similar local The quantitative anatomical studies indicating
circuit complexity is encountered in the reorgan- that the overwhelming number of new synapses
ized dentate gyrus. Except for the occasional in- formed by sprouted mossy fibers in the inner mo-
terneuron dendrite that receives dense innervation lecular layer form new recurrent excitatory circuits
by mossy fibers in the normal dentate gyrus (Ribak are not incompatible with the fact that many
and Peterson, 1991), the terminal field of mossy mossy fiber terminals in the hilus of normal rats
fibers does not project to the inner molecular layer form recurrent inhibitory circuits (Acsady et al.,
and this region is therefore devoid of recurrent 1998), and that these circuits and potentially new
circuits. Seizures alter the terminal field and the mossy fiber connections to inhibitory interneurons
synaptic targets of mossy fibers in the inner mo- formed in the hilus by sprouted axons may con-
lecular layer by sprouting and growth of new tribute to inhibition in the reorganized dentate
mossy fiber axons which predominantly contact gyrus. The important point, however, is that new
excitatory principal cells and to a much lesser ex- recurrent excitatory circuits are formed in the
tent form synapses with interneurons (Frotscher inner molecular layer, a region that normally lacks
and Zimmer, 1983; Franck et al., 1995; Okazaki any recurrent excitatory connections among gran-
et al., 1995; Kotti et al., 1997; Wenzel et al., 2000; ule cells. It is noteworthy that some studies have
Pierce and Milner, 2001; Buckmaster et al., 2002; reported expression of the 67 kDa isoform of glut-
Cavazos et al., 2003; Sloviter et al., 2006) (Figs. 2 amic acid decarboxylase (GAD67, the synthetic
and 3). Quantitative analyses of the relative fre- enzyme for GABA) in mossy fiber terminals after
quency of synapses have provided unequivocal status epilepticus (Sloviter et al., 1996), suggesting
evidence that the overwhelming number of new that sprouted asymmetric MF synapses may
synapses are on granule cells and form excitatory contribute to inhibition in some circumstances
circuits. In the most comprehensive study of (Gutierrez and Heinemann, 2001), although the
synaptic targets of sprouted mossy fibers, 96% of overall physiological significance of this alteration
sprouted terminals contacted granule cell dendrites remains unclear. So while mossy fibers in the nor-
in the inner molecular layer and thus formed mal and reorganized dentate gyrus form regionally
recurrent excitatory circuits (Buckmaster et al., specific and heterogeneous synaptic connections
2002) (Fig. 3). In another study, the number with both CA3 pyramidal neurons and interneu-
of asymmetric excitatory axo-spinous synapses rons in the hilus, seizures importantly induce
characteristic of granule cell synapses exceeded a new and predominantly excitatory circuit among
asymmetric axo-dendritic shaft synapses putatively granule cells that does not exist in the normal
on inhibitory dendrites by nearly 2 to 1 in the dentate gyrus.
reorganized inner molecular layer of kainic acid-
treated rats (Cavazos et al., 2003). Perforated axo-
spinous synapses on granule cells are observed The paradox of epileptic circuits: episodic
in kindled rats (Geinisman et al., 1992; Cavazos dysfunction and synchronization in association with
et al., 2003), as well as axo-somatic synapses which permanent structural reorganization
are not typically observed in normal rats (Cavazos
et al., 2003). While some authors have emphasized Despite the abundant evidence for mossy fiber
the very occasional to rare synapses formed by sprouting and a variety of chronic structural and
sprouted mossy fibers on interneurons (Kotti molecular alterations in the epileptic hippocam-
et al., 1997; Sloviter et al., 2006), the overwhelm- pus, it has been surprisingly difficult to detect
ing anatomical evidence from the most compre- evidence for functional abnormalities in epileptic
hensive and quantitative analyses clearly indicates circuitry in normal physiological conditions both
that the predominant circuit formed by seizure- in vitro and in vivo, including in the resected
induced sprouting in the inner molecular layer is human epileptic hippocampus (Cronin and
a recurrent excitatory circuit (Coulter, 2002; Dudek, 1988; Cronin et al., 1992; Golarai and
Sutula, 2002; Dudek and Shao, 2004). Sutula, 1996; Wuarin and Dudek, 1996, 2001;
548

Fig. 3. Sprouted mossy fibers synapse with GABA-negative and GABA-positive targets identified by EM immunohistochemistry. The
predominant circuit formed by sprouted mossy fibers in the supragranular and inner molecular layer is a recurrent excitatory circuit.
(A) A biocytin-labeled axon (black) forms synaptic contacts (arrowheads) with two GABA-negative spines. Nearby GABA-positive
structures are labeled with 10 nm-diameter colloidal gold particles, and a GABA-positive axon terminal forms a symmetric synapse
(arrow) with a granule cell body. (B) A biocytin-labeled axon (black) in the molecular layer forms a synaptic contact (arrowhead) with a
GABA-positive dendritic shaft. (C) Sprouted mossy fibers synapse preferentially with GABA-negative dendritic spines. Example of a
reconstructed sprouted mossy fiber with synaptic contacts indicated by markers. Squares indicate that the postsynaptic target was a
dendritic spine; circles indicate a dendritic shaft. Open markers indicate that the postsynaptic target was GABA-negative; filled markers
indicate GABA-positive. Borders between strata (h, hilus; gcl, granule cell layer; ml, molecular layer) are indicated by lines. The vast
majority of synapses (93–96% in the granule cell layer and inner molecular layer respectively) form synapses with GABA-negative
structures demonstrating that sprouted mossy fibers in the supragranular and inner molecular predominantly form recurrent excitatory
circuits. Adapted with permission from Buckmaster et al., 2002.
549

Buckmaster and Dudek, 1997a, b, 1999; Lynch not detect simple direct relationships between
and Sutula, 2000; Sutula, 2002). Membrane prop- sprouting and outcome variables such as sponta-
erties and evoked responses in resected human neous seizure frequency, and led some to conclude
epileptic hippocampal slices in standard physio- that sprouting and other cellular alterations in
logical conditions appear surprisingly normal the reorganized epileptic hippocampus have neg-
(Kim et al., 1993; Dudek et al., 1995). Initial anal- ligible functional importance for epileptogenesis
yses of evoked field potentials in the dentate gyrus (Benardo, 2002).
of hippocampal slices from kainic acid-treated rats Contrary to this viewpoint, synaptic transmis-
and resected hippocampal slices from patients with sion in the terminal field of the sprouted mossy
temporal lobe epilepsy demonstrated an associa- fiber pathway has been directly demonstrated in
tion between the duration and complexity of vivo by current source density analysis in kindled
antidromically evoked extracellular field potentials rats (Golarai and Sutula, 1996) (Fig. 4A–E).
and the extent of mossy fiber sprouting examined An inward current (sink) corresponding spatially
by the Timm method (Tauck and Nadler, 1985; to the terminal field of the sprouted mossy fiber
Masukawa et al., 1992), but other studies revealed terminals in the inner molecular layer of
only minimal evidence for spontaneous epileptic the dentate gyrus develops in kindled rats at a
burst discharges in hippocampal slices from kainic latency consistent with disynaptic transmission
acid-treated epileptic rats (Cronin and Dudek, in response to perforant path stimulation. The
1988; Cronin et al., 1992; Nissinen et al., 2001). difficulty of detecting functional alterations in
Attempts to correlate the extent of mossy fiber chronically reorganized epileptic circuitry despite
sprouting with spontaneous seizures in chronic in multiple neuronal and circuit alterations and
vivo models and human epileptic hippocampus functional synaptic transmission in the terminal
also failed to reveal a direct relationship between field of the sprouted mossy fiber pathway might
sprouting and seizure frequency, and were inter- appear to be a paradox. This paradox should not
preted as evidence that mossy fiber sprouting does be entirely surprising, however, given the clinical
not contribute to abnormal excitation and epilepsy fact that patients with epilepsy manifest seizures
(Longo and Mello, 1998, 1999; Pitkanen et al., only sporadically and briefly, so continuous
2000; Nissinen et al., 2001). Studies indicating that evidence for network synchronization and un-
prevention of mossy fiber sprouting by co-admin- derlying imbalance of excitation and inhibition
istration of the protein synthesis inhibitor cyclo- should not be expected. Although chronic sus-
heximide did not prevent development of seizures ceptibility to recurring network synchronization
after pilocarpine (Longo and Mello, 1998, 1999), and seizures is the defining feature of epilepsy,
which were subsequently not replicated (Williams seizures are an infrequent emergent event even in
et al., 2002; Toyoda and Buckmaster, 2005), also intractable patients, and may not be associated
were cited as evidence that sprouting is not related with continuously detectable functional abnor-
to epileptogenesis. Other studies noting that malities (Litt et al., 2001; Litt and Echauz, 2002;
seizures could be induced in the absence of sprout- Worrell et al., 2004). The paradox of infrequent
ing similarly were interpreted as evidence that expression of spontaneous functional abnormal-
sprouting is not necessary for seizure induction by ity despite the presence of permanent structural
kindling (Armitage et al., 1998; Mohapel et al., alterations has implications for design and in-
2000), which should not be surprising given that terpretation of experiments seeking to identify
previous studies demonstrated that kindled functional abnormalities associated with epilep-
seizures can be evoked from limbic pathways tic circuits reorganized by sprouting and a vari-
when granule cells in the dentate gyrus have been ety of cellular alterations. The paradox has
selectively lesioned by colchicine (Dasheiff and produced confusion that is in part based on lack
McNamara, 1982; Sutula et al., 1986). Skepticism of understanding about emergent functional
about the potential importance of mossy fiber properties in complex systems such as neural
sprouting thus developed as multiple studies did circuits.
Fig. 4. Synaptic transmission in the terminal field of the sprouted mossy fiber pathway and unmasking of functional abnormalities in
the reorganized dentate gyrus with mossy fiber sprouting. (A) Surface plot of current source density (CSD) as function of depth and
time in the dentate gyrus of a rat that experienced three afterdischarges but prior to development of mossy fiber sprouting. Inward
currents (sinks) are upward and outward currents (sources) are downward. (B) Corresponding CSD contour plot demonstrates that the
inward current sink (solid lines) develops in the middle and outer molecular layer at short latency, and is followed by a lower amplitude
sink developing in the inner molecular layer at 16 ms. (C) Surface plot of CSD in the dentate gyrus of a rat with mossy fiber sprouting
after 105 class V seizures demonstrates a prominent inward current sink developing in the inner molecular layer at a latency of 9–10 ms
(arrow) corresponding to the terminal field of the sprouted mossy fiber pathway. (D) Corresponding CSD contour plot also dem-
onstrates the prominent inward current sink developing in the inner molecular layer at a latency of 9–10 ms (arrow). (E) Schematic
summary of the spatial and temporal features of inward current sinks corresponding to the terminal field of sprouted mossy fiber
pathway compared to normal control rats without sprouting. Black areas in the schematic dendrites are inward currents of highest
amplitude and gray areas are inward currents of lower amplitude. The initial inward current at 3 ms in both the normal and sprouted
dentate gyrus is in the distal dendrites (black area), corresponding to monosynaptic transmission in the perforant path. At 9–12 ms in
the normal DG, inward current are not yet developed in the proximal dendrites of the inner molecular layer. In the reorganized dentate
gyrus with sprouting, a large inward current develops in the proximal dendrites at 9–12 ms, which colocalizes with the laminar
distribution of sprouted mossy fiber terminals. This current is consistent with disynaptic transmission in the sprouted pathway in
response to perforant path stimulation. (F) Spontaneous epileptic bursts are unmasked in the dentate gyrus of hippocampal slices from
rats with mossy fiber sprouting in 10–30 mM bicuculline. The spontaneous bursts consist of prolonged negative shifts in the extra-
cellular field potential and synchronous action potentials. Slices without sprouting had small intracellular depolarizations followed by
hyperpolarizations and positive-going extracellular fields; no action potentials or extracellular population spikes were observed. In
slices with sprouting, large spontaneous depolarizations that evoked intracellular action potentials could occur synchronously with
slow extracellular negative shifts and population spikes. Arrows show expansion of traces. (G) Unmasking of functional abnormalities
in the reorganized dentate gyrus with mossy fiber sprouting by altering the extracellular ionic environment. In standard bathing
solution (3.5 mM K+), antidromic stimulation evoked epileptiform activity only in the presence of robust recurrent mossy fiber
growth. Increasing Timm scores indicate increased density of mossy fiber sprouting. For each value of the Timm score, raising [K+]o
increased the percentage of slices that responded with epileptiform activity. In slices with a Timm score of 1, epileptiform activity
appeared only when [K+]o was increased. Adapted with permission from Golarai and Sutula (1996; panels A–E), Cronin et al. (1992;
panel F), and Hardison et al. (2002; panel G).
551

Episodic paroxysmal abnormalities in epileptic inhibitory processes, and complicates efforts to


circuitry: the phenomena of unmasking and directly identify the contributions of new recurrent
emergent functional properties in reorganized excitatory circuits in the epileptic dentate gyrus.
circuitry How and when functional effects emerge at the
network and behavioral levels as a consequence
Evidence for functional abnormalities in brain of the abnormal excitatory disynaptic activation of
slices from experimental models of epilepsy exam- granule cell dendrites associated with sprouted
ined by in vitro physiological methods generally terminal field in the inner molecular layer is a crit-
emerges only when the reorganized circuitry ical question. The phenomena of unmasking and
is perturbed by alterations in the extracellular dynamic emergent properties must be considered
environment such as elevated [K+]o, or reduced in the design and interpretation of experiments
[Ca2+]o, or by disinhibition by GABAA antago- seeking to identify the functional consequences of
nists (Cronin et al., 1992; Wuarin and Dudek, sprouting and other circuit alterations associated
1996; Patrylo and Dudek, 1998; Molnar and with epilepsy. Emergent functional effects of new
Nadler, 1999; Patrylo et al., 1999; Hardison recurrent excitatory circuits formed by mossy fiber
et al., 2000; Lynch and Sutula, 2000; Wuarin and sprouting will be influenced by the state of inhi-
Dudek 2001; Sutula, 2002) (Fig. 4F and G). Sur- bition in the dentate gyrus, which must remain
gically resected human epileptic hippocampus sufficiently robust even in intractable epilepsy to
similarly appears normal unless the extracellular prevent runaway excitation and continuous syn-
environment is altered directly or in response to chronization. Dynamic alterations in inhibition
tetanic stimulation (Masukawa et al., 1989, 1992; are therefore a key factor in assessing func-
Wuarin et al., 1990; Kim et al., 1993; Dudek et al., tional effects of sprouting in reorganized epileptic
1995). These observations suggest that the reor- circuity of the dentate gyrus.
ganized circuits are a substrate for dysfunction,
but functional abnormality becomes unmasked
and emerges only in the context of some other The state of inhibition in the reorganized dentate
perturbation or transient alterations. These obser- gyrus: a critical variable for emergence of recurrent
vations in hippocampal slices and the fact that excitation
spontaneous seizures occur sporadically and
unpredictably in both experimental and human The state of GABAergic inhibition thus plays an
epilepsy indicate that seizures and underlying important and indeed critical role in regulating
epileptic functional abnormalities are emergent emergence of network synchronization and other
properties of epileptic neural circuitry altered functional abnormalities in both the normal and
by primary pathologies or by seizure-induced re- reorganized dentate gyrus. GABAergic neuro-
organization. transmission is not a unitary or simple process
but rather represents a diverse range of potentially
complex and dynamic cellular and physiological
Unmasking emergent functional abnormalities in mechanisms which undergo activity-dependent
recurrent sprouted circuits by modifying modification (McCarren and Alger, 1985;
GABAergic neurotransmission Brooks-Kayal et al., 1998; Coulter, 2001; Cossart
et al., 2005). In some circumstances, such as in
Functional abnormalities in reorganized epileptic early postnatal development and during acute
circuits including spontaneous synchronous net- conditions when intracellular Cl increases, GAB-
work discharge become unmasked by reducing Aergic neurotransmission can contribute to exci-
GABAergic inhibition. The fact that epileptic net- tation as well as cellular inhibition (Staley et al.,
work synchronization can emerge even in normal 1995; Khalilov et al., 1999; Leinekugel et al., 1999;
neural circuits when inhibition is reduced indicates Dzhala and Staley, 2003). Assessing the functional
the narrow dynamic balance of excitatory and effects of sprouting therefore requires analysis of
552

alterations in both inhibitory and excitatory with the GABA-dependent Cl mediated conduct-
synaptic transmission and their dynamic interac- ance change and provide an indirect measure
tions in circuits that have undergone reorganiza- of GABA-dependent inhibition of evoked granule
tion as a result of seizures and primary injury or cell discharge under appropriately controlled stim-
pathology. Dynamic alterations in excitatory and ulation and timing parameters (Tuff et al., 1983a,
inhibitory synaptic transmission and seizure-in- b; de Jonge and Racine, 1987; Stringer and
duced formation of new circuits vary significantly Lothman, 1989; Sayin et al., 2003). The increase
in different regions of the hippocampus and inde- in inhibition implied by the indirect measure of
pendently as a function of time after seizures. The paired pulse inhibition has been directly confirmed
time course of alterations in synaptic transmission by demonstration of increased frequency and am-
evolve over periods of as long as months after the plitude of IPSCs in granule cells (Otis et al., 1994;
last episode of network synchronization in circuits Buhl et al., 1996; Nusser et al., 1998). It is also
undergoing remodeling and reorganization, so ex- clear that inhibition after status epilepticus is
periments seeking to understand the emergence of increased in the dentate gyrus after an initial
functional abnormalities associated with sprouting period of depression (Hellier et al., 1999; Gorter
also need to evaluate how inhibition and altera- et al., 2002; Harvey and Sloviter, 2005; Sloviter
tions in connectivity systematically vary across et al., 2006).
these prolonged periods in order to understand Although inhibition in the dentate gyrus is
functional effects of sprouting and phenomena initially increased by kindling, inhibition is even-
such as latent periods after initial injuries. tually lost after 100 evoked Class V seizures
The normal dentate gyrus has powerful systems in association with emergence of spontaneous sei-
of inhibition as demonstrated by the fact that zures, as directly confirmed by reduction in
granule cells are resistant to repetitive discharge in amplitude and alterations in kinetics of the
response to synaptic activation, infrequently gen- monosynaptic IPSC measured by single electrode
erate spontaneous activity, and typically do not voltage clamp techniques (Sayin et al., 2003). The
generate repetitive discharges even when GABAA loss of inhibition and emergence of spontaneous
inhibition is blocked (Fricke and Prince, 1984; seizures is associated with reduction of CCK and
Lynch and Sutula, 2000; Lynch et al., 2000). The GAT-1 subpopulations of interneurons, which
state of inhibition in the dentate gyrus is predict- provide axo-somatic and axo-axonic inhibitory
ably modified by episodic network synchroniza- terminals on granule cells. These observations are
tion, specifically as a function of both the timing consistent with the view that recurrent excitatory
and duration of seizures. A variety of physiolog- circuits which are progressively formed by sprout-
ical and pharmacological methods both in vivo ing in kindled rats gradually increase capacity to
and in vitro have demonstrated that repeated brief generate recurrent excitation in the dentate gyrus
seizures evoked by kindling increase inhibition in but are insufficient at early stages of seizure-
the dentate gyrus (Tuff et al., 1983a, b; de Jonge induced reorganization to generate spontaneous
and Racine, 1987; Stringer and Lothman, 1989; seizures unless inhibition is reduced. With seizure-
Otis et al., 1994; Buhl et al., 1996; Nusser et al., induced loss of interneurons and critical axo-
1998; Sayin et al., 2003), while sustained seizures somatic and axo-axonic inhibitory terminals at
as during status epilepticus initially have opposite advanced stages of kindling, recurrent excitation
effects and produce rundown or acute reduction of generated by sprouted circuits among granule
inhibition (Kapur et al., 1989c; Kapur and Mac- cells, while still for the most part checked by
donald, 1997), which is later followed by increases inhibition, periodically becomes sufficient to drive
in inhibition in granule cells (Cohen et al., 2003). spontaneous emergent network synchronization.
After a few repeated brief seizures evoked during These studies in rats experiencing brief seizures
the early stages of kindling, paired pulse inhibition evoked by kindling and status epilepticus indicate
in the dentate gyrus is acutely increased at inter- that the state of inhibition in the dentate gyrus
pulse intervals of 15–40 ms, which are associated varies both acutely and chronically in response to
553

ongoing episodes of network synchronization as a Direct evidence for recurrent excitation unmasked
function of the duration, frequency, and number by disinhibition in the reorganized dentate gyrus
of seizures. Efforts to assess the potential func-
tional properties of recurrent circuits formed by The functional abnormalities that become
sprouting must be considered with recognition of unmasked in the reorganized dentate gyrus by
these dynamic alterations in inhibition as a con- reducing inhibition or altering the extracellular
sequence of timing and duration of recent seizures, ionic environment include direct evidence for
and pursued with a variety of electrophysiological development of recurrent excitation in association
techniques. with mossy fiber sprouting. Evidence for recurrent
In addition to the effects of timing and duration excitation in association with mossy fiber sprout-
of seizures, it is also clear from in vivo and in vitro ing has been demonstrated by focal stimulation by
studies that the effects of seizures on network glutamate application using microdrop or flash
inhibition vary significantly as function of location photolysis techniques (Wuarin and Dudek, 1996,
in the hippocampus, and that results in other 2001; Molnar and Nadler, 1999; Lynch and
regions of hippocampal and neocortical circuitry Sutula, 2000). Focal application of glutamate
do not necessarily apply to the dentate gyrus. microdrops to dendrites and cell bodies of gran-
The effects of seizures in the dentate gyrus specifi- ule cells remote from the recorded granule cell
cally contrast with CA1, where in vivo methods in hippocampal slices from normal rats evokes
have demonstrated reduction in inhibition after no responses when inhibition is blocked, but mi-
brief evoked seizures and status epilepticus (Kapur crodrop application in disinhibited slices with
et al., 1989a–c; Michelson et al., 1989; Gorter sprouting evokes EPSPs at long and variable
et al., 2002). In further contrast to the dentate latency (Wuarin and Dudek, 1996; Lynch and
gyrus, GABAA receptor dependent currents Sutula, 2000). In kindled rats with mossy fiber
in granule cells are increased after status epilep- sprouting, trains of EPSPs and population dis-
ticus induced by pilocarpine but are reduced charges can be evoked by glutamate microstimu-
in pyramidal neurons of CA1 (Gibbs et al., lation remote from the recording site at one week
1997), and dendritic GABAergic inhibition after induction of kindled seizures, when sprouting
in CA1 as measured by analysis of frequency is first detectable by Timm histochemistry, but are
and amplitude of IPSCs is reduced while somatic not evoked in hippocampal slices from kindled rats
inhibition is increased (Cossart et al., 2001). examined at 24 h after a single afterdischarge prior
Audiogenic kindling reduces GABAA depend- to the development of sprouting (Lynch and
ent IPCSs in the inferior colliculus (Evans et al., Sutula, 2000). Stimulation by flash photolysis of
2006), and increases amplitude and duration of caged glutamate at sites remote from the recording
mIPSCs in piriform cortex (Gavrilovici et al., site in granule cells also evokes EPSCs in granule
2006). These contrasting observations in the cells (Molnar and Nadler, 1999; Wuarin and
dentate gyrus, CA1, and other regions clearly Dudek, 2001) which increase in correlation with
demonstrate that assessment of inhibitory and ex- the development of sprouting (Fig. 5A and B). The
citatory processes in both normal and reorgan- long and variable latency of responses evoked by
ized circuitry cannot be simply characterized focal glutamate stimulation in these studies sug-
by a single technique or in a single model or in gested that recurrent excitation was generated by
a single region of hippocampal circuitry, multisynaptic rather than monosynaptic circuits,
and that multiple techniques are required to but monosynaptic EPSPs can be evoked at short
systematically characterize the effects of seizures latency (2.670.36 ms) between blades of the dent-
on these synaptic and circuit properties. This is ate gyrus under conditions in which recurrent
an important perspective for efforts to assess inhibitory circuits are blocked by bicuculline,
the effects of sprouting or any other putative polysynaptic activity is suppressed by 10 mM
cellular causes of seizures and chronic Ca2+ in the bathing medium, and perforant path
epilepsy. activation is prevented by knife cuts (Lynch and
554

Fig. 5. Formation of recurrent excitatory circuits in the reorganized dentate gyrus with mossy fiber sprouting. (A) Evidence for
development of excitatory connectivity in the reorganized dentate gyrus with mossy fiber sprouting. Photostimulation of caged
glutamate in the granule cell layer in conditions of reduced inhibition (bicuculline, 30 mM) and elevated [K+]o (6 mM) which evokes no
responses in the normal dentate gyrus evokes epileptiform bursts of action potentials in a granule cell from a rat 33 weeks after kainate
treatment. The patch pipette in the diagram indicates the position of the recorded granule cell in the outer blade. The numbers show the
locations of the photostimulations in the diagram and the corresponding evoked bursts of action potentials. The granule cell was
recorded in the whole cell current-clamp configuration at resting membrane potential. Arrowheads show stimulation artifact produced
by the flash. (B) Excitatory connectivity as revealed by flash photolysis increases with time after treatment with kainic acid in
association with increasing mossy fiber sprouting. Plot of the percentage of granule cells from saline- and kainate-injected animals
responding to photostimulation with an increase in EPSCs as a function of time after treatment. The number of cells tested is indicated
for each age group. (C) Monosynaptic excitatory connections between blades of the dentate gyrus in a hippocampal slice from a rat
treated with kainic acid. The transected transverse hippocampal slice contained only the infrapyramidal and suprapyramidal blades of
the dentate gyrus, the intervening hilus, a small sector of CA3c, and CA1. The transection removed the crest of the dentate gyrus,
which severed perforant path connections between the blades. In bathing solution containing 10 mM bicuculline to suppress inhibitory
postsynaptic potentials and 10 mM [Ca2+]o to suppress polysynaptic activity, stimulation of the molecular layer of the infrapyramidal
blade (site indicated by arrows) with a 100 ms constant-voltage pulse of 7 V evoked an EPSP in a granule cell in the suprapyramidal
blade (location indicated by the arrowhead). The latency of the EPSP was 2.7 ms. Focal electrical microstimulation evoked EPSPs in 5
of 18 suprapyramidal granule cells in kainic acid-treated rats with an average latency of 2.5970.36 ms. In contrast, stimulation of the
infrapyramidal blade in hippocampal slices from normal rats failed to evoke EPSPs in 15 suprapyramidal granule cells. Calibration
bars: 2 mV, 8 ms. (D) Monosynaptic connections directly demonstrated between granule cells in slices from rats with mossy fiber
sprouting. Recordings from a pair of simultaneously recorded granule cells are shown in (5D1): presynaptic neuron (top), postsynaptic
neuron (bottom). Intracellular current (a 150 ms rectangular current pulse; start and end of the pulse are marked by the dots) triggered
an action potential (AP) in the presynaptic cell. Immediately thereafter, a small depolarization occurred in the postsynaptic cell. An
arrow marks the capacitative artifact of the presynaptic cell’s AP. Calibration: presynaptic cell, 20 mV, 30 ms; postsynaptic cell, 4 mV,
30 ms. (5D2): recordings from the same pair of neurons with higher gain. Several postsynaptic responses are overlapped to show the
variability in the response to the presynaptic AP. Calibration: presynaptic cell, 20 mV, 4 ms; postsynaptic cell, 3 mV, 4 ms. (5D3): in a
different pair of granule cells, tonic intracellular current was used to depolarize both the putative presynaptic (top) and postsynaptic
(bottom) cells. A spontaneous AP in the presynaptic cell triggered an AP in the second cell. Membrane potentials: top, 55 mV;
bottom, 54 mV. Calibration: 15 mV, 25 ms. Adapted with permission from Wuarin and Dudek (2001; panel A), Hardison et al. (2000;
panel B), Lynch et al. (2000; panel C), and Scharfman et al. (2003; panel D).
555

Sutula, 2000) (Fig. 5C). These studies from mul- pathological process and the subsequent processes
tiple laboratories are evidence that seizures induce of seizure-induced plasticity following an initial
recurrent excitatory connectivity in the dentate episode of network synchronization are sustained
gyrus which emerges only under conditions of re- or permanent, recurring seizures that define epi-
duced inhibition or alterations in the extracellular lepsy can emerge. The strength and dynamic bal-
ionic environment. ance of excitatory and inhibitory transmission,
This physiological evidence for formation of which undergo activity-dependent alterations after
recurrent excitatory connections has been con- seizures in reorganizing circuitry with sprouting,
firmed by dual recordings from pairs of granule will evolve and vary as a function of time and
cells in the dentate gyrus reorganized by mossy duration of recent seizures. These potentially com-
fiber sprouting (Fig. 5D1–3). Examination of 903 plex interactions and the temporal relationships
granule cell pairs in hippocampal slices from of acute and chronic seizure-induced plasticity
epileptic pilocarpine-treated rats revealed monosy- support the view that recurrent excitation and net-
naptic connections in 1/150 granule cell pairs work synchronization be considered as emergent
which were not observed in any of 285 pairs from circuit properties in different experimental models.
normal animals (Scharfman et al., 2003). While the As the neurobiological phenomena of plasticity
number of pairs was small, the results directly associated with repeated network synchronization
confirm the presence of recurrent excitatory and seizures in neural circuits, the gradually evolv-
circuits in epileptic dentate gyrus when inhibition ing time course of circuit alteration induced
was intact, and leave open the possibility that by kindling has been informative for assessment
reduction in inhibition might reveal additional of how mossy fiber sprouting contributes to
evidence for recurrent excitatory circuits masked network dysfunction and epileptogenesis. Unlike
by powerful inhibition present in both the normal models of status epilepticus which produce
and reorganized dentate gyrus. massive initial damage followed by mossy fiber
sprouting, kindling induces gradually progressive
sprouting accompanied by incremental but cumu-
Dynamic and evolving plasticity in reorganized lative neuronal loss. These features of gradually
excitatory and inhibitory circuits of the epileptic evolving cumulative circuit alterations after brief
dentate gyrus: implications for assessment of the seizures evoked by kindling have enabled physio-
functional effects of sprouting logical analysis at different time points after
a range of induced seizures, and thereby provide
The molecular and cellular processes underlying an experimental opportunity to examine both
network inhibition and excitation may be altered acute and chronic effects of repeated network syn-
by primary pathologies that cause epilepsy. In chronization in normal and reorganized circuits
addition, network inhibition and excitation are with sprouting. Exploitation of these features
systematically altered as a consequence of repeated of kindling has enabled recognition of the distinct
network synchronization and seizures, i.e. by sei- contributions of increased NMDA dependent
zure-induced plasticity or kindling. Inhibition and synaptic currents, fading of inhibition in associa-
excitation undergo independent acute and long- tion with seizure-induced loss of interneuron sub-
term alterations that extend for as long as months populations, and identification of recurrent
after seizures. For example, in addition to the excitatory circuits at both early and advanced
rewiring of synaptic connections by mossy fiber stages of circuit reorganization (Golarai and
sprouting, enhancement of kainate-receptor and Sutula, 1996; Sayin et al., 1999, 2003; Lynch
NMDA-receptor dependent glutamatergic trans- et al., 2000; Lynch and Sutula, 2000). Similar
mission also results in alterations in excitatory experimental designs have been employed for
synaptic properties that may contribute to net- analysis of structural and functional alterations
work synchronization (Sayin et al., 1999; Behr in models of status epilepticus (Houser and
et al., 2001; Epsztein et al., 2005). If the primary Esclapez, 1996; Buckmaster and Dudek, 1997a, b,
556

1999; Cossart et al., 2001; Austin and Buck- component parts. The functional properties of
master, 2004; Peng et al., 2004; Peng and Houser, complex systems typically include emergent events
2005; Ang et al., 2006). Attempts to define the which are context dependent.
functional effects of sprouting in both kindling The heterogeneous etiologies of epilepsy and the
that gradually evolves into spontaneous seizures diverse variety of underlying molecular and cellu-
and after status epilepticus that more quickly lar mechanisms are consistent with the view that
induces spontaneous seizures after a latent period epilepsy is a ‘‘complex systems’’ disorder which
have provided some straightforward lessons about cannot be fully explained by simple understanding
the pitfalls of simple approaches in complex sys- of one or perhaps even a few processes or com-
tems. For example, efforts to characterize the ponents. Multiple molecular and physiological
functional effects of sprouting anticipating simple mechanisms work alone or together to promote
associations with the presence or absence of epileptogenesis (Dudek et al., 2002; Sutula, 2002;
recurrent excitatory or inhibitory circuits formed Dudek and Shao, 2004), as demonstrated by the
by sprouted mossy fibers which ignore independ- rapidly growing list of transgenic mice demon-
ently evolving activity-dependent plasticity in strating epilepsy. An example is the interaction
excitatory and inhibitory synaptic transmission between cellular mechanisms studied in an isolated
are subject to significant misinterpretation (Harvey neuron, and the profound functional alterations
and Sloviter, 2005; Sloviter et al., 2006). At both that may emerge when neurons with subtle abnor-
early and advanced stages of reorganization with malities are components of a neural network.
minimal or extensive sprouting, recurrent excita- Alterations that appear subtle at one level may
tion and seizures are emergent events in a complex produce cumulative and profound alterations
system. How recurrent excitatory circuits formed manifesting as emergent properties at another
by mossy fiber sprouting contribute to recurrent level, such as recurrent excitation and behavioral
excitation needs to be considered from the seizures.
perspective of complex systems. Emergent properties in complex systems may
not be associated with continuously detectable
abnormalities (Gallagher and Appenzeller, 1999;
Epilepsy as a ‘‘complex systems’’ disorder Sole and Goodwin, 2000). Application of reduc-
tionistic approaches to ‘‘complex systems’’ often
There is increasing awareness that biological sys- result in causal ‘‘gaps’’ between one level of
tems often behave as ‘‘complex systems’’, and that understanding and the next (Sole and Goodwin,
neural circuitry of the brain and dentate gyrus 2000). Phenomena at one level cannot be viewed in
fulfill criteria of a ‘‘complex system’’ (Koch and isolation, but need to be considered together with
Laurent, 1999). As the circuits formed by sprout- other processes at the same and at other levels.
ing are but one component of a complex system of These features apply to the interactions of inhibi-
molecular and cellular pathways in the reorganized tion and excitation generated at both synaptic and
dentate gyrus, when and how these new circuits are network levels in the circuitry of the normal and
activated and contribute to network function can reorganized dentate gyrus and hippocampus.
be anticipated to be governed by principles of Observations that spontaneous seizures are
‘‘complex systems’’. The phenomena of complex infrequent even in intractable epilepsy and func-
systems are well-known in engineering design, but tional abnormalities that become unmasked by
have received attention in biology only relatively disinhibition or extracellular ionic alterations in
recently (Gallagher and Appenzeller, 1999). Com- hippocampal slices from epileptic rats and human
plex systems are difficult to fully explain through epileptic temporal lobe fulfill the defining criteria
reductionistic understanding of their component for emergent properties of a complex system. Un-
parts, and typically demonstrate functional prop- derstanding of constituent parts of a complex sys-
erties that are context dependent and cannot tem, for example the functional properties
be readily predicted from isolated analysis of generated by the sprouted mossy fiber pathway,
557

requires not only precise knowledge of the parts, time dependent dynamic changes in neural circuits,
but also the context in which the part operates. but in addition manipulation of multiple variables
These principles are highly relevant to the inter- to isolate context dependent processes. This level
play of activity-dependent alterations in inhibition, of complexity demands detailed quantitative char-
unmasking of recurrent excitation, generation acterization of the structural features of sprouting
of episodic network synchronization, and recur- and reorganized circuitry, physiological assess-
ring behavioral seizures which are the defining ment using multiple techniques, and experimental
features of epilepsy and are prototypical emergent designs which recognize evolving and time
events in the complex system of neural circuitry in dependent alterations in neurons and circuits sub-
the dentate gyrus. ject to activity dependent modification across
intervals spanning milliseconds to months or
more. Analysis of relationships between structural
Recurrent excitation in the dentate gyrus alterations such as mossy fiber sprouting in the
reorganized by sprouting: an emergent property of a inner molecular layer of the dentate gyrus and
complex system context dependent emergent events such as recur-
rent excitation will be flawed if analysis is pursued
While the heterogeneity and complexity of epilepsy without appreciation of regional variations and
at the systems biology level is obvious, studies of critical quantitative features such as numbers
potential underlying mechanisms for seizures and of excitatory and inhibitory synapses and targets.
chronic epilepsies frequently employ experimental Attempting to define functional effects of sprout-
designs which anticipate linear relationships be- ing based on simple association, correlation,
tween a given alteration such as an ion channel or with anticipation of linear relationships will
mutation with altered biophysical properties and invariably overlook important context depend-
emergent properties such as network synchroniza- ent emergent properties and will be subject to
tion and behavioral seizures. At the circuit level, misinterpretation and flawed perspectives.
numerous past studies of specific molecular or cel- The preceding sections have reviewed some of
lular alterations associated with epilepsy have been the anatomical and physiological considerations
pursued as possible mechanistic ‘‘causes’’ of the that need to be addressed in order to characterize
disorder, and then dismissed when it is recognized the functional effects of sprouting in the reorgan-
that sporadically expressed emergent phenomena, ized dentate gyrus. Recurrent excitatory circuits
such as network synchronization or behavioral between granule cells are the predominant recur-
seizures, are not universally or linearly related to rent connection formed by sprouted mossy fiber
the specific defect (Dudek, 2002). This interpretive axons in the inner molecular layer of the reorgan-
flaw is common in epilepsy research, and has been ized dentate gyrus. The available quantitative
contributed to skepticism about the function and characterizations of the synaptic connections
importance of mossy fiber sprouting (Armitage of the sprouted mossy fiber pathway, assessment
et al., 1998; Longo and Mello, 1998, 1999; of the functional features of sprouted circuitry
Timofeeva and Peterson, 1999; Mohapel et al., using multiple physiological measures and exper-
2000; Nissinen et al., 2001; Romcy-Pereira and imental designs, and the perspective of complex
Garcia-Cairasco, 2003; Sloviter et al., 2006). systems analysis of neural circuits strongly support
Experiments seeking to define the functional the conclusion that mossy fiber sprouting induced
effects of mossy sprouting in the complex system in the inner molecular layer by seizures or injury
of the reorganized dentate gyrus must include forms predominantly recurrent excitatory circuits
design and interpretive perspectives recognizing in this region whose functional effects emerge only
that sprouting is but one alteration among the conditionally and intermittently. The episodic
many molecular and cellular alterations in epilep- emergence of recurrent excitation in this region
tic neural circuitry. Characterization of its effects of circuitry that is commonly involved in human
will require more than correlational assessment of focal and limbic epilepsy is a potentially important
558

functional property of circuitry reorganized by Anderson, A.E., Hrachovy, R.A., Antalffy, B.A., Armstrong,
primary pathologies and recurring seizures, D.L. and Swann, J.W. (1999) A chronic focal epilepsy with
mossy fiber sprouting follows recurrent seizures induced by
but should not be anticipated to be a necessary
intrahippocampal tetanus toxin injection in infant rats. Ne-
condition or absolute requirement for network uroscience, 92: 73–82.
synchronization underlying the heterogeneous Ang, C.W., Carlson, G.C. and Coulter, D.A. (2006) Massive
variety of epileptic syndromes. and specific dysregulation of direct cortical input to the hip-
pocampus in temporal lobe epilepsy. J. Neurosci., 26:
11850–11856.
Recurrent excitation as a functional effect of mossy Armitage, L.L., Mohapel, P., Jenkins, E.M., Hannesson, D.K.
fiber sprouting: experimental challenges and and Corcoran, M.E. (1998) Dissociation between mossy fiber
therapeutic opportunities sprouting and rapid kindling with low-frequency stimulation
of the amygdala. Brain Res., 781: 37–44.
Austin, J.E. and Buckmaster, P.S. (2004) Recurrent excitation
The dentate gyrus and hippocampus are promi-
of granule cells with basal dendrites and low interneuron
nently involved in the most common form of density and inhibitory postsynaptic current frequency in the
intractable human epilepsy. The recognition dentate gyrus of macaque monkeys. J. Comp. Neurol., 476:
of mossy fiber sprouting in experimental models 205–218.
and in human epileptic temporal lobe and appre- Behr, J., Heinemann, U. and Mody, I. (2001) Kindling induces
transient NMDA receptor-mediated facilitation of high-fre-
ciation of its capacity to episodically generate
quency input in the rat dentate gyrus. J. Neurophysiol., 85:
recurrent excitation has been a milestone for 2195–2202.
epilepsy research and a major influence on under- Benardo, L.S. (2002) Evidence against a pathogenic role for
standing of fundamental aspects of the epilepsies. mossy fiber sprouting. Epilepsy Curr., 2: 96–97.
The perspective that mossy fiber sprouting con- Ben-Ari, Y. (1985) Limbic seizure and brain damage produced
by kainic acid: mechanisms and relevance to human temporal
tributes conditionally to emergence of recurrent
lobe epilepsy. Neuroscience, 14(2): 375–403.
excitation provides a conceptual framework Ben-Ari, Y., Tremblay, E., Ottersen, O.P. and Meldrum, B.S.
for understanding how injury and seizure-induced (1980) The role of epileptic activity in hippocampal and ‘‘re-
circuit reorganization may contribute to paroxys- mote’’ cerebral lesions induced by kainic acid. Brain Res.,
mal network synchronization, epileptogenesis, and 191: 79–97.
Ben-Ari, Y., Tremblay, E., Ottersen, O.P. and Naquet, R.
the consequences of repeated seizures. The accom-
(1979) Evidence suggesting secondary epileptogenic lesions
plishments of nearly two decades of investigation after kainic acid: pre-treatment with diazepam reduces dis-
on mossy fiber sprouting set the stage for efforts tant but not local damage. Brain Res., 165: 362–365.
to modify seizure-induced sprouting and processes Bender, R.A., Dube, C., Gonzalez-Vega, R., Mina, E.W. and
of plasticity in the reorganizing dentate gyrus as Baram, T.Z. (2003) Mossy fiber plasticity and enhanced hip-
pocampal excitability, without hippocampal cell loss or al-
a major translational and therapeutic opportunity
tered neurogenesis, in an animal model of prolonged febrile
for epilepsy research. seizures. Hippocampus, 13: 399–412.
Bengzon, J., Kokaia, Z., Elmer, E., Nanobashvili, A., Ko-
kaia, M. and Lindvall, O. (1997) Apoptosis and prolifer-
References ation of dentate gyrus neurons after single and intermittent
limbic seizures. Proc. Natl. Acad. Sci. U.S.A., 94:
Acsady, L., Kamondi, A., Sik, A., Freund, T. and Buzsaki, G. 10432–10437.
(1998) GABAergic cells are the major postsynaptic targets of Bonthius, D.J., Lothman, E.W. and Steward, O. (1995) The
mossy fibers in the rat hippocampus. J. Neurosci., 18: role of extracellular ionic changes in upregulating the mRNA
3386–3403. for glial fibrillary acidic protein following spreading depres-
Adams, B., Lee, M., Fahnestock, M. and Racine, R.J. (1997) sion. Brain Res., 674: 314–328.
Long-term potentiation trains induce mossy fiber sprouting. Brooks-Kayal, A.R., Shumate, M.D., Jin, H., Rikhter, T.Y.
Brain Res., 775: 193–197. and Coulter, D.A. (1998) Selective changes in single cell
Amaral, D.G. (1979) Synaptic extensions from the mossy fibers GABA(A) receptor subunit expression and function in tem-
of the fascia dentata. Anat. Embryol., 155: 241–251. poral lobe epilepsy. Nat. Med., 4: 1166–1172.
Amaral, D.G. and Dent, J.A. (1981) Development of the mossy Buckmaster, P.S. and Dudek, F.E. (1997a) Network properties
fibers of the dentate gyrus: I. A light and electron microscopic of the dentate gyrus in epileptic rats with hilar neuron loss
study of the mossy fibers and their expansions. J. Comp. and granule cell axon reorganization. J. Neurophysiol., 77:
Neurol., 195: 51–86. 2685–2696.
559

Buckmaster, P.S. and Dudek, F.E. (1997b) Neuron loss, gran- Coulter, D.A. (2002) Sprouted mossy fibers form primarily
ule cell axon reorganization, and functional changes in the excitatory connections. Epilepsy Curr., 2: 194–195.
dentate gyrus of epileptic kainate-treated rats. J. Comp. Cronin, J. and Dudek, F.E. (1988) Chronic seizures and col-
Neurol., 385: 385–404. lateral sprouting of dentate mossy fibers after kainic acid
Buckmaster, P.S. and Dudek, F.E. (1999) In vivo intracellular treatment in rats. Brain Res., 474: 181–184.
analysis of granule cell axon reorganization in epileptic rats. Cronin, J., Obenaus, A., Houser, C.R. and Dudek, F.E. (1992)
J. Neurophysiol., 81: 712–721. Electrophysiology of dentate granule cells after kainate-
Buckmaster, P.S., Zhang, G.F. and Yamawaki, R. (2002) Axon induced synaptic reorganization of the mossy fibers. Brain
sprouting in a model of temporal lobe epilepsy creates a pre- Res., 573: 305–310.
dominantly excitatory feedback circuit. J. Neurosci., Cross, D.J. and Cavazos, J.E. (2007) Synaptic reorganization
22: 6650–6658. in subiculum and CA3 after early-life status epilepticus
Buhl, E.H., Otis, T.S. and Mody, I. (1996) Zinc-induced col- in the kainic acid rat model. Epilepsy Res.,
lapse of augmented inhibition by GABA in a temporal lobe 73(2): 156–165.
epilepsy model. Science, 271: 369–373. Dalby, N.O., West, M. and Finsen, B. (1998) Hilar somatosta-
Cavazos, J.E., Das, I. and Sutula, T.P. (1994) Neuronal loss tin-mRNA containing neurons are preserved after perforant
induced in limbic pathways by kindling: evidence for induc- path kindling in the rat. Neurosci. Lett., 255: 45–48.
tion of hippocampal sclerosis by repeated brief seizures. J. Dasheiff, R.M. and McNamara, J.O. (1982) Intradentate col-
Neurosci., 14: 3106–3121. chicine retards the development of amygdala kindling. Ann.
Cavazos, J.E., Golarai, G. and Sutula, T.P. (1991) Mossy fiber Neurol., 11: 347–352.
synaptic reorganization induced by kindling: time course Dolorfo, C.L. and Amaral, D.G. (1998) Entorhinal cortex of
of development, progression, and permanence. J. Neurosci., the rat: topographic organization of the cells of origin of the
11: 2795–2803. perforant path projection to the dentate gyrus. J. Comp.
Cavazos, J.E., Golarai, G. and Sutula, T.P. (1992) Septotem- Neurol., 398: 25–48.
poral variation of the supragranular projection of the mossy Dudek, F.E. and Shao, L.R. (2004) Mossy fiber sprouting and
fiber pathway in the dentate gyrus of normal and kindled recurrent excitation: direct electrophysiologic evidence and
rats. Hippocampus, 2: 363–372. potential implications. Epilepsy Curr., 4: 184–187.
Cavazos, J.E., Jones, S.M. and Cross, D.J. (2004) Sprouting Dudek, F.E., Staley, K.J. and Sutula, T.P. (2002) The search
and synaptic reorganization in the subiculum and CA1 region for animal models of epileptogenesis and pharmacoresist-
of the hippocampus in acute and chronic models of partial- ance: are there biologic barriers to simple validation strate-
onset epilepsy. Neuroscience, 126: 677–688. gies? Epilepsia, 43: 1275–1277.
Cavazos, J.E. and Sutula, T.P. (1990) Progressive neuronal loss Dudek, F.E., Wuarin, J.P., Tasker, J.G., Kim, Y.I. and Pea-
induced by kindling: a possible mechanism for mossy fiber cock, W.J. (1995) Neurophysiology of neocortical slices
synaptic reorganization and hippocampal sclerosis. Brain resected from children undergoing surgical treatment for
Res., 527: 1–6. epilepsy. J. Neurosci. Methods, 59: 49–58.
Cavazos, J.E., Zhang, P., Qazi, R. and Sutula, T.P. (2003) Dzhala, V.I. and Staley, K.J. (2003) Excitatory actions of
Ultrastructural features of sprouted mossy fiber synapses endogenously released GABA contribute to initiation of ictal
in kindled and kainic acid-treated rats. J. Comp. Neurol., epileptiform activity in the developing hippocampus. J.
458: 272–292. Neurosci., 23: 1840–1846.
Claiborne, B.J., Amaral, D.G. and Cowan, W.M. (1986) A light El Bahh, B., Lespinet, V., Lurton, D., Coussemacq, M., Le Gal
and electron microscopic analysis of the mossy fibers of the La Salle, G. and Rougier, A. (1999) Correlations between
rat dentate gyrus. J. Comp. Neurol., 246: 435–458. granule cell dispersion, mossy fiber sprouting, and hippo-
Cohen, A.S., Lin, D.D., Quirk, G.L. and Coulter, D.A. (2003) campal cell loss in temporal lobe epilepsy. Epilepsia, 40:
Dentate granule cell GABA(A) receptors in epileptic hippo- 1393–1401.
campus: enhanced synaptic efficacy and altered pharmacol- Epsztein, J., Represa, A., Jorquera, I., Ben-Ari, Y. and Crepel,
ogy. Eur. J. Neurosci., 17: 1607–1616. V. (2005) Recurrent mossy fibers establish aberrant kainate
Cossart, R., Bernard, C. and Ben-Ari, Y. (2005) Multiple facets receptor-operated synapses on granule cells from epileptic
of GABAergic neurons and synapses: multiple fates rats. J. Neurosci., 25: 8229–8239.
of GABA signalling in epilepsies. Trends Neurosci., Esclapez, M., Hirsch, J.C., Ben-Ari, Y. and Bernard, C. (1999)
28: 108–115. Newly formed excitatory pathways provide a substrate for
Cossart, R., Dinocourt, C., Hirsch, J.C., Merchan-Perez, A., hyperexcitability in experimental temporal lobe epilepsy. J.
De Felipe, J., Ben-Ari, Y., Esclapez, M. and Bernard, C. Comp. Neurol., 408: 449–460.
(2001) Dendritic but not somatic GABAergic inhibition Evans, M.S., Cady, C.J., Disney, K.E., Yang, L. and Lag-
is decreased in experimental epilepsy. Nat. Neurosci., uardia, J.J. (2006) Three brief epileptic seizures reduce in-
4: 52–62. hibitory synaptic currents, GABA(A) currents, and
Coulter, D.A. (2001) Epilepsy-associated plasticity in gamma- GABA(A)-receptor subunits. Epilepsia, 47: 1655–1664.
aminobutyric acid receptor expression, function, and inhib- Franck, J.E., Pokorny, J., Kunkel, D.D. and Schwartzkroin,
itory synaptic properties. Int. Rev. Neurobiol., 45: 237–252. P.A. (1995) Physiologic and morphologic characteristics of
560

granule cell circuitry in human epileptic hippocampus. Epile- Hardison, J.L., Okazaki, M.M. and Nadler, J.V. (2000) Mod-
psia, 36: 543–558. est increase in extracellular potassium unmasks effect
Fricke, R.A. and Prince, D.A. (1984) Electrophysiology of of recurrent mossy fiber growth. J. Neurophysiol.,
dentate gyrus granule cells. J. Neurophysiol., 51: 195–209. 84: 2380–2389.
Frotscher, M. and Zimmer, J. (1983) Lesion-induced mossy Harvey, B.D. and Sloviter, R.S. (2005) Hippocampal granule
fibers to the molecular layer of the rat fascia dentata: iden- cell activity and c-Fos expression during spontaneous sei-
tification of postsynaptic granule cells by the Golgi-EM zures in awake, chronically epileptic, pilocarpine-treated rats:
technique. J. Comp. Neurol., 215: 299–311. implications for hippocampal epileptogenesis. J. Comp.
Gallagher, R. and Appenzeller, T. (1999) Beyond reductionism. Neurol., 488: 442–463.
Science, 284: 79. Hellier, J.L., Patrylo, P.R., Dou, P., Nett, M., Rose, G.M. and
Gavrilovici, C., D’Alfonso, S., Dann, M. and Poulter, M.O. Dudek, F.E. (1999) Assessment of inhibition and epilepti-
(2006) Kindling-induced alterations in GABA(A) receptor- form activity in the septal dentate gyrus of freely behaving
mediated inhibition and neurosteroid activity in the rat rats during the first week after kainate treatment. J. Neuro-
piriform cortex. Eur. J. Neurosci., 24: 1373–1384. sci., 19: 10053–10064.
Geinisman, Y., Morrell, F. and deToledo-Morrell, L. (1992) Holahan, M.R., Rekart, J.L., Sandoval, J. and Routtenberg, A.
Increase in the number of axospinous synapses with (2006) Spatial learning induces presynaptic structural
segmented postsynaptic densities following hippocampal remodeling in the hippocampal mossy fiber system of two
kindling. Brain Res., 569: 341–347. rat strains. Hippocampus, 16: 560–570.
Gibbs III, J.W., Shumate, M.D. and Coulter, D.A. (1997) Holmes, G.L., Gairsa, J.L., Chevassus-Au-Louis, N. and Ben-
Differential epilepsy-associated alterations in postsynaptic Ari, Y. (1998) Consequences of neonatal seizures in the rat:
GABA(A) receptor function in dentate granule and CA1 morphological and behavioral effects. Ann. Neurol.,
neurons. J. Neurophysiol., 77: 1924–1938. 44: 845–857.
Golarai, G., Cavazos, J.E. and Sutula, T.P. (1992) Activation of Holmes, G.L., Sarkisian, M., Ben-Ari, Y. and Chevassus-Au-
the dentate gyrus by pentylenetetrazol evoked seizures Louis, N. (1999) Mossy fiber sprouting after recurrent sei-
induces mossy fiber synaptic reorganization. Brain Res., zures during early development in rats. J. Comp. Neurol.,
593: 257–264. 404: 537–553.
Golarai, G. and Sutula, T.P. (1996) Functional alterations in Houser, C.R. and Esclapez, M. (1996) Vulnerability and plas-
the dentate gyrus after induction of long-term potentiation, ticity of the GABA system in the pilocarpine model of spon-
kindling, and mossy fiber sprouting. J. Neurophysiol., taneous recurrent seizures. Epilepsy Res., 26: 207–218.
75: 343–353. Houser, C.R., Miyashiro, J.E., Swartz, B.E., Walsh, G.O.,
Gombos, Z., Spiller, A., Cottrell, G.A., Racine, R.J. and Rich, J.R. and Delgado-Escueta, A.V. (1990) Altered
McIntyre Burnham, W. (1999) Mossy fiber sprouting induced patterns of dynorphin immunoreactivity suggest mossy fiber
by repeated electroconvulsive shock seizures. Brain Res., reorganization in human hippocampal epilepsy. J. Neurosci.,
844: 28–33. 10: 267–282.
Gonzales, R.B., DeLeon Galvan, C.J., Rangel, Y.M. and Clai- Jiang, M., Lee, C.L., Smith, K.L. and Swann, J.W. (1998) Spine
borne, B.J. (2001) Distribution of thorny excrescences on loss and other persistent alterations of hippocampal pyram-
CA3 pyramidal neurons in the rat hippocampus. J. Comp. idal cell dendrites in a model of early-onset epilepsy. J.
Neurol., 430: 357–368. Neurosci., 18: 8356–8368.
Gorter, J.A., van Vliet, E.A., Aronica, E. and Lopes da Silva, de Jonge, M. and Racine, R.J. (1987) The development and
F.H. (2002) Long-lasting increased excitability differs in decay of kindling-induced increases in paired-pulse depres-
dentate gyrus vs. CA1 in freely moving chronic epileptic rats sion in the dentate gyrus. Brain Res., 412: 318–328.
after electrically induced status epilepticus. Hippocampus, Kapur, J., Bennett Jr., J.P., Wooten, G.F. and Lothman, E.W.
12: 311–324. (1989a) Evidence for a chronic loss of inhibition in the hip-
Gutierrez, R. and Heinemann, U. (2001) Kindling induces pocampus after kindling: biochemical studies. Epilepsy Res.,
transient fast inhibition in the dentate gyrus — CA3 projec- 4: 100–108.
tion. Eur. J. Neurosci., 13: 1371–1379. Kapur, J. and Macdonald, R.L. (1997) Rapid seizure-induced
Haas, K.Z., Sperber, E.F., Opanashuk, L.A., Stanton, P.K. and reduction of benzodiazepine and Zn2+ sensitivity of hippo-
Moshe, S.L. (2001) Resistance of immature hippocampus campal dentate granule cell GABA(A) receptors. J. Neuro-
to morphologic and physiologic alterations following status sci., 17: 7532–7540.
epilepticus or kindling. Hippocampus, 11: 615–625. Kapur, J., Michelson, H.B., Buterbaugh, G.G. and Lothman,
Hampson, R.E., Simeral, J.D. and Deadwyler, S.A. (1999) E.W. (1989b) Evidence for a chronic loss of inhibition in
Distribution of spatial and nonspatial information in dorsal the hippocampus after kindling: electrophysiological studies.
hippocampus. Nature, 402: 610–614. Epilepsy Res., 4: 90–99.
Hansen, A., Jorgensen, O.S., Bolwig, T.G. and Barry, D.I. Kapur, J., Stringer, J.L. and Lothman, E.W. (1989c) Evidence
(1990) Hippocampal kindling alters the concentration of glial that repetitive seizures in the hippocampus cause a lasting
fibrillary acidic protein and other marker proteins in rat reduction of GABAergic inhibition. J. Neurophysiol.,
brain. Brain Res., 531: 307–311. 61: 417–426.
561

Khalilov, I., Dzhala, V., Ben-Ari, Y. and Khazipov, R. (1999) antidromically evoked field responses of the dentate gyrus
Dual role of GABA in the neonatal rat hippocampus. Dev. and mossy fiber reorganization in temporal lobe epileptic
Neurosci., 21: 310–319. patients. Brain Res., 579: 119–127.
Khurgel, M. and Ivy, G.O. (1996) Astrocytes in kindling: Matveeva, E.A., Vanaman, T.C., Whiteheart, S.W. and Slevin,
relevance to epileptogenesis. Epilepsy Res., 26: 163–175. J.T. (2006) Asymmetric accumulation of hippocampal 7S
Kim, Y.I., Peacock, W.J. and Dudek, F.E. (1993) Properties SNARE complexes occurs regardless of kindling paradigm.
and synaptic mechanisms of bicuculline-induced epileptiform Epilepsy Res., 73(3): 266–274.
bursts in neocortical slices from children with intractable McCarren, M. and Alger, B.E. (1985) Use-dependent depres-
epilepsy. J. Neurophysiol., 70: 1759–1766. sion of IPSPs in rat hippocampal pyramidal cells in vitro. J.
Koch, C. and Laurent, G. (1999) Complexity and the nervous Neurophysiol., 53: 557–571.
system. Science, 284: 96–98. Messenheimer, J.A., Harris, E.W. and Steward, O. (1979)
Kotloski, R., Lynch, M., Lauersdorf, S. and Sutula, T. (2002) Sprouting fibers gain access to circuitry transsynaptically al-
Repeated brief seizures induce progressive hippocampal neu- tered by kindling. Exp. Neurol., 64: 469–481.
ron loss and memory deficits. Prog. Brain Res., 135: 95–110. Michelson, H.B., Kapur, J. and Lothman, E.W. (1989) Reduc-
Kotti, T., Riekkinen Sr., P.J. and Miettinen, R. (1997) Char- tion of paired pulse inhibition in the CA1 region of the hip-
acterization of target cells for aberrant mossy fiber collaterals pocampus by pilocarpine in naive and in amygdala-kindled
in the dentate gyrus of epileptic rat. Exp. Neurol., rats. Exp. Neurol., 104: 264–271.
146: 323–330. Mikkonen, M., Soininen, H., Kalvianen, R., Tapiola, T., Ylin-
Lehmann, T.N., Gabriel, S., Eilers, A., Njunting, M., Kovacs, en, A., Vapalahti, M., Paljarvi, L. and Pitkanen, A. (1998)
R., Schulze, K., Lanksch, W.R. and Heinemann, U. (2001) Remodeling of neuronal circuitries in human temporal lobe
Fluorescent tracer in pilocarpine-treated rats shows wide- epilepsy: increased expression of highly polysialylated neural
spread aberrant hippocampal neuronal connectivity. Eur. J. cell adhesion molecule in the hippocampus and the ent-
Neurosci., 14: 83–95. orhinal cortex. Ann. Neurol., 44: 923–934.
Leinekugel, X., Khalilov, I., McLean, H., Caillard, O., Gaiarsa, Mohapel, P., Armitage, L.L., Gilbert, T.H., Hannesson, D.K.,
J.L., Ben-Ari, Y. and Khazipov, R. (1999) GABA is the Teskey, G.C. and Corcoran, M.E. (2000) Mossy fiber sprout-
principal fast-acting excitatory transmitter in the neonatal ing is dissociated from kindling of generalized seizures in the
brain. Adv. Neurol., 79: 189–201. guinea-pig. Neuroreport, 11: 2897–2901.
Lim, C., Blume, H.W., Madsen, J.R. and Saper, C.B. (1997) Molnar, P. and Nadler, J.V. (1999) Mossy fiber-granule cell
Connections of the hippocampal formation in humans: I. The synapses in the normal and epileptic rat dentate gyrus studied
mossy fiber pathway. J. Comp. Neurol., 385: 325–351. with minimal laser photostimulation. J. Neurophysiol., 82:
Litt, B. and Echauz, J. (2002) Prediction of epileptic seizures. 1883–1894.
Lancet Neurol., 1: 22–30. Nadler, J.V., Perry, B.W. and Cotman, C.W. (1980a) Selective
Litt, B., Esteller, R., Echauz, J., D’Alessandro, M., Shor, R., reinnervation of hippocampal area CA1 and the fascia den-
Henry, T., Pennell, P., Epstein, C., Bakay, R., Dichter, M. tata after destruction of CA3-CA4 afferents with kainic acid.
and Vachtsevanos, G. (2001) Epileptic seizures may begin Brain Res., 182: 1–9.
hours in advance of clinical onset: a report of five patients. Nadler, J.V., Perry, B.W., Gentry, C. and Cotman, C.W.
Neuron, 30: 51–64. (1980b) Degeneration of hippocampal CA3 pyramidal cells
Longo, B.M. and Mello, L.E. (1998) Supragranular mossy fiber induced by intraventricular kainic acid. J. Comp. Neurol.,
sprouting is not necessary for spontaneous seizures in the 192: 333–359.
intrahippocampal kainate model of epilepsy in the rat. Nadler, J.V., Perry, B.W., Gentry, C. and Cotman, C.W. (1981)
Epilepsy Res., 32: 172–182. Fate of the hippocampal mossy fiber projection after de-
Longo, B.M. and Mello, L.E. (1999) Effect of long-term spon- struction of its postsynaptic targets with intraventricular
taneous recurrent seizures or reinduction of status epilepticus kainic acid. J. Comp. Neurol., 196: 549–569.
on the development of supragranular mossy fiber sprouting. Niquet, J., Jorquera, I., Faissner, A., Ben-Ari, Y. and Represa,
Epilepsy Res., 36: 233–241. A. (1995) Gliosis and axonal sprouting in the hippocampus
Lynch, M., Sayin, U., Golarai, G. and Sutula, T. (2000) of epileptic rats are associated with an increase of tenascin-C
NMDA receptor-dependent plasticity of granule cell spiking immunoreactivity. J. Neurocytol., 24: 611–624.
in the dentate gyrus of normal and epileptic rats. J. Nissinen, J., Lukasiuk, K. and Pitkanen, A. (2001) Is mossy
Neurophysiol., 84: 2868–2879. fiber sprouting present at the time of the first spontaneous
Lynch, M. and Sutula, T. (2000) Recurrent excitatory connec- seizures in rat experimental temporal lobe epilepsy? Hippo-
tivity in the dentate gyrus of kindled and kainic acid-treated campus, 11: 299–310.
rats. J. Neurophysiol., 83: 693–704. Nusser, Z., Hajos, N., Somogyi, P. and Mody, I. (1998)
Masukawa, L.M., Higashima, M., Kim, J.H. and Spencer, Increased number of synaptic GABA(A) receptors underlies
D.D. (1989) Epileptiform discharges evoked in hippocampal potentiation at hippocampal inhibitory synapses. Nature,
brain slices from epileptic patients. Brain Res., 493: 168–174. 395: 172–177.
Masukawa, L.M., Uruno, K., Sperling, M., O’Connor, M.J. Okazaki, M.M., Evenson, D.A. and Nadler, J.V. (1995) Hip-
and Burdette, L.J. (1992) The functional relationship between pocampal mossy fiber sprouting and synapse formation after
562

status epilepticus in rats: visualization after retrograde trans- Ribak, C.E., Seress, L., Weber, P., Epstein, C.M., Henry, T.R.
port of biocytin. J. Comp. Neurol., 352: 515–534. and Bakay, R.A. (1998) Alumina gel injections into the tem-
Otis, T.S., De Koninck, Y. and Mody, I. (1994) Lasting poral lobe of rhesus monkeys cause complex partial seizures
potentiation of inhibition is associated with an increased and morphological changes found in human temporal lobe
number of gamma-aminobutyric acid type A receptors acti- epilepsy. J. Comp. Neurol., 401: 266–290.
vated during miniature inhibitory postsynaptic currents. Ribak, C.E., Tran, P.H., Spigelman, I., Okazaki, M.M. and
Proc. Natl. Acad. Sci. U.S.A., 91: 7698–7702. Nadler, J.V. (2000) Status epilepticus-induced hilar basal
Parent, J.M., Janumpalli, S., McNamara, J.O. and Lowenstein, dendrites on rodent granule cells contribute to recurrent
D.H. (1998) Increased dentate granule cell neurogenesis fol- excitatory circuitry. J. Comp. Neurol., 428: 240–253.
lowing amygdala kindling in the adult rat. Neurosci. Lett., Romcy-Pereira, R.N. and Garcia-Cairasco, N. (2003) Hippo-
247: 9–12. campal cell proliferation and epileptogenesis after audiogenic
Patrylo, P.R. and Dudek, F.E. (1998) Physiological unmasking kindling are not accompanied by mossy fiber sprouting
of new glutamatergic pathways in the dentate gyrus of hip- or Fluoro-Jade staining. Neuroscience, 119: 533–546.
pocampal slices from kainate-induced epileptic rats. J. van Rooyen, F., Young, N.A., Larson, S.E. and Teskey, G.C.
Neurophysiol., 79: 418–429. (2006) Hippocampal kindling leads to motor map expansion.
Patrylo, P.R., Schweitzer, J.S. and Dudek, F.E. (1999) Abnor- Epilepsia, 47: 1383–1391.
mal responses to perforant path stimulation in the dentate Salin, P., Tseng, G.F., Hoffman, S., Parada, I. and Prince, D.A.
gyrus of slices from rats with kainate-induced epilepsy and (1995) Axonal sprouting in layer V pyramidal neurons
mossy fiber reorganization. Epilepsy Res., 36: 31–42. of chronically injured cerebral cortex. J. Neurosci.,
Peng, Z. and Houser, C.R. (2005) Temporal patterns of fos 15: 8234–8245.
expression in the dentate gyrus after spontaneous seizures Sayin, U., Osting, S., Hagen, J., Rutecki, P. and Sutula, T.
in a mouse model of temporal lobe epilepsy. J. Neurosci., (2003) Spontaneous seizures and loss of axo-axonic and axo-
25: 7210–7220. somatic inhibition induced by repeated brief seizures in kin-
Peng, Z., Huang, C.S., Stell, B.M., Mody, I. and Houser, C.R. dled rats. J. Neurosci., 23: 2759–2768.
(2004) Altered expression of the delta subunit of the Sayin, U., Rutecki, P. and Sutula, T. (1999) NMDA-dependent
GABA(A) receptor in a mouse model of temporal lobe currents in granule cells of the dentate gyrus contribute to
epilepsy. J. Neurosci., 24: 8629–8639. induction but not permanence of kindling. J. Neurophysiol.,
Pierce, J.P. and Milner, T.A. (2001) Parallel increases in the 81: 564–574.
synaptic and surface areas of mossy fiber terminals following Scharfman, H.E., Sollas, A.L., Berger, R.E. and Goodman, J.H.
seizure induction. Synapse, 39: 249–256. (2003) Electrophysiological evidence of monosynaptic excita-
Pitkanen, A., Nissinen, J., Lukasiuk, K., Jutila, L., Paljarvi, L., tory transmission between granule cells after seizure-induced
Salmenpera, T., Karkola, K., Vapalahti, M. and Ylinen, A. mossy fiber sprouting. J. Neurophysiol., 90: 2536–2547.
(2000) Association between the density of mossy fiber sprout- Scharfman, H.E., Sollas, A.L. and Goodman, J.H. (2002)
ing and seizure frequency in experimental and human tem- Spontaneous recurrent seizures after pilocarpine-induced sta-
poral lobe epilepsy. Epilepsia, 41: S24–S29. tus epilepticus activate calbindin-immunoreactive hilar cells
Pretel, S., Applegate, C.D. and Piekut, D. (1997) Apoptotic and of the rat dentate gyrus. Neuroscience, 111: 71–81.
necrotic cell death following kindling induced seizures. Acta Sloviter, R.S., Dichter, M.A., Rachinsky, T.L., Dean, E.,
Histochem., 99: 71–79. Goodman, J.H., Sollas, A.L. and Martin, D.L. (1996) Basal
Qiao, X. and Noebels, J.L. (1993) Developmental analysis of expression and induction of glutamate decarboxylase and
hippocampal mossy fiber outgrowth in a mutant mouse with GABA in excitatory granule cells of the rat and monkey
inherited spike-wave seizures. J. Neurosci., 13: 4622–4635. hippocampal dentate gyrus. J. Comp. Neurol., 373: 593–618.
Ramirez-Amaya, V., Escobar, M.L., Chao, V. and Bermudez- Sloviter, R.S., Zappone, C.A., Harvey, B.D. and Frotscher, M.
Rattoni, F. (1999) Synaptogenesis of mossy fibers induced by (2006) Kainic acid-induced recurrent mossy fiber innervation
spatial water maze overtraining. Hippocampus, 9: 631–636. of dentate gyrus inhibitory interneurons: possible anatomical
Represa, A. and Ben-Ari, Y. (1992) Kindling is associated with substrate of granule cell hyper-inhibition in chronically
the formation of novel mossy fibre synapses in the CA3 epileptic rats. J. Comp. Neurol., 494: 944–960.
region. Exp. Brain Res., 92: 69–78. Smith, B.N. and Dudek, F.E. (2001) Short- and long-term
Represa, A., Le Gall La Salle, G. and Ben-Ari, Y. (1989a) changes in CA1 network excitability after kainate treatment
Hippocampal plasticity in the kindling model of epilepsy in in rats. J. Neurophysiol., 85: 1–9.
rats. Neurosci. Lett., 99: 345–350. Sole, R. and Goodwin, B. (2000) Signs of Life — How Com-
Represa, A., Robain, O., Tremblay, E. and Ben-Ari, Y. (1989b) plexity Invades Biology. Basic Books, New York.
Hippocampal plasticity in childhood epilepsy. Neurosci. Staley, K.J., Soldo, B.L. and Proctor, W.R. (1995) Ionic mech-
Lett., 99: 351–355. anisms of neuronal excitation by inhibitory GABA(A)
Ribak, C.E. and Peterson, G.M. (1991) Intragranular mossy receptors. Science, 269: 977–981.
fibers in rats and gerbils form synapses with the somata Stanfield, B.B. (1989) Excessive intra- and supragranular mossy
and proximal dendrites of basket cells in the dentate gyrus. fibers in the dentate gyrus of tottering (tg/tg) mice. Brain
Hippocampus, 1: 355–364. Res., 480: 294–299.
563

Staubli, U., Gall, C. and Lynch, G. (1984) The distribution of gyrus in kainate-treated rats. J. Comp. Neurol., 390:
the commissural-associational afferents of the dentate gyrus 578–594.
after perforant path lesions in one-day-old rats. Brain Res., Tauck, D.L. and Nadler, J.V. (1985) Evidence of functional
292: 156–159. mossy fiber sprouting in hippocampal formation of kainic
Steward, O. (1976) Reinnervation of dentate gyrus by homol- acid-treated rats. J. Neurosci., 5: 1016–1022.
ogous afferents following entorhinal cortical lesions in adult Timofeeva, O.A. and Peterson, G.M. (1999) Dissociation of
rats. Science, 194: 426–428. mossy fiber sprouting and electrically-induced seizure sensi-
Steward, O. (1992) Lesion-induced synapse reorganization in the tivity: rapid kindling versus adaptation. Epilepsy Res., 33:
hippocampus of cats: sprouting of entorhinal, commissural/ 99–115.
associational, and mossy fiber projections after unilateral Toyoda, I. and Buckmaster, P.S. (2005) Prolonged infusion
entorhinal cortex lesions, with comments on the normal of cycloheximide does not block mossy fiber sprouting in
organization of these pathways. Hippocampus, 2: 247–268. a model of temporal lobe epilepsy. Epilepsia, 46: 1017–1020.
Steward, O., Cotman, C. and Lynch, G. (1976) A quantitative Tuff, L.P., Racine, R.J. and Adamec, R. (1983a) The effects
autoradiographic and electrophysiological study of the rein- of kindling on GABA-mediated inhibition in the dentate
nervation of the dentate gyrus by the contralateral entorhinal gyrus of the rat. I. Paired-pulse depression. Brain Res.,
cortex following ipsilateral entorhinal lesions. Brain Res., 277: 79–90.
114: 181–200. Tuff, L.P., Racine, R.J. and Mishra, R.K. (1983b) The effects
Steward, O., Cotman, C.W. and Lynch, G.S. (1973) Re-estab- of kindling on GABA-mediated inhibition in the dentate
lishment of electrophysiologically functional entorhinal cor- gyrus of the rat. II. Receptor binding. Brain Res., 277: 91–98.
tical input to the dentate gyrus deafferented by ipsilateral Vaidya, V.A., Siuciak, J.A., Du, F. and Duman, R.S. (1999)
entorhinal lesions: innervation by the contralateral ent- Hippocampal mossy fiber sprouting induced by chronic elec-
orhinal cortex. Exp. Brain Res., 18: 396–414. troconvulsive seizures. Neuroscience, 89: 157–166.
Steward, O., Cotman, C.W. and Lynch, G.S. (1974) Growth Wenzel, H.J., Woolley, C.S., Robbins, C.A. and Schwartzkroin,
of a new fiber projection in the brain of adult rats: re-inner- P.A. (2000) Kainic acid-induced mossy fiber sprouting and
vation of the dentate gyrus by the contralateral entorhinal synapse formation in the dentate gyrus of rats. Hippocam-
cortex following ipsilateral entorhinal lesions. Exp. Brain pus, 10: 244–260.
Res., 20: 45–66. Williams, P.A., Wuarin, J.P., Dou, P., Ferraro, D.J. and
Steward, O. and Messenheimer, J.A. (1978) Histochemical Dudek, F.E. (2002) Reassessment of the effects of cyclohexi-
evidence for a post-lesion reorganization of cholinergic affer- mide on mossy fiber sprouting and epileptogenesis in the
ents in the hippocampal formation of the mature cat. J. pilocarpine model of temporal lobe epilepsy. J. Ne-
Comp. Neurol., 178: 697–709. urophysiol., 88: 2075–2087.
Stringer, J.L. (1996) Repeated seizures increase GFAP and Worrell, G.A., Parish, L., Cranstoun, S.D., Jonas, R., Baltuch,
vimentin in the hippocampus. Brain Res., 717: 147–153. G. and Litt, B. (2004) High-frequency oscillations and seizure
Stringer, J.L. and Lothman, E.W. (1989) Repetitive seizures generation in neocortical epilepsy. Brain, 127: 1496–1506.
cause an increase in paired-pulse inhibition in the dentate Wuarin, J.P. and Dudek, F.E. (1996) Electrographic seizures
gyrus. Neurosci. Lett., 105: 91–95. and new recurrent excitatory circuits in the dentate gyrus
Sur, M. and Rubenstein, J.L. (2005) Patterning and plasticity of of hippocampal slices from kainate-treated epileptic rats. J.
the cerebral cortex. Science, 310: 805–810. Neurosci., 16: 4438–4448.
Sutula, T. (2002) Seizure-induced axonal sprouting: assessing Wuarin, J.P. and Dudek, F.E. (2001) Excitatory synaptic input
connections between injury, local circuits, and epileptogen- to granule cells increases with time after kainate treatment. J.
esis. Epilepsy Curr., 2: 86–91. Neurophysiol., 85: 1067–1077.
Sutula, T., Cascino, G., Cavazos, J., Parada, I. and Ramirez, L. Wuarin, J.P., Kim, Y.I., Cepeda, C., Tasker, J.G., Walsh, J.P.,
(1989) Mossy fiber synaptic reorganization in the epileptic Peacock, W.J., Buchwald, N.A. and Dudek, F.E. (1990)
human temporal lobe. Ann. Neurol., 26: 321–330. Synaptic transmission in human neocortex removed for
Sutula, T., Harrison, C. and Steward, O. (1986) Chronic epile- treatment of intractable epilepsy in children. Ann. Neurol.,
ptogenesis induced by kindling of the entorhinal cortex: the 28: 503–511.
role of the dentate gyrus. Brain Res., 385: 291–299. Zhang, L.X., Smith, M.A., Li, X.L., Weiss, S.R.B. and Post,
Sutula, T., He, X.X., Cavazos, J. and Scott, G. (1988) Synaptic R.M. (1998) Apoptosis of hippocampal neurons after
reorganization in the hippocampus induced by abnormal amygdala kindled seizures. Mol. Brain Res., 55: 198–208.
functional activity. Science, 239: 1147–1150. Zhang, N. and Houser, C.R. (1999) Ultrastructural localization
Sutula, T., Zhang, P., Lynch, M., Sayin, U., Golarai, G. and of dynorphin in the dentate gyrus in human temporal lobe
Rod, R. (1998) Synaptic and axonal remodeling of mossy epilepsy: a study of reorganized mossy fiber synapses. J.
fibers in the hilus and supragranular region of the dentate Comp. Neurol., 405: 472–490.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 30

A behavioral analysis of dentate gyrus function

Raymond P. Kesner

University of Utah, Department of Psychology, 380 S. 1530 E., Room 502, Salt Lake City, UT 84121, USA

Abstract: Computational models of the dentate gyrus (DG) have suggested based on anatomical, electro-
physiological, and computer simulation data that the DG plays an important role in learning and memory
by processing and representing spatial information on the basis of conjunctive encoding, pattern sepa-
ration, and encoding of spatial information in conjunction with the CA3. Behavioral evidence supports a
role for the DG in mnemonic processing of spatial information based on the operation of conjunctive
encoding of multiple sensory inputs, pattern separation of spatial (especially metric) information, and
subsequent encoding in cooperation with CA3. A potential role of the DG in mediating processes, such as
recall of sequential information and short-term memory as well as temporal order for remote memory, are
also discussed.

Keywords: conjunctive encoding; spatial pattern separation; dentate gyrus; encoding; retrieval

Introduction Conjunctive encoding

The anatomy and neural circuitry of the dentate It can be shown that the DG receives multiple
gyrus (DG) has been described in detail in previ- sensory inputs including vestibular, olfactory, vis-
ous chapters in this volume (Chapter 1, Amaral, ual, auditory, and somatosensory from the peri-
Scharfman and Lavenex). Based on the anatomy rhinal cortex and lateral entorhinal cortex in
of the DG, its input and output pathways, and the conjunction with spatially organized grid cells
development of a computational model, Rolls from the medial entorhinal cortex (Hafting et al.,
(1996) and Rolls and Kesner (2006) have sug- 2005) to represent metric spatial representations.
gested that the DG has three major functions in- The perforant path input in the DG can be divided
cluding conjunctive encoding of multiple sensory into a medial and lateral component. The medial
inputs, spatial pattern separation, and facilitation component processes spatial information and the
of encoding of spatial information based on its lateral component processes non-spatial (e.g. ob-
outputs to CA3. This is accomplished by a com- jects, odors) information (Witter et al., 1989a;
petitive learning network with Hebb-like modifi- Hargreaves et al., 2005). Based on the idea that the
ability to remove redundancy from the inputs and medial perforant path input into the DG mediates
produce a more orthogonal, sparse, and catego- spatial information via activation of NMDA re-
rized set of outputs. ceptors and the lateral perforant path input into
the DG mediates visual object information via
activation of opioid receptors, the following ex-
Corresponding author. Tel.: +801 581 7430; periment was conducted. Using a paradigm deve-
Fax: +801 581 5841; E-mail: rpkesner@behsci.utah.edu loped by Poucet (1989), rats were tested for the

DOI: 10.1016/S0079-6123(07)63030-1 567


568

detection of a novel spatial change and the detec- extent that DG acts to produce separate represen-
tion of a novel visual object change under the in- tations of different but similar places, it is pre-
fluence of direct infusions of AP5 (an NMDA dicted that the DG will be especially important
antagonist) or naloxone (a m opiate antagonist) when memories must be formed about similar
into the DG. The results indicated that naloxone places.
infusions into the DG disrupted both novelty de- Rolls’ model proposes that pattern separation is
tection of a spatial location and a visual object, facilitated by sparse connections in the mossy-fiber
whereas AP5 infusions into the DG disrupted only system, which connects DG granular cells to CA3
detection of a novel spatial location, but had no pyramidal neurons. The separation of patterns is
effect on detection of a novel object. In contrast, accomplished based on the low probability that
infusions of AP5 into the CA1 region disrupts only any two CA3 neurons will receive mossy fiber in-
the detection of a spatial location change, but not put synapses from a similar subset of DG cells.
a visual object change, whereas naloxone into the The mossy fiber inputs to CA3 from DG are sug-
CA1 region disrupts only the detection of a visual gested to be essential during learning and may in-
object change, but not a spatial location change fluence which CA3 neurons will fire based on the
(Hunsaker et al., 2007). These data suggest that distributed activity in the DG. The cells of the DG
the DG uses conjunctive encoding of visual object are suggested to act as a competitive learning
and spatial information to provide for a spatial network with Hebb-like modifiability to reduce
representation, which I will show below might be redundancy and produce sparse, orthogonal out-
based on metric information. puts. O’Reilly and McClelland (1994) and Shapiro
and Olton (1994) also suggest that the mossy fiber
connections between the DG and CA3 may sup-
Spatial pattern separation port pattern separation. Rolls (1996) notes addi-
tional characteristics of the mossy fiber projection
It can clearly be demonstrated that single cells system that may promote pattern separation in the
within the hippocampus are activated by most DG-CA3 system and hence, efficient information
sensory inputs, including vestibular, olfactory, vis- storage in CA3. First, mossy fiber synapses are
ual, auditory, and somatosensory as well as higher very large and terminate close to the soma of the
order integration of sensory stimuli (Cohen and CA3 pyramidal neurons in the pyramidal layer.
Eichenbaum, 1993). The question of importance is Therefore, the mossy fiber synapses will be rela-
whether these sensory inputs via conjunctive en- tively powerful in activating the CA3 cell. The
coding have a memory representation within the projections from layer II of entorhinal cortex to
hippocampus. One possible role for the hippo- CA3 terminate in the lacunosum moleculare layer
campus in processing all sensory information of the CA3 pyramidal cells, which is much further
might be to provide sensory markers to demar- from the soma. Second, the firing activity of
cate a spatial location, so that the hippocampus granule cells within the DG is sparse (Jung and
can more efficiently mediate spatial information. McNaughton, 1992) and coupled with the small
Thus, it is possible that one of the main processing number of connections to CA3 cells, which would
functions of the hippocampus is to encode and produce a sparse signal that may be transformed
separate spatial events from each other. This into an even sparser signal in CA3. It also has been
would ensure that new highly processed sensory demonstrated that the place fields of DG cells
information is organized within the hippocampus (Mizumori et al., 1990) and specifically granular
and enhances the possibility of encoding and tem- cells (Jung and McNaughton, 1992) are small and
porarily remembering one place as separate from highly reliable, which may support the role of DG
another place. This may be accomplished via pat- in pattern separation. In addition, the mossy fibers
tern separation of event information, so that sim- demonstrate non-associative plasticity (Brown
ilar spatial events can be separated from each et al., 1989), which may enhance the signal-
other and spatial interference is reduced. To the to-noise ratio such that the mossy fiber cell would
569

produce a non-linearly amplified current in the measured short-term memory for spatial location
CA3 cell. Due to this type of plasticity in the DG, information as a function of spatial similarity bet-
a particular stimulus such as spatial location ween two spatial locations (Gilbert et al., 1998).
would be likely to activate the same population Specifically, the study was designed to examine the
of CA3 neurons across subsequent presentations, role of the DG subregion in discriminating spatial
which may result in economical storage, because locations when rats were required to remember a
any given CA3 cell could be used in different spatial location based on distal environmental cues
memories. and differentiate between the to-be-remembered
If disruption of DG function results in ineffi- location and a distractor location with different
cient pattern separation, then deficits on spatial degrees of similarity or overlap among the distal
tasks may occur when there is increased overlap or cues. Animals were tested using a cheeseboard
similarity among distal cues and presumably in- maze apparatus on a delayed-match-to-sample for
creased similarity among representations within a spatial location task. Animals were trained to
the DG. Remembering a specific location in an displace an object that was randomly positioned to
8-arm maze, a water maze, or a spatial context in cover a baited food well in 1 of 15 locations along
fear conditioning may be influenced by the degree a row of food wells. Following a short delay, the
of overlap among critical distal spatial cues. Rats animals were required to choose between two ob-
with lesions in DG have been tested on a working jects identical to the sample phase object. One ob-
memory version of the radial 8-arm maze. The re- ject was in the same location as the sample phase
sults demonstrated that a lesion of the DG resulted object and the second object was in a different
in deficits similar to complete hippocampal lesions location along the row of food wells. An animal
(Walsh et al., 1986; Tilson et al., 1987; McLamb was rewarded for displacing the object in the same
et al., 1988; Emerich and Walsh, 1989). In addi- spatial location as the sample phase object (correct
tion, rats with DG lesions were tested on the choice) but received no reward for displacing the
Morris water maze task and showed deficits com- foil object (incorrect choice). Five spatial separa-
parable to rats with complete hippocampal lesions tions, from 15 cm to 105 cm, were used to separate
when the start location varied on each trial the correct object and the foil object during the
(Sutherland et al., 1983; Nanry et al., 1989; choice phase. The results showed that rats with
Xavier et al., 1999; Jeltsch et al., 2001). Lee and DG lesions were significantly impaired at short
Kesner (2004a) tested rats with DG lesions on ac- spatial separations; however, the performance of
quisition and retrieval of contextual fear condi- the DG lesioned animals increased as a function of
tioning. Rats with DG lesions showed initial increased spatial separation between the correct
impairments in freezing behavior during acquisi- object and the foil on the choice phases. The per-
tion of the task but eventually reached the level of formance of rats with DG lesioned matched con-
freezing of controls with subsequent testing. When trols at the largest spatial separation. The graded
retrieval of contextual fear was examined in rats nature of the impairment and the significant linear
with DG lesions 24 h after acquisition, the animals improvement in performance as a function of in-
showed a significant deficit in freezing compared creased separation illustrate the deficit in pattern
to controls. Based on these studies it is clear that separation. Based on these results, it was con-
DG lesions impair spatial memory. The DG le- cluded that lesions of the DG decrease the effi-
sioned animals may be impaired as a result of ciency of spatial pattern separation, which resulted
impaired pattern separation; however, it is difficult in impairments on trials with increased spatial
to determine if a particular memory process is proximity and increased spatial similarity among
impaired in these animals using these three working memory representations.
paradigms. Based on the evidence that the hippocampus,
To examine the contribution of the DG to spa- and especially the DG, receives inputs from all
tial pattern separation, Gilbert et al. (2001) tested sensory modalities, there is a possibility that the
rats with DG lesions using a paradigm that DG uses sensory markers to demarcate a spatial
570

location, allowing the DG to more efficiently rep- possible to examine the detection of a topological
resent spatial information. Thus, one function of spatial change. In the topological manipulation
the DG may be to encode and separate events in condition, rats were allowed to explore four differ-
space resulting in spatial pattern separation. Spa- ent visual objects that were positioned in a square
tial pattern separation would ensure that new on the cheeseboard maze. The rats again were al-
highly processed sensory information is organized lowed to explore the objects and eventually habitu-
within the hippocampus, which in turn enhances ated to the objects with subsequent exposure.
the possibility of encoding and temporarily re- However, following habituation, the locations of
membering one spatial location as separate from two of the objects were switched and the time the
another location. Based on the observations that rat spent exploring each object was recorded. The
cells in the CA3 and CA1 region respond to results showed that rats with CA1 lesions impair,
changes in the metric and topological aspects of whereas DG or CA3, lesions do not impair the
the environment (O’Keefe and Burgess, 1996; detection of the topological manipulation. The re-
Jeffery and Anderson, 2003), one can ask the sults suggest that neurons in the DG may be crit-
question whether these different features of the ically involved in processing spatial information
spatial environment are processed via the DG and on a metric scale, but may not be necessary for
subsequently transferred to the CA3 subregion or representing topological space. The results of both
are these features communicated via the direct experiments provide for empirical validation of the
perforant path projection to the CA3 subregion? role of DG in spatial pattern separation and sup-
In both cases the information may then be trans- port the predictions of computational models
ferred to the CA1 subregion. To answer this ques- (Rolls, 1996; Rolls and Kesner, 2006).
tion, Goodrich-Hunsaker et al. (2005) examined There are some studies in the literature that have
the contribution of the DG to memory for metric demonstrated that hippocampal lesions, including
and topological spatial relationships. Using a the DG, can result in deficits for spatial tasks that
modified version of an exploratory paradigm de- can be interpreted to be a function of increased
veloped by Poucet (1989), rats with DG, CA3, and interference and an impairment in the utilization
CA1 lesions and controls were tested on tasks in- of a spatial pattern separation process. A few ex-
volving a metric spatial manipulation. In this task, amples will suffice. Because rats are started in
a rat was allowed to explore two different visual different locations in the standard water maze
objects that were separated by a specific distance task, there is potential interference among similar
on a cheeseboard maze. On the initial presentation and overlapping spatial patterns. Thus, the obser-
of the objects, the rat explored each object. How- vation that hippocampal lesioned rats are im-
ever, across subsequent presentations of the ob- paired in learning and subsequent consolidation of
jects in the same spatial locations, the rat important spatial information in this task could be
habituated and eventually spent less time explor- due to difficulty to separate spatial patterns re-
ing the objects. Once the rat had habituated to the sulting in enhanced spatial interference. Support
objects in their locations, the metric spatial dis- for this idea comes from the observation of
tance between the two objects was manipulated so Eichenbaum et al. (1990) who demonstrated that
the two objects were either closer together or fur- when fimbria-fornix lesioned rats are trained on
ther apart. The time the rat spent exploring each the water maze task from only a single starting
moved object was recorded. The results showed position (less spatial interference) there are hardly
that DG lesions impaired detection of the metric any learning deficits, whereas training from many
distance change, in that rats with DG lesions spent different starting points resulted in learning diffi-
significantly less time exploring the two objects culties. In a somewhat similar study, it was shown
that were displaced relative to controls. Rats with that total hippocampal lesioned rats learned or
CA3 or CA1 lesions displayed smaller reductions consolidated rather readily that only one spatial
in re-exploration and the DG was reliably im- location was correct on an 8-arm maze (Hunt
paired relative to controls and CA1. It is also et al., 1994). In a different study, McDonald and
571

White (1995) used a place preference procedure in neurons exhibited lower overlap in their activity
an 8-arm maze. In this procedure food is placed at between the two environments compared to CA1
the end of one arm and no food is placed at neurons. This could be consistent with the com-
the end of another arm. In a subsequent preference putational point that if CA3 is an autoassociator,
test normal rats prefer the arm that contained the the pattern representations within it should be as
food. In this study fornix lesioned rats acquired orthogonal as possible to maximize memory ca-
the place preference task as quickly as controls if pacity and minimize interference. The actual
the arm locations were opposite each other, pattern separation may be performed, the com-
but the fornix lesioned rats were markedly im- putational model holds, as a result of the
paired if the locations were adjacent to each other. operation of the dentate granule cells as a
Clearly, there would be greater spatial interference competitive net and the nature of the mossy
when the spatial locations are adjacent to each fiber connections to CA3 cells. Thus, CA3 may
other rather than far apart. In a different task, rats represent different environments relatively ortho-
were trained on a pair-wise visual spatial discrim- gonally. In the computational account, each
ination of vertical lines that were not very far environment would be a separate chart, and the
apart. Compared to controls, rats with hippocam- number of charts that could be stored in CA3
pal lesions were not able to learn the task (Lazaris would be high if the representations in each chart
et al., 2006). Thus, spatial pattern separation may were relatively orthogonal to those in other charts
play an important role in the acquisition of new and further, charts could operate independently
spatial information and there is a good possibility (Stringer et al., 2004). Any one chart of a given
that the DG may have been the subregion respon- spatial environment can be understood as a con-
sible for the impairments in the various tasks de- tinuous attractor network with place cells with
scribed above. Gaussian-shaped place fields that overlap contin-
Do other subregions of the hippocampus en- uously with each other. In different charts (differ-
gage in spatial pattern separation? It appears that ent spatial environments) the same neurons may
the CA3 region may also engage spatial pattern represent very different regions of space, and neu-
separation processes. For example, Tanila (1999) rons representing close places in one environment
showed that CA3 place cells were able to maintain may represent distant locations in another envi-
distinct representations of two visually identical ronment (chart). It appears that the DG mediated
environments and selectively reactivate either one pattern separation is based on reducing interfer-
of the representation patterns depending on the ence between discrete spatial locations based on
experience of the rat. Also, Leutgeb et al. (2004) metric representations of distance between the
recently showed that when rats experienced a spatial locations, whereas the CA3 mediated pat-
completely different environment, CA3 place cells tern separation is based on reducing interference
developed orthogonal representations of those across global environments based on representa-
different environments by changing their firing tion of unique charts. Whether the actual pattern
rates between the two environments, whereas CA1 separation in CA3 is performed as a result of the
place cells maintained similar responses. In a operation of the dentate granule cells as a com-
different study Vazdarjanova and Guzowski petitive net and the nature of the mossy fiber
(2004) placed rats in two different environments connections to CA3 cells remains to be investi-
separated by 30 min. The two environments gated. An alternative possibility is that pattern
differed greatly in that different objects were lo- separation observed for the global environment is
cated in each room. The authors were able to separate from pattern separation observed for
monitor the time course of activations of ensem- discrete elements in the DG. The dissociation bet-
bles of neurons in both CA3 and CA1, using a ween these two forms of pattern separation could
new immediate-early gene-based brain-imaging be achieved by assuming that the perforant path
method (Arc/H1a catFISH). When the two envi- input into CA3 is critical for representing a global
ronments were significantly different, CA3 environment or context. The perforant path into
572

CA3 can also be divided into a medial and lateral Encoding vs. retrieval of spatial information
component. Because there is associative LTP in
CA3 between the medial or lateral perforant path Rolls’ (1996) computational model postulates that
and the intrinsic commissural/associational-CA3 the dentate/mossy fiber system is necessary for
synapses (Martinez et al., 2002), either place, ob- setting up the appropriate conditions for optimal
ject or contextual cues can be processed by the storage of new information in the CA3 system
associative medial and lateral perforant path con- (which could be called encoding). In order to test
nections to CA3 cells. This is consistent with the the model rats with DG or CA3 lesions were ad-
finding that disruption of the perforant path input ministered 10 learning trials per day in a Hebb-
impairs the multiple trial acquisition of an object- Williams maze. The results based on a within-day
place paired associate task [perhaps because re- analysis indicated that DG and CA3 lesions im-
trieval of what has been previously learned is im- paired the acquisition of this task, consistent with
paired (unpublished observations)]. an encoding or learning impairment (Lee and
In addition, Martinez et al. (2002) demonstrated Kesner, 2004b; Jerman et al., 2006). However,
associative (cooperative) LTP between the medial when tested using a between days analysis, re-
and lateral perforant path inputs to the CA3 neu- trieval of what had been learned previously was
rons. This could provide a mechanism for object not impaired by DG lesions (Lee and Kesner,
(lateral perforant path) — place (medial perforant 2004b; Jerman et al., 2006). In the Hebb-Williams
path) associative learning in conjunction with a learning task, one can measure improvement in
global context. Further support for this idea is performance within a day, reflecting the operation
based on testing rats for the detection of a novel of encoding of new information based in part on
spatial change and the detection of a novel visual short-term memory representations and one can
object change under the influence of direct infu- also measure improvement in performance bet-
sions of AP5 or naloxone into the CA3. The ween days, reflecting the operation of retrieval of
results indicated that naloxone or AP5 infusions information either based on intermediate-term
into CA3 disrupted both novelty detection of a memory representations mediated by synaptic con-
spatial location and a visual object, suggesting solidation and/or access to stored information. It
that the perforant path input into CA3 could pro- is recognized that the separation of encoding from
vide the context for integration of both spatial and retrieval processes is extremely difficult. Therefore,
visual object information (Hunsaker et al., 2007). it is assumed that during acquisition within a day,
These studies thus indicate that both the DG there will be a greater involvement of encoding
and CA3 play an important role in spatial pattern compared to retrieval processes, and that during
separation for spatial memory tasks. Other forms retention across days, there will be a greater in-
of pattern separation do not involve the DG sub- volvement of retrieval compared to encoding
region of the hippocampus. For example, temporal processes. It is also assumed that encoding en-
pattern separation is mediated by the CA1, but not compasses spatial pattern separation processes in
the DG (Gilbert et al., 2001). Furthermore, hippo- conjunction with associative processes and repre-
campal lesions including DG do not produce a sentations within short-term memory. Even
deficit for pattern separation of reward values, though there is likely to be some retrieval from
visual objects, or motor responses (Gilbert and short-term memory that may also occur during
Kesner, 2002, 2003b, 2006; Saksida et al., 2006). acquisition, it is assumed not to be the dominant
Instead, the perirhinal cortex subserves pattern factor governing performance within the first 10
separation for visual objects (Bussey et al., 2002; trials on Day 1. Furthermore, it is assumed that
Gilbert and Kesner, 2003), the amygdala subserves retrieval 24 h later encompasses associative proc-
pattern separation for reward value (Gilbert and esses as well as representations within intermedi-
Kesner, 2002) and the caudate nucleus subserves ate-term memory. Even though there is likely to be
pattern separation for motor responses (Kesner some encoding that may also occur during re-
and Gilbert, 2006). trieval, it is assumed not to be the most critical
573

determinant of performance within the first five projections forming the primary output of the DG.
trials on Day 2. Based on characteristics of the mossy fiber system,
It is therefore of interest that the results indicate Rolls (1996) suggests that pattern separation may
that both DG and CA3 lesions disrupted encod- be a function of the DG and its mossy fiber pro-
ing, but not retrieval, even though CA3 lesions did jections to CA3 and thus may facilitate the encod-
produce more errors during acquisition than DG ing of spatial information via an interaction
lesions. The DG lesion data are consistent with the between the DG and CA3. However, recent stud-
results of (Lassalle et al., 2000), who showed that ies have shown that the functions of these two
in a water maze learning task, lesions of the mossy hippocampal subregions can be dissociated using
fibers disrupted encoding, but not retrieval. The behavioral tasks. For example, Gilbert and Kesner
CA3 lesion data are consistent with the idea that (2003a) demonstrated that DG lesioned animals
short-term encoding of new information is medi- were able to learn object-place and odor-place
ated by CA3. Other studies involving alteration of paired-associate tasks as quickly as controls. How-
CA3 function support this idea (Lee and Kesner, ever, rats with CA3 lesions showed significant
2002, 2003, 2004a; Nakazawa et al., 2003). Thus, learning impairments on both tasks. Furthermore,
the data suggest a possible cooperation between a lesion of the perforant path input into the CA3
the DG and CA3 in encoding of information in a also disrupted object-place paired associate learning
Hebb-Williams maze, in that there is a deficit in (unpublished observations). These results could in-
encoding of information following either a DG or dicate that the CA3, but not the DG, subregion is
CA3 lesion. involved in associative learning. Furthermore, DG,
The present data appear to be in conflict with but not CA3, lesioned rats produce a deficit in the
the results of Lee and Kesner (2004b), wherein a acquisition of the standard version water maze
lesion to the perforant path into CA3 caused a (Sutherland et al., 1983; Nanry et al., 1989; Xavier
deficit in retrieval but no effect for encoding. There et al., 1999; Lassalle et al., 2000; Jeltsch et al., 2001).
is the possibility that the lesion to the perforant The dissociations that arise between DG and
path may have disrupted the Schaffer collateral CA3 are primarily due to the CA3 subregion of the
output from CA3 into CA1. Since the Lee and hippocampus having two major inputs with a di-
Kesner paper was published (2004b), Hebb- rect connection from the DG via the mossy fibers
Williams maze data for CA1 lesioned animals and a direct input from the perforant path that
from our laboratory has shown that animals with bypasses the DG. An alternate explanation is pos-
lesions to dorsal CA1 show no deficit in encoding sible since a lesion of the perforant path disrupts
but a significant deficit in retrieval relative to a retrieval, but not encoding (Lee and Kesner,
control group (Vago et al., 2007). If the Schaffer 2004b). The differences between DG and CA3 in
collateral input into CA1 were disrupted by the the above mentioned tasks could be simply a func-
perforant path lesion performed by Lee and tion of differential intrinsic processing of similar
Kesner (2004b), then the effects seen in that study spatial information. As an another example
would be more indicative of a CA1 effect rather Gilbert et al. (2001) and Gilbert and Kesner
than a CA3 effect. This would be more consistent (2006) tested rats with DG or CA3 lesions using
with the present results indicating that a lesion to a paradigm that measured one-trial short-term
CA3 does not result in a deficit for retrieval of memory for spatial location information as a
information — only encoding. function of spatial similarity between two spatial
Given that there was an encoding, but not a re- locations. The results showed that rats with DG
trieval deficit for the DG and CA3, could there be lesions were significantly impaired at short spatial
an interaction in terms of the encoding processes separations; however, the performance of the DG
between the DG and the CA3? This cooperation lesioned rats improved as a function of increased
between DG and CA3 could theoretically derive spatial separation between the correct object and
from the observation that the DG granule cells the foil on the choice phases. In contrast, CA3
project to CA3 pyramidal neurons via mossy fiber lesioned rats were impaired for all spatial
574

separations, suggesting a disruption of a short- concerning subregional specificity of CA3 and DG


term memory process subserved by CA3 as an lesions). Thus, the data suggest that the DG and
intrinsic contribution to the spatial pattern sepa- CA3 cooperate and interact with each other in the
ration task. Therefore, there is no guarantee that encoding of new spatial information in the Hebb-
any potential cooperation between DG and CA3 Williams maze without affecting retrieval.
in encoding information in the Hebb-Williams
maze is due to the direct connection between these
two regions. To examine this issue, a disconnection Additional potential functions of the dentate gyrus
study was carried out using an ipsilateral lesion
(DG and CA3 lesion on one side only) group vs. a Based on the observation that neurogenesis occurs
contralateral lesion (DG on one side and CA3 on in the DG and that new DG granule cells can be
the other side) group. It is assumed the right and formed across time, it has been proposed that the
left hemispheres operate in parallel. Crossed le- DG mediates a temporal pattern separation mech-
sions (i.e. unilateral lesions in contralateral hemi- anism that can generate patterns of episodic mem-
spheres), therefore, would disrupt communication ories within remote memory (Aimone et al., 2006).
within each of the two hemispheres, thus func- There are currently no behavioral data available to
tionally disconnecting the two brain regions. If the assess the proposal. It should be noted that the
DG and CA3 subregions of the hippocampus in- CA1, but not DG, is involved in temporal ordering
teract, then crossed lesioned rats should be mark- of spatial information within an intermediate
edly impaired relative to animals with lesions on memory system, but remote memory has not been
the same side. If these two regions do not interact, investigated (Kesner et al., 2004).
then crossed lesions would not produce a deficit, There are some data that suggest that disruption
suggesting that each subregion produces a deficit of neurogenesis using MAM (an anti-mitotic
in encoding for different reasons such as spatial agent) impairs the learning of trace eye-blink con-
pattern separation problems vs. associative or ditioning, but has no effect on delay eye-blink
short-term memory problems. conditioning, water maze spatial navigation, or
The results indicate that rats with an ipsilateral contextual fear conditioning (Shors, 2002). Fur-
lesion of both DG and CA3 do not produce a thermore, trace eye-lid conditioning prevents the
deficit in either encoding or retrieval in the Hebb- loss of newly developed granule cells (Shors, 2004).
Williams maze. In contrast, rats with a DG lesion These data suggest that the DG may play a role in
on one side of the brain and a CA3 lesion on the temporal processing, but it appears to be based on
other side of the brain disrupted encoding, but not a different mechanism than the one proposed by
retrieval, of information on the Hebb-Williams Aimone et al. (2006).
maze. Thus, the DG and CA3 interact to support Based on the observation that the DG has a re-
encoding of spatial information in the present task. current network system in which DG granule cells
There is always the possibility that following a DG excite mossy cells which feed back modifiable ex-
lesion on one side and a CA3 lesion on the other citatory connections to the granule cells and that
side, there would be a significant amount of degen- CA3 pyramidal cells have a feedback connection to
eration within CA3 following the DG lesion, so that the dentate network (see Scharfman’s Chapter 34,
there would be a lesion of the CA3 lesion on the this volume), Lisman (1999) has suggested that this
other side in effect creating a bilateral lesion of CA3 recurrent circuit can provide for short-term mem-
and similarly a significant amount of degeneration ory across seconds and thus can play an important
within DG following the CA3 lesion in effect cre- role in providing accurate recall of sequences.
ating a bilateral DG lesion. Against this argument From a behavioral perspective, there is evidence
is that upon histological analysis, there is no clear that the DG contributes to short-term or working
loss of the contralateral CA3 following the DG le- memory (Xavier et al., 1999; Lee and Kesner,
sion and no loss of the contralateral DG following 2003), which could be a function of both the CA3
the CA3 lesion (see Jerman et al., 2005 for data and DG recurrent circuits. However, there is
575

currently no behavioral evidence to demonstrate Gilbert, P.E. and Kesner, R.P. (2006) The role of dorsal CA3
that the DG contributes to sequence recall. hippocampal subregion in spatial working memory and pat-
tern separation. Behav. Brain Res., 169: 142–149.
In summary, based on the development of com-
Gilbert, P.E., Kesner, R.P. and DeCoteau, W.E. (1998) The
putational models of DG and behavioral evidence role of the hippocampus in mediating spatial pattern sepa-
based on dysfunction of DG, it appears that the ration. J. Neurosci., 18: 804–810.
DG mediates mnemonic processing of spatial in- Gilbert, P.E., Kesner, R.P. and Lee, I. (2001) Dissociating
formation. The processes subserved by DG include hippocampal subregions: a double dissociation between the
(a) the operation of conjunctive encoding of mul- dentate gyrus and CA1. Hippocampus, 11: 626–636.
Goodrich-Hunsaker, N.J., Hunsaker, M.R. and Kesner, R.P.
tiple sensory inputs, implying an integration of (2005) Effects of hippocampus sub-regional lesions for metric
sensory inputs to determine a spatial representa- and topological spatial information processing. Society for
tion, (b) pattern separation of spatial (especially Neuroscience Abstracts, Number 647.1.
metric) information, involving the reduction of in- Hafting, T., Fyhn, M., Molden, S., Moser, M.B. and Moser,
E.I. (2005) Microstructure of a spatial map in the entorhinal
terference between similar spatial locations, and
cortex. Nature, 436: 801–806.
(c) subsequent encoding in cooperation with CA3, Hargreaves, E.L., Rao, G., Lee, I. and Knierim, J.J. (2005)
suggesting an important role in the acquisition of Major dissociation between medial and lateral entorhinal in-
spatial information. A potential role of the DG in put to dorsal hippocampus. Science, 308: 1792–1794.
recall of sequential information requires more em- Hunsaker, M.R., Mooy, G.G., Swift, J. and Kesner, R.P.
pirical evidence. (2007) Dissociations of the role of the medial and lateral
perforant path projections into dorsal DG, CA3 and CA1 for
spatial and nonspatial (visual object) information processing.
References Behav. Neuroscience (in press).
Hunt, M.E., Kesner, R.P. and Evans, R.B. (1994) Memory for
spatial location: functional dissociation of entorhinal cortex
Aimone, J.B., Wiles, J. and Gage, F.H. (2006) Potential role for
adult neurogenesis in the encoding of time in new memories. and hippocampus. Psychobiology, 22: 186–194.
Nat. Neurosci., 9: 723–727. Jeffery, K.J. and Anderson, M.I. (2003) Dissociation of the
geometric and contextual influences on place cells. Hippo-
Brown, T.H., Ganong, A.H., Kairiss, E.W., Keenan, C.L. and
Kelso, S.R. (1989) Long-term potentiation in two synaptic campus, 13: 868–872.
systems of the hippocampal brain slice. In: Byrne J.H. and Jeltsch, H., Bertrand, F., Lazarus, C. and Cassel, J.-C. (2001)
Berry W.O. (Eds.), Neural Models of Plasticity. Academic Cognitive performances and locomotor activity following
dentate granule cell damage in rats: role of lesion extent
Press, San Diego, CA, pp. 266–306.
Bussey, T.J., Saksida, L.M. and Murray, E.A. (2002) Perirhinal and type of memory tested. Neurobiol. Learn. Mem., 76:
cortex resolves and feature ambiguity in complex discrimi- 81–105.
nations. Eur. J. Neurosci., 15: 365–374. Jerman, T., Kesner, R.P. and Hunsaker, M.R. (2006) Discon-
nection analysis of CA3 and DG in mediating encoding but
Cohen, N.J. and Eichenbaum, H.B. (1993) Memory, Amnesia,
and Hippocampal Function. MIT Press, Cambridge. not retrieval in a spatial maze learning task. Learn. Mem., 13:
Eichenbaum, H., Stewart, C. and Morris, R.G.M. (1990) Hip- 458–464.
Jerman, T.S., Kesner, R.P., Lee, I. and Berman, R.F. (2005)
pocampal representation in spatial learning. J. Neurosci., 10:
331–339. Patterns of hippocampal cell loss based on subregional le-
Emerich, D.F. and Walsh, T.J. (1989) Selective working mem- sions of the hippocampus. Brain Res., 1065: 1–7.
ory impairments following intradentate injection of col- Jung, M.W. and McNaughton, B.L. (1992) Spatial selectivity of
unit activity in the hippocampal granular layer. Hippocam-
chicine: attenuation of the behavioral not the
neuropathological effects by gangliosides GM1 and AGF2. pus, 3: 165–182.
Physiol. Behav., 45: 93–101. O’Keefe, J. and Burgess, N. (1996) Geometric determinants of
Gilbert, P.E. and Kesner, R.P. (2002) The amygdala but not the the place field of hippocampal neurons. Nature, 381:
425–428.
hippocampus is involved in pattern separation based on re-
ward value. Neurobiol. Learn. Mem., 77: 338–353. Kesner, R.P. and Gilbert, P.E. (2006) The role of the medial
Gilbert, P.E. and Kesner, R.P. (2003a) Localization of function caudate nucleus, but not the hippocampus, in a matching-to
within the dorsal hippocampus: the role of the dorsal CA3 sample task for a motor response. Eur. J. Neurosci., 23:
1888–1894.
subregion in paired-associate learning. Behav. Neurosci., 117:
1385–1394. Kesner, R.P., Lee, I. and Gilbert, P. (2004) A behavioral as-
Gilbert, P.E. and Kesner, R.P. (2003b) Recognition memory sessment of hippocampal function based on a subregional
analysis. Rev. Neurosci., 15: 333–351.
for complex visual discrimination is influenced by stimulus
interference in rodents with perirhinal cortex damage. Learn. Lassalle, J.M., Bataille, T. and Halley, H. (2000) Reversible
Mem., 10: 525–530. inactivation of the hippocampal mossy fiber synapses in mice
576

impairs spatial learning, but neither consolidation nor mem- Rolls, E.T. (1996) A theory of hippocampal function in mem-
ory retrieval, in the Morris navigation task. Neurobiol. ory. Hippocampus, 6: 601–620.
Learn. Mem., 73: 243–257. Rolls, E.T. and Kesner, R.P. (2006) A computational theory of
Lazaris, A., Talpos, J.C., Dias, R., Saksida, L.M. and Bussey, T.J. hippocampal function, and empirical tests of the theory.
(2006) Lesions of the hippocampus impair a pair-wise spatial Prog. Neurobiol., 79: 1–48.
discrimination task in a touchscreen-equipped operant box. 5th Saksida, L.M., Bussey, T.J., Buckmaster, C.A. and Murray,
Forum of European Neuroscience, Program No. A160.9 Poster E.A. (2006) No effect of hippocampal lesions on perirhinal
board 390. FENS Forum Abstracts, Vienna, Austria. cortex-dependent feature-ambiguous visual discriminations.
Lee, I. and Kesner, R.P. (2002) Differential contribution of Hippocampus, 16: 421–430.
NMDA receptors in hippocampal subregions to spatial Shapiro, M.L. and Olton, D.S. (1994) Hippocampal func-
working memory. Nat. Neurosci., 5: 162–168. tion and interference. In: Schacter D.L. and Tulving E.
Lee, I. and Kesner, R.P. (2003) Differential role of dorsal hip- (Eds.), Memory Systems 1994. MIT Press, London,
pocampus subregions in spatial working memory with short pp. 141–146.
versus intermediate delay. Behav. Neurosci., 117: 1044–1053. Shors, T.J. (2002) Neurogenesis may relate to some but not all
Lee, I. and Kesner, R.P. (2004a) Differential contributions of types of hippocampal-dependent learning. Hippocampus, 12:
dorsal hippocampal subregions to memory acquisition and 578–584.
retrieval in contextual fear-conditioning. Hippocampus, 14: Shors, T.J. (2004) Memory traces of trace memories: neuro-
301–310. genesis, synaptogenesis and awareness. Trends Neurosci., 27:
Lee, I. and Kesner, R.P. (2004b) Encoding versus retrieval of 250–256.
spatial memory: double dissociation between the dentate Stringer, S.M., Rolls, E.T. and Trappenberg, T.P. (2004) Self-
gyrus and the perforant path inputs into CA3 in the dorsal organising continuous attractor networks with multiple ac-
hippocampus. Hippocampus, 14: 66–76. tivity packets, and the representation of space. Neurl. Net-
Leutgeb, S., Leutgeb, J.K., Treves, A., Moser, M.B. and Moser, wk., 17: 5–27.
E.L. (2004) Distinct ensemble codes in hippocampal areas Sutherland, R.J., Whitshaw, I.Q. and Kolb, B. (1983) A be-
CA3 and CA1. Science, 305: 1295–1298. havioural analysis of spatial localization following electro-
Lisman, J.E. (1999) Relating hippocampal circuitry to function: lytic, kainate- or colchicine-induced damage to the
recall of memory sequences by reciprocal dentate-CA3 inter- hippocampal formation in the rat. Behav. Brain Res., 7:
actions. Neuron, 22: 233–242. 133–153.
Martinez, C.O., Do, V.H., Martinez Jr., J.L. and Derrick, B.E. Tanila, H. (1999) Hippocampal place cells can develop distinct
(2002) Associative long-term potentiation (LTP) among ex- representations of two visually identical environments.
trinsic afferents of the hippocampal CA3 region in vivo. Hippocampus, 9: 235–246.
Brain Res., 940: 86–94. Tilson, H.A., Rogers, B.S., Crimes, L., Harry, J.G., Peterson,
McDonald, R.J. and White, N.M. (1995) Hippocampal and N.J., Hong, J.S. and Dryer, R.S. (1987) Time-dependent ne-
nonhippocampal contributions to place learning in rats. Be- urobiological effects of colchicine administered directly into
hav. Neurosci., 109: 579–593. the hippocampus of rats. Brain Res., 408: 163–172.
McLamb, R.L., Mundy, W.R. and Tilson, H.A. (1988) Intra- Vago D.R., Bevan, A. and Kesner, R.P. (2007) The role of the
dentate colchicine disrupts the acquisition and performance direct perforant path input to the CA1 subregion of the dor-
of a working memory task in the radial arm maze. Neuro- sal hippocampus in memory retention and retrieval. Hippo-
toxicology, 9: 521–528. campus (in press).
Mizumori, S.J.Y., Perez, G.M., Alvarado, M.C., Barnes, C.A. Vazdarjanova, A. and Guzowski, J.F. (2004) Differences in
and McNaughton, B.L. (1990) Reversible inactivation of the hippocampal neuronal population responses to modifications
medial septum differentially affects two forms of learning in of an environmental context: evidence for distinct, yet com-
rats. Brain Res., 528: 12–20. plementary, functions of CA3 and CA1 ensembles. J. Neu-
Nakazawa, K., Sun, L.D., Quirk, M.C., Rondi-Reig, L., Wil- rosci., 24: 6489–6496.
son, M.A. and Tonegawa, S. (2003) Hippocampal CA3 Walsh, T.J., Schulz, D., Tilson, H.A. and Schmechel, D.E.
NMDA receptors are crucial for memory acquisition of one- (1986) Colchicine-induced granule cell loss in rat hippocam-
time experience. Neuron, 38: 305–315. pus: selective behavioral and histological alterations. Brain
Nanry, K.P., Mundy, W.R. and Tilson, H.A. (1989) Colchicine- Res., 389: 23–36.
induced alternations of reference memory in rats: role of Witter, M.P., Groenewegen, H.J., Lopes da Silva, F.H. and
spatial versus non-spatial task components. Behav. Brain Lohman, A.H. (1989) Functional organization of the extrin-
Res., 35: 45–53. sic and intrinsic circuitry of the parahippocampal region.
Poucet, B. (1989) Object exploration, habituation, and response Prog. Neurobiol., 33: 161–253.
to a spatial change in rats following septal or medial frontal Xavier, G.F., Oliveira-Filho, F.J.B. and Santos, A.M.G.
cortical damage. Behav. Neurosci., 103: 1009–1016. (1999) Dentate gyrus-selective colchicine lesion and
O’Reilly, R.C. and McClelland, J.L. (1994) Hippocampal con- disruption of performance in spatial tasks: difficulties in
junctive encoding, storage, and recall: avoiding a trade- off. ‘‘place strategy’’ because of a lack of flexibility in the use of
Hippocampus, 4: 661–682. environmental cues? Hippocampus, 9: 668–681.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 31

Models, structure, function: the transformation


of cortical signals in the dentate gyrus

László Acsády1, and Szabolcs Káli1,2

1
Institute of Experimental Medicine, Hungarian Academy of Sciences, PO Box 67, 1450 Budapest, Hungary
2
HAS-PPCU-SU Neurobionics Research Group, Budapest, Hungary

Abstract: Our central question is why the hippocampal CA3 region is the only cortical area capable of
forming interference-free representations of complex environmental events (episodes), given that appar-
ently all cortical regions have recurrent excitatory circuits with modifiable synapses, the basic substrate for
autoassociative memory networks. We review evidence for the radical (but classic) view that a unique
transformation of incoming cortical signals by the dentate gyrus and the subsequent faithful transfer of the
resulting code by the mossy fibers are absolutely critical for the appropriate association of memory items by
CA3 and, in general, for hippocampal function.
In particular, at the gate of the hippocampal formation, the dentate gyrus possesses a set of unusual
properties, which selectively evolved for the task of code transformation between cortical afferents and the
hippocampus. These evolutionarily conserved anatomical features enable the dentate gyrus to translate the
noisy signal of the upstream cortical areas into the sparse and specific code of hippocampal formation,
which is indispensable for the efficient storage and recall of multiple, multidimensional memory items.
To achieve this goal the mossy fiber pathway maximally utilizes the opportunity to differentially regulate
its postsynaptic partners. Selective innervation of CA3 pyramidal cells and interneurons by distinct ter-
minal types creates a favorable condition to differentially regulate the short-term and long-term plasticity
and the motility of various mossy terminal types. The utility of this highly dynamic system appears to be the
frequency-dependent fine-tuning of excitation and inhibition evoked by the large and the small mossy
terminals respectively. This will determine exactly which CA3 cell population is active and induces per-
manent modification in the autoassociational network of the CA3 region.

Introduction The learning of arbitrary associations of complex,


multimodal items during a single exposure (i.e.,
The hippocampus, a peculiar cortical structure, episodic memory) is irreversibly compromised fol-
has long been known to be involved in higher or- lowing hippocampal lesions, but the acquisition of
der cognitive functions, most notably, memory simple associative pairings (e.g., classical condi-
formation and spatial navigation (Scoville and tioning) and motor learning remain intact (Squire,
Milner, 1957; O’Keefe and Dostrovsky, 1971). Not 1992). Hippocampal-dependent memory traces
all kinds of memory depend on the hippocampus. can be used flexibly, i.e., they can be activated in
a context different from the one where they were
Corresponding author. Tel.: +36-1-210-9413; learned. Clinical studies in humans, as well as a
Fax: +36-1-210-9412; E-mail: acsady@koki.hu large number of behavioral and physiological

DOI: 10.1016/S0079-6123(07)63031-3 577


578

experiments in other mammalian species (mainly CA3 pyramidal cells, as well as those between py-
rodents) have also implicated the hippocampus in ramidal cells in CA3 and their primary down-
the formation and flexible use of world-centered stream target CA1, are subject to associative and
(allocentric) spatial representations (O’Keefe and cooperative long-term potentiation (LTP), a pos-
Nadel, 1978). It has been argued repeatedly that sible molecular mechanism underlying memory
these two domains share several important char- formation (Debanne et al., 1998). Interestingly,
acteristics, and may have fairly similar computa- however, many features of CA3 pyramidal cells
tional requirements. In particular, both episodic are shared by pyramidal cells of other cortical ar-
memory and spatial navigation may require the eas. Neocortical pyramidal cells have just as pro-
fast storage and subsequent recall of specific con- fuse recurrent local collaterals and their synapses
junctions of environmental stimuli. It has been are subject to long-term plastic changes. CA3 and
suggested that the bulk of neocortex, which is as- layer II–III cells also have many intrinsic electro-
sumed to be involved in continuously creating and physiological characteristics in common. Thus,
refining an internal representation of the general autoassociative networks with plastic synapses are
structure of the observed world, may not be well abundant in the cortex. Therefore, it is not imme-
suited for the rapid, interference-free acquisition of diately obvious why the formation of complex
such specific memory traces (McClelland et al., memory traces is restricted to the hippocampal
1995). On the other hand, the hippocampus might formation and cannot be performed by other mul-
be optimized for exactly this operation, and could timodal associational cortical areas.
thus complement the generic learning capabilities Two features of the CA3 area, however, clearly
of the neocortex. distinguish it from other cortical regions. CA3 py-
However, the hippocampus constitutes only a ramidal cells receive a prominent excitatory input
tiny fraction of the cortical areas, and it has a rel- to their proximal apical dendrites (Ramon y Cajal,
atively simple structure. It receives input from and 1911; Claiborne et al., 1986), which enables an
sends information back to multimodal associatio- unusually strong spike coupling between an up-
nal cortical areas. The question is why neocortex stream region — the dentate gyrus (DG) — and
needs the hippocampal loop to implement rapid CA3 (Henze et al., 2002). The faithfulness of this
learning of arbitrary complex associations? Why is synaptic transmission is unparalleled in excitatory
it that other cortical areas with multiple cellular cortical circuits. The second feature (shared by the
layers cannot do the same job? What is so special principal cells of the other hippocampal subre-
about hippocampus that enables it to establish the gions) is the way in which the firing pattern of
most complex memory traces? In short, what is the pyramidal cell codes the environmentally relevant
‘‘trick’’ of the hippocampus? stimuli. Most pyramidal cells in other cortical
For a long time, the central role for memory regions are characterized by higher spontaneous
formation was assigned to the hippocampal CA3 firing rates, and a lower rate of modulation by the
region (Marr, 1971; McNaughton and Morris, appropriate stimuli. In sharp contrast, back-
1987; Rolls, 1989). This hippocampal subfield was ground activity in hippocampal principal cells,
favored by most computational neuroscientists and granule cells in particular, is very low (Muller
and electrophysiologists since its principal cells, et al., 1987; Barnes et al., 1990; Quirk et al., 1992;
the CA3 pyramidal cells, form a so-called ‘‘auto- Jung and McNaughton, 1993). However, when
associative memory network’’ with their abundant, hippocampal principal cells participate in infor-
local recurrent collaterals (Li et al., 1994). In com- mation transfer (e.g., place cells), their activity in-
putational models these types of network were creases enormously. In computational terms, the
found to be optimal for the efficient storage of a hippocampus uses a sparse code, whereas coding
large number of memory items and for reactiva- in other cortical areas is denser. Apparently, the
tion of complete memory traces if only part of the hippocampus and the other cortical regions
trace was provided, which are key components of ‘‘speak’’ different neuronal languages. Since the
episodic memory. Moreover, the synapses among environmentally specific information reaches the
579

hippocampus via other cortical areas, the imme- subsequently, the rest of neocortex. This operation
diate consequence of the difference in coding strat- is referred to as pattern completion, or autoasso-
egies is that densely encoded cortical information ciative memory function. Finally, the hippocam-
reaching the hippocampus needs to be translated pus may also be capable of autonomously
into a sparser hippocampal code. In other words, reactivating stored memory traces in the absence
an interface is needed between the hippocampus of specific retrieval cues, thereby reactivating com-
and neocortex. The main task of this interface plete cortical memory representations during ‘‘off-
would be the translation of the neocortical code line’’ behavioral states (e.g., slow-wave sleep).
into a hippocampal one, and to transfer the new Such replay may contribute in various ways to
code as efficiently as possible to the next station of the consolidation (transfer to a final repository)
information processing. According to the classic and maintenance of episodic and semantic mem-
view, this interface is the DG, the first step of the ory (Kali and Dayan, 2004).
trisynaptic hippocampal loop. But do we have any reason to believe that the
We argue that the DG is the ‘‘odd-man-out’’ hippocampus is even capable of carrying out these
among cortical regions. In particular, at the gate of operations, and if so, that it is in some sense opt-
the hippocampal formation, the DG possesses a imized for exactly these tasks? The most widely
set of unusual properties that selectively evolved cited evidence is the existence of the extensive re-
for the task of code transformation between cor- current collateral network of pyramidal neurons in
tical afferents and the hippocampus. These evolu- area CA3. Such recurrent networks are the classic
tionarily conserved anatomical features enable the examples of autoassociative memory devices,
DG to translate the noisy signal of the upstream whose properties have been extensively investi-
cortical areas to the sparse and specific code of gated by theoretical means and computer simula-
hippocampal formation, which is indispensable for tions (Marr, 1971; McNaughton and Morris, 1987;
the formation of multiple, multidimensional mem- Willshaw and Buckingham, 1990; Treves and
ory items. Rolls, 1992; Samsonovich and McNaughton,
1997; Kali and Dayan, 2000). In particular, the
storage capacity of such networks, i.e., the number
Computational requirements for the formation of patterns they can store and retrieve reliably, has
of episodic memories been determined (Treves and Rolls, 1992), and was
found to depend substantially on the properties of
There is a sort of general consensus about how the the set of input patterns that we attempt to store.
hippocampus could contribute to cortical memory Capacity is roughly inversely proportional to the
functions (Alvarez and Squire, 1994; Treves and sparsity of individual patterns (i.e., the proportion
Rolls, 1994; McClelland et al., 1995; Kali and of active units in a pattern), and generally in-
Dayan, 2004; Rolls and Kesner, 2006). First, the creases as the overlap between different stored
hippocampus is assumed to be capable of rapidly patterns decreases. Therefore, it is reasonable to
creating and storing a memory trace, which is dis- assume that the need to maximize storage capacity
tinct from all existing traces. The memory trace is is an important reason for the conspicuously low
associated with a snapshot of activity in medial activity levels of principal cell populations in the
temporal neocortex (particularly entorhinal cor- hippocampus (the proportion of active principal
tex), which, in turn, is thought to represent a com- cells in any hippocampal subfield at any given
pressed version of activity in the rest of neocortex. moment is thought to be on the order of a few
Second, the hippocampus is thought to be capable percent (Barnes et al., 1990; Quirk et al., 1992;
of retrieving particular stored traces if the ent- Jung and McNaughton, 1993; Treves and Rolls,
orhinal activity pattern provides only a partial or 1994 #5013).
noisy version of the corresponding original pat- However, activity patterns in most areas of neo-
tern. Using this cue, the hippocampus reinstates cortex, including entorhinal cortex, appear to be
the original pattern in the entorhinal cortex and, much denser (involving a larger proportion of
580

neurons at any given time), which is thought to be patterns, which constitute its only major cortical
beneficial for generalization (McClelland et al., input? Since no direct experimental investigation
1995), an important characteristic of the kind of of this issue has been undertaken, we need to rely
representational learning that the neocortex may on indirect evidence and the results of computa-
be engaged in. On the other hand, if the optimal tional studies to try to answer this question. At an
type of activity pattern (dense vs. sparse) is differ- abstract level, well-known computational algo-
ent in the neocortex and the hippocampus, then rithms exist which create sparse, pattern-separated
the information contained in the entorhinal dense representations from distributed input patterns. A
code needs to be ‘‘translated’’ into a sparse code simple example is the competitive-learning pattern
when hippocampal representations are created classification device described by Rumelhart and
(and vice versa). In principle, such translation Zipser (1986), which has been shown to be capable
could be implemented directly by a single set of of generating hippocampal place field-like activity
projections from the input area to the autoassoci- patterns from input patterns resembling neocorti-
ative network (i.e., by the direct perforant path cal sensory representations (Sharp, 1991). This
input from entorhinal cortex to the CA3 region), algorithm operates on binary input patterns and
without an intervening specific interface for spars- generates binary output patterns. For each pres-
ification. However, as argued on theoretical entation of an input pattern, the current values of
grounds by Treves and Rolls (1992), a projection the synaptic weights (which are initially set to ran-
with the characteristics of the perforant path — a dom values) are used to determine the feedforward
large number of relatively weak, associatively activation of units in the output layer. Then the
modifiable synapses on each target cell — may output unit with the highest level of feedforward
be optimal during retrieval, but a different type of input is allowed to become active (this step is
input to the CA3 recurrent network — one with assumed to reflect the action of feedforward and
a small number of individually strong synapses feedback inhibitory circuits), and the incoming
per cell — is probably required for the storage of weights of this unit are allowed to change.
new memory traces. The mossy fiber pathway, the This weight change is assumed to be Hebbian in
projection to CA3 from granule cells, meets the nature: weights from active input units are in-
requirements for this second type of input. Addi- creased, while weights from inactive units are de-
tional theoretical and computer simulation studies creased in such a way that the sum of all incoming
by O’Reilly and McClelland (1994) indicated that weights remains constant (and identical to the sum
the two-stage pathway from EC to DG to CA3 of incoming weights to all other output units). This
could perform very effective pattern separation, way, the ‘‘winning’’ unit will have an even higher
provided that individual mossy fiber connections level of feedforward activation the next time the
were sufficiently strong to transfer the benefits of same input pattern appears. It will also show an
pattern separation in the DG to CA3. They further increased response to other similar input patterns;
argued that the coexistence of this indirect path- on the other hand, its response to patterns that are
way with the direct EC–CA3 connection enables very different (with a low degree of overlap) will
the hippocampus to avoid an inherent conflict be- diminish. As a result, different output units will
tween pattern separation and pattern completion, eventually respond to different kinds of patterns,
both of which are important for the efficient and end up partitioning the input space among
operation of the hippocampal autoassociator. In themselves into non-overlapping groups of similar
summary, these studies suggest that a possible role patterns. The algorithm performs both sparsificat-
of the DG is to form sparse, pattern-separated ion — since only a single output unit becomes ac-
representations of entorhinal activity patterns, and tive for any given (distributed) input pattern —
transmit this sparse representation reliably for and orthogonalization (pattern separation) —
subsequent storage in the CA3 recurrent network. since relatively dissimilar, but still overlapping
But how does the DG create sparse representa- input patterns end up activating different output
tions from the relatively dense entorhinal activity units (zero overlap).
581

However, is there any experimental evidence memory task where the proximity of relevant
that sparsification requires a separate relay station, locations could be varied systematically. They
and that the DG has unique features — besides its found that DG lesions impaired performance at
well-known ‘‘detonator’’ type of terminals — for small, but not at large separations, consistent with
code conversion and reliable transmission? In the the role of the DG in spatial pattern separation.
following pages we review behavioral, morpholog- Lassalle et al. (2000) examined the consequences of
ical, and physiological data relevant to these selective and reversible inactivation of mossy fiber
questions. synapses in CA3 in mice during various stages of a
reference memory task in the Morris water maze.
They found that the mossy fiber input from DG to
Lesion studies CA3 was essential during learning, but not during
the retrieval phase of the task, or in the period
Let us first examine whether the putative role of directly following learning (the early stages of
the DG as described above is consistent with be- consolidation). Similarly, Lee and Kesner (2004)
havioral data on the effects of specific lesions. In attempted to distinguish encoding and retrieval
general, we expect that some, but not all tasks that deficits during the acquisition of a navigation task
are sensitive to global hippocampal lesions will in the Hebb-Williams maze following lesions of
also be sensitive to more specific lesions of the DG, various type. They found that lesions of the DG
and might hope that the pattern of impairments impaired learning, but not retrieval, in this task;
that occur after selective DG lesions sheds light on conversely, lesions of the perforant path input to
the role of the DG in hippocampal processing. CA3 affected retrieval, but not learning.
Most lesion studies have taken advantage of the In summary, the available behavioral data from
fact that intrahippocampal injections of colchicine rodents with lesions to dentate granule cells or their
cause a fairly selective destruction of the granule mossy fiber output are generally consistent with a
cells of the DG, but other techniques, such as crucial role of the DG in providing input to area
neonatal X-ray irradiation and adrenalectomy, CA3 during the acquisition of allocentric spatial
which also cause a similar pattern of damage, have information, and provide some support for a more
also been used. Perhaps the most consistent find- specific function in (spatial) pattern separation. So
ing after DG lesions in rodents has been a severe what are the morphological peculiarities of the DG
impairment in the acquisition of the reference that support its role in orthogonalization and make
memory task in the Morris water maze (Suther- it indispensable for proper CA3 function?
land et al., 1983; McNaughton et al., 1989; Conrad
and Roy, 1993; Xavier et al., 1999). Working
Morphological arguments — heterogeneous
memory versions of the task are also affected,
terminal types of the mossy fibers
although perhaps to a lesser extent (Xavier et al.,
1999; Jeltsch et al., 2001). Reference and working
Let’s see first how Cajal described of the unusual
memory performance in the radial arm maze are
terminals types of the mossy fibers:
also compromised (McNaughton et al., 1989;
Jeltsch et al., 2001). In some more recent studies, ‘‘ythere arise either short and thick
DG lesions were also found to cause impairment divergent appendages or quite long
in a delayed-matching-to-place task, in a temporal fine threads that end in a swelling. Thus,
task (Costa et al., 2005), and in a task which we reproduced here the arrangement
required the detection of metric distance change (although less distinctly) that we
between objects (Goodrich-Hunsaker et al., 2005). described in certain branched fibers of
Recently, there have been some attempts to test the cerebellum the mossy fibers. There-
more directly the proposed contributions of the fore without further ado let us apply the
DG to hippocampal function. In particular, same name to the axons of the granules
Gilbert et al. (2001) designed a short-term spatial of the fascia dentata.’’
582

It is clear from the above description that the takes place at the perforant path input of the cod-
investigator who named the mossy fibers by ex- ing (granule) cells. However, feedback inhibition is
amining Golgi-stained material clearly identified well-known in all cortical regions. Is there any
that this peculiar fiber system has more than one reason to suppose that feedback inhibition is more
terminal type. The well-recognizable giant endings powerful in the DG than in other cortical regions,
(the large mossy terminals) gave rise to thin fila- which makes it especially useful for pattern sepa-
mentous structures that ended in terminal-like ration?
swellings. Apparently the name ‘‘mossy fiber’’ is In cortical regions, interneurons constitute
based on the presence of these filopodial exten- 10–20% of all the neurons. Cortical pyramidal
sions. Later electron microscopic work identified neurons innervate their postsynaptic principal and
that filopodial terminals indeed establish asym- interneuron targets in a quasi-random manner,
metrical synapses (Amaral, 1979; Claiborne et al., i.e., the incidence of the targets is determined by
1986) (Fig. 1). In addition, a third terminal type the relative distribution of the neuron types
has been described, small ‘‘en passant’’ boutons, (Gulyas et al., 1993; Sik et al., 1993). Thus, the
which resemble the most the conventional axonal estimated ratio of interneurons among the postsy-
varicosities of cortical pyramidal cells (Claiborne naptic targets of cortical pyramidal cells is around
et al., 1986) (Fig. 1). Together the two smaller 10%.
mossy terminal types far outnumber their larger In the DG granule cells have axon collaterals
counterpart, but still, the total number of termi- only in the hilus below the granule cell layer not
nals is actually very low (Claiborne et al., 1986; in stratum granulosum or stratum moleculare
Acsady et al., 1998). A single granule cell has no (Claiborne et al., 1986; Acsady et al., 1998). As a
more than 200 terminals, which is at least two consequence only inhibitory neurons having so-
orders of magnitude less than the number of vari- mata and/or dendrites in the hilus can participate
cosities along the axonal arbor of a cortical py- in feedback inhibition. Close to 50% of the hilar
ramidal cell. How can these few but variable neurons are GABAergic (Houser and Esclapez,
terminals account for the code conversion and 1994) which suggests that inhibitory cells may be
efficient transmission as outlined above? abundant among the postsynaptic targets of gran-
ule cells. The 5–8 hilar collaterals of the granule
cells possess around 7–12 large mossy boutons and
Unique features for code conversion 102–147 small terminals (filopodial and en passant
boutons) (Acsady et al., 1998). The postsynaptic
The computational studies summarized above sug- targets of the small terminal types are almost ex-
gested that memory formation in the hippocampus clusively interneurons, whereas targets of the large
may be a two-step process: the sparsification of the mossy terminals are mainly excitatory mossy cells
entorhinal signal by granule cells, followed by an (Acsady et al., 1998) (Fig. 2). Thus, due to the
association of the now sparse code in the CA3 surprising target selectivity of granule cell terminal
network. Modeling studies suggest that the ‘‘win- types, interneurons may constitute up to 90% of
ner-take-all’’ method of sparsification requires the the postsynaptic targets of the mossy fibers in the
recruitment of strong feedback inhibition. Accord- hilus, in contrast to the 10–20% GABAergic tar-
ing to this scheme, the output of a small popula- gets in other cortical regions. Many of these hilar
tion of granule cells that become active in a given neurons provide feedback inhibitory control of the
environmental context (e.g., during a couple of granule cells, suggesting that proportionally stron-
theta cycles as the rat passes through their place ger feedback inhibition is recruited here than in
fields) activate GABAergic interneurons which other cortical regions.
exert fast and strong feedback inhibition on the A characteristic cell type of the hilus, the
somata and dendrites of the non-coding granule somatostatin-immunoreactive interneuron, pro-
cells, shunting their entorhinal inputs and preclud- vides a good example of the strong recurrent
ing their firing. As a result, synaptic plasticity only inhibition operating in this system. The
583

Fig. 1. Electron micrographs of different terminal types along the mossy fibers in the CA3 region (A–C, E) and of a CA3 pyramidal
cell terminal (D) for comparison. All electron micrographs have the same magnification. A, B, A small en passant terminal establishes a
single asymmetrical synapse on a dendritic shaft, showing the characteristic long perforated postsynaptic density of small terminal
types on interneurons (arrows). C, A filopodial extension of a mossy fiber terminal forms a synapse (arrow) with a Substance P
receptor-immunoreactive interneuron. D, A CA3 pyramidal cell establishes asymmetrical synapse on a simple spine of a CA1 py-
ramidal neuron. E, A large, double-headed mossy fiber terminal forms multiple contacts (arrows) with thorny excrescences of a CA3
pyramidal cell. All active zones converged on the same pyramidal cell. The individual release sites are short. Scale bars: A–D, 0.5 mm;
E, 1 mm. Reprinted with permission from Acsady et al., 1998, Society for Neuroscience.

somatostatin-containing hilar neurons restrict they are also called HIPP cells, i.e., HIlar
their entire dendritic arbor to the hilus and inner- Interneurons with Perforant Path associated
vate the dendritic segment of granule cells in the axon terminal) (Han et al., 1993; Sik et al.,
zone where entorhinal afferents terminate (hence 1997). Unlike many other interneuron types that
584

several thousand granule cells on any one of these


interneurons. The axons of HIPP cells terminate in
the same layers as the perforant path, and there-
fore are in a critical position to modulate the
information transfer from the entorhinal cortex to
the DG (Han et al., 1993). A single HIPP cell may
form an unusually large number of axon terminals
in the molecular layer (up to 80,000 compared to
5000–10,000 in the case of a basket cell), the vast
majority of which innervate granule cells (Sik
et al., 1997). The little data available about the
activity of HIPP cells suggest that these neurons
are not fast-firing cells; rather, their activity is
principally driven by the firing pattern of granule
cells, a key feature for appropriately timed feed-
back inhibition (Buckmaster and Schwartzkroin,
Fig. 2. Filopodial extensions of mossy fiber terminals are spe- 1995).
cialized to innervate GABAergic cells. Artistic rendition of two
large mossy terminals, each equipped with four filopodial ex-
Another unique morphological feature of the
tensions (large arrowheads). The mossy fibers were labeled by connectivity in DG is the lack of interaction
intracellular injection of biocytin into two neighboring granule among the major inhibitory basket cell classes
cells. All filopodial terminals were examined in the electron and among basket cells and other interneurons
microscope (not shown) and all contacted the dendrites or (Acsady et al., 2000). In other hippocampal re-
spines of altogether six GABAergic neurons. Four of the GAB-
Aergic neurons were identified by their Substance P receptor-
gions as well as in other cortical regions, basket
content and two of them by ultrastructural characteristics. Five cells densely innervate other basket- and non-
of the six postsynaptic interneurons were spiny cells. Arrows basket-type interneurons (Sik et al., 1995; Cobb et
point to the main axons. Reprinted with permission from Ac- al., 1997; Tamas et al., 1998) forming an interact-
sady et al., 1998, Society for Neuroscience. ing local GABAergic network. For example, in the
CA1 region the somatic region of parvalbumin-
are characterized by a smooth dendritic surface, positive basket cells is contacted by more GAB-
the dendrites of HIPP cells are densely covered Aergic terminals than the somatic region of py-
with thousands of long thin, elaborated spines ramidal cells (Gulyas et al., 1999; Megias et al.,
(Baude et al., 1993). In contrast to the compart- 2001). In sharp contrast, GABAergic cells in the
mentalized spines of the pyramidal cells, which hilus of the DG receive, on average, 15–40 times
usually receive a single terminal, these spines are less input from local basket cells than mossy cells,
contacted by multiple asymmetrical synapses (up the principal excitatory cell type of this region
to 8–10) (Acsady et al., 1998). Thus, these spines (Acsady et al., 2000). Since GABAergic cells in-
increase the total synaptic input of the HIPP cells hibit each other, this connectivity pattern suggests
enormously. Small terminals of granule cells have minimal disinhibitory influence in the hilus and
been described to contact these spines (at least different GABAergic network dynamics.
30% of the total synaptic output of granule cells But what are the physiological properties of the
contacts HIPP cells) (Acsady et al., 1998), whereas granule cell–interneuron synapses? Examination of
the axons of CA3 pyramidal cells that project back synaptic transmission at the granule cell–basket
to the hilus selectively avoided HIPP cells (Wittner cell synapses demonstrated very fast kinetics: the
et al., 2006), suggesting that the majority of the postsynaptic conductance of the unitary current
excitatory input of these cells originates from demonstrated submillisecond rise and decay
granule cells. Since the number of contact between (Geiger et al., 1997). This effect was largely at-
a granule cell and a HIPP cell is only 1 or 2 these tributed to the high synchrony of transmitter re-
morphological data indicate the convergence of lease and the rapid time course of AMPA receptor
585

deactivation. The fast postsynaptic response efficacy of the input in driving the postsynaptic
allows rapid activation of feedback inhibition, cell to threshold. Recent in vitro and in vivo data
which supports a role in pattern separation. confirm this assumption. Monosynaptic AMPA/
These data suggest that the basic principles of kainate receptor-mediated EPSCs from granule
connectivity between excitatory and inhibitory cell–pyramidal cell pairs had a mean peak ampli-
neurons in the DG are significantly different from tude of 163.0723 pA at 70 mV in organotypic
those in any other cortical region. The peculiar slice cultures which displayed morphological prop-
arrangement of excitatory and inhibitory connec- erties similar to the in vivo condition (Mori et al.,
tions in the DG suggests an unusually strong re- 2004). Strong excitatory action has been described
cruitment of inhibition that can be used to earlier between granule cells and their excitatory
suppress the activity of granule cells in a compet- targets in the hilus, the mossy cells (Scharfman
itive manner. As a result, only granule cells with et al., 1990). In the in vivo anesthetized preparation,
the strongest entorhinal excitatory drive will par- repetitive firing in a single granule cell reliably
ticipate in the information transfer to CA3. This induced action potentials in monosynaptically
small population of active granule cells could connected CA3 pyramidal cells (Henze et al.,
effectively prevent large granule cell populations 2002). It has to be emphasized that such strong
from reaching firing threshold via the strong feed- coupling is extremely rare among excitatory cells
back inhibitory system outlined above. Only those in cortical circuits. The general rule is that a large
entorhinal inputs will be potentiated which contact number of excitatory inputs have to be simultane-
active granule cells. This would further strengthen ously active to reach postsynaptic spike threshold.
the competitive process, which could lead to the But how sparse is the detonator signal in mor-
sparsification of the entorhinal signal resulting, phological terms?
e.g., in the sharp and focused place fields of the The number of large mossy terminals along the
granule cells which are in sharp contrast to the single unbranching axon of granule cells within
grid-like entorhinal signal (see below). The next area CA3 is very low (average: 12.3; range: 10–18)
step is to faithfully transmit this recoded cortical (Acsady et al., 1998). If one considers that a single
signal to the associational station, the CA3 region, CA3 pyramidal cell may have up to 60,000 termi-
and to ensure that only the activated subset of nals, it is straightforward to conclude that granule
CA3 pyramidal cells would be included in the cells are a specific cortical cell type designed to
autoassociation network responsible for memory transfer the sparse code generated in DG very
storage. effectively to only a restricted set of postsynaptic
pyramidal cells.
Two recent studies described additional surpris-
Unique features for efficient and sparse transmission ing features of the mossy fibers. These features not
only support faithful transmission through this
The giant mossy terminals display all the morpho- pathway, but also demonstrate the computational
logical features of a classic ‘‘detonator’’ or power of the axon terminals in unexpected ways.
‘‘driver’’ type terminal. No other cortical synapse The first study (Engel and Jonas, 2005) describes
is comparable to them, but several subcortical the active properties of the large mossy terminals,
structures (e.g., thalamus, cerebellum) utilize sim- which facilitates reliable transmission of high
ilar synaptic arrangements to secure faithful frequency trains of action potentials. Apparently,
synaptic transmission (Sherman and Guillery, mossy terminals have a very high density of spe-
1998). A single large mossy terminal may estab- cialized Na+ channels with faster activation and
lish up to 30–40 release sites, all converging on the inactivation kinetics than somatic Na+ channels.
proximal dendrite of a single postsynaptic pyram- These Na+ channels enable reliable action poten-
idal cell (Chicurel and Harris, 1992; Acsady et al., tial invasion into large mossy terminals and in-
1998). One would predict that the short electro- crease presynaptic Ca2+ influx, resulting in up to
tonic distance from the soma would maximize the 16-fold increase of transmitter release. In addition,
586

modeling studies suggest that this active property activating 120 CA3 pyramidal cells, each of
of the terminals is absolutely necessary to induce which has around 30,000 terminals within the CA3
Ca2+ influx into the filopodial extensions to trig- (Li et al., 1994). Thus, other solutions are needed
ger glutamate release at the mossy fiber–interneu- to solve Problem 1.
ron synapse. This mechanism may underlie the Similar to the hilus, the number of small mossy
high release probability of the filopodial connec- fiber terminals in the CA3 region exceeds the
tion compared to the mossy fiber–CA3 pyramidal number of large mossy fiber terminals by a factor
cell synapse (Jonas et al., 1993; Lawrence et al., of at least four (Acsady et al., 1998). As in the
2004). hilus, the small terminals selectively innervate in-
The second study (Alle and Geiger, 2006) dem- hibitory cells. All examined inhibitory cell classes
onstrates that mossy fibers transmit not only ac- (perisomatic, dendritic, and interneuron-selective)
tion potentials, but also postsynaptic potentials were among the postsynaptic elements of granule
originating in the soma-dendritic compartment. cells. Since a single mossy fiber rarely innervates a
These ‘‘excitatory presynaptic potentials’’ alter the postsynaptic interneuron via multiple contacts
transmitter release of the subsequent action po- (Acsady et al., 1998), granule cell firing activates
tential (within a 10–20 ms delay), thus enabling the at least four times as many inhibitory as excitatory
axon to integrate subthreshold and suprathreshold cells. Physiological data confirm strong activation
signals provided they are temporally proximal. of this feedforward inhibitory circuit. In paired
In sum, mossy terminals fulfill all criteria for a recordings, a single action potential in the granule
classical, sparse ‘‘detonator’’ synapse. The strong cell induced a biphasic response in the postsynap-
granule cell–pyramidal cell connection, however, tic CA3 pyramidal cell: a brief EPSC followed by a
poses two major problems for the autoassociative pronounced IPSC (Mori et al., 2004). The direc-
network of the CA3 region. tion of the summated charge transfer was outward,
indicating a net inhibitory synaptic response. The
Problem 1. Spontaneously active granule cells, authors calculated that approximately four inter-
which represent only background noise, may in- neurons fired together to evoke the measured in-
duce irrelevant spiking of CA3 pyramidal cells, hibitory responses, which corresponds well to the
resulting in non-coding CA3 networks which may anatomical data. Reliable activation of interneu-
overlap with the coding networks. rons was also observed in vivo. Multiple granule
Problem 2. Spontaneously active CA3 pyramidal cell spikes induced firing in monosynaptically cou-
cells may generate action potentials coincident pled interneurons with quite high probability
with the CA3 spikes evoked by granule cell firing, (Henze et al., 2002) (Fig. 3D), suggesting that al-
which represent the given environment. These non- though the granule cell–interneuron contact rarely
coding EPSPs will be also strengthened in the contains multiple release sites, the single active
autoassociative CA3 network, which would ruin zone of the small terminals is still efficient enough
the orthogonalized, sparse code. to induce postsynaptic firing of the GABAergic
cell. Recruitment of more inhibitory than excita-
A first solution to Problem 1 is that granule cells tory cells suggests that the net effect of the exci-
have very low spontaneous firing rates tatory mossy fiber system on the CA3 pyramidal
(0.1–0.01 spike/s) (Jung and McNaughton, 1993), cell population is, counter-intuitively, inhibitory.
probably due to their hyperpolarized resting mem- A peculiar EEG transient, the dentate spike, can
brane potential (80 mV in vivo) (Penttonen et al., be utilized to demonstrate the impact of granule
1997) and/or strong inhibitory control. Still, if we cell activation on the postsynaptic cell populations
consider that there are one million granule cells per along the dentate–CA3 axis during a ‘‘natural’’
hippocampus in rodents (Seress, 1988) and we take stimulus. Dentate spikes are short-duration, large-
the low end of their spontaneous discharge fre- amplitude field potentials caused by synchronous
quency (0.01 Hz), we can calculate that at least 10 activation of the entorhinal input, which occur
granule cells will fire in any one millisecond, during behavioral immobility and slow-wave sleep
587

C D 0.12
0.03

0.08
Probability

0.02

0.01 0.04

0 0.00
0 2 4 6 8 10 0 2 4 6 8 10

E 1.0 100 Hz F 0.3


0.8 60 Hz
40 Hz
Probability

0.2
Probability

0.6
20 Hz
0.4
10 Hz
0.1
0.2

0.0
0.0
0 0.2 0.4 0.6 0.8 1 1 3 5 7 9
Time (s) Spike number

Fig. 3. Spike transmission dynamics between a granule cell and its interneuron and pyramidal cell targets in CA3c in vivo. (A) Camera
lucida reconstruction of the extracellular electrode track and biocytin-labeled granule cell. Inset, a higher-power view of the mossy fiber
axon near the probe track. Arrowheads, mossy fiber boutons. (B) Superimposed (n ¼ 60) intracellularly evoked action potentials in a
granule cell (bottom traces) and simultaneously recorded extracellular units (filtered 0.8–8 kHz). Note the time-locked response of a
putative pyramidal cell to the granule cell action potentials. (C–D) Cross-correlograms between the evoked granule cell action
potentials and the activity of a putative CA3c pyramidal cell (C) or interneuron (D). The values are shuffle corrected and expressed as
probability (number of unit spikes per bin/total number of granule cell spikes). Arrowhead, peak time of the granule cell action
potential. (E) Representative results of the effect of intratrain frequency on spike transmission probability for a putative pyramidal cell.
(F) Spike transmission probability (shuffle-corrected probability of spike in 6 ms following granule cell spike) as a function of spike
number in evoked 100 Hz train (7s.e.m). Solid line, putative interneurons (n ¼ 24); dotted line, putative pyramidal cells (n ¼ 21). Scale
bars, (A) 50 mm; inset 20 mm; (B) 1 ms, 25 mV, 75 mV. m, molecular layer; g, granule cell layer; h, hilus; IC, intracellular electrode track;
EC, extracellular electrode track.

(Bragin et al., 1995). Extracellular recordings dur- 1997). It is worth mentioning that this pattern of
ing dentate spikes demonstrated increased unit activity is in sharp contrast to the neuronal be-
activity in hilar neurons (many of which are GAB- havior that can be observed during the other major
Aergic), but suppressed multiunit activity in the excitatory field transient in the hippocampus, the
CA3 region (Bragin et al., 1995). Intracellular sharp wave, which originates in CA3. Since in this
studies confirmed depolarization of granule cells case there is no inhibitory ‘‘barrier’’ comparable
and hilar interneurons but hyperpolarization in to the dentate GABAergic network which could
CA3 and CA1 pyramidal cells (Penttonen et al., block the spread of excitation, sharp waves
588

propagate not only to the CA1 region but also to (half duration, 27 ms) (Miles and Wong, 1986) and
parahippocampal cortical regions via polysynaptic accidentally the mossy fiber input can arrive
activation (Chrobak and Buzsaki, 1994, 1996). together with their peak, thus these early back-
Most recently the efficacy of the dentate inhib- ground (unwanted) EPSPs can be potentiated be-
itory ‘‘barrier’’ was demonstrated by comparing fore the disynaptic feed forward inhibition arrives.
the correlation of intracellular activity of neurons A recent report provides a possible solution for
in various cortical fields with the UP and DOWN this problem (Kobayashi and Poo, 2004). This
states in the EEG during slow cortical oscillation study describes LTP at the recurrent synapses of
(Isomura et al., 2006). Surprisingly, all cortical the CA3 network induced by paired stimulation of
regions (including entorhinal cortex and DG) mossy fiber input and the associational/commis-
changed intracellular activity coherently with the sural input. First, this study provides direct and
EEG states with the exception of CA3 region. In elegant evidence for the role of mossy fibers in
contrast to basically all cortical cells CA3 pyram- changing the synaptic strength of the autoassoci-
idal cells were not active during the UP states. ative CA3 matrix. Second, an interesting observa-
Apparently the inhibitory network launched by tion from this study helps resolve the problem of
the dentate during the UP states is able to shield early EPSPs. The results demonstrate that the po-
CA3 even from the most synchronous excitatory tentiation of associational input depends on the
events of the cortical mantle. relative timing of mossy fiber and associational
In summary, anatomical and physiological data spike trains. The potentiation was smaller if addi-
provide convergent evidence for the conclusion tional associational spikes were added before the
that the output of the granule cells activates an paired stimulation, compared to the protocol that
unusually strong feedforward inhibition to area added spikes after the paired (associational/mossy
CA3. Since even a single basket cell is able to block fiber) pulses. The effect depended on mGluR1 ac-
action potential generation in a large number of tivation. Thus, apparently the system favors the
pyramidal cells (Miles et al., 1996), feedforward potentiation of CA3 EPSPs arriving coincidentally
inhibition will effectively reduce the number of or after the dentate signal. Since the mossy fiber
active CA3 pyramidal cells during dentate activa- input is able to induce CA3 spiking in vivo, these
tion. In addition, in the CA3 region, specialized EPSPs will mostly represent the spiking of CA3
interneuron types exist which restrict their dendri- pyramidal cells evoked by the sparse, orthogonali-
tic (Gulyas et al., 1991) or axonal (Vida and zed dentate input. In this way only the associat-
Frotscher, 2000) arbor to stratum lucidum of CA3, ional EPSPs representing dentate activity will be
suggesting selective control of this pathway. Thus, potentiated, but EPSPs arriving earlier, represent-
the likely solution of Problems 1 and 2 is that ing spontaneous CA3 activity, will not.
‘‘background’’ or ‘‘non-coding’’ EPSPs and action
potentials in CA3 are actively inhibited during
dentate–CA3 information transfer by the strong Target-dependent plasticity — the meaning of
feedforward inhibition. In this way only the small various terminals types
population of CA3 cells receiving the decorrelated
sparse dentate signal will be active, and following If we consider the strong feedforward inhibition
the Hebbian rule, only the recurrent synapses be- operating along the dentate–CA3 axis the follow-
tween the activated CA3 cells will be potentiated. ing, third problem arises:
According to the modeling studies (see above) the Problem 3. How to overcome the strong feedfor-
matrix of potentiated synapses will represent the ward inhibition when the specific dentate pattern
memory trace. has to activate the pyramidal cell?
However, what happens to the EPSPs generated
by the spontaneous activity of CA3 pyramidal cells Apparently the solution to Problem 3 is that
immediately before the dentate input arrives? short- and long-term plasticity at large and
The EPSPs among pyramidal cells are quite slow small types of mossy fiber ending are different.
589

The physiological data clearly demonstrate that observed during low frequency trains switched to
the small terminals are not only distinct morpho- excitatory dominant PSPs at high frequencies.
logical units, evolved to contact a large number of At 40 Hz EPSPs dominated the response already
interneurons, but discrete compartments which after the third granule cell action potentials,
harbor distinct molecular machinery for plasticity whereas at 10 Hz the response remained inhibi-
different from their giant cousins. tory even at the 15th action potential.
Small and large mossy terminals display differ- Frequency-dependent facilitation of mossy fiber
ent types of short-term plasticity. The giant mossy transmission was demonstrated in monosynaptic-
terminal–pyramidal cell connection express very ally coupled granule cell — pyramidal cell pairs
strong short-term facilitation (Salin et al., 1996; in vivo as well (Henze et al., 2002) (Fig. 3). The
Toth et al., 2000), reaching up to threefold ampli- probability of postsynaptic pyramidal cell spikes
tude increase on average at the fifth EPSC in case rapidly increased with increasing granule cell firing
of 20 Hz stimulation. (This is highly unusual in the and reached 0.8 at 100 Hz (Fig. 3E), which is an
case of giant ‘‘driver’’-like terminals; e.g., in the extremely high value in cortical circuits and results
thalamus excitatory terminals with a similar in an almost one-to-one relay of the granule cell
synaptic arrangement show strong depression in activity. Within the spike train the probability of
relay cells; Reichova and Sherman, 2004). In con- CA3 pyramidal spikes increased with the number
trast the small mossy terminal–interneuron con- of presynaptic granule cell spikes. The maximum
nection shows short-term depression after spike transmission probability was reached after
repetitive mossy fiber stimulation in about 50% the 4–5th spike (Fig. 3F). In contrast, in the case of
of the interneurons. Others show modest facilita- interneurons, the probability of transmission did
tion at 20 Hz (Toth et al., 2000), but at 40 Hz, all not increase with an increasing number of presy-
responses are depressed after the 8th action naptic spikes at 100 Hz and remained lower than
potential (Mori et al., 2004). The third synapse that of the pyramidal cell (Fig. 3F). These in vitro
in the feedforward inhibitory circuit, the interneu- and in vivo data indicate that differential short-
ron–CA3 pyramidal cell contact, expresses pro- term plasticity in this feed-forward circuit result
nounced short-term depression (Mori et al., 2004) in a frequency-dependent shift of the polarity of
at all frequencies tested. postsynaptic response.
This variability in short-term plasticity at differ- In the freely moving condition, granule cells
ent mossy fiber synapses favors conditions for have very low spontaneous firing rates, which can
a single granule cell action potential to induce rapidly increase to 40 Hz (Jung and McNaughton,
weaker excitation, but relatively strong feedfor- 1993) as the animal enters the place field of the
ward inhibition. Repetitive firing, however, rapidly neuron. As a consequence of the frequency-
increases the effect of excitation and at the same dependent switch from excitation to inhibition de-
time decreases the efficacy of inhibition. Thus, the scribed above, the probability of spike transmis-
net effect of mossy fiber activation on CA3 py- sion between the DG and CA3 is very low at low
ramidal cells depends heavily on the frequency of firing rates, despite the ‘‘detonator’’ nature of the
granule cell firing. The system appears to act as a synapse. When granule cells code the specific
high-pass filter, where spike transmission from information of the environment, the probability
granule cell to pyramidal cell is blocked at low of spike transmission to CA3 becomes very high.
frequencies but favored when the frequency of Thus, granule cells can act as a ‘‘conditional
granule cell discharge increases. Recently this as- detonator’’ as suggested by Henze et al. (2002).
sumption was tested directly in in vitro paired This mechanism solves both the problem of
recordings of granule cells and pyramidal cells filtering out non-coding spontaneous activity
(Mori et al., 2004). The postsynaptic potentials (Problem 1), and resolves the issue that strong
evoked in pyramidal cells by granule cell stimula- feedforward inhibition must be overcome when
tion were measured at increasing frequencies specific information must be transmitted from
(10–40 Hz). The inhibitory dominant PSPs dentate to CA3 (Problem 3).
590

The delicate frequency-dependent balance be- excitation may change and/or the steepness of the
tween excitation and inhibition substantially in- high-pass filter cutoff may increase. In this way,
creases the computational power of the mossy presynaptic mossy fiber–LTP may help the
fibers. It raises the possibility that they not only orthogonalization process performed by the dent-
participate in the faithful transmission of an or- ate–CA3 circuit and creates a good opportunity
thogonalized dentate signal but they themselves for the faithful activation of the same CA3 circuit
participate in creating the non-overlapping repre- in case of repetitive presentation.
sentations in the CA3 region. Recent data suggest Finally, morphological plasticity of the small
(Leutgeb et al., 2007, see below) that the repre- terminals provides yet another way to fine-tune the
sentation of the environment is more orthogonali- balance of excitation and inhibition in the mossy
zed in CA3 than in the DG, which questions the fiber pathway. The structure of the filopodial ter-
role of dentate as the sole contributor to this minals strongly resembles rapidly advancing and
process. Due to its high-pass filter nature, the retracting axonal filopodia observed during axonal
dentate–CA3 circuit may refine the dentate code, development and synapse formation. Recent stud-
and, e.g., participate in creating the single recep- ies of slice cultures indeed demonstrated that over
tive field observed in CA3 place cells as opposed to one-third of the filopodia are highly active
the multiple place fields of dentate place cells. The (De Paola et al., 2003; Tashiro et al., 2003). In
critical variable in this process will be the exact addition, in mature slices, approximately 9% of
firing pattern of the dentate granule cell. For the small en passant boutons were also labile (half-
instance, assume that a granule cell fires at 40 Hz life, approximately 1 day), in contrast to the sta-
in one of its receptive fields but only 10 Hz in the bility of large terminals. Interestingly, in mature
other. Based on the data discussed above, only cultures, the total number of synapses remained
the higher firing rate will induce firing in the CA3 stable in the presence of substantial turnover of
pyramidal cell, the lower activity will be filtered individual terminal structures. Motility of the
out, and the CA3 pyramidal cell will display a filopodial extensions was observed not only in
single receptive field as a result. slice cultures but also in the acute whole-mount
What about long-term changes? One presenta- hippocampal preparation and acute hippocampal
tion of the dentate code may not be sufficient to slices. This actin-based motility can be regulated
induce long-term changes in the CA3 recurrent by brain-derived neurotrophic factor (BDNF),
network. Multiple presentations (e.g., crossing the AMPA and/or kainate receptors, in a cAMP de-
place field several times) or autonomous replay of pendent manner (De Paola et al., 2003; Tashiro
the memory trace in absence of the original con- et al., 2003). These data strongly suggest that tar-
dition, may be necessary to induce long-term plas- get selectivity of the small terminal types of the
ticity. Replay of a given firing pattern may occur mossy fiber is based on the dynamic morpholog-
during different EEG states, as has been shown for ical properties of these axonal elements. In addi-
theta and sharp wave activity (Nadasdy et al., tion, they suggest that even in mature animals
1999), or during the subsequent sleep episodes activity change, traumatic injury or cell loss may
(Wilson and Mcnaughton, 1993). induce rapid remodeling of the mossy fiber-to-
Tetanic stimulation of mossy fibers induces LTP interneuron connections from granule cells. In-
in pyramidal neurons, but is either without effect, deed, ischemic damage, which results in the loss of
or it induces depression at synapses onto inter- hilar and stratum lucidum interneurons, dramat-
neurons (Maccaferri et al., 1998). Since the mossy ically reduces the number of filopodia (Arabadzisz
fiber–LTP onto pyramidal cells critically depends and Freund, 1999).
on cAMP, the effect can be explained by the In summary, the mossy fiber pathway maxim-
absence of the adenylyl cyclase-cAMP cascade ally utilizes the opportunity to differentially regu-
from the filopodial and ‘‘en passant’’ small termi- late its postsynaptic partners. Selective innervation
nals. As a result of this differential LTP, the crit- of pyramidal cells and interneurons by distinct
ical frequency at which inhibition switches to terminal types creates a favorable condition to
591

differentially regulate short-term and long-term granule cells showed position-selective or position-
plasticity and the motility of various mossy termi- independent reward site responses, whereas an-
nal types. The ‘‘bottom-line’’ of this highly other population showed pure place responses
dynamic system appears to be the fine-tuning of (Tabuchi et al., 2003).
the balance between the excitation evoked by the When rats are allowed to explore a novel envi-
large ‘‘detonator’’ terminals and the feedforward ronment for the first time, both dentate granule
inhibition activated by the small terminals. This cells and CA1 pyramidal cells acquire distinct
will determine exactly which dentate firing patterns spatial preferences within the first few minutes
will induce permanent modification of the auto- (Nitz and McNaughton, 2004). However, concur-
associational network of the CA3 region. The final rently recorded interneurons showed very different
test for all predictions regarding dentate function behavior in the two regions: while most CA1
is the examination of firing activity in the freely interneurons transiently decreased their activity
moving animal. while the animal was exploring a novel environ-
ment, the majority of interneurons in the DG sig-
nificantly increased their activity during the same
Neuronal activity patterns in vivo and the processing epoch. These data are congruent with the role of
of entorhinal input by the dentate gyrus strong feedback inhibition in establishing sparse
and decorrelated output in the DG.
Compared to the vast amount of data available on In order to understand the nature of the com-
spatial (and non-spatial) representations in area putations performed by the DG, it is essential to
CA1 of the hippocampus, relatively little is known have an accurate description of the cortical input it
about the in vivo firing properties of neurons in the receives. The last few years have witnessed a major
DG under various behavioral conditions. During revolution in our understanding of spatial repre-
exploration, granule cells display spatially selective sentations in entorhinal cortex. Until a few years
activity, and their firing behavior is, at least under ago, existing data on EC representations indicated
most conditions studied so far, qualitatively quite that, although neurons in EC were spatially selec-
similar to that of pyramidal cells (place cells) in tive, their place fields were much larger and less
areas CA1 and CA3. In particular, in the radial clearly defined than those in hippocampal areas
arm maze, granule cells have directional place (Quirk et al., 1992; Frank et al., 2000). Indeed, this
fields, although the average number of subfields is was perhaps the most direct experimental evidence
somewhat higher and the average size of these for the assumption that sparse, orthogonal, ‘‘hip-
subfields is slightly smaller than in CA3 place cells pocampal-type’’ representations are created first in
(Jung and McNaughton, 1993). The firing activity the DG. However, when spatial firing patterns
of dentate granule cells is modulated by local field were measured in the part of medial entorhinal
potential oscillations in the theta frequency range, cortex (mEC) which projects to the dorsal HC
and the timing of individual action potentials (where place fields are normally recorded), much
changes during traversals of the place field similar smaller place fields, similar in size to correspond-
to phase precession observed in CA1 pyramidal ing hippocampal place fields, were found (Fyhn
cells (Skaggs et al., 1996). When different spatial et al., 2004), while areas of mEC with large place
cues were put in conflict by manipulating the en- fields projected to more ventral parts of the HC,
vironment, sudden coherent transitions (known as which itself was found to have large place fields
reference frame shifts) could be observed in the (Maurer et al., 2005). From these data, it appeared
activity of the set of simultaneously recorded that there was in fact no major transformation of
granule cells, also analogous to the behavior of spatial representations from EC to the DG (and
the CA1 cell population under these conditions the rest of the hippocampus), although subtler
(Gothard et al., 2001). In an experiment where an differences (especially in response to environmen-
explicit attempt was made to identify non-spatial tal manipulations) could not be ruled out. How-
as well as spatial responses, a subpopulation of ever, it soon emerged that if spatial firing patterns
592

are recorded over a larger spatial scale, entorhinal information content, but instead, carry more
and hippocampal representations are again fun- information about relevant objects (Hargreaves
damentally different. In particular, EC place fields et al., 2005). Such object-based information is
were found to repeat periodically at the vertices of probably sufficient to disambiguate different envi-
a regular hexagonal lattice, and different EC ‘‘grid ronments, and create distinct codes in the HC.
cells’’ differ in the center location (phase), orien- Hippocampal pattern separation between envi-
tation, and scale of the grid (Hafting et al., 2005). ronments of varying degrees of similarity has re-
In contrast, hippocampal place cells have at most a cently been investigated in a series of experiments
few discrete place fields even in these larger envi- (Leutgeb et al., 2004, 2005a,b), and the initial
ronments, and their fields typically do not have analysis in areas CA3 and CA1 has now been par-
any special geometrical relationship. Thus, it cur- tially extended to mEC and the DG (Leutgeb
rently appears that hippocampal spatial represen- et al., 2007; Hafting et al., 2006). By analyzing the
tations are in fact different in nature, and in spatial firing fields of neurons within environments
particular, much sparser over a large environment of varying degrees of similarity (manipulations in-
than representations in the (medial) entorhinal cluded switching between different recording loca-
cortex, and a transformation (in fact, a pattern tions, as well as changes in the shape and/or the
separation operation) by the DG is still required. color of the enclosure), Moser and colleagues
Indeed, if we consider two locations in the envi- found that firing patterns of neurons in all hippo-
ronment which are separated by the grid period campal areas distinguished between different en-
(assumed to be relatively invariable) in the area of vironments much better than those of grid cells in
entorhinal cortex which projects to a given part of mEC. Essentially no pattern separation was
the HC, the activity of the grid cell population will detected in EC, as population firing patterns in
be quite similar, while the hippocampal activity enclosures of different shapes, colors, or even in
pattern will be distinct, reflecting the outcome of different rooms were not significantly different
some kind of pattern separation process. In fact, (any observed changes occurred coherently in all
the need for pattern separation becomes even more recorded neurons; Hafting et al., 2006). However,
obvious if we consider two environments. The they also found that the basic properties of pattern
root of the problem is the observation that the separation were different between different sub-
relative grid parameters (phase and orientation) of fields of the hippocampus. In area CA3, the phe-
different grid cells appear to be fixed in all envi- nomenon termed ‘‘rate remapping’’ was observed
ronments, so, in principle, there must be corre- when the shape of the enclosure was varied con-
sponding locations (and orientations) in two tinuously: the firing rate pattern of the active cell
distinct environments where the activity of the en- population changed continuously as the environ-
tire grid cell population is identical. This raises an ment was gradually transformed, while place field
obvious question: how distinct spatial representa- locations remained constant (Leutgeb et al.,
tions of the two environments could be formed in 2005a). In the end, environments of clearly dis-
the HC based on a single entorhinal grid cell rep- tinct shapes (e.g., circle vs. square) activated CA3
resentation. One possible answer to this question is pyramidal cell populations with relatively little
based on the fact that the entorhinal grids are not overlap — indeed, when recordings were made in
perfectly regular; for instance, peaks in the grid two different rooms, the two populations of active
vary in amplitude, so that, in principle, there could CA3 neurons appeared to be chosen independ-
be sufficient spatial information present in the ently, a phenomenon referred to as ‘‘global
amplitudes to distinguish different environments. remapping’’ (Leutgeb et al., 2005b).
Another, perhaps more plausible explanation is Interestingly, preliminary data from recent ex-
that the hippocampus (and, in particular, the DG) periments have revealed a radically different type
receives, in addition to input from medial EC, of pattern separation in the DG. First, unlike in
where grid cells are located, input from lateral EC, CA3, the same dentate cells were found to be ac-
where neuronal firing patterns have a lower spatial tive in different environments (Leutgeb et al.,
593

2007). Second, DG neurons typically had multiple formation of hippocampal representations is also
place fields even in a single environment, consist- indicated by recordings of neuronal activity pat-
ent with earlier data (Jung and McNaughton, terns following selective lesions. In particular, fol-
1993). Finally, the peak firing rate within any one lowing colchicine lesions of dentate granule cells as
place field of a single cell varied with even small described above, place cell representations could
changes in the shape of the enclosure, independ- still be observed in area CA1 (McNaughton et al.,
ently of rate changes in other fields of the same cell 1989). Similarly, the spatial representation in area
(Leutgeb et al., 2007). These results show that the CA1 was largely intact after surgical separation
DG performs a pattern separation operation on its from all other hippocampal areas, which left it
entorhinal inputs, but pattern separation appears with the direct projection from entorhinal cortex
to work differently from what was previously as- as its sole cortical input (Brun et al., 2002). There-
sumed. The observation that the proportion of si- fore, hippocampal areas other than the DG must
multaneously active cells is much lower in the DG be capable of creating sparse, distributed, place-
than in EC (a fact that was confirmed by the data field-like representations on their own under some
described above, and independently by another circumstances, probably based on their direct
recent study, which measured immediate early entorhinal inputs. However, it has to be kept in
gene expression in the DG following spatial expe- mind that despite the presence of a proper place
rience; Chawla et al., 2005) suggested that different cell representation in CA1, navigation memory
environments might activate different sets of neu- was compromised following both types of lesion,
rons in the DG, thereby implementing a particu- suggesting that utilization of the spatial code at
larly efficient type of pattern separation. The the behavioral level requires intact dentate–CA3
results of Moser and colleagues now suggest that interaction.
this might not be the case, and different environ- In summary, these data suggest that the DG
ments (as well as different locations within these does perform a pattern separation of the ent-
environments) might be encoded by different firing orhinal signal, which, however, needs further
rate patterns in essentially the same population of processing to achieve the sparse and decorrelated
active DG neurons. This latter scheme potentially activity pattern observed in the CA3 region.
also allows a fine discrimination of different envi-
ronments (and locations) based on the DG pop-
ulation activity pattern (especially since the firing The dentate gyrus: an evolutionary-developmental
rates of DG neurons appear to be rather sensitive perspective
to changes in environmental features). However,
since the CA3 code appears to be more efficiently Apparently the basic organization of the dent-
orthogonalized than the DG code, an additional ate–CA3 network has the deepest phylogenetic
processing step may be needed. The physiological root among cortical regions. Before the divergence
data reviewed above indicate that the mossy fiber of reptilian-mammalian lineages which apparently
projection may be utilized to arrive at the sparser, preceded the mass extinction of the Permian pe-
more completely pattern-separated representation riod (250 million years ago) neurons in all cor-
recorded in area CA3, but further contributions tical areas were most probably packed in a single
from other sources (temporo-ammonic pathway cellular layer. During the ontogenesis they likely
and local network connections) cannot be ex- followed an outside-in pattern of histogenesis,
cluded. where newly generated neurons settle below the
The different behavior of spatial representations older ones, like in extant reptiles (Goffinet et al.,
at different stages of hippocampal processing 1986). Their main excitatory afferents entered and
clearly argues against the simple view that all terminated in the embryonic marginal zone, above
properties of hippocampal place cells originate in the cortical plate, i.e., in the same zone where the
the DG, and downstream areas simply inherit apical dendrites of their main targets (excitatory
these properties. A more complex view of the principal cells) were present (Ten Donkelaar,
594

1998). As suggested recently, this arrangement some areal segregation, and the original reptilian
does not support the evolution of a cortical struc- dorsomedial cortex became CA3, CA2, CA1, and
ture with multiple cellular layers and distinct areas subiculum (Amaral et al., 1990). Again, the DG
(Super and Uylings, 2001). showed no areal segregation.
As the mammalian nervous system evolved, the But why did the DG–CA3 connection remain
dorsal and lateral cortex of the ancient reptilian essentially unchanged during the course of evolu-
cortex underwent a significant modification that tion, when the rest of the cortical mantle under-
opened up tremendous opportunities for areal and went a dramatic reorganization? Here we would
cellular diversification (Super and Uylings, 2001). like to propose that the reason behind the pro-
This included changing the pattern of histogenesis tracted evolutionary pattern of the hippocampus,
to the inside-out pattern, where newly generated and especially the DG, is the structural constraints
neurons settle above the older ones, and changing of hippocampal function. Apparently, the forma-
the pattern of axonal ingrowth (for a review, see tion of freely accessible multidimensional memory
Super and Uylings, 2001). In the mammalian neo- traces can only be performed in a two-step process
cortex, most of the afferents fibers enter not above that includes a segregation of input followed by an
but below the cortical plate, via the subplate associative step. The reciprocally coupled, multi-
(Allendoerfer and Shatz, 1994; Molnar, 2000). layered cortical structures evolved for a different
This new developmental scheme allowed the de- role (for a more complex interaction with the
velopment of multilayered cortical structures with environment). Apparently, the basic plan of the
highly variable cell types, rich reciprocal interac- two-step information processing through the hip-
tion with the thalamus and the establishment of pocampal formation remained essentially the same
new cortical regions. from lizard to human. Similarly to mammals, in
Little developmental change occurred, however, reptiles, the dentate-equivalent medial cortex
in the medial and dorsomedial cortex, which be- receives the cortical input. This cortical region
came the mammalian allocortex. In the dorsome- lacks an autoassociative network and projects
dial cortex that is homologous with the Cornu the recoded information unidirectionally, to the
Ammonis, the pattern of histogenesis changed to reptilian analogue of the Cornu Ammonis, the
the inside-out pattern, but the pattern of axonal dorsomedial cortex (Lopez-Garcia and Martinez-
ingrowth did not (Super et al., 1998). However, in Guijarro, 1988; Martinez-Guijarro et al., 1991; de
the most ‘‘stubborn’’ structure, the DG, the basic la Iglesia et al., 1994). Association functions in the
reptilian developmental pattern was retained reptile may take place in the next step here in the
(Bayer, 1980) histogenesis follows the outside-in dorsomedial cortex, where an extensive recurrent
pattern and the main excitatory afferents enter collateral system exists, similar to the CA3 region
above the cortical plate. in mammals (Martinez-Guijarro et al., 1984).
Neocortex has evolved rapidly to become a Interestingly, damage to the reptilian homologue
highly complex multilayered structure, with exten- of hippocampus causes similar learning problems
sive intracortical and thalamo-cortical reciprocal as hippocampal lesions do in mammals (Rodriguez
connections (Butler, 1994; Nieuwenhuys, 1994; et al., 2002), suggesting an analogous structural-
Rakic, 1995; Northcutt and Kaas, 1995). Allocor- functional relationship.
tex remained a unilayered structure with a basi- Has anything changed in DG during the mam-
cally unidirectional information flow, DG being malian evolution? In primates, the volumetric ratio
the only cortical structure without thalamic input of the DG and CA3 has changed in favor of the
(Amaral and Witter, 1989). Neocortex has been first. Indeed, DG is ‘‘dentate’’ sensu stricto only in
significantly expanded laterally and was parceled primates, where it includes numerous infoldings.
into numerous functionally segregated areas, but Among these areas the hilus showed the largest
the hippocampus retained the original two major relative increase in volume and cell number
subfields, the DG and Cornu Ammonis. For these (Seress, 1988) underlying the importance of the
two regions, only the Cornu Ammonis showed region, where most of the peculiarities in
595

microcircuits have been noticed. Apparently, the the role of dentate spikes is not explored. It is
feed back regulation of granule cells became tempting to speculate that they may participate in
more elaborated with the increasing complexity memory replay, like the sharp wave in the
of information to be categorized by the system. CA3–CA1 network (Buzsaki, 1989), but definitive
proof is currently unavailable. From the compu-
tational perspective, it is not clear how the lack of
Conclusions, unresolved questions connectivity among granule cells and the relative
paucity of the direct backprojection from the CA3
The DG appears to be an ancient cortical structure to the DG (Li et al., 1994) helps the categorization
from the phylogenetic perspective, yet it displays a function in DG. In conclusion, this peculiar neo-
number of unique features not found in other cor- cortex–archicortex interface will likely keep us
tical regions. Its morphological and physiological busy for a long time.
properties are highly specialized, and these can
explain, at least in part, the role assigned to the
DG and CA3 by computational theories. In par- Acknowledgment
ticular, mossy fibers are characterized by distinct
mechanisms of signal transfer at excitatory vs. This work was supported by the Wellcome Trust
inhibitory targets, and unusually strong activation (A.L. is the recipient of a Wellcome Trust Inter-
of GABAergic circuits. This arrangement allows national Senior Fellowship), the Institut de Cerv-
a delicate balance of excitation and inhibition, eau et de la Moelle épiniere, the Hungarian
which is utilized for code conversion and sparsifi- Scientific Research Fund (OTKA T 049100) and
cation at the entorhinal–dentate connection and the EU Framework 6.
for frequency-dependent spike transfer at the dent-
ate–CA3 connection.
Several issues, however, remain unresolved. References
More comparable data are needed from freely
moving animals to understand the precise compu- Acsady, L., Kamondi, A., Sik, A., Freund, T. and Buzsaki, G.
tation that takes place in the DG and in CA3. (1998) GABAergic cells are the major postsynaptic targets of
Since DG–CA3 transmission appears to be de- mossy fibers in the rat hippocampus. J. Neurosci., 18:
pendent on the frequency of granule cell discharge, 3386–3403.
Acsady, L., Katona, I., Martinez-Guijarro, F.J., Buzsaki, G.
DG firing patterns should be carefully analyzed, and Freund, T.F. (2000) Unusual target selectivity of peri-
and particularly with respect to multiple receptive somatic inhibitory cells in the hilar region of the rat hippo-
fields. The role of two major excitatory inputs of campus. J. Neurosci., 20: 6907–6919.
the granule cells, not discussed here, the mossy cell Alle, H. and Geiger, J.R. (2006) Combined analog and action
potential coding in hippocampal mossy fibers. Science, 311:
input and the supramammillary afferents, is nec-
1290–1293.
essary to clarify DG information processing com- Allendoerfer, K.L. and Shatz, C.J. (1994) The subplate, a tran-
prehensively. Both mossy cells and supramam- sient neocortical structure—its role in the development of
millary afferents contact the proximal dendrites of connections between thalamus and cortex. Annu. Rev.
granule cells, and therefore are likely to exert a Neurosci., 17: 185–218.
powerful influence. Mossy cells of the hilus have Alvarez, P. and Squire, L.R. (1994) Memory consolidation and
the medial temporal lobe: a simple network model. Proc.
highly divergent axons, and thus may link distant Natl. Acad. Sci. U.S.A., 91: 7041–7045.
DG populations involved in coding similar envi- Amaral, D.G. (1979) Synaptic extensions from the mossy fibers
ronmental events, whereas the supramammillary of the fascia dentata. Anat. Embryol. (Berl), 155: 241–251.
input may mediate the modulation of granule cells Amaral, D.G., Ishizuka, N. and Claiborne, B. (1990) Neurons,
numbers and the hippocampal network. Understanding the
by the theta rhythm. The role of rhythmic EEG
Brain Through the Hippocampus, 83.
activities (theta, gamma) in mediating signal trans- Amaral, D.G. and Witter, M.P. (1989) The three-dimensional
fer, code conversion, and the short- and long-term organization of the hippocampal formation: a review of an-
plasticity is unclear at the present time. Similarly, atomical data. Neuroscience, 31: 571–591.
596

Arabadzisz, D. and Freund, T.F. (1999) Changes in excitatory cells in the rat hippocampus [published erratum appears in
and inhibitory circuits of the rat hippocampus 12–14 months Neuroscience 1997 Oct; 80(3): 971]. Neuroscience, 79:
after complete forebrain ischemia [In Process Citation]. 629–648.
Neuroscience, 92: 27–45. Conrad, C.D. and Roy, E.J. (1993) Selective loss of hippocam-
Barnes, C.A., McNaughton, B.L., Mizumori, S.J., Leonard, pal granule cells following adrenalectomy: implications for
B.W. and Lin, L.H. (1990) Comparison of spatial and tem- spatial memory. J. Neurosci., 13: 2582–2590.
poral characteristics of neuronal activity in sequential stages Costa, V.C., Bueno, J.L. and Xavier, G.F. (2005) Dentate
of hippocampal processing. Prog. Brain Res., 83: 287–300. gyrus-selective colchicine lesion and performance in temporal
Baude, A., Nusser, Z., David, J., Roberts, B., Mulvihill, E., and spatial tasks. Behav. Brain Res., 160: 286–303.
McIlhinney, R.A.J. and Somogyi, P. (1993) The metabo- Debanne, D., Gahwiler, B.H. and Thompson, S.M. (1998)
tropic glutamate receptor (mGluR1Ó) is concentrated at Long-term synaptic plasticity between pairs of individual
perisynaptic membrane of neuronal subpopulations as CA3 pyramidal cells in rat hippocampal slice cultures. J.
detected by immunogold reaction. Neuron, 11: 771–787. Physiol., 507(Pt 1): 237–247.
Bayer, S.A. (1980) Development of the hippocampal region in De Paola, V., Arber, S. and Caroni, P. (2003) AMPA receptors
the rat. I. Neurogenesis examined with 3H-thymidine auto- regulate dynamic equilibrium of presynaptic terminals in
radiography. J. Comp. Neurol., 190: 87–114. mature hippocampal networks. Nat. Neurosci., 6: 491–500.
Bragin, A., Jando, G., Nadasdy, Z., van Landeghem, M. and Engel, D. and Jonas, P. (2005) Presynaptic action potential
Buzsaki, G. (1995) Dentate EEG spikes and associated amplification by voltage-gated Na+ channels in hippocam-
interneuronal population bursts in the hippocampal hilar pal mossy fiber boutons. Neuron, 45: 405–417.
region of the rat. J. Neurophysiol., 73: 1691–1705. Frank, L.M., Brown, E.N. and Wilson, M. (2000) Trajectory
Brun, V.H., Otnass, M.K., Molden, S., Steffenach, H.A., encoding in the hippocampus and entorhinal cortex. Neuron,
Witter, M.P., Moser, M.B. and Moser, E.I. (2002) Place cells 27: 169–178.
and place recognition maintained by direct entorhinal- Fyhn, M., Molden, S., Witter, M.P., Moser, E.I. and Moser,
hippocampal circuitry. Science, 296: 2243–2246. M.B. (2004) Spatial representation in the entorhinal cortex.
Buckmaster, P.S. and Schwartzkroin, P.A. (1995) Interneurons Science, 305: 1258–1264.
and inhibition in the dentate gyrus of the rat in vivo. J. Geiger, J.R., Lubke, J., Roth, A., Frotscher, M. and Jonas, P.
Neurosci., 15: 774–789. (1997) Submillisecond AMPA receptor-mediated signaling
Butler, A.B. (1994) The evolution of the dorsal thalamus of at a principal neuron-interneuron synapse. Neuron, 18:
jawed vertebrates, including mammals: cladistic analysis and 1009–1023.
a new hypothesis. Brain Res. Brain Res. Rev., 19: 29–65. Gilbert, P.E., Kesner, R.P. and Lee, I. (2001) Dissociating hip-
Buzsaki, G. (1989) Two-stage model of memory trace forma- pocampal subregions: double dissociation between dentate
tion: a role for ‘noisy’ brain states. Neuroscience, 310: gyrus and CA1. Hippocampus, 11: 626–636.
551–570. Goffinet, A.M., Daumerie, C., Langerwerf, B. and Pieau, C.
Chawla, M.K., Guzowski, J.F., Ramirez-Amaya, V., Lipa, P., (1986) Neurogenesis in reptilian cortical structures: 3H-
Hoffman, K.L., Marriott, L.K., Worley, P.F., McNaughton, thymidine autoradiographic analysis. J. Comp. Neurol., 243:
B.L. and Barnes, C.A. (2005) Sparse, environmentally selec- 106–116.
tive expression of Arc RNA in the upper blade of the rodent Goodrich-Hunsaker, N.J., Hunsaker, M.R. and Kesner, R.P.
fascia dentata by brief spatial experience. Hippocampus, 15: (2005) Dissociating the role of the parietal cortex and dorsal
579–586. hippocampus for spatial information processing. Behav.
Chicurel, M.E. and Harris, K.M. (1992) 3-dimensional analysis Neurosci., 119: 1307–1315.
of the structure and composition of CA3 branched dendritic Gothard, K.M., Hoffman, K.L., Battaglia, F.P. and McN-
spines and their synaptic relationships with mossy fiber aughton, B.L. (2001) Dentate gyrus and ca1 ensemble activity
boutons in the rat hippocampus. J. Comp. Neurol., 325: during spatial reference frame shifts in the presence and
169–182. absence of visual input. J. Neurosci., 21: 7284–7292.
Chrobak, J.J. and Buzsaki, G. (1994) Selective activation of Gulyas, A.I., Megias, M., Emri, Z. and Freund, T.F. (1999) Total
deep layer (V–VI) retrohippocampal cortical neurons during number and ratio of excitatory and inhibitory synapses con-
hippocampal sharp waves in the behaving rat. J. Neurosci., verging onto single interneurons of different types in the CA1
14: 6160–6170. area of the rat hippocampus. J. Neurosci., 19: 10082–10097.
Chrobak, J.J. and Buzsaki, G. (1996) High-frequency oscilla- Gulyas, A.I., Miettinen, R., Jacobowitz, D.M. and Freund,
tions in the output networks of the hippocampal-entorhinal T.F. (1991) Calretinin-immunoreacti cells in the rat hippo-
axis of the freely behaving rat. J. Neurosci., 16: 3056–3066. campus. I. A new type of neuron — specifically associated
Claiborne, B.J., Amaral, D.G. and Cowan, W.M. (1986) A light with the mossy fibre system — revealed. Eur. J. Neurosci.
and electron microscopic analysis of the mossy fibers of the Suppl., 4: 158.
rat dentate gyrus. J. Comp. Neurol., 246: 435–458. Gulyas, A.I., Miles, R., Hájos, N. and Freund, T.F. (1993)
Cobb, S.R., Halasy, K., Vida, I., Nyiri, G., Tamas, G., Buhl, Precision and variability in postsynaptic target selection
E.H. and Somogyi, P. (1997) Synaptic effects of identified of inhibitory cells in the hippocampal CA3 region. Eur. J.
interneurons innervating both interneurons and pyramidal Neurosci., 5: 1729–1751.
597

Hafting, T., Fyhn, M., Molden, S., Moser, M.B. and Moser, memory retrieval, in the Morris navigation task. Neurobiol.
E.I. (2005) Microstructure of a spatial map in the entorhinal Learn. Mem., 73: 243–257.
cortex. Nature, 436: 801–806. Lawrence, J.J., Grinspan, Z.M. and McBain, C.J. (2004) Quan-
Hafting, T., Fyhn, M., Treves, A., Moser, E.I. and Moser, M.B. tal transmission at mossy fibre targets in the CA3 region of
(2006) Coherent reallignment of entorhinal grid cells the rat hippocampus. J. Physiol., 554: 175–193.
coincides global remapping in the hippocampus. FENS, Lee, I. and Kesner, R.P. (2004) Encoding versus retrieval of
Vienna. spatial memory: double dissociation between the dentate
Han, Z.S., Buhl, E.H., Lorinczi, Z. and Somogyi, P. (1993) A gyrus and the perforant path inputs into CA3 in the dorsal
high degree of spatial selectivity in the axonal and dendritic hippocampus. Hippocampus, 14: 66–76.
domains of physiologically identified local-circuit neurons in Leutgeb, J.K., Leutgeb, S., Moser, M.B. and Moser, E.I. (2007)
the dentate gyrus of the rat hippocampus. Eur. J. Neurosci., Pattern separation in the dentate gyrus and CA3 of the hip-
5: 395–410. pocampus. Science, 315: 961–966.
Hargreaves, E.L., Rao, G., Lee, I. and Knierim, J.J. (2005) Leutgeb, J.K., Leutgeb, S., Treves, A., Meyer, R., Barnes, C.A.,
Major dissociation between medial and lateral entorhinal in- McNaughton, B.L., Moser, M.B. and Moser, E.I. (2005a)
put to dorsal hippocampus. Science, 308: 1792–1794. Progressive transformation of hippocampal neuronal repre-
Henze, D.A., Wittner, L. and Buzsaki, G. (2002) Single granule sentations in ‘‘morphed’’ environments. Neuron, 48:
cells reliably discharge targets in the hippocampal CA3 net- 345–358.
work in vivo. Nat. Neurosci., 5: 790–795. Leutgeb, S., Leutgeb, J.K., Barnes, C.A., Moser, E.I., McN-
Houser, C.R. and Esclapez, M. (1994) Localization of mRNAs aughton, B.L. and Moser, M.B. (2005b) Independent codes
encoding two forms of glutamic acid decarboxylase in the rat for spatial and episodic memory in hippocampal neuronal
hippocampal formation. Hippocampus, 4: 530–545. ensembles. Science, 309: 619–623.
de la Iglesia, J.A., Martinez-Guijarro, F.I. and Lopez-Garcia, Leutgeb, S., Leutgeb, J.K., Treves, A., Moser, M.B. and Moser,
C. (1994) Neurons of the medial cortex outer plexiform layer E.I. (2004) Distinct ensemble codes in hippocampal areas
of the lizard Podarcis hispanica: Golgi and immunocyto- CA3 and CA1. Science, 305: 1295–1298.
chemical studies. J. Comp. Neurol., 341: 184–203. Li, X.G., Somogyi, P., Ylinen, A. and Buzsaki, G. (1994) The
Isomura, Y., Sirota, A., Ozen, S., Montgomery, S., Mizuseki, hippocampal ca3 network — an in vivo intracellular labeling
K., Henze, D.A. and Buzsaki, G. (2006) Integration and study. J. Comp. Neurol., 339: 181–208.
segregation of activity in entorhinal-hippocampal subregions Lopez-Garcia, C. and Martinez-Guijarro, F.J. (1988) Neurons
by neocortical slow oscillations. Neuron, 52: 871–882. in the medial cortex give rise to Timm-positive boutons in the
Jeltsch, H., Bertrand, F., Lazarus, C. and Cassel, J.C. (2001) cerebral cortex of lizards. Brain Res., 463: 205–217.
Cognitive performances and locomotor activity following Maccaferri, G., Toth, K. and McBain, C.J. (1998) Target-spe-
dentate granule cell damage in rats: role of lesion extent and cific expression of presynaptic mossy fiber plasticity. Science,
type of memory tested. Neurobiol. Learn. Mem., 76: 81–105. 279: 1368–1370.
Jonas, P., Major, G. and Sakmann, B. (1993) Quantal compo- Marr, D. (1971) Simple memory: a theory for archicortex. Phi-
nents of unitary EPSCs at the mossy fibre synapse on CA3 los. Trans. R. Soc. Lond. B Biol. Sci., 262: 23–81.
pyramidal cells of rat hippocampus. J. Physiol. Paris Lon- Martinez Guijarro, F.J., Berbel, P.J., Molowny, A. and Lopez
don, 472: 615–663. Garcia, C. (1984) Apical dendritic spines and axonic termi-
Jung, M.W. and McNaughton, B.L. (1993) Spatial selectivity of nals in the bipyramidal neurons of the dorsomedial cortex of
unit activity in the hippocampal granular layer. Hippocam- lizards (Lacerta). Anat. Embryol., 170: 321–326.
pus, 3: 165–182. Martinez-Guijarro, F.J., Soriano, E., Del Rio, J.A. and Lopez-
Kali, S. and Dayan, P. (2000) The involvement of recurrent Garcia, C. (1991) Zinc-positive boutons in the cerebral cortex
connections in area CA3 in establishing the properties of of lizards show glutamate immunoreactivity. J. Neurocytol.,
place fields: a model. J. Neurosci., 20: 7463–7477. 20: 834–843.
Kali, S. and Dayan, P. (2004) Off-line replay maintains declar- Maurer, A.P., Vanrhoads, S.R., Sutherland, G.R., Lipa, P. and
ative memories in a model of hippocampal-neocortical inter- McNaughton, B.L. (2005) Self-motion and the origin of
actions. Nat. Neurosci., 7: 286–294. differential spatial scaling along the septo-temporal axis of
O’Keefe, J. and Dostrovsky, J. (1971) The hippocampus as a the hippocampus. Hippocampus, 15: 841–852.
spatial map. Preliminary evidence from unit activity in the McClelland, J.L., McNaughton, B.L. and O’Reilly, R.C. (1995)
freely-moving rat. Brain Res., 34: 171–175. Why there are complementary learning systems in the hip-
O’Keefe, J. and Nadel, L. (1978) The hippocampus as a cog- pocampus and neocortex: insights from the successes and
nitive map. Clarendon, Oxford. failures of connectionist models of learning and memory.
Kobayashi, K. and Poo, M.M. (2004) Spike train timing-de- Psychol. Rev., 102: 419–457.
pendent associative modification of hippocampal CA3 recur- McNaughton, B.L., Barnes, C.A., Meltzer, J. and Sutherland,
rent synapses by mossy fibers. Neuron, 41: 445–454. R.J. (1989) Hippocampal granule cells are necessary for nor-
Lassalle, J.M., Bataille, T. and Halley, H. (2000) Reversible mal spatial learning but not for spatially-selective pyramidal
inactivation of the hippocampal mossy fiber synapses in mice cell discharge. Experimental brain research. Experimentelle
impairs spatial learning, but neither consolidation nor Hirnforschung, 76: 485–496.
598

McNaughton, B.L. and Morris, N.G. (1987) Hippocampal Rolls, E.T. (1989) Functions of neuronal networks in the hip-
synaptic enhancement and information storage within a dis- pocampus and cerebral cortex in memory. In: Cotterill
tributed memory system. TINS, 10: 408–415. R.M.J. (Ed.), Models of Brain Function. Cambridge
Megias, M., Emri, Z., Freund, T.F. and Gulyas, A.I. (2001) University Press, Cambridge, pp. 15–33.
Total number and distribution of inhibitory and excitatory Rolls, E.T. and Kesner, R.P. (2006) A computational theory of
synapses on hippocampal CA1 pyramidal cells. Neurosci- hippocampal function, and empirical tests of the theory.
ence, 102: 527–540. Prog. Neurobiol., 79: 1–48.
Miles, R., Toth, K., Gulyas, A.I., Hájos, N. and Freund, T.F. Rumelhart, D.E. and Zipser, D. (1986) Feature discovery by
(1996) Differences between somatic and dendritic inhibition competitive learning. In: Rumelhart D.E., McClelland J. and
in the hippocampus. Neuron, 16: 815–823. Group. P.R. (Eds.), Parallel Distributed Processing: Explo-
Miles, R. and Wong, R.K. (1986) Excitatory synaptic interac- rations in the Microstructure of Cognition. Vol. 1. Founda-
tions between CA3 neurones in the guinea-pig hippocampus. tions. MIT Press, Cambridge, MA, pp. 151–193.
J. Physiol., 373: 397–418. Salin, P.A., Scanziani, M., Malenka, R.C. and Nicoll, R.A.
Molnar, Z. (2000) Development and evolution of thalamocor- (1996) Distinct short-term plasticity at two excitatory
tical interactions. Eur. J. Morphol., 38: 313–320. synapses in the hippocampus. Proc. Natl. Acad. Sci.
Mori, M., Abegg, M.H., Gahwiler, B.H. and Gerber, U. (2004) U.S.A., 93: 13304–13309.
A frequency-dependent switch from inhibition to excitation Samsonovich, A. and McNaughton, B.L. (1997) Path integra-
in a hippocampal unitary circuit. Nature, 431: 453–456. tion and cognitive mapping in a continuous attractor neural
Muller, R.U., Kubie, J.L. and Ranck Jr., J.B. (1987) Spatial network model. J. Neurosci., 17: 5900–5920.
firing patterns of hippocampal complex-spike cells in a fixed Scharfman, H.E., Kunkel, D.D. and Schwartzkroin, P.A.
environment. J. Neurosci., 7: 1935–1950. (1990) Synaptic Connections of Dentate Granule Cells and
Nadasdy, Z., Hirase, H., Czurko, A., Csicsvari, J. and Buzsaki, Hilar Neurons — Results of Paired Intracellular Recordings
G. (1999) Replay and time compression of recurring spike and Intracellular Horseradish Peroxidase Injections. Neuro-
sequences in the hippocampus [in process citation]. J. Ne- science, 37: 693–707.
urosci., 19: 9497–9507. Scoville, W.B. and Milner, B. (1957) Loss of recent memory
Nieuwenhuys, R. (1994) The neocortex. An overview of its ev- after bilateral hippocampal lesions. J. Neurol. Neurosurg.
olutionary development, structural organization and syna- Psychiatry, 20: 11–21.
ptology. Anat. Embryol., 190: 307–337. Seress, L. (1988) Interspecies comparison of the hippocampal
Nitz, D. and McNaughton, B. (2004) Differential modulation formation shows increased emphasis on the regio superior in
of CA1 and dentate gyrus interneurons during exploration of the Ammon’s horn of the human brain. Journal fur Hirn-
novel environments. J. Neurophysiol., 91: 863–872. forschung, 29: 335–340.
Northcutt, R.G. and Kaas, J.H. (1995) The emergence and ev- Sharp, P.E. (1991) Computer simulation of hippocampal place
olution of mammalian neocortex. Trends Neurosci., 18: cells. Psychobiology, 19: 103–115.
373–379. Sherman, S.M. and Guillery, R.W. (1998) On the actions that
Penttonen, M., Kamondi, A., Sik, A., Acsady, L. and Buzsaki, one nerve cell can have on another: distinguishing ‘‘drivers’’
G. (1997) Feed-forward and feed-back activation of the from ‘‘modulators.’’ Proc. Natl. Acad. Sci. U.S.A., 95:
dentate gyrus in vivo during dentate spikes and sharp wave 7121–7126.
bursts. Hippocampus, 7: 437–450. Sik, A., Penttonen, M. and Buzsaki, G. (1997) Interneurons in
Quirk, G.J., Muller, R.U., Kubie, J.L. and Ranck Jr., J.B. the hippocampal dentate gyrus: an in vivo intracellular study.
(1992) The positional firing properties of medial entorhinal Eur. J. Neurosci., 9: 573–588.
neurons: description and comparison with hippocampal place Sik, A., Penttonen, M., Ylinen, A. and Buzsaki, G. (1995)
cells. J. Neurosci., 12: 1945–1963. Hippocampal CA1 interneurons: an in vivo intracellular
Rakic, P. (1995) A small step for the cell, a giant leap for man- labeling study. J. Neurosci., 15(10): 6651–6665.
kind: a hypothesis of neocortical expansion during evolution. Sik, A., Tamamaki, N. and Freund, T.F. (1993) Complete axon
Trends Neurosci., 18: 383–388. arborization of a single CA3 pyramidal cell in the rat hip-
Ramon y Cajal, S. (1911) Histologie de systeme nerveux de pocampus, and its relationship with postsynaptic parvalbu-
l’homme et des vertebres tomme II. Maloine, Paris. min-containing interneurons. Eur. J. Neurosci., 5:
Reichova, I. and Sherman, S.M. (2004) Somatosensory co- 1719–1728.
rticothalamic projections: distinguishing drivers from modu- Skaggs, W.E., McNaughton, B.L., Wilson, M.A. and Barnes,
lators. J. Neurophysiol., 92: 2185–2197. C.A. (1996) Theta phase precession in hippocampal neuronal
O’Reilly, R.C. and McClelland, J.L. (1994) Hippocampal con- populations and the compression of temporal sequences.
junctive encoding, storage, and recall: avoiding a trade-off. Hippocampus, 6: 149–172.
Hippocampus, 4: 661–682. Squire, L.R. (1992) Memory and the hippocampus: a synthesis
Rodriguez, F., Lopez, J.C., Vargas, J.P., Gomez, Y., Broglio, from findings with rats, monkeys, and humans. Psychol.
C. and Salas, C. (2002) Conservation of spatial memory Rev., 99: 195–231.
function in the pallial forebrain of reptiles and ray-finned Super, H., Martinez, A., Del Rio, J.A. and Soriano, E. (1998)
fishes. J. Neurosci., 22: 2894–2903. Involvement of distinct pioneer neurons in the formation of
599

layer-specific connections in the hippocampus. J. Neurosci., at three types of mossy fiber synapse. J. Neurosci., 20:
18: 4616–4626. 8279–8289.
Super, H. and Uylings, H.B. (2001) The early differentiation of Treves, A. and Rolls, E.T. (1992) Computational constraints
the neocortex: a hypothesis on neocortical evolution. Cere- suggest the need for two distinct input systems to the hip-
bral cortex New York, N.Y., 11: 1101–1109. pocampal CA3 network. Hippocampus, 2: 189–199.
Sutherland, R.J., Whishaw, I.Q. and Kolb, B. (1983) A behav- Treves, A. and Rolls, E.T. (1994) Computational analysis of the
ioural analysis of spatial localization following electrolytic, role of the hippocampus in memory. Hippocampus, 4:
kainate- or colchicine-induced damage to the hippocampal 374–391.
formation in the rat. Behav. Brain Res., 7: 133–153. Vida, I. and Frotscher, M. (2000) A hippocampal interneuron
Tabuchi, E., Mulder, A.B. and Wiener, S.I. (2003) Reward associated with the mossy fiber system. Proc. Natl. Acad. Sci.
value invariant place responses and reward site associated U.S.A., 97: 1275–1280.
activity in hippocampal neurons of behaving rats. Hippo- Willshaw, D.J. and Buckingham, J.T. (1990) An assessment of
campus, 13: 117–132. Marr theory of the hippocampus as a temporary memory
Tamas, G., Somogyi, P. and Buhl, E.H. (1998) Differentially store. Phil. Trans. R. Soc. London, 329: 205–215.
interconnected networks of GABAergic interneurons in the Wilson, M.A. and Mcnaughton, B.L. (1993) Dynamics of the
visual cortex of the cat. J. Neurosci., 18: 4255–4270. hippocampal ensemble code for space. Science, 261:
Tashiro, A., Dunaevsky, A., Blazeski, R., Mason, C.A. and 1055–1058.
Yuste, R. (2003) Bidirectional regulation of hippocampal Wittner, L., Henze, D.A., Zaborszky, L. and Buzsaki, G. (2006)
mossy fiber filopodial motility by kainate receptors: a two- Hippocampal CA3 pyramidal cells selectively innervate
step model of synaptogenesis. Neuron, 38: 773–784. aspiny interneurons. Eur. J. Neurosci., 24: 1286–1298.
Ten Donkelaar, H. (1998) Reptiles. In: Nieuwenhuys R., Ten Xavier, G.F., Oliveira-Filho, F.J. and Santos, A.M. (1999)
Donkelaar H.J. and Nicholson C. (Eds.), The Central Nerv- Dentate gyrus-selective colchicine lesion and disruption of
ous System of Vertebrates. Springer, Berlin. performance in spatial tasks: difficulties in ‘‘place strategy’’
Toth, K., Suares, G., Lawrence, J.J., Philips-Tansey, E. and because of a lack of flexibility in the use of environmental
McBain, C.J. (2000) Differential mechanisms of transmission cues? Hippocampus, 9, 668–681.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 32

The dentate gyrus as a filter or gate: a look back and


a look ahead

David Hsu

Department of Neurology, University of Wisconsin, 600 Highland Avenue, H6/526, Madison, WI 53792, USA

Abstract: The idea of the dentate gyrus as a gate or filter at the entrance to the hippocampus, blocking or
filtering incoming excitation from the entorhinal cortex, has been an intriguing one. Here we review the
historical development of the idea, and discuss whether it may be possible to be more specific in defining
this gate. We propose that dentate function can be understood within a context of Hebbian association and
competition: hilar mossy cells help the dentate granule cells to recognize incoming entorhinal patterns of
activity (Hebbian association), after which patterns that are consistently and repetitively presented to the
dentate gyrus are passed through, while random, more transient patterns are blocked (non-associative
Hebbian competition). Translamellar inhibition as well as translamellar potentiation can be understood in
this context. The dentate-hilar complex thus plays the role of a ‘‘pattern excluder’’, not a pattern completer.
The unique role of pattern exclusion may explain the peculiar qualities of dentate granule cells and hilar
mossy cells.

Keywords: dentate gate; dentate filter; pattern excluder

Introduction 1992; Staley et al., 1992; Williamson et al., 1993),


and strong GABA receptor-mediated inhibition
The dentate gyrus, sitting between the entorhinal (Mody et al., 1992; Coulter, 1999; Nusser and
cortex and area CA3, is both anatomically well Mody, 2002; Stell and Mody, 2002; Cohen et al.,
positioned and physiologically predisposed to play 2003; Mody, 2005).
the role of a gate, blocking or filtering excitatory
activity from the entorhinal cortex and controlling
the amount of excitation that gets through to the History of the idea
hippocampus. Normal adult granule cells rarely
generate action potentials. In part this is because Data that the dentate gyrus may serve as a gate
there is little direct interconnectivity between appeared in at least as early as 1966 in work by
dentate granule cells under normal conditions Andersen et al. (1966). In these experiments, the
(reviewed in Chapter 1 of this volume). In addi- hippocampal formation of adult rabbits was ex-
tion, granule cells have a high resting membrane posed by removal of the overlying neocortex and
potential (Fricke and Prince, 1984; Scharfman, corpus callosum. Stimulating electrodes were
placed into entorhinal cortex or in the perforant
Corresponding author. Tel.: +1 608 263 8551; pathway. Extracellular as well as intracellular re-
Fax: +1 608 263 0412; E-mail: hsu@neurology.wisc.edu cordings were made from electrodes placed in the

DOI: 10.1016/S0079-6123(07)63032-5 601


602

transverse plane of the hippocampus, piercing perforant pathway in behaving rats, and compared
CA1 and both blades of the dentate gyrus. Perfo- dentate granular layer population spike ampli-
rant pathway stimulation resulted in a negative tudes and molecular layer EPSPs in slow wave and
wave reflecting granule cell excitatory postsynaptic rapid eye movement (REM) sleep vs. the alert and
potentials (EPSPs) generated in the middle third of still state. The authors found that slow wave and
the dentate gyrus molecular layer. The EPSP was REM sleep states are associated with larger pop-
followed by a large and slow inhibitory postsy- ulation spikes but smaller EPSPs compared to the
naptic potential (IPSP), persisting for 100–150 ms. still, alert state. To explain these findings, the au-
That EPSPs were followed by IPSPs persisting thors hypothesized that there was relative hyper-
for 100–150 ms suggested that perforant pathway polarization of the dentate granule cell membrane
stimulation at frequencies higher than 10 Hz potential during the still, alert state compared to
should result in a decremental granule cell popu- slow-wave sleep. The relative hyperpolarization
lation spike response (habituation). For pulse was interpreted as a gating mechanism.
stimulus durations of less than 1 s, this decrement- Collins et al. (1983) studied the role of the dent-
al response was indeed observed. However, for ate gyrus in seizure propagation in behaving rats.
longer pulse durations of 6 and 9 s, there was a Stimulation with either focal chemoconvulsant in-
gradual incremental granule cell response (facili- jection into or electrical stimulation of entorhinal
tation), beginning after a few seconds of repetitive cortex produced a graded response. If focal
stimulation. The degree of facilitation increased chemoconvulsant injection induced fewer than 10
with increasing stimulation frequency up to a spikes per minute in entorhinal cortex, no or min-
maximum of 10 Hz, and was abolished by a stim- imal behavioral changes were noted, and no met-
ulation frequency higher than 20 Hz (Andersen et abolic changes were noted on later sectioning and
al., 1966). Although not stated explicitly by the deoxyglucose autoradiography. If 10–30 spikes per
authors, these experiments established one way minute are induced in entorhinal cortex, then there
that the dentate gyrus appears to act as a gate. is increased deoxyglucose uptake in the entorhinal
Short-duration stimuli carried at a certain fre- cortex and in the dentate gyrus, but restricted in
quency are blocked, but longer duration stimuli dentate gyrus to the molecular layer. Behaviors
carried at the same frequency are facilitated. Later associated with weak seizure activity were ob-
studies demonstrated additional ways the dentate served. However, if greater than 40 spikes per
gyrus acts as a gate. minute were induced, moderate seizures devel-
In 1976, Alger and Teyler applied repetitive per- oped. Metabolic changes then spread to the entire
forant pathway stimulation to rat hippocampal dentate gyrus, areas CA3 and CA1, the septal nu-
slices at 1 Hz for a total duration of 10 s (Alger and cleus, and occasionally to the amygdala, nucleus
Teyler, 1976; Teyler and Alger, 1976). They found accumbens, and ventral pallidum-lateral preoptic
an incremental EPSP and population spike re- area. Spread to the contralateral side also oc-
sponse with each succeeding stimulus (facilitation) curred. Because metabolic changes were initially
in CA3 and CA1 but a decremental response restricted to the molecular layer of the dentate
(habituation) in the dentate gyrus. In contrast, gyrus, and only with further increases of chemical
stimulation of the perforant pathway at 15 Hz for or electrical stimulation were these changes able to
15 s produced potentiation of EPSP and population spread to other parts of the hippocampus and to
spike responses of dentate gyrus, CA3 and CA1. extrahippocampal structures, the authors con-
These results are similar to those of Andersen et al. cluded that ‘‘the sequential changes in [deoxyglu-
(1966), showing that the dentate gyrus suppresses cose] metabolism suggest that the dentate gyrus
flow of excitation from perforant pathway input but acts as a restrictive gateway for seizure spread
that this suppression can be reversed with longer from entorhinal cortex to the rest of the limbic
duration stimuli in an appropriate frequency range. systemy.’’ (Collins et al., 1983).
Winson and Abzug (1977, 1978) placed stimu- Further work by Lothman and coworkers was
lating electrodes into the angular bundle of the reviewed in 1992 (Lothman et al., 1992). Working
603

in vivo with unanesthetized rats, Stringer et al. point is breached, actually acts as a ‘‘promoter’’ or
(1989) and Stringer and Lothman (1992) devel- ‘‘amplifier’’ of seizural discharges.
oped the concept of maximal dentate activation Walther et al. (1986) studied rat brain slices in
(MDA). Their experiments involved stimulation of superfusate containing a low concentration of
the perforant pathway or area CA3 until a max- magnesium, which was shown to induce repetitive
imal, apparently saturating level of activity was burst discharges in their slices. The entorhinal cor-
recorded in the dentate gyrus. This state was tex demonstrated prolonged epilepiform dis-
defined as MDA. MDA was most easily obtained charges, lasting minutes at a time. The subiculum
with a stimulating frequency between 10–40 Hz. was also capable of spontaneous discharges, last-
The onset of MDA occurred with a pronounced ing up to 9 s. In contrast, isolated minislices of area
negative shift of the DC potential, reflecting a de- CA3 were only capable of brief spontaneous tran-
polarization of the granule cell layer. In addition, sients, and the dentate gyrus demonstrated no ac-
there was an abrupt rise in extracellular K+ con- tivity at all when connections to the entorhinal
centration, and the appearance of bursts of large cortex were disrupted. Because even prolonged
amplitude population spikes. If stimulation was epileptiform discharges in the entorhinal cortex
triggered above the threshold for MDA, afterdis- elicited only brief transient activity in the dentate
charges were observed, which could persist for a gyrus, the authors suggested that ‘‘the dentate
short time after stimulation ceased. MDA in the gyrus y may serve as a filter which reduces the
intact rat was always bilateral and associated with excitatory load into CA3 and hence into CA1’’
synchronous epileptiform discharges in bilateral (Walther et al., 1986). Further details of the phar-
CA3, CA1, subiculum, and entorhinal cortex. macology and electrophysiology of these slices
Lesioning the entorhinal cortex on one side can were reviewed in Heinemann et al. (1992).
block MDA on that side, but the contralateral A striking visual demonstration of dentate
side retained the ability to reach MDA. If stimu- gating was presented by Iijima et al. (1996) using
lation were applied to the perforant pathway, optical imaging with fluorescence voltage-sensitive
MDA occurred in the dentate gyrus before activity dye in rat brain slices. After superfusing their slices
was recorded in CA1. Thus, it appeared that with an antagonist of GABA-A receptors, electri-
transmission flowed from entorhinal cortex to cal stimulation in the superficial layers of the
dentate gyrus to CA3 and CA1, and not by the entorhinal cortex led to signals reflecting robust
direct entorhinal to CA1 (temporoammonic) excitation of the entorhinal cortex. This first
pathway. spread throughout the superficial layers of the
Lothman et al. (1992) and Stringer and Loth- entorhinal cortex, then involved the deep layers.
man (1992) interpreted their data in the following At 33.6 ms after stimulation, excitation invaded
way: ‘‘(1) MDA serves to initiate and sustain re- the hippocampus, lasting until 151.2 ms after stim-
verberatory seizure activity in [the] hippocam- ulation. The entorhinal cortex remained active
pal–parahippocampal loop; (2) this reverberatory through this time, and activity reverberated within
seizure activity bombards extrahippocampal struc- the entorhinal cortex for the next 200 ms. In this
tures; (3) MDA can directly (within the hippo- period, the hippocampus showed only weak and
campal–parahippocampal loop) and indirectly partial activation. A second stimulus, delivered
(by influencing sites outside the hippocam- 352.8 ms after the first, caused further reverbera-
pal–parahippocampal loop) modulate the length tory activity in the entorhinal cortex, which then
of electrographic seizures and, in turn, their prop- again penetrated to the hippocampus and led to
agation, thereby affecting the expression of vari- hippocampal excitation lasting about 70 ms. A
ous types of behavioral seizures; (4) MDA can be second experiment was also performed in normal
accessed from any point within the hippocam- solution (without the GABA-A receptor antago-
pal–parahippocampal loop; (5) MDA can also be nist), using 1 Hz repetitive stimulation instead
accessed from points outside this loop....’’. Thus of single stimuli. Each stimulation resulted in
the dentate gyrus, when its function as a control increased activity in the entorhinal cortex, but it
604

was not until the seventh stimulus that activity where it has been studied. At 15 min after a
penetrated to the hippocampus. These results 1–2 min spontaneous behavioral seizure in epilep-
demonstrated that entorhinal cortex activation, tic mice, c-fos labeling appeared in dentate granule
even when robust, does not easily penetrate into cells, spread throughout the entire extent of the
the hippocampus, presumably due to gating at the dentate gyrus but not involving the interneurons of
level of the entrance to the hippocampus, i.e., at the dentate-hilar border or the dentate molecular
the dentate gyrus. layer. At 30 min, c-fos staining was intense in the
Behr et al. (1998) used kindled animals to dem- dentate gyrus, involving both granule cells and
onstrate the dentate gate and its ‘‘breakdown’’. dentate-hilar border interneurons, and increased
The authors first superfused entorhinal-hippocam- c-fos staining also spread to the rest of the hippo-
pal brain slices in low-magnesium solution. Spon- campus. At 1–2 h, c-fos staining began to fade in
taneous seizure-like activity in both entorhinal the dentate granule cells but was intense in inter-
cortex and in area CA3 was then recorded, but neurons of the dentate-hilar border and the dent-
activity was significantly larger in amplitude and ate molecular layer. At 4 h, c-fos staining was
longer in duration in area CA3 of kindled slices lighter throughout the hippocampus, including the
compared to control slices. Transecting the perfo- dentate gyrus, compared to controls. These data
rant pathway greatly diminished epileptiform ac- suggested that dentate granule cell activation was
tivity in area CA3, but transecting the subiculum- likely to have been an early event in spontaneous
to-entorhinal pathway did not. These results dem- seizures. The authors commented that the rate of
onstrated that epileptiform activity can be passed c-fos expression varies between cell types, so that
from entorhinal cortex through the dentate gyrus c-fos expression occurred first in dentate granule
into area CA3, and that passage is made more cells and later in dentate interneurons is suggestive,
likely after kindling. The authors then devised an but not proof that activity in granule cells pre-
experiment so that only the entorhinal cortex was ceded that in dentate interneurons.
locally perfused with both a GABA-A receptor The breakdown of dentate gating and its pre-
antagonist and elevated K+ to induce epileptiform sumed relationship to epileptogenesis motivated
activity in a spatially-specific manner. Electrical much of the research described above, at least as
stimulation of the entorhinal cortex resulted in a early as the work by Collins et al. (1983). It was
much stronger response in the dentate gyrus of recognized that the dentate gyrus is normally re-
kindled slices. Furthermore, in a separate experi- sistant to the propagation of discharges from the
ment, spontaneous entorhinal interictal activity entorhinal cortex. In the setting of limbic epilepsy
failed to trigger epileptiform discharges in dentate (i.e., temporal lobe epilepsy), the gate is thought to
gyrus in 8 out of 8 control slices, but did trigger be compromised, so that seizure activity from ent-
them in 7 out of 9 kindled slices. Taken together, orhinal cortex is allowed into the hippocampus,
these results show that, in normal brain, the dent- and propagated in a reverberatory cycle back to
ate gyrus appears to prevent epileptiform activity entorhinal cortex again (Stringer and Lothman,
in the entorhinal cortex from reaching the hippo- 1992). This suggestion has led to many studies,
campus, but after epileptogenesis (exemplified by which have focused on reasons why the dentate
kindling), the dentate gyrus no longer functions as gyrus gate may ‘‘breakdown’’ in temporal lobe
a gate. epilepsy. Based primarily on animal models of
Evidence for dentate gating of seizures was also temporal lobe epilepsy, the results have suggested
found by monitoring the level of c-fos protein ex- that the dentate gyrus gate may breakdown be-
pression after the development of spontaneous cause of a change in the balance of excitation
seizures in a pilocarpine model of epilepsy in mice and inhibition of dentate gyrus granule cells. De-
(Peng and Houser, 2005). The expression of c-fos creased inhibition of granule cells could develop
protein is a marker for neuronal activity. Increased because of seizure-induced loss of GABAergic
c-fos protein levels are evident 20–40 min after neurons or altered expression of GABAA recep-
c-fos activation, at least in many neuronal types tors, among many other reasons (Mody et al.,
605

1992; Mody, 2005). Increased excitation may de- Alger, 1985; Deisz and Prince, 1989; Thompson
velop because of mossy fiber sprouting, as well as and Gahwiler, 1989a, b, c; Scanziani et al., 1991;
other factors (Stringer and Lothman, 1992; Sutula Mott and Lewis, 1992; Thomson et al., 1993). For
et al., 1992; Jackson and Scharfman, 1996; Scharf- instance, in area CA3, repeated stimulation of py-
man, 2004). A more complex interplay between ramidal neurons leads to decreased IPSCs via two
initial hyperexcitability followed by chronic hyper- mechanisms: (a) prolonged activation of GABA-A
inhibition has also been suggested (Sloviter et al., receptors, which leads to a chloride influx into the
2006). principal neuron, which leads to a decrease in
driving force for chloride-mediated GABAergic
inhibition, and (b) presynaptic negative feedback
The idea of dentate gate vs. filter of GABA onto GABA-B receptors, which leads to
decreased presynaptic GABA release (Thompson
In summary, there is now a series of studies, which and Gahwiler, 1989a, b, c). Further details on
suggest that activity in the entorhinal cortex is of- complex GABAergic responses have been studied
ten halted, delayed, or diminished at the dentate and reviewed (Kaila, 1994; Kaila et al., 1997;
gyrus. The decreased excitability appears to be re- Staley, 2004).
lated at least in part to the strong, prolonged Given that the dentate gyrus can function as a
dentate granule cell IPSP first described by And- gate, is there also a way in which the dentate gyrus
ersen et al. (1966). Further, as also found in that might act as a filter? The prolonged IPSP in effect
study, repetitive perforant pathway stimulation for acts as a high-frequency filter, but is there a more
a prolonged period of time (a few seconds or specific way in which the dentate gyrus might act
longer) results in facilitation of succeeding stimu- as a filter of information? We would like to suggest
lations. Thus if the dentate gyrus is a gate, it is a that the dentate gyrus does indeed filter informa-
gate that can be opened if one is persistent, i.e., if tion in a specific way. We prepare for discussion of
one keeps ‘‘knocking’’ on it. dentate filtering function with the following com-
Why does the dentate gate open with repeated ments on hippocampal anatomy, mossy cell func-
stimulation? Meticulous simultaneous intracellular tion, the relation of translamellar inhibition and
recordings seem to show that dentate and hilar potentiation to associative Hebbian learning, and
interneurons respond strongly and faithfully to the role of synaptic scaling in non-associative
dentate granule cell discharges (Scharfman et al., Hebbian competition.
1990). However, with repeated granule cell dis- A detailed review of hippocampal anatomy ap-
charge, the interneuronal response switch from pears in other chapters of this volume. We note
action potentials to EPSPs — the interneurons still here only that the hippocampus has a striking
hear the command to fire but stop firing. Con- lamellar structure, with the projections of the
versely, dentate and hilar interneuronal discharges perforant pathway, the mossy fibers, the Schaffer
produce IPSPs in dentate granule cells, but with collaterals and the alvear pathway, all appearing
many failures. Interestingly, failures are more to be on nearly a plane (‘‘lamella’’) perpendicular
likely after a large IPSP. These results suggest to the longitudinal axis of the hippocampus
that at least part of the gating function resides in (Andersen et al., 1969, 1971; Amaral and Witter,
the local granule cell and interneuron circuitry, 1989). A point to be emphasized here for the dis-
and that this inhibitory circuit is tuned down in cussion below is that hilar mossy cells project
efficacy with repeated activation (Scharfman et al., maximally onto granule cells that are septally and
1990). temporally displaced from the mossy cells of
Such activity-dependent disinhibition, involving origin, not onto the granule cells from which the
principal neurons with their local interneuronal mossy cells receive input. That is, mossy cells pro-
circuitry, appears to be an important recurring jections are unique in being preferentially perpen-
theme in other brain areas as well (Ben-Ari et al., dicular to the plane of the lamellae (Amaral and
1980; Wong and Watkins, 1982; McCarren and Witter, 1989).
606

Why do mossy cells project in this way? A de- that are systematically scaled up or potentiated.
tailed discussion of mossy cell function appears in We refer to synaptic scaling as being representative
a separate chapter of this volume. We summarize of a type of non-associative Hebbian competition.
the principal features and suggest its role in dent- Thus we suggest that translamellar potentiation
ate-hilar function as follows: (1) mossy cells can be viewed in terms of Hebbian associative learn-
mediate both translamellar inhibition as well as ing, with the additional twist that double input
potentiation; (2) translamellar potentiation as me- from both the perforant and associative pathways
diated by mossy cells is weak, and an individual results in more effective potentiation. Indeed, in-
mossy cell, by itself, cannot cause a dentate gran- put from only one source, e.g., the association
ule cell to discharge; (3) optimal potentiation of pathway only, may actually result in depotentiat-
dentate granule cells requires near-simultaneous ion through Hebbian competition. To see this,
stimulation of both the perforant and association consider repetitive perforant pathway input that
pathways (which we refer to as ‘‘double input’’), to arrives at dentate granule cells in a certain number
within a time interval on the order of about 5 ms. of lamellae (Fig. 1). The dentate granule cells in
A caveat is that the functional width of a lamella these lamellae fire multiple action potentials. These
at the level of the dentate gyrus may be as thick as granule cells cause mossy cells downstream in the
2.5 mm (Zappone and Sloviter, 2004). same lamellae to fire multiple action potentials as
We also propose that granule cells scale their well. These mossy cells then send signals to many
activity to pass only the most favored or most po- other lamellae. Some of these other lamellae re-
tentiated dentate patterns. That is, individual ceive near-simultaneous perforant and association
granule cells evaluate not only whether individual pathway input, and some do not. For lamellae
synapses are favorable or not (the traditional, as- that do receive near-simultaneous perforant and
sociative Hebbian LTP or LTD), but also whether association pathway input, one expects the mossy
the number of action potentials averaged over
some period of time is too low or too high. If the
average activity of one particular neuron is too EC
low, all synaptic strengths of this neuron are scaled
up; and if too high, all synaptic strengths are
scaled down, so as to preserve the average activity
within some characteristic range (LeMasson et al.,
1993; Turrigiano and Nelson, 2000). Experimental DG
evidence for this kind of activity-dependent
homeostatic synaptic scaling in other brain areas
exists (Royer and Pare, 2003; Wierenga et al.,
2005). Hebbian systems that are not capable of
similar homeostasis of activity evolve inevitably
into a state of tonic hyperactivity or global silence MC
(Miller, 1996, Marder and Prinz, 2002).
As a consequence of activity-dependent synaptic Fig. 1. Dentate-hilar potentiation is mediated by double input
from both entorhinal cortex neurons and from hilar mossy cells.
scaling, the establishment of potentiated input
EC ¼ entorhinal cortex; DG ¼ dentate gyrus granule cells;
patterns causes the response of a neural system to MC ¼ mossy cells of hilus. Repetitive stimulation of dentate
non-potentiated patterns to be scaled down, even granule cells by entorhinal cortex neurons causes transmission
in the absence of specific LTD mechanisms for the of excitation to mossy cells in the hilus. The mossy cells then
non-potentiated patterns. That is, for synapses to stimulate extralamellar granule cells (plus interneurons near
those granule cells). Those granule cells that receive input from
survive in competition with other synapses, it is
both entorhinal cortex and from mossy cells become potenti-
not enough that they not be specifically identified ated, while those that do not, become depotentiated. Potenti-
as being unfavorable; synapses will nonetheless be ated connections are represented by thick arrows.
scaled down in strength if there are other synapses Depotentiated connections are represented by dashed arrows.
607

cell-to-granule cell connection to be potentiated. the role of the mossy cells is to fill in missing
However, Hebbian competition then requires that components of perforant pathway input. For ex-
non-potentiated connections be scaled down in ample, if the dentate-hilar complex learns to rec-
strength. Thus lamellae that do not receive simul- ognize a pattern involving co-activation of granule
taneous perforant and association pathway input cells in lamellae A, B, and C, but later receives
will find their mossy cell input weakened. perforant pathway input only at lamellae A and B,
What is the timescale for translamellar inhibi- then the mossy cells via association pathway po-
tion? A scaling mechanism should take place on tentiation will nonetheless stimulate granule cells
a timescale that is much longer than the baseline in lamella C to fire. This type of pattern recogni-
dentate granule cell firing interval, because a scal- tion is often referred to as ‘‘pattern completion’’.
ing mechanism requires monitoring and averaging It is tuned to be sensitive but not specific. The
the firing rate over some period of time. One set of mossy cells will cause granule cell co-activation in
experiments (Zappone and Sloviter, 2004) found a remembered pattern, if the input pattern is
translamellar inhibition to appear on a timescale ‘‘close enough’’ to the remembered pattern. That
of 200 s. A timescale this long is consistent with mossy cell projections are perpendicular to the
a scaling mechanism, and is not consistent with perforant pathway and project preferentially to
direct connectivity-related effects (i.e., disynaptic distant granule cells, sets up an ideal geometry for
mossy cell to basket cell to granule cell transmis- the mossy cells to play an associative role. It was
sions). not clear to these authors why associations be-
Translamellar potentiation and inhibition can tween distant granule cells should be mediated by
now be put together in a consistent scheme for a separate cell population (the mossy cells), but it
system learning. Translamellar potentiation allows was conjectured that this arrangement allowed for
associative learning, while translamellar inhibi- independent influences to act on granule cells and
tion helps maintain dynamical system stability. mossy cells separately, and that nearest-neighbor
Consistently successful double input from both granule cell co-activations may be discouraged,
perforant and association pathways results in thus preventing a dangerous accretion of co-local-
translamellar potentiation (Steward et al., 1977; ized excitation (Buckmaster and Schwartzkroin,
Buzsaki and Eidelberg, 1982; Strowbridge et al., 1994).
1992; Hetherington et al., 1994; Strowbridge and Alternatively, considering that granule cells are
Schwartzkroin, 1996; Kleschevnikov and Rout- difficult to activate while mossy cells are easily ac-
tenberg, 2003), while multiple inputs from mossy tivated, Jackson and Scharfman (1996) proposed
cells without concomitant perforant pathway in- that mossy cells act as a switch: ‘‘By keeping
put result in translamellar inhibition (Zappone activity in the granule cells either above or below a
and Sloviter, 2004). Translamellar inhibition and threshold for potentiation of synapses on pyram-
potentiation are thus complementary mechanisms, idal cells, mossy cells could create a bistable
both necessary for a stable system capable of con- system, and thus form a gate to control whether
tinual learning. or not information will be stored in the down-
stream elements of the trisynaptic circuit’’.
Other influences on the mossy cells could presum-
Dentate-hilar filtering function: a hypothesis ably determine whether the switch is turned on or
off.
Various ways in which mossy cells can help the We offer yet another explanation of the dentate
dentate gyrus function as a filter have been pro- filter, closer to that of Buckmaster and Schwa-
posed. Buckmaster and Schwartzkroin (1994) have rtzkroin (1994) but differing in an important way.
suggested a granule cell association hypothesis, Mossy cells can help bring granule cells closer
wherein the mossy cells help to link subpopulat- to threshold but rarely trigger granule cell action
ions of granule cells. They suggested that the dent- potentials by themselves (Hetherington et al.,
ate-hilar role is one of pattern recognition, where 1994; Scharfman, 1995; Kleschevnikov and
608

Routtenberg, 2003). Such weak mossy cell associ- What is the optimal frequency for opening the
ation does not lend itself to high-sensitivity pattern dentate gate? Comparing repetitive stimulation at
recognition, as missing components of a perforant 0.5, 4, 7, 10, and 20 Hz, Andersen et al. (1966)
pathway pattern are not likely to be filled in by found that 10 Hz was optimal. Comparing repet-
mossy cell collateral input. The finding that tem- itive stimulation at 0.5, 5, and 100 Hz, Mott and
poral to septal association is absent or very weak Lewis (1992) found that 5 Hz was optimal. The
in rats (Hetherington et al., 1994) would also lead frequencies 5–10 Hz fall in the theta–alpha band,
to very poor pattern completion capabilities, as which is known to be prominent in limbic struc-
essentially there is no pattern completion capabil- tures including the hippocampus (Bland, 1986;
ity in the temporal to septal direction. Further- Freund and Buzsaki, 1996; Buzsaki, 2002; Buzsaki
more, association pathway potentiation, as et al., 2003). We therefore conjecture that theta
discussed in the previous section, is best triggered oscillations indicate activity requiring opening of
with double input from both the perforant and the dentate gate.
association pathways (Fig. 1). Inconsistent pattern We comment that the dentate-hilar complex
input may not simply be ignored, but may lead to combines a sluggish but powerful excitatory
loss of potentiation and possibly even inhibition of source (the dentate granule cells) with a highly la-
granule cell response. Thus, we agree that the bile but weaker component (the mossy cells). The
dentate-hilar complex is a pattern recognition sluggishness of the granule cells and the weakness
complex, but we propose that it represents an of mossy cell output allow the granule cells to
unforgiving pattern recognizer, one that is specific maintain high specificity, but the lability of the
but not necessarily sensitive. We propose that the mossy cells nonetheless allows highly responsive
dentate-hilar complex is not a pattern completer, associative learning. This dual need for specificity
but a pattern excluder. Its job is to exclude input and association may explain why dentate-hilar as-
patterns that are not exactly right. sociation is mediated by a specialized cell popula-
Dentate-hilar function, in our conjecture, thus tion (the mossy cells), while in CA3 and in the
consists of the following steps: (1) patterns are neocortex, association occurs directly between
presented to the dentate-hilar complex via repet- principal cells. The dentate-hilar complex is unique
itive perforant pathway input, (2) mossy cells in being tuned for specificity.
strengthen granule cell responses to patterns that
are repeated in a consistent and persistent way
(translamellar potentiation), and weaken random A look ahead
or erratically presented patterns (translamellar
inhibition), (3) Hebbian competition, through Our conjecture for dentate-hilar function is spec-
non-associative mechanisms, scales granule cell ulative, but testable. The simplest type of exper-
firing thresholds to fire only with the most highly iment would be to monitor all EPSPs from the
potentiated pattern or patterns, (4) with future dentate granular layer in one lamella, and to cal-
repetitions, the most highly potentiated pattern or culate the initial slope of each EPSP, denoted E in
patterns are allowed to pass through, while more units of volts per second. One would also keep
random patterns are blocked. track of which E’s result in population spikes.
In this model, the dentate gyrus functions as Then a distribution function can be constructed,
both gate and filter. It is a gate that can be opened G(E), giving the probability of observing each
by persistent, repetitive stimulation, and it is a fil- value of E (Fig. 2, filled squares). One can also
ter in that it prefers that the stimulation be con- determine the threshold E0, defined as the smallest
sistent, in terms of the pattern of the stimulation as E above which a population spike is likely (e.g.,
distributed along the longitudinal axis. That is, with a likelihood greater than 95%). Alternatively,
one must ‘‘knock’’ on the gate not only many one can define various threshold functions with
times, but in nearly exactly the same way each parameters that can be extracted from experiment.
time. Random knocks are ignored. For instance, one can hypothesize a threshold
609

0.030

0.025

0.020
G(E) and Eo

0.015

0.010

0.005

0.000

0 2 4 6 8 10
E

Fig. 2. Hypothetical distributions of G(E) and E0 with E in arbitrary units. Filled squares: baseline G(E). The single solid square at E
¼ 5.28 represents the threshold E0 such that 95% of the distribution has EoE0. Open triangles: hypothetical G(E) after potentiation.
The threshold is at E ¼ 6.72. Open circles: hypothetical G(E) in chronic epilepsy. The threshold is at E ¼ 6.00.

function of the form higher E, representing potentiated EPSPs, with


threshold E0 between this new peak and the old
PðEÞ ¼ A expðBðE  E 0 ÞÞ,
peak (Fig. 2, open triangles). EPSPs from the new
where P(E) is the probability that a given E results high-E peak are passed by the dentate gyrus, while
in a population spike. Here A is a normalization those from the old peak are blocked. The more
constant, E0 is the firing threshold for E, and B distinct is this new peak, the easier it becomes to
controls how sharp is the transition to firing. The exclude incorrect patterns. Small fluctuations in
parameters B and E0 can be extracted from exper- the value of E0 would not greatly affect specificity.
iment, by plotting log P(E) vs. E. The idea of an What happens to G(E) and E0 in chronic epi-
adjustable firing threshold E0 is key to our model lepsy? With loss of mossy cells, one might hypoth-
for dentate-hilar function. esize that it becomes more difficult to create a
What happens to G(E) and E0 after potentiat- distinct high-E peak. One might hypothesize, for
ion? If one stimulates the perforant pathway re- instance, that only a high-E shoulder is created
petitively at a frequency fS with a certain number (Fig. 2, open circles). The function of the dentate-
of repetitions NR, one expects potentiation of the hilar complex as a pattern excluder would thus be
dentate response to future stimulations. The ideal degraded. The threshold E0, furthermore, would
frequency for potentiation may be 5–10 Hz for have to be placed on a steeper part of the curve for
6–9 s (Andersen et al., 1966; Mott and Lewis, G(E). Any slight fluctuation of E0 would cause
1992), as discussed above. The relevant E to mon- dentate activity either to be overly inhibited
itor may be the median or maximal value over the (E0 too high) or overly excitable (E0 too low).
NR repetitions. After potentiation, one may repeat The effect of recurrent excitatory mossy fiber
the procedure for constructing G(E), and see how collaterals (reviewed in Chapter 29 by Sutula and
this distribution function has changed. One may Dudek in this volume) on G(E) and E0 should
hypothesize that G(E) develops a new peak at also be very interesting. One might expect either a
610

high-E shoulder or a separate higher-E peak, sim- The stimulation strength necessary to produce
ilar to that seen due to potentiation in the normal EPSPs and population spikes is also of interest.
state discussed above. However, unlike the high-E One expects a threshold effect for the stimulation
contribution mediated by mossy cells, the high-E strength S (in units of volts), wherein a minimal
contribution from recurrent mossy fiber collaterals value of this necessary before population spikes
carries no useful information, because the associ- are seen. A distribution function can be defined,
ated EPSPs are the result of local recurrent exci- D(E,S), giving the probability of observing a given
tations. Furthermore, if the high-E contributions E and S. It would be of interest to know what
from recurrent mossy fiber collaterals and from happens to D(E,S), E0 and S0 after potentiation,
mossy cells overlap, then the filtering function of the and in the context of chronic epilepsy.
dentate-hilar complex may be severely degraded. Finally, if it turns out to be true that the dent-
A more ambitious experimental goal would be ate-hilar complex is a pattern excluder, one may
to determine the functional width of a lamella, and then employ similar arguments as developed in this
to develop techniques to stimulate and to record chapter to speculate on the function of the auxil-
from individual lamellae reliably. This goal is iary pathways, e.g., the direct pathways from ent-
likely to be technically challenging. The simplest orhinal cortex to area CA3 (Hjorth-Simonsen and
alternative would be to place a single stimulating Jeune, 1972; Steward and Scoville, 1976; Witter
electrode into the angular bundle, one in each and Amaral, 1991). One possible function for the
hemisphere, and a single recording electrode into auxiliary pathways may be to help validate signal
the dentate granular layer, again one in each hem- passed into the hippocampus. We discuss this
isphere. Presumably, one is guaranteed in this ar- point at a little greater length below.
rangement to have one pair of stimulating and By the activity-dependent homeostasis hypoth-
recording electrodes in each of two distinct, non- esis (LeMasson et al., 1993; Turrigiano and Nel-
overlapping lamellae (one in each hemisphere). son, 2000), all principal neurons have a preferred
However many distinct lamellae are accessible target firing rate, with homeostatic mechanisms to
to experiment, one may then stimulate a subset of return to this rate over some period of time if per-
them simultaneously and repetitively, at a certain turbed away from it. If we assume that this hy-
frequency of repetition for a certain number of pothesis applies to granule cells, then there must
repetitions, NR. This frequency may again be be some rate at which granule cells fire action
taken in the range of 5–10 Hz (Andersen et al., potentials spontaneously, even in the absence of
1966; Mott and Lewis, 1992). The spatial pattern perforant pathway stimulation. This rate is low
of the stimuli represents the pattern to be learned. but cannot be zero. How can CA3 principal neu-
A distribution function G(E) can then be con- rons downstream from granule cells know if the
structed for this system, with one G(E) for each signals they receive from granule cells are due to
lamella. spontaneous granule cell activity or due to perfo-
Of interest would be the number of repetitions, rant pathway stimulated activity? There should
NR, needed to teach a given target pattern, and be some way to ignore the inevitable (if rare)
whether the train of NR repetitions need to be re- spontaneous granule cell discharge, while not com-
peated a certain number of times. After potentiat- promising a faithful response to legitimate, stim-
ion of the target pattern is achieved, a test of ulated granule cell discharge.
sensitivity would be to present, simultaneously, the The answer may be that CA3 principal neurons
target spatial pattern plus a random pattern of have their own G(E) distribution function, and
variable amplitude, and then see if the random they scale their firing thresholds to fire only with
component is blocked while the target pattern is the most highly potentiated inputs. Thus if a set of
allowed through. A test of specificity would be to CA3 principal neurons have been trained to expect
present only a part of the target spatial pattern, ‘‘double input’’ from granule cells and from ent-
and see how close the presented pattern has to be orhinal neurons via the auxiliary entorhinal-to-
to the target pattern to be passed through. CA3 pathway, then these CA3 principal neurons
611

are less likely to fire if they receive input only from Ben-Ari, Y., Krnjevic, K., et al. (1980) Lability of synaptic
granule cells. Thus, auxiliary pathways may play a inhibition of hippocampal pyramidal cells. J. Physiol., 298:
36P–37P [proceedings].
crosschecking role, validating information arriving
via the main pathways. If a confirmatory signal Bland, B.H. (1986) The physiology and pharmacology of hip-
pocampal formation theta rhythms. Prog. Neurobiol., 26(1):
does not arrive via an auxiliary pathway, then in- 1–54.
formation arriving via the main pathway may be Buckmaster, P.S. and Schwartzkroin, P.A. (1994) Hippocampal
ignored. The additional input from the auxiliary mossy cell function: a speculative view. Hippocampus, 4(4):
pathway may be needed to push a principal neuron 393–402.
above the firing threshold. Buzsaki, G. (2002) Theta oscillations in the hippocampus.
Neuron, 33(3): 325–340.
In summary, the current wealth of experimental
Buzsaki, G., Buhl, D.L., et al. (2003) Hippocampal network
data on the dentate gyrus shows that the dentate patterns of activity in the mouse. Neuroscience, 116(1):
gyrus does function as a gate. We further conjec- 201–211.
ture that it also functions as a highly specific pat- Buzsaki, G. and Eidelberg, E. (1982) Direct afferent excitation
tern recognizer, or filter. The data to date do not and long-term potentiation of hippocampal interneurons. J.
Neurophysiol., 48(3): 597–607.
directly address this conjecture. We suggest future
Cohen, A.S., Lin, D.D., et al. (2003) Dentate granule cell
experiments that may help to prove or disprove the GABA(A) receptors in epileptic hippocampus: enhanced
filtering conjecture. Even if the specifics of our synaptic efficacy and altered pharmacology. Eur. J. Neuro-
conjecture are wrong, we hope these experiments sci., 17(8): 1607–1616.
will deepen our insight into the structure and Collins, R.C., Tearse, R.G., et al. (1983) Functional anatomy of
limbic seizures: focal discharges from medial entorhinal cor-
function of the dentate gyrus and hippocampus.
tex in rat. Brain Res., 280(1): 25–40.
Coulter, D.A. (1999) Chronic epileptogenic cellular alterations
in the limbic system after status epilepticus. Epilepsia,
Acknowledgments 40(Suppl 1): S23–S33 discussion S40–S41.
Deisz, R.A. and Prince, D.A. (1989) Frequency-dependent de-
I am grateful to Drs. Helen Scharfman and pression of inhibition in guinea-pig neocortex in vitro by
GABAB receptor feed-back on GABA release. J. Physiol.,
Thomas Sutula for helpful advice and for the op-
412: 513–541.
portunity to write this chapter. I also thank the Freund, T.F. and Buzsaki, G. (1996) Interneurons of the hip-
American Epilepsy Society for support, and the pocampus. Hippocampus, 6(4): 347–470.
National Institutes of Health and National Center Fricke, R.A. and Prince, D.A. (1984) Electrophysiology of
for Research Resources K12 Roadmap, Project dentate gyrus granule cells. J. Neurophysiol., 51(2): 195–209.
Heinemann, U., Beck, H., et al. (1992) The dentate gyrus as a
number 8K12RR023268-02.
regulated gate for the propagation of epileptiform activity.
Epilepsy Res. Suppl., 7: 273–280.
Hetherington, P.A., Austin, K.B., et al. (1994) Ipsilateral asso-
References ciational pathway in the dentate gyrus: an excitatory feed-
back system that supports N-methyl-D-aspartate-dependent
Alger, B.E. and Teyler, T.J. (1976) Long-term and short-term long-term potentiation. Hippocampus, 4(4): 422–438.
plasticity in the CA1, CA3, and dentate regions of the rat Hjorth-Simonsen, A. and Jeune, B. (1972) Origin and termina-
hippocampal slice. Brain Res., 110(3): 463–480. tion of the hippocampal perforant path in the rat studied by
Amaral, D.G. and Witter, M.P. (1989) The three-dimensional silver impregnation. J. Comp. Neurol., 144(2): 215–232.
organization of the hippocampal formation: a review of an- Iijima, T., Witter, M.P., et al. (1996) Entorhinal-hippocampal
atomical data. Neuroscience, 31(3): 571–591. interactions revealed by real-time imaging. Science,
Andersen, P., Bliss, T.V., et al. (1969) Lamellar organization of 272(5265): 1176–1179.
hippocampal excitatory pathways. Acta Physiol. Scand., Jackson, M.B. and Scharfman, H.E. (1996) Positive feedback
76(1): 4A–5A. from hilar mossy cells to granule cells in the dentate gyrus
Andersen, P., Bliss, T.V., et al. (1971) Lamellar organization of revealed by voltage-sensitive dye and microelectrode record-
hippocampal pathways. Exp. Brain Res., 13(2): 222–238. ing. J. Neurophysiol., 76(1): 601–616.
Andersen, P., Holmqvist, B., et al. (1966) Entorhinal activation Kaila, K. (1994) Ionic basis of GABAA receptor channel func-
of dentate granule cells. Acta Physiol. Scand., 66(4): 448–460. tion in the nervous system. Prog. Neurobiol., 42(4): 489–537.
Behr, J., Lyson, K.J., et al. (1998) Enhanced propagation of Kaila, K., Lamsa, K., et al. (1997) Long-lasting GABA-medi-
epileptiform activity through the kindled dentate gyrus. J. ated depolarization evoked by high-frequency stimulation in
Neurophysiol., 79(4): 1726–1732. pyramidal neurons of rat hippocampal slice is attributable to
612

a network-driven, bicarbonate-dependent K+ transient. J. Sloviter, R.S., Zappone, C.A., et al. (2006) Kainic acid-induced
Neurosci., 17(20): 7662–7672. recurrent mossy fiber innervation of dentate gyrus inhibitory
Kleschevnikov, A.M. and Routtenberg, A. (2003) Long-term interneurons: possible anatomical substrate of granule cell
potentiation recruits a trisynaptic excitatory associative net- hyperinhibition in chronically epileptic rats. J. Comp.
work within the mouse dentate gyrus. Eur. J. Neurosci., Neurol., 494(6): 944–960.
17(12): 2690–2702. Staley, K.J. (2004) Role of the depolarizing GABA response in
LeMasson, G., Marder, E., et al. (1993) Activity-dependent epilepsy. Adv. Exp. Med. Biol., 548: 104–109.
regulation of conductances in model neurons. Science, Staley, K.J., Otis, T.S., et al. (1992) Membrane properties of
259(5103): 1915–1917. dentate gyrus granule cells: comparison of sharp microelec-
Lothman, E.W., Stringer, J.L., et al. (1992) The dentate gyrus trode and whole-cell recordings. J. Neurophysiol., 67(5):
as a control point for seizures in the hippocampus and 1346–1358.
beyond. Epilepsy Res. Suppl., 7: 301–313. Stell, B.M. and Mody, I. (2002) Receptors with different affin-
Marder, E. and Prinz, A.A. (2002) Modeling stability in neuron ities mediate phasic and tonic GABA(A) conductances in
and network function: the role of activity in homeostasis. hippocampal neurons. J. Neurosci., 22(10): RC223.
Bioessays, 24: 1145–1154. Steward, O. and Scoville, S.A. (1976) Cells of origin of ent-
McCarren, M. and Alger, B.E. (1985) Use-dependent depres- orhinal cortical afferents to the hippocampus and fascia
sion of IPSPs in rat hippocampal pyramidal cells in vitro. dentata of the rat. J. Comp. Neurol., 169(3): 347–370.
J. Neurophysiol., 53(2): 557–571. Steward, O., White, W.F., et al. (1977) Potentiation of the
Miller, K.D. (1996) Synaptic economics: competition and excitatory synaptic action of commissural, associational and
cooperation in synaptic plasticity. Neuron, 17: 371–374. entorhinal afferents to dentate granule cells. Brain Res.,
Mody, I. (2005) Aspects of the homeostaic plasticity of GAB- 134(3): 551–560.
AA receptor-mediated inhibition. J. Physiol., 562(Pt 1): Stringer, J.L. and Lothman, E.W. (1992) Reverberatory seizure
37–46. discharges in hippocampal-parahippocampal circuits. Exp.
Mody, I., Otis, T.S., et al. (1992) The balance between excita- Neurol., 116(2): 198–203.
tion and inhibition in dentate granule cells and its role in Stringer, J.L., Williamson, J.M., et al. (1989) Induction of
epilepsy. Epilepsy Res. Suppl., 9: 331–339. paroxysmal discharges in the dentate gyrus: frequency
Mott, D.D. and Lewis, D.V. (1992) GABAB receptors mediate dependence and relationship to afterdischarge production.
disinhibition and facilitate long-term potentiation in the J. Neurophysiol., 62(1): 126–135.
dentate gyrus. Epilepsy Res. Suppl., 7: 119–134. Strowbridge, B.W., Buckmaster, P.S., et al. (1992) Potentiation
Nusser, Z. and Mody, I. (2002) Selective modulation of tonic of spontaneous synaptic activity in rat mossy cells. Neurosci.
and phasic inhibitions in dentate gyrus granule cells. J. Lett., 142(2): 205–210.
Neurophysiol., 87(5): 2624–2628. Strowbridge, B.W. and Schwartzkroin, P.A. (1996) Transient
Peng, Z. and Houser, C.R. (2005) Temporal patterns of fos potentiation of spontaneous EPSPs in rat mossy cells induced
expression in the dentate gyrus after spontaneous seizures in by depolarization of a single neurone. J. Physiol., 494(Pt 2):
a mouse model of temporal lobe epilepsy. J. Neurosci., 493–510.
25(31): 7210–7220. Sutula, T.P., Golarai, G., et al. (1992) Assessing the functional
Royer, S. and Pare, D. (2003) Conservation of total synaptic significance of mossy fiber sprouting. Epilepsy Res. Suppl., 7:
weight through balanced synaptic depression and potentiat- 251–259.
ion. Nature, 422(6931): 518–522. Teyler, T.J. and Alger, B.E. (1976) Monosynaptic habituation
Scanziani, M., Gahwiler, B.H., et al. (1991) Paroxysmal inhib- in the vertebrate forebrain: the dentate gyrus examined in
itory potentials mediated by GABAB receptors in partially vitro. Brain Res., 115(3): 413–425.
disinhibited rat hippocampal slice cultures. J. Physiol., 444: Thompson, S.M. and Gahwiler, B.H. (1989a) Activity-depend-
375–396. ent disinhibition. I. Repetitive stimulation reduces IPSP driv-
Scharfman, H.E. (1992) Differentiation of rat dentate neurons ing force and conductance in the hippocampus in vitro.
by morphology and electrophysiology in hippocampal slices: J. Neurophysiol., 61(3): 501–511.
granule cells, spiny hilar cells and aspiny ‘fast-spiking’ cells. Thompson, S.M. and Gahwiler, B.H. (1989b) Activity-depend-
Epilepsy Res. Suppl., 7: 93–109. ent disinhibition. II. Effects of extracellular potassium,
Scharfman, H.E. (1995) Electrophysiological evidence that furosemide, and membrane potential on ECl- in hippocam-
dentate hilar mossy cells are excitatory and innervate both pal CA3 neurons. J. Neurophysiol., 61(3): 512–523.
granule cells and interneurons. J. Neurophysiol., 74(1): Thompson, S.M. and Gahwiler, B.H. (1989c) Activity-depend-
179–194. ent disinhibition. III. Desensitization and GABAB receptor-
Scharfman, H.E. (2004) Functional implications of seizure-in- mediated presynaptic inhibition in the hippocampus in vitro.
duced neurogenesis. Adv. Exp. Med. Biol., 548: 192–212. J. Neurophysiol., 61(3): 524–533.
Scharfman, H.E., Kunkel, D.D., et al. (1990) Synaptic connec- Thomson, A.M., Deuchars, J., et al. (1993) Single axon exci-
tions of dentate granule cells and hilar neurons: results of tatory postsynaptic potentials in neocortical interneurons
paired intracellular recordings and intracellular horseradish exhibit pronounced paired pulse facilitation. Neuroscience,
peroxidase injections. Neuroscience, 37(3): 693–707. 54(2): 347–360.
613

Turrigiano, G.G. and Nelson, S.B. (2000) Hebb and home- sion varies with behavioral state. Science, 196(4295):
ostasis in neuronal plasticity. Curr. Opin. Neurobiol., 10(3): 1223–1225.
358–364. Winson, J. and Abzug, C. (1978) Neuronal transmission
Walther, H., Lambert, J.D., et al. (1986) Epileptiform activity through hippocampal pathways dependent on behavior.
in combined slices of the hippocampus, subiculum and J. Neurophysiol., 41(3): 716–732.
entorhinal cortex during perfusion with low magnesium Witter, M.P. and Amaral, D.G. (1991) Entorhinal cortex of the
medium. Neurosci. Lett., 69(2): 156–161. monkey: V. Projections to the dentate gyrus, hippocampus,
Wierenga, C.J., Ibata, K., et al. (2005) Postsynaptic expression and subicular complex. J. Comp. Neurol., 307(3): 437–459.
of homeostatic plasticity at neocortical synapses. J. Neuro- Wong, R.K. and Watkins, D.J. (1982) Cellular factors influ-
sci., 25(11): 2895–2905. encing GABA response in hippocampal pyramidal cells.
Williamson, A., Spencer, D.D., et al. (1993) Comparison be- J. Neurophysiol., 48(4): 938–951.
tween the membrane and synaptic properties of human and Zappone, C.A. and Sloviter, R.S. (2004) Translamellar
rodent dentate granule cells. Brain Res., 622(1-2): 194–202. disinhibition in the rat hippocampal dentate gyrus after
Winson, J. and Abzug, C. (1977) Gating of neuronal seizure-induced degeneration of vulnerable hilar neurons.
transmission in the hippocampus: efficacy of transmis- J. Neurosci., 24(4): 853–864.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 33

Role of the dual entorhinal inputs to hippocampus:


a hypothesis based on cue/action (non-self/self)
couplets

John E. Lisman

Department of Biology and Volen Center for Complex Systems, Brandeis University, Waltham, MA 02454, USA

Abstract: The hippocampus sits at the highest level of memory processing circuits and receives two major
inputs, one coming from the lateral entorhinal cortex and one coming from the medial entorhinal cortex.
This duality must be of fundamental importance, but its functional meaning remains unclear. A compu-
tational model used for robot navigation (Verschure, P.F., et al. (2003). Nature, 425: 620–624) has a dual
information structure that may provide insight. In this model, information is stored as couplets consisting
of information about the current sensory cues and information about the current action of the robot.
Sequences of such couplets are stored in a short-term memory buffer and transferred to a long-term
memory store whenever a goal is found. The overall system enhances the ability of the robot to find reward
sites because stored sequences enable the robot to retrace the path to a goal site whenever any of the cues
along the path to a goal is subsequently encountered. A review of the literature suggests that the idea of cue/
action couplets can be usefully mapped onto the function of the entorhinal cortex. Cue information may be
supplied by the lateral entorhinal cortex whereas action (motor) information may be supplied by the medial
entorhinal cortex. However, given that self-position information is prominent in the medial pathway and
that this is not directly related to action, a modified formulation of the duality is proposed in which the
fundamental distinction is between information about non-self vs. information about self. According to this
view, the lateral entorhinal pathway carries information about external (non-self) cues and their positions
(in egocentric coordinates) whereas the medial entorhinal pathway carries information about the organism
itself, including its position (in allocentric coordinates), motor actions and goals.

Keywords: short-term memory; reward; navigation; sequence memory; dopamine

Overview and medial regions. These provide excitatory input


to several parts of the hippocampus, but the du-
The duality (Fig. 1) of the connections from the ality of their structure is most obvious in the dent-
entorhinal cortex to the hippocampus is anatom- ate gyrus. The outermost dendritic region of
ically striking (Burwell et al., 1995). The entorhinal dentate granule cells receives exclusive input from
cortex contains two distinct regions, the lateral the lateral entorhinal cortex whereas the middle
dendritic region receives exclusive input from the
Corresponding author. Tel.: +1 781 736 3145 or 3148; medial entorhinal cortex. Each granule cell re-
Fax: +1 781 736 3107; E-mail: Lisman@brandeis.edu ceives both types of inputs, thereby integrating the

DOI: 10.1016/S0079-6123(07)63033-7 615


616

Fig. 1. Wiring diagram of the hippocampus (interneurons excluded) showing dual inputs from the lateral and medial entorhinal cortex.
The layer 2 cortical inputs to the dentate and CA3 diverge fan out (f) widely over these networks and then provide convergent input to
individual dentate granule cells. In contrast, the layer 3 cortical inputs are specialized for individual subregions of CA1. This is an
example of a point-to-point (p-p) connection. Within the hippocampus, granule cells provide input, via mossy fibers to the mossy cells
of the dentate and to CA3 cells. CA3 cells make feedback connections to themselves and to the mossy cells of the dentate. These, in
turn, provide excitatory input to granule cells in the inner third of the granule cell dendritic tree. (See Color Plate 33.1 in color plate
section.)

two lines of cortical information. The purpose of But the robot becomes much more efficient at
this review is to discuss the possible functional finding reward sites if there are additional circuits.
basis of this mysterious duality. These make it possible for a cue that is far away
A priori, what are the grand dualities that might from a reward site to specify a complex path that,
be considered? Here is a list of some possibilities: according to previous experience, led to the reward
what/where; specific/context; sensory/motor; past/ site. The circuitry that allows the robot to do this
present; conscious/unconscious; rewarded/punished; (Fig. 2) involves a ‘‘cortical’’ multi-item ‘‘circular’’
stimulus/response. short-term memory (STM) buffer that stores the
My interest in the last of these, stimulus/re- last five salient events, and a long-term ‘‘hippo-
sponse, was stimulated by a paper on robotic con- campal’’ memory that stores sequences of events in
trol (Verschure et al., 2003). In that paper, the long-term memory (LTM). When the robot for-
authors describe a ‘‘brain-like’’ computer program ages and accidentally finds a reward, it incorpo-
that enables a robot to efficiently find sites at rates the entire content of the buffer into its LTM
which reward is located. The authors postulated store. Importantly, each event is defined as a stim-
several levels of control. At the lowest levels, cir- ulus/response couplet (Verschure and Voegtlin,
cuits support classical conditioning. In this way, 1998; Verschure et al., 2003). The ‘‘stimulus’’ part
the robot learns to associate a previously neutral of the couplet corresponds to a prototype of the
stimulus with a reward that is close to the robot. sensations from the external world at a particular
617

Fig. 2. Diagram of circuits responsible for storing events that led to the discovery of goals according to the model of Verschure et al.
(2003). Stimulus (C) and motor action (A) for each event are stored in a circular (first in, first out) short-term memory buffer. This
multi-item buffer can store five C/A couplets. When a goal is reached, the sequence in the short-term memory buffer is transferred to
the long-term memory buffer together with a last item representing the goal (G). As illustrated, the long-term memory store has stored
six sequences.

location. The response part of the couplet corre- Sutton, 1981; Hasselmo, 2005). This form of nav-
sponds to the motor act that was executed to get igation is less flexible than map-based navigation,
from that cue to the next location (cue) in the se- which may have developed later in the evolution.
quence. ‘‘Stimulus’’ and ‘‘response’’ have a partic- Importantly, the use of C/A sequences to specify
ular meaning in classical conditioning that is not the route to an accidentally discovered reward site
appropriate in the current context; I will therefore absolutely relies on one-trial learning, a property
use the terminology cue/action (C/A) to describe a that remains a defining property of human hippo-
couplet. Thus the utilization of stored couplets can campal episodic memory. Indeed, the development
be described as follows. If the robot comes across a of a multi-item STM buffer may have been the
previously encountered cue, Cn, it can get to the crucial evolutionary change that made one-trial
reward site using a simple algorithm: execute the learning possible rather than a change in synaptic
action An, that is associated with Cn in LTM; when plasticity mechanisms themselves. This is because
Cn+1 is found, execute An+1, etc. such buffers capture one-trial, brief events and
Several comments are in order. First, Verschure then, upon command, provide the repetitive activ-
(personal communication) was not aware of the ity patterns to LTM networks that are necessary to
dual inputs to the hippocampus; the dual repre- produce stable synaptic modifications.
sentation was utilized simply because it enhanced In the following sections I will address two
the robot’s efficiency at finding goals. Second the questions:
use of a sequence of cue-driven actions to find a
goal site is an experimentally observed mode of 1. Is the idea of storing C/A couplets potentially
animal navigation (Collett et al., 2003) and is re- applicable to understanding the dual cortical
lated to ideas incorporated into previous rein- inputs to the hippocampus? In particular, is
forcement driven models of behavior (Barto and there any way of mapping this duality onto
618

the known properties of the medial and lat- only when the behavioral paradigm has conse-
eral entorhinal inputs to the hippocampus? quences for reward (Smith and Mizumori, 2006).
2. What is the status of the evidence for the The influence of goals on hippocampal responses
various building blocks postulated in the is not limited to rats, but is also evident in humans
model of Fig. 2? (Ekstrom et al., 2003).
There is also strong experimental support for
the entry of sensory-specific information into the
Evaluation of the cue/action (C/A) couplet hippocampus. Odor-specific cells are found in the
hypothesis rat hippocampus (Wood et al., 1999), and neurons
that are scene-dependent have been observed in
Evidence that the hippocampus receives action monkey hippocampus (Wirth et al., 2003). Rat
(motor) and cue (sensory) information hippocampal neurons become sensitive to tones
after aversive conditioning (Berger et al., 1976;
The idea that the hippocampus might be influ- Moita et al., 2003). Particularly clear evidence for
enced by the motor system is one that is not com- cells responsive to particular faces has been
monly discussed, but support for this idea is obtained in humans (Quiroga et al., 2005).
actually quite long-standing. Vanderwolf (1969) It thus seems clear that both motor and sensory
observed that the frequency and amplitude of hip- information can get to the hippocampus. Moreo-
pocampal theta oscillations is modulated by the ver, particularly important from the perspective of
speed of a rat [reviewed in Bland and Oddie the hypothesis developed here is the finding of cells
(2001)]. Other work shows the influence of motor in both rat (Ranck, 1973) and monkey (Wirth
set and running speed on place cell firing et al., 2003) that jointly encode sensory and motor
(McNaughton et al., 1983; Foster et al., 1989; information. For instance, in the monkey work,
Wiener et al., 1989). Ranck (1973) found hippo- particular pictures are rewarded for movements in
campal neurons that fired on ‘‘approach’’ to a particular direction; as learning becomes appar-
objects. More recent work provides direct evi- ent, some hippocampal neurons fire selectively
dence that ‘‘motor intent’’ influences hippocampal when a particular picture is presented and the
function. Specifically, when rats run in a T-maze, monkey performs the correct learned movement.
the firing of neurons that occurs when the rat is in
the vertical stem of the T can be different depend-
ing on which direction the rat will turn at the top Evidence that the hippocampus is necessary for
of the stem (Frank et al., 2000; Wood et al., 2000; sensory/motor associations
Bower et al., 2005). Because the sensory stimula-
tion in the stem is always identical, it is hard to An important series of experiments has addressed
escape the conclusion that goals or action are what whether the hippocampus is necessary for learning
cause the difference in firing. It might be argued the association between sensory cues and motor
that what is represented is the goal, in some ab- responses. It was found that lesions to the fornix, a
stract sense not related to the motor action needed major input/output tract of the hippocampus, pre-
to get to goal, but this is not the case. In exper- vent a monkey from learning which direction to
iments using a plus maze, rats were started in the move in response to a sensory cue (Rupniak and
north or south arm and had to reach a goal in the Gaffan, 1987), and interferes with the ability of a
east arm; when the rat reached the east arm, many monkey to report the previous movement that it
cells fired differentially depending on whether made (Gaffan, 1985). Recent work (Brasted et al.,
the approach was from the north or south 2005) extends these observations by showing that
(Ferbinteanu and Shapiro, 2003). Thus, although fornix lesions prevent learning of arbitrary sen-
a common goal was involved, information about sory/motor tasks in single trials, the fast learning
the specific path taken was encoded. Recent results that is required to enhance foraging (see above).
emphasize that pathway-specific firing may occur Related work on instrumental conditioning in rats
619

demonstrates a similar requirement for hippocam- interfere with place learning (Ferbinteanu et al.,
pal involvement (Corbit and Balleine, 2000). 1999).
Taken together, these observations indicate that
hippocampal cells code for much more than po-
Validity of the C/A duality
sition and are influenced by the goal and actions
required to get there. Moreover, the association of
In mapping the C/A duality onto the entorhinal
these two types of information in the hippocampus
cortex, we would have to suppose that the lateral
appears necessary for behavioral responses based
region is sensory (cue) and that the medial region
on this association. I now turn to the issue of
represents ‘‘action’’. However, the most abundant
which region of the entorhinal cortex supplies each
cells in the medial region are grid cells, represent-
type of information.
ing the position of the rat in the environment, and
such information does not seem directly tied
to action. At best one could argue that motor in-
Sensory properties of the lateral entorhinal cortex
formation may be used to compute position
(McNaughton et al., 2006); the computation is
The perirhinal cortex, often termed inferotemporal
not dramatically altered when important sensory
cortex, is involved in high-level aspects of object
cues, such as vision, are taken away, probably
recognition (Brown and Aggleton, 2001). This re-
because a path integration mechanism can com-
gion provides strong input to the lateral entorhinal
pute position using head direction and velocity
cortex, but little to the medial entorhinal (Burwell,
information, both of which are motor-related.
2000). Physiological evidence for sensory input to
A significant number of layer 3 cells are direction-
the lateral entorhinal cortex is its responsiveness to
dependent, providing additional suggestive evi-
stimulation of olfactory cortex in rat (Gnatkovsky
dence that goal/motor information is present in
et al., 2004) and the presence of novelty-sensitive
the medial entorhinal cortex (Sargolini et al.,
responses to visually presented objects in monkey
2006). Thus, to some extent, positional cells could
(Fahy et al., 1993). Thus, the idea that the lateral
be described as related to action, but this argument
perforant pathway is sensory driven has experi-
is not compelling.
mental support. Consistent with a sensory role,
Another major theory of the duality (what/where)
lesions of the hippocampal inputs from the lateral
also faces difficulties. If one views grid cells as pri-
entorhinal cortex produces decreased investigation
marily representing ‘‘where’’, it at first seems sen-
of novel objects (Myhrer, 1988).
sible to conclude (Hargreaves et al., 2005) that the
dual inputs to the hippocampus form a what/where
couplet, not a C/A couplet. However, does the
Motor/goal properties of the medial entorhinal
‘‘where’’ in this formulation refer to the organism
cortex
or to objects in the environment? The brain needs to
keep track of both, and it is unclear how this would
In contrast to the cells of the lateral entorhinal
be done in the context of a what/where duality.
cortex, which do not have spatial properties
(Hargreaves et al., 2005), layers 2 and 3 of the
medial entorhinal cortex contain grid cells with Cue/action as a special case of a non-self/self
robust spatial properties (Fyhn et al., 2004; Hafting duality
et al., 2005; Sargolini et al., 2006). These cells are
coded in an allocentric coordinate system (i.e. with To deal with these difficulties, I suggest a refor-
respect to an absolute reference frame in the envi- mulation of the duality as non-self information
ronment, as e.g. north/south). Consistent with the vs. self-information. According to this view, the
presence of such allocentric spatial information, lateral entorhinal pathway would carry information
lesions of the hippocampal inputs from the medial about what is in the environment (non-self) and
entorhinal (but not the lateral entorhinal cortex) the ‘‘where’’ of those objects. The coding of the
620

‘‘where’’ component would presumably be done in transfer the buffer content to the LTM store when a
the egocentric (relative to the observer) coordinates reward site is found. I will now review what is
of the parietal lobe. The medial entorhinal pathway known about the existence of such components.
would encode information about various aspects of
self that includes a general specification of where
Evidence for the circular multi-item STM buffer in
one is in the environment (in allocentric coordinates
cortex
— relative to a fixed map of the environment) and
the action being taken towards achieving goals.
A requirement of reward-dependent storage of
Importantly, the allocentric coding evident in the
long paths leading to the reward is the existence of
medial entorhinal cortex could provide an excellent
a multi-item ‘‘circular’’ buffer. When the reward
way of specifying directional action (e.g. turn
site is found, the contents of the buffer (the C/A
North) because the correct action can be triggered
sequences leading to the reward site) are trans-
by a cue approached from any direction. In con-
ferred into the LTM store. The existence of a
trast, egocentric specification (e.g. turn Right)
multi-item STM in humans has been deduced from
would work only if the cue is approached from
psychophysical studies (Atkinson and Shiffrin,
the same angle as during initial learning.
1968). This buffer is thought to have limited ca-
To give a specific example of how non-self/self
pacity (772 items) and to be ‘‘circular’’. Here
information might be encoded by a lateral/medial
‘‘circular’’ means that when the buffer is full, the
entorhinal couplet, consider the following descrip-
next arriving item knocks out the item that has
tion of a moment in my morning commute. When
been in the buffer the longest (i.e. first in, first out).
I see the following non-self information: [McDo-
In the context of the model shown in Fig. 2, the
nald’s and, to its right, Burger King (complex cue
circular property ensures that when the goal is
specified in egocentric coordinates)], this would be
reached, transfer of buffer information into LTM
coupled to the following self information: [given that
will incorporate information about the last five
I am on my way to work (my goal), when I’m near
events that occurred before the goal was found.
Brandeis (my place approximated in allocentric co-
A physiologically plausible model of a multi-
ordinates), I turn West (my action in allocentric
item STM buffer has been developed (Lisman and
coordinates)].
Idiart, 1995) and a recent variant has first in, first
Parenthetically, one can see from this example
out properties (Koene and Hasselmo, 2006). fMRI
that purely landmark-based navigation would fail
evidence points to the temporal lobe as a site of a
if there are many similar landmarks, as is the case
multi-item working memory buffer, as evidenced
for fast food establishments, and that even a crude
by the load dependence of the fMRI signal (Fie-
allocentric system (‘‘near Brandeis’’) can therefore
bach et al., 2006). There is evidence that the ent-
be helpful in disambiguating landmark cues.
orhinal cortex can maintain STM information
It would seem that the non-self/self formulation
(Otto and Eichenbaum, 1992) [reviewed in Jensen
is a plausible one and fits the data somewhat better
and Lisman (2005); Hasselmo and Stern (2006)].
than the C/A formulation. In the final sections I
Recent experiments indicate that the persistent fir-
will turn to the question of how this formulation
ing is due to a cholinergically enhanced intrinsic
might be tested.
conductance that produces an after depolarization
(Fransen et al., 2006), as postulated in the theo-
retical models (Lisman and Idiart, 1995).
Evidence for the building blocks required by the
model (Fig. 2)
Evidence for reward-mediated fixation of STM
The general model of Verschure et al. requires sev- content into hippocampal LTM
eral important building blocks: a circular, multi-
item STM buffer, a LTM capable of holding cou- It is generally thought that reward signaling is
plet sequences, and a control system that would mediated by the dopaminergic cells of the ventral
621

tegmental area and substantia nigra. It was orig- successful transfer of sequences from a multi-item
inally thought that dopamine did not affect the cortical STM buffer to hippocampal LTM.
hippocampus, but recent work indicates that It may be instructive to consider how such
there is dopaminergic innervation of the hippo- processes might transfer C/A sequences from STM
campus and that it can profoundly affect trans- to LTM and later use the stored information to
mission and enhance LTP [reviewed in Lisman recall them. During learning, the elements of C
and Grace (2005)]. In the context of the model of from lateral entorhinal cortex and the elements of
Fig. 2, one might imagine that it is the dopamine A from medial entorhinal cortex converge on sub-
reward signal that stimulates reward-dependent sets of granule cells; each granule cell that fires
transfer of information from STM to LTM, but forms an element of the C/A representation. The
no experiments have addressed this directly, so output of active granule cells excites a subset of
not much can be said about this possibility at the CA3 pyramidal cells, and these are combined into
present time. One relevant observation (Foster an associative representation of the C/A couplet
and Wilson, 2006) is that when rats reach a re- by LTP in the recurrent connections of the active
ward site, a replay of the sequence of events CA3 cells. The CA3 output is then sent back to the
(places) that led the rat to the reward occurs in dentate, where the synapses onto granule cells
the hippocampus (the replay occurs in reverse representing elements of the next C/A couplet in
order). Although it is suspected that this replay the sequence become potentiated. These feedback
occurs as a result of processes within the hippo- synapses thus form linkages between sequential
campus itself, the possibility that it actually oc- events. In this way, the entire sequence can be
curs because of input from the cortex needs to be transferred from a cortical buffer to the hippo-
directly tested. campal memory [see Lisman et al. (2005) for de-
Although the role of dopamine in transfer, is tails]. Now let us assume that the organism
uncertain, there is quite strong evidence that do- subsequently comes across C2, and the corre-
pamine is necessary for the late stage of LTP (re- sponding input is supplied to the dentate, which
viewed in Lisman and Grace, 2005). In this way, then provides input to CA3. Through autoassoci-
dopamine may contribute to the reward-dependent ative memory of the stored couplet, the A2 com-
storage of sequences envisioned in Fig. 2. ponent of the couplet will be evoked. The
completed C2/A2 couplet is sent back to the dent-
ate, where it evokes the C3/A3 couplet (this is
Evidence for hippocampal LTM stores capable of termed a chaining step). The vagaries of synaptic
storing sequences transmission will, however, lead to a slightly cor-
rupted representation of the couplet, which, if used
There have been many computational models of the to chain to the next element, would lead to an even
hippocampus (Burgess et al., 2001; Kunec more corrupted version of C4/A4. However, the
et al., 2005; Rolls and Kesner, 2006), but most reciprocal connections between the dentate and
deal exclusively with autoassociational memories CA3 can serve to prevent such concatenation of
that encode events at a given moment, and so can- errors (Lisman, 1999). Specifically, C3A3 sent to
not be used to account for sequences of events. CA3 where the autoassociative properties of CA3
Models developed in my laboratory, which have correct the errors in the C3/A3 representation and
been improved over time (Jensen and Lisman, 1996; it is this corrected version that is sent from CA3 to
Lisman, 1999; Lisman et al., 2005), deal with how the dentate, where it triggers C4/A4. Thus, through
the hippocampus can store and recall sequences (see sequential chaining and correction processes, the
also Levy (1996)]. This class of models shows how dentate and CA3 can lead to accurate recall of the
gamma and theta oscillations, standard NMDA entire sequence of couplets leading to a goal. Other
receptor-mediated plasticity processes and known brain circuits could then use this stored informa-
anatomical connections (including the feedback tion to make a decision about whether this goal is
connections from CA3 to dentate) can mediate worthwhile in the current context and, if so, to
622

execute the specific motor plans stored in the cou- Representation of self-action on the T-maze in the
plet sequence. medial entorhinal cortex
Although most computational models of the
hippocampus posit that the sole function of the It should be possible to test the hypothesis pro-
dentate is to produce orthogonalization (the gen- posed here by recording from the entorhinal cortex
eration of very different firing patterns to slightly during the T- and plus-maze paradigms that dem-
different input patterns), it is clear from the anat- onstrate motor/action information in the hippo-
omy that a second function, the mixing of lateral campus. The specific prediction is that the medial,
and medial inputs, must also be important. If the but not the lateral entorhinal cortex should encode
ideas proposed here are correct, dentate and CA3 motor/goal information. Preliminary evidence in-
cells should display a unique code that is a mixture dicates that this prediction is correct (Lipton et al.,
of sensory cues and action. 2006).
Importantly, the mixing of cortical inputs to
CA1 is already quite different; the region close to
CA3 gets exclusive input from the medial ent- Unique conjunctions of non-self/self in the dentate/
orhinal cortex whereas the region closer to the CA3; cue information in the lateral entorhinal cortex
subiculum gets exclusive input from the lateral
entorhinal input (Fig. 1). The same segregation A second set of predictions has to do with the
holds for the CA1 projections back to cortex. This changes in representation. The dentate and CA3
pattern of segregation is different from dentate/ should have cells that represent conjunction of
CA3, where all cells receive both lateral and medial non-self and self (e.g. cue and action) and this
information, thus presumably mixing them into a conjunction should not be present in layers 2 and 3
joint representation. CA1 does not appear to be of the entorhinal cortex. An example of such
part of this coding system; it is thus likely to have a a conjunction would be a cell that responded to
role in converting the dentate/CA3 code back into scene, but only in conjunction with a cued action
the separate codes of the lateral and medial ent- (Wirth et al., 2003). In these experiments, the scene
orhinal cortex (Lisman, 1999). and cue information should be encoded in the lat-
eral entorhinal; the action taken (which may
or may not be consistent with the cue) should be
Concluding remarks and predictions of the model encoded in the medial entorhinal.

My goal in this discussion has been to examine the


ideas incorporated into the model of Fig. 2 and to Conditioned fearful stimuli vs. the emotion of fear
assess whether they can be usefully mapped onto
the architecture of the hippocampus (Fig. 1). The output of the basolateral amygdala represents
My general conclusion is that the correspondence a sensory signal, the conditioned stimulus (CS),
is promising. The basic building blocks utilized in and may well be source of the hippocampal re-
the model have a reasonable correspondence to the sponse to the CS that develops after conditioning
capabilities of the cortex and hippocampus. (Berger et al., 1976). Because these signals are non-
The more specific question, whether any of the self similar signals would be predicted in the lateral
functional dualities inspired by the model (C/A, entorhinal cortex. In contrast the further process-
non-self/self) underlie the structural duality of ent- ing in the amygdala that occurs in the central nu-
orhinal inputs, cannot yet be answered with any cleus (Pare et al., 2004; Balleine and Killcross,
certainty. The idea that the medial entorhinal 2006) is thought to represent the emotion of fear
encodes aspects of self whereas the lateral ent- and to evoke fear-related responses in the hypo-
orhinal encodes the outside world (non-self) leads thalamus and brainstem. Such emotional re-
to experimentally testable predictions, as outlined sponses are a property of self and would be
in the following paragraphs. predicted to occur in the medial entorhinal cortex.
623

Connectivity with areas mediating personal/extra- Barto, A.G. and Sutton, R.S. (1981) Landmark learning: an
personal attention illustration of associative search. Biol. Cybern., 42: 1–8.
Berger, T.W., Alger, B. and Thompson, R.F. (1976) Neuronal
substrate of classical conditioning in the hippocampus. Sci-
If there is a fundamental segregation of information ence, 192: 483–485.
about self and non-self in the entorhinal cortex, Bland, B.H. and Oddie, S.D. (2001) Theta band oscillation and
might one expect to find functionally related sub- synchrony in the hippocampal formation and associated
divisions at earlier stages of cortical processing? In structures: the case for its role in sensorimotor integration.
this regard it is interesting that very recent work Behav. Brain Res., 127: 119–136.
Bower, M.R., Euston, D.R. and McNaughton, B.L. (2005) Se-
(Committeri et al., 2007; Ortigue et al., 2006) points quential-context-dependent hippocampal activity is not nec-
to two very different forms of hemi-neglect due to essary to learn sequences with repeated elements. J.
brain injuries. One form, termed ‘‘extra-personal’’, Neurosci., 25: 1313–1323.
leads to neglect of one half of the external world; Brasted, P.J., Bussey, T.J., Murray, E.A. and Wise, S.P.
(2005) Conditional motor learning in the nonspatial do-
the other form, termed ‘‘personal’’ leads to neglect
main: effects of errorless learning and the contribution of
of half of the body. These forms can be doubly the fornix to one-trial learning. Behav. Neurosci., 119:
dissociated and involve injuries to different brain 662–676.
regions. It will be of interest to determine whether Brown, M.W. and Aggleton, J.P. (2001) Recognition memory:
these regions have selective connections to the me- what are the roles of the perirhinal cortex and hippocampus?
dial and lateral entorhinal cortex. Nat. Rev. Neurosci., 2: 51–61.
Buckner, R.L. and Carroll, D.C. (2007) Self-projection and the
brain. Trends Cogn. Sci., 11: 49–57.
Burgess, N., Becker, S., King, J.A. and O’Keefe, J. (2001)
Note added in proof Memory for events and their spatial context: models and ex-
periments. Philos. Trans. R. Soc. Lond. B Biol. Sci., 356:
I have recently become aware of other works on 1493–1503.
cortical and subcortical processing that posits a Burwell, R.D. (2000) The parahippocampal region: corticocor-
separate neural system for self-referential processing tical connectivity. Ann. N.Y. Acad. Sci., 911: 25–42.
Burwell, R.D., Witter, M.P. and Amaral, D.G. (1995) Per-
(Buckner and Carroll, 2007; Northoff et al., 2006). irhinal and postrhinal cortices of the rat: a review of the
neuroanatomical literature and comparison with findings
from the monkey brain. Hippocampus, 5: 390–408.
Acknowledgments Collett, T.S., Graham, P. and Durier, V. (2003) Route learning
by insects. Curr. Opin. Neurobiol., 13: 718–725.
I thank Andre Fenton, Menno Witter, Bruce McN- Committeri, G., Pitzalis, S., Galati, G., Patria, F.,
aughton, Edvard Moser, James Knierim, Steven Pelle, G., Sabatini, U., Castriota-Scanderbeg, A., Piccardi,
L., Guariglia, C. and Pizzamiglio, L. (2007) Neural bases of
Wise, Howard Eichenbaum, David Redish, and
personal and extrapersonal neglect in humans. Brain, 130:
Matthew Shapiro for useful conversations. I thank 431–441.
Paul Verschure for comments on the manuscript. Corbit, L.H. and Balleine, B.W. (2000) The role of the hippo-
This work was supported by NIH grants R01 campus in instrumental conditioning. J. Neurosci., 20:
NH027337, P50 MH060450-07A1, and R01 4233–4239.
Ekstrom, A.D., Kahana, M.J., Caplan, J.B., Fields, T.A.,
MH61975-01A2. Special thanks also to the Dart
Isham, E.A., Newman, E.L. and Fried, I. (2003) Cellular
Foundation and the Marine Biological Laboratory. networks underlying human spatial navigation. Nature, 425:
184–188.
Fahy, F.L., Riches, I.P. and Brown, M.W. (1993) Neuronal
References activity related to visual recognition memory: long-term
memory and the encoding of recency and familiarity infor-
Atkinson, R. and Shiffrin, R. (1968) Human memory: a pro- mation in the primate anterior and medial inferior temporal
posed system and its control processes. In: Spence K. (Ed.), and rhinal cortex. Exp. Brain Res., 96: 457–472.
The psychology of learning and motivation: advances in re- Ferbinteanu, J., Holsinger, R.M. and McDonald, R.J. (1999)
search and theory, Vol. 2. Academic Press, New York. Lesions of the medial or lateral perforant path have different
Balleine, B.W. and Killcross, S. (2006) Parallel incentive effects on hippocampal contributions to place learning and
processing: an integrated view of amygdala function. Trends on fear conditioning to context. Behav. Brain Res., 101:
Neurosci., 29: 272–279. 65–84.
624

Ferbinteanu, J. and Shapiro, M.L. (2003) Prospective and ret- Levy, W.B. (1996) A sequence predicting CA3 is a flexible
rospective memory coding in the hippocampus. Neuron, 40: associator that learns and uses context to solve hippocampal-
1227–1239. like tasks. Hippocampus, 6: 579–590.
Fiebach, C.J., Rissman, J. and D’Esposito, M. (2006) Lipton, P.A., White, J.A. and Eichenbaum, H. (2006) Di-
Modulation of inferotemporal cortex activation during sambiguation of overlapping spatial sequences in medial ent-
verbal working memory maintenance. Neuron, 51: orhinal cortex. In: Annual Society for Neuroscience Meeting.
251–261. Atlanta, GA: Society for Neuroscience, Program #66.6/
Foster, D.J. and Wilson, M.A. (2006) Reverse replay of be- Poster #Z23.
havioural sequences in hippocampal place cells during the Lisman, J.E. (1999) Relating hippocampal circuitry to function:
awake state. Nature, 440: 680–683. recall of memory sequences by reciprocal dentate-CA3 inter-
Foster, T.C., Castro, C.A. and McNaughton, B.L. (1989) Spa- actions. Neuron, 22: 233–242.
tial selectivity of rat hippocampal neurons: dependence on Lisman, J.E. and Grace, A.A. (2005) The hippocampal-VTA
preparedness for movement. Science, 244: 1580–1582. loop: controlling the entry of information into long-term
Frank, L.M., Brown, E.N. and Wilson, M. (2000) Trajectory memory. Neuron, 46: 703–713.
encoding in the hippocampus and entorhinal cortex. Neuron, Lisman, J.E. and Idiart, M.A. (1995) Storage of 7+/ 2 short-
27: 169–178. term memories in oscillatory subcycles. Science, 267:
Fransen, E., Tahvildari, B., Egorov, A.V., Hasselmo, M.E. and 1512–1515.
Alonso, A.A. (2006) Mechanism of graded persistent cellular Lisman, J.E., Talamini, L.M. and Raffone, A. (2005) Recall of
activity of entorhinal cortex layer v neurons. Neuron, 49: memory sequences by interaction of the dentate and CA3: a
735–746. revised model of the phase precession. Neural Netw., 18:
Fyhn, M., Molden, S., Witter, M.P., Moser, E.I. and Moser, 1191–1201.
M.B. (2004) Spatial representation in the entorhinal cortex. McNaughton, B.L., Barnes, C.A. and O’Keefe, J. (1983) The
Science, 305: 1258–1264. contributions of position, direction, and velocity to single
Gaffan, D. (1985) Hippocampus: memory, habit and voluntary unit activity in the hippocampus of freely-moving rats. Exp.
movement. Philos. Trans. R. Soc. Lond. B Biol. Sci., 308: Brain Res., 52: 41–49.
87–99. McNaughton, B.L., Battaglia, F.P., Jensen, O., Moser, E.I.
Gnatkovsky, V., Uva, L. and de Curtis, M. (2004) Topographic and Moser, M.B. (2006) Path integration and the neural
distribution of direct and hippocampus-mediated entorhinal basis of the ‘cognitive map’. Nat. Rev. Neurosci., 7:
cortex activity evoked by olfactory tract stimulation. Eur. J. 663–678.
Neurosci., 20: 1897–1905. Moita, MA., Rosis, S., Zhou, Y., LeDoux, J.E. and Blair, H.T.
Hafting, T., Fyhn, M., Molden, S., Moser, M.B. and Moser, (2003) Hippocampal place cells acquire location-specific re-
E.I. (2005) Microstructure of a spatial map in the entorhinal sponses to the conditioned stimulus during auditory fear
cortex. Nature, 436: 801–806. conditioning. Neuron, 37: 485–497.
Hargreaves, E.L., Rao, G., Lee, I. and Knierim, J.J. Myhrer, T. (1988) Exploratory behavior and reaction to novelty
(2005) Major dissociation between medial and lateral ent- in rats with hippocampal perforant path systems disrupted.
orhinal input to dorsal hippocampus. Science, 308: Behav. Neurosci., 102: 356–362.
1792–1794. Northoff, G., Heinzel, A., de Greck, M., Bermpohl, F., Do-
Hasselmo, M.E. (2005) A model of prefrontal cortical mech- browolny, H. and Panksepp, J. (2006) Self-referential
anisms for goal-directed behavior. J. Cogn. Neurosci., 17: processing in our brain – a meta-analysis of imaging studies
1115–1129. on the self. Neuroimage, 31: 440–457.
Hasselmo, M.E. and Stern, C.E. (2006) Mechanisms underlying Ortigue, S., Megevand, P., Perren, F., Landis, T. and Blanke,
working memory for novel information. Trends. Cogn. Sci., O. (2006) Double dissociation between representational per-
10: 487–493. sonal and extrapersonal neglect. Neurology, 66: 1414–1417.
Jensen, O. and Lisman, J.E. (1996) Hippocampal CA3 region Otto, T. and Eichenbaum, H. (1992) Complementary roles of
predicts memory sequences: accounting for the phase preces- the orbital prefrontal cortex and the perirhinal-entorhinal
sion of place cells. Learn. Mem., 3: 279–287. cortices in an odor-guided delayed-nonmatching-to-sample
Jensen, O. and Lisman, J.E. (2005) Hippocampal sequence-en- task. Behav. Neurosci., 106: 762–775.
coding driven by a cortical multi-item working memory Pare, D., Quirk, G.J. and Ledoux, J.E. (2004) New vistas on
buffer. Trends Neurosci., 28: 67–72. amygdala networks in conditioned fear. J. Neurophysiol., 92:
Koene, R.A. and Hasselmo, M.E. (2006) First-in-first-out 1–9.
item replacement in a model of short-term memory Quiroga, R.Q., Reddy, L., Kreiman, G., Koch, C. and Fried, I.
based on persistent spiking. Cereb. Cortex (epub ahead of (2005) Invariant visual representation by single neurons in
print). the human brain. Nature, 435: 1102–1107.
Kunec, S., Hasselmo, M.E. and Kopell, N. (2005) En- Ranck Jr., J.B. (1973) Studies on single neurons in dorsal hip-
coding and retrieval in the CA3 region of the hippocampus: pocampal formation and septum in unrestrained rats. I. Be-
a model of theta-phase separation. J. Neurophysiol., 94: havioral correlates and firing repertoires. Exp. Neurol., 41:
70–82. 461–531.
625

Rolls, E.T. and Kesner, R.P. (2006) A computational theory of representations applied to a behaving real-world device: Dis-
hippocampal function, and empirical tests of the theory. tributed Adaptive Control III. Neural Netw., 11: 1531–1549.
Prog. Neurobiol., 79: 1–48. Verschure, P.F., Voegtlin, T. and Douglas, R.J. (2003) Envi-
Rupniak, N.M. and Gaffan, D. (1987) Monkey hippocampus ronmentally mediated synergy between perception and
and learning about spatially directed movements. J. Neuro- behaviour in mobile robots. Nature, 425: 620–624.
sci., 7: 2331–2337. Wiener, S.I., Paul, C.A. and Eichenbaum, H. (1989) Spatial and
Sargolini, F., Fyhn, M., Hafting, T., McNaughton, B.L., behavioral correlates of hippocampal neuronal activity. J.
Witter, M.P., Moser, M.B. and Moser, E.I. (2006) Conjunc- Neurosci., 9: 2737–2763.
tive representation of position, direction, and velocity in Wirth, S., Yanike, M., Frank, L.M., Smith, A.C., Brown, E.N.
entorhinal cortex. Science, 312: 758–762. and Suzuki, W.A. (2003) Single neurons in the monkey hip-
Smith, D.M. and Mizumori, S.J. (2006) Learning-related de- pocampus and learning of new associations. Science, 300:
velopment of context-specific neuronal responses to places 1578–1581.
and events: the hippocampal role in context processing. J. Wood, E.R., Dudchenko, P.A. and Eichenbaum, H. (1999) The
Neurosci., 26: 3154–3163. global record of memory in hippocampal neuronal activity.
Vanderwolf, C.H. (1969) Hippocampal electrical activity and Nature, 397: 613–616.
voluntary movement in the rat. Electroencephalogr. Clin. Wood, E.R., Dudchenko, P.A., Robitsek, R.J. and Eichen-
Neurophysiol., 26: 407–418. baum, H. (2000) Hippocampal neurons encode information
Verschure, P.F. and Voegtlin, T. (1998) A bottom up approach about different types of memory episodes occurring in the
towards the acquisition and expression of sequential same location. Neuron, 27: 623–633.
Plate 33.1. Wiring diagram of the hippocampus (interneurons excluded) showing dual inputs from the lateral and medial entorhinal
cortex. The layer 2 cortical inputs to the dentate and CA3 diverge fan out (f) widely over these networks and then provide convergent
input to individual dentate granule cells. In contrast, the layer 3 cortical inputs are specialized for individual subregions of CA1. This is
an example of a point-to-point (p-p) connection. Within the hippocampus, granule cells provide input, via mossy fibers to the mossy
cells of the dentate and to CA3 cells. CA3 cells make feedback connections to themselves and to the mossy cells of the dentate. These,
in turn, provide excitatory input to granule cells in the inner third of the granule cell dendritic tree. (For B/W version, see page 616 in
the volume.)

Plate 36.1. (A) Two representative granule cells filled with 5,6 carboxyfluorescein from the dentate gyrus of a 24-month-old rat. (B)
Histograms showing the numbers of carboxyfluorescein injections that resulted in single, double, triple, or greater numbers of granule
cells filled with dye. Aged rats showed significantly increased electrotonic coupling compared to young animals. This may account for
the increased excitability of old granule cells. Adapted with modifications from Barnes et al. (1987). (For B/W version, see page 665 in
the volume.)
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 34

The CA3 ‘‘backprojection’’ to the dentate gyrus

Helen E. Scharfman

Department of Pharmacology and Neurology, Columbia University, New York, NY 10032, USA and the Center for
Neural Recovery and Rehabilitation Research, Helen Hayes Hospital, New York State Department of Health,
Rte 9W, West Haverstraw, NY 10993-1195, USA

Abstract: The hippocampus is typically described in the context of the trisynaptic circuit, a pathway that
relays information from the perforant path to the dentate gyrus, dentate to area CA3, and CA3 to area
CA1. Associated with this concept is the assumption that most hippocampal information processing occurs
along the trisynaptic circuit. However, the entorhinal cortex may not be the only major extrinsic input to
consider, and the trisynaptic circuit may not be the only way information is processed in hippocampus.
Area CA3 receives input from a variety of sources, and may be as much of an ‘‘entry point’’ to hippo-
campus as the dentate gyrus. The axon of CA3 pyramidal cells targets diverse cell types, and has com-
missural projections, which together make it able to send information to much more of the hippocampus
than granule cells. Therefore, CA3 pyramidal cells seem better designed to spread information through
hippocampus than the granule cells. From this perspective, CA3 may be a point of entry that receives
information which needs to be ‘‘broadcasted,’’ whereas the dentate gyrus may be a point of entry that
receives information with more selective needs for hippocampal processing. One aspect of the argument
that CA3 pyramidal cells have a widespread projection is based on a part of its axonal arbor that has
received relatively little attention, the collaterals that project in the opposite direction to the trisynaptic
circuit, ‘‘back’’ to the dentate gyrus. The evidence for this ‘‘backprojection’’ to the dentate gyrus is strong,
particularly in area CA3c, the region closest to the dentate gyrus, and in temporal hippocampus. The
influence on granule cells is indirect, through hilar mossy cells and GABAergic neurons of the dentate
gyrus, and appears to include direct projections in the case of CA3c pyramidal cells of ventral hippo-
campus. Physiological studies suggest that normally area CA3 does not have a robust excitatory influence
on granule cells, but serves instead to inhibit it by activating dentate gyrus GABAergic neurons. Thus,
GABAergic inhibition normally controls the backprojection to dentate granule cells, analogous to the way
GABAergic inhibition appears to control the perforant path input to granule cells. From this perspective,
the dentate gyrus has two robust glutamatergic inputs, entorhinal cortex and CA3, and two ‘‘gates,’’ or
inhibitory filters that reduce the efficacy of both inputs, keeping granule cells relatively quiescent. When
GABAergic inhibition is reduced experimentally, or under pathological conditions, CA3 pyramidal cells
activate granule cells reliably, and do so primarily by disynaptic excitation that is mediated by mossy cells.
We suggest that the backprojection has important functions normally that are dynamically regulated by
nonprincipal cells of the dentate gyrus. Slightly reduced GABAergic input would lead to increased po-
lysynaptic associative processing between CA3 and the dentate gyrus. Under pathological conditions as-
sociated with loss of GABAergic interneurons, the backprojection may support reverberatory excitatory

Corresponding author. Tel.: +1 845 398 5427;


Fax: +1 845 398 5422; E-mail: hscharfman@nki.rfmh.org

DOI: 10.1016/S0079-6123(07)63034-9 627


628

activity between CA3, mossy cells, and granule cells, possibly enhanced by mossy fiber sprouting. In this
case, the backprojection could be important to seizure activity originating in hippocampus, and help
explain the seizure susceptibility of ventral hippocampus.

Keywords: pyramidal cell; dentate gyrus; hilus; mossy cell; interneuron; granule cell; recurrent excitation

Introduction
hippocampus is not to be doubted here. The point
of emphasis here is, instead, that the trisynaptic
Most explanations of hippocampal circuitry begin
pathway is not the only circuit to consider. One
with the identification of the major subfields and
reason for hesitation is that a considerable number
the trisynaptic circuit. As shown in Fig. 1, the
of studies have now shown that nonprincipal cells
hippocampus of mammals is composed of three
of hippocampus play critical roles in modulating
subfields: area CA1, containing primarily CA1 py-
the trisynaptic pathway. They are innervated by
ramidal cells, area CA3, also composed primarily
each of the fiber pathways mentioned above, and
of pyramidal cells, and the dentate gyrus, where
have diverse projections to all hippocampal prin-
the primary principal cell is the granule cell. The
cipal cells. For example, within the dentate gyrus,
trisynaptic circuit is composed of three sequential
the perforant path and the mossy fibers innervate
glutamatergic synapses:
numerous types of GABAergic neurons, as well as
1) perforant path axons of layer II neurons in glutamatergic mossy cells of the hilus. The GAB-
entorhinal cortex project to the outer two- Aergic ‘‘interneurons’’ and mossy cells project to
thirds of the dentate gyrus molecular layer, the granule cells at the level of the soma, axon hillock,
location of the distal granule cell dendrites, and at every part of the dendritic tree. Indeed, it
2) mossy fiber axons of granule cells project to has been proposed that the primary targets of
proximal dendrites of area CA3 pyramidal mossy fibers are GABAergic neurons, and the pri-
cells, and mary effect of mossy fiber transmission under
3) the Schaffer collateral axons of CA3 pyram- normal conditions is inhibitory (Acsady et al.,
idal cells project to stratum radiatum of CA3, 1998). This appears to be the case in recordings of
where the apical dendrites of area CA1 py- CA3 neurons in slices, where they dominant re-
ramidal cells are located (Amaral and Witter, sponse to mossy fiber stimulation is an inhibitory
1989). postsynaptic potential (IPSP) (Scharfman, 1993a,
b), despite the large unitary excitatory postsynap-
This projection is clearly fundamental, and tic potential (EPSP) produced by granule cells
its importance to information processing in (Scharfman et al., 1990) (for review, see Jaffe and

Fig. 1. The ‘‘trisynapto-centric’’ vs. ‘‘CA3-centric’’ view of hippocampal information processing. (A) The trisynaptic circuit is di-
agrammed schematically for a horizontal section through the adult rodent hippocampus. Cell layers are in gray. (B) The axonal arbor
of CA3 pyramidal cells is schematically presented to illustrate the perspective that these neurons may be a central point of information
processing in the hippocampus.
629

Gutierrez, this volume). It has been suggested that Ventral CA3c pyramidal cells collateralized
other characteristics of mossy fiber transmission approximately 3–4  as much in the hilus as dor-
may be even more important than their ability to sal CA3c pyramidal cells, or CA3b cells.
depolarize postsynaptic pyramidal cells (Urban Li et al. (1994) also noted that CA3 pyramidal
et al., 2001). cell axons entered the granule cell layer and col-
In this chapter an alternative to the trisynapto- lateralized in the inner molecular layer, and this
centric view is emphasized. It involves the axons of was predominantly a characteristic of ventral
CA3 pyramidal cells that project in the opposite CA3c pyramidal cells. These axons exhibited nu-
direction of the trisynaptic circuit (‘‘back projec- merous varicosities, suggesting that they inner-
tion’’), into the dentate gyrus (Fig. 1B). vated dentate gyrus granule cells, and possibly
other cell types. Therefore, the CA3c population
of ventral hippocampus appears to be a subset of
Anatomical evidence for a ‘‘backprojection’’ CA3 pyramidal cells with the highest potential for
dentate gyrus interactions.
What is the evidence for a ‘‘backprojection?’’ First These studies were the first to define hilar col-
we will consider anatomical arguments and then lateralization of CA3 neurons unequivocally.
physiological data. Historically, there have been However, one caveat was that the sample sizes
only rare suggestions that CA3 pyramidal cells were small, and it remains unclear exactly how
might project into the dentate gyrus (Zimmer, representative the data were of the entire CA3
1971; Schwerdtfeger and Sarvey, 1983). These did population. It also is unclear whether the data
not seem to attract much attention, possibly be- from adult rat might vary across age and species.
cause some of the experimental approaches had Due to the labor-intensive nature of the experi-
technical limitations. For example, the most com- ments, these questions remain unanswered.
mon approach for tract tracing at the time relied Comparisons of CA3a, b, and c pyramidal cells
on tracer injection, but CA3 is difficult to label are interesting to review in light of the findings that
selectively with this technique. their axons appear to have a different preference
Newer approaches provided the requisite selec- for the dentate gyrus. Anatomically, CA3c pyram-
tivity. These techniques involved dye injection, typ- idal cells are heterogeneous neurons (Scharfman,
ically biocytin, into individual CA3 pyramidal cells 1993; Turner et al., 1995), sometimes reflecting
using intracellular microelectrodes, followed by stereotypical pyramidal cell morphology, but in
axon reconstruction to evaluate the axon arbor. other cases they may appear more similar to non-
Using this approach, Ishizuka et al. (1990) filled pyramidal cells. Some of the electrophysiological
individual CA3 neurons in hippocampal slices of characteristics of CA3c pyramidal cells appear to
the rat, and emphasized several aspects of the axon be distinct from CA3a and b, and this may be
of CA3 neurons. They found evidence for axon important because it could lead to unique physi-
collateralization within the hilus in many of their ological attributes to the backprojection. CA3c
sampled cells, particularly those with a cell body in neurons are not as easily activated by mossy fiber
‘‘proximal’’ CA3 (near the dentate gyrus, i.e., CA3b stimulation as CA3b neurons in slices, suggesting
and CA3c). Li et al. (1994) injected biocytin into that the GABAergic neurons which are targets of
CA3 pyramidal cells in vivo, also using adult rats, mossy fibers innervate CA3c more than CA3b. All
and found even greater evidence for a backprojec- else being equal, such preferential inhibition might
tion. They found that individual CA3 neurons had quiet the backprojection normally, relative to
collaterals in the hilus whether the cell body of or- other CA3 projection systems.
igin was in CA3a, b, or c. Like Ishizuka et al. In comparisons of CA3b and CA3c pyramidal
(1990), they found the most hilar collateralization cells of the adult male rat, almost all CA3c neu-
from cells within CA3c, and they specifically iden- rons demonstrated burst (phasic) firing in response
tified the ventral portion of CA3c region as the area to current injection, whereas CA3b pyramidal cells
with the greatest collateralization in the hilus. were much more heterogeneous, often exhibiting
630

trains of action potential (tonic firing) only. Var- studies in hippocampal slices, which showed that a
iation in CA3a and b in this respect has been re- stimulating electrode placed in the hilus could ac-
ported (Bilkey and Schwartzkroin, 1990). If CA3c tivate an antidromic population spike, recorded
neurons do have a greater percentage of neurons extracellularly in the CA3 pyramidal cell layer.
which intrinsically discharge in bursts, the infor- Subsequent studies identified what cell types in
mation processed by backprojecting axons may the dentate gyrus could be influenced by the CA3
differ in the way it influences granule cells from the projection. Several approaches established that hi-
way information is transferred to other targets of lar neurons could be activated readily, but granule
CA3 neurons. cells could not.
CA3c neurons demonstrated the longest time First, fimbria stimulation was used to activate
constant, which is noteworthy because this char- CA3 pyramidal cells instead of the hilus, because
acteristic might lead to a greater capacity for in- antidromic action potentials are easily elicited in
tegration of synaptic inputs from distinct locations CA3 neurons by stimulation at this site, with no
along the dendrites (Scharfman, 1993b). However, evidence for direct activation of hilar neurons and
CA3c neurons appear to have a relatively short granule cells. Direct activation would be expected
electrotonic length, which would lead to rapid de- because hilar cells that project commissurally have
cay before integration of dendritic inputs would an axon that descends into stratum oriens of
reach the soma. Together the data would suggest a CA3b, presumably on its way to the fimbria, but in
specialization of the dendrite as a separable our slices, we have found that it is severed before
processing unit, perhaps more than other CA3 reaching the fimbria. Direct activation of granule
neurons. It also would suggest that somatic inputs cells would only be possible if current spread en-
would have a relatively greater influence on action tered stratum lucidum, which recordings show
potential output than dendritic inputs. does not occur after fimbria stimulation using the
CA3c is more vulnerable to certain types of ex- methods we have employed.
citotoxic stimulation than CA3a or b pyramidal cells Extracellular recordings showed that such fi-
(Chang and Dyer, 1985; Freund et al., 1992; Miet- mbria stimulation led to antidromic and or-
tinen et al., 1998; Haas et al., 2001). It was proposed thodromic activation of CA3 b and c pyramidal
that CA3 may be an area where burst generation cells, followed by an extracellular negative wave in
occurs in models of epilepsy, but others have sug- the granule cell layer and inner molecular layer
gested CA3a/b is that location (Knowles et al., 1987; that reversed polarity in the middle molecular
Colom and Saggau, 1994). Selective lesions to CA3a layer (Fig. 2). Due to the negative resting potential
vs. b vs. c are difficult to make without injuring of granule cells, the negative potential could reflect
adjacent areas, so the functional relevance of the an EPSP or depolarized IPSP, and it was shown
differences between CA3a, b, and c remain unclear. subsequently by intracellular recording that it was
In summary, the anatomical data suggest that a depolarized GABAA receptor-mediated IPSP
ventral hippocampus and CA3c neurons make the (Fig. 3; Scharfman, 1994a, b, c).
most robust backprojection in normal adult rats, This fimbria stimulation site activated hilar
and there is collateralization in the hilus, granule mossy cells and hilar GABAergic neurons disy-
cell layer and inner molecular layer. Specific phys- naptically, and was sensitive to the AMPA recep-
iological characteristics of CA3c neurons may im- tor antagonist CNQX but not the muscarinic
part unique functional consequences to the antagonist atropine (Scharfman, 1993a). At a
backprojection. longer latency, the granule cell IPSP began. These
data suggesting that fimbria activation of CA3
pyramidal cells led to hilar neuron activation, pre-
Physiological evidence for a ‘‘backprojection’’ sumably by hilar collaterals of CA3 pyramidal
cells. Comparing the latency to the response of
Some of the first physiological evidence that CA3 CA3 pyramidal cells and hilar cells supported this
axons innervated the dentate gyrus came from hypothesis, because CA3 pyramidal cells were
631

Fig. 2. Recordings in hippocampal slices show evidence of trisynaptic and backprojecting pathways. (A) Extracellular recordings of
evoked responses to five stimulation sites (1–5) and four recording sites (A–D). Recordings were made sequentially in the same slice in
response to a fixed stimulus for each stimulus site. (B) A schematic is used to allow better comparisons of fimbria-evoked and outer
molecular layer-evoked responses recorded at four sites in the same slices: the outer molecular layer, granule cell layer, CA3c cell layer,
and CA3b cell layer. X marks recording site locations. (C) Comparison of evoked responses to fimbria stimulation in the same slice
demonstrate the responses in the granule cell layer following pyramidal cell activation by the fimbria, but do so with a delay.
Superimposition of responses illustrates that the onset of the granule cell layer field potential begins approximately 1–2 ms after the
antidromic population spike recorded in area CA3b. This delay suggests an intermediary synapse, probably in CA3c or the hilus,
between CA3b and granule cells. Indeed, the granule cell field potential begins immediately after the CA3c orthodromic population
spike, which probably is due to CA3b recurrent excitation of CA3c pyramidal cells. At the same time, hilar cells are also activated and
innervate granule cells (see later figures), although CA3c pyramidal cells could innervate granule cells also (Li et al., 1994). The bottom
trace is an IPSP recorded intracellularly from a granule cell in the same slice in response to the same stimulus. It shows that the onset of
the IPSP is similar to the onset of the granule cell layer field potential, which likely reflects the average of many IPSPs in granule cells
situated around the extracellular electrode.

activated at short latency and hilar cells approx- GABAergic neurons may not only involve those
imately 2 ms later. The same stimulus evoked a in the hilus, but also the ‘‘basket cells’’ of the
long latency IPSP in granule cells, consistent with granule cell layer (Scharfman, 1994a, c; Kneisler
a CA3-hilar neuron-granule cell pathway. In the and Dingledine, 1995).
presence of bicuculline, a GABAA receptor antag- Use of GABAA receptor antagonists also al-
onist, fimbria stimulation evoked EPSPs in granule lowed another approach to examine the effects of
cells, and did so with a similar latency as the CA3 on the dentate gyrus, one that involved no
IPSPs, suggesting a CA3-mossy cell-granule cell stimulation to activate CA3 neurons. In the pres-
pathway is normally masked by the activation of ence of GABAergic receptor antagonists such as
GABAergic interneurons by CA3 hilar collaterals penicillin, picrotoxin, or bicuculline, CA3 neurons
(Fig. 3). Additional studies revealed that the discharge rhythmically in bursts of action
632

Fig. 3. Physiological evidence for a CA3 backprojection mediated by hilar neurons. (A) A schematic illustrates the backprojection
supported by physiological evidence collected to date. It shows that CA3 pyramidal cells innervate GABAergic and glutamatergic
mossy cells of the hilus, which in turn innervate granule cells. (B) Recordings illustrate the physiological correlates of the schematic in
A. Intracellular recordings from granule cells in response to fimbria stimulation in an adult male rat slice illustrate an evoked IPSP that
reverses at 70 mV, indicated a GABAA receptor-mediated IPSP is primarily evoked under normal conditions. After the GABAA
receptor antagonist bicuculline was bath-applied, the evoked response was an EPSP followed by an IPSP that reversed at approx-
imately 80 mV, suggesting an EPSP followed by a GABAB receptor-mediated IPSP is normally masked by GABAA receptor-
mediated inhibition. From Scharfman (1994a). Reprinted with permission from the Society for Neuroscience.

potentials (Schwartzkroin and Prince, 1977; The most parsimonious explanation was that CA3
Knowles et al., 1987; Scharfman, 1994a, c). There- pyramidal cells innervated both types of hilar
fore, one can examine simultaneously any other neurons, and evidence was obtained for this
cell type in the slice to determine when and if they by simultaneous intracellular recordings from mo-
are concurrently influenced. Simultaneous record- nosynaptically connected pyramidal cells and
ings showed that immediately after the onset of mossy cells (Scharfman, 1994b). The same ap-
spontaneous bursts in CA3 pyramidal cells, hilar proach showed that mossy cells innervated granule
mossy cells and GABAergic interneurons also cells and interneurons in the hilus (Scharfman,
demonstrated bursts (Scharfman, 1994a, c). Gran- 1995), suggesting that multiple polysynaptic cir-
ule cells depolarized and could reach threshold, cuits could be set in motion by CA3 pyramidal
but did so at a greater delay from the CA3 burst cells, and ultimately influence dentate granule
than hilar cells. The hilar activity and granule cell cells. Current source density has also provided evi-
excitation, but not CA3 bursts, were blocked by dence of CA3–dentate gyrus interactions similar to
transecting the border between the dentate gyrus those discussed above (Wu et al., 1998).
and CA3b, or by application of CNQX (Fig. 4; In summary, there is strong evidence from stud-
Scharfman, 1994a). When the timing of the burst ies in hippocampal slices that the CA3 backpro-
discharges was examined more closely, the onset of jection innervates hilar neurons, both mossy cells
the depolarization of CA3 neurons immediately and GABAergic neurons. Currently there is no
preceded the onset in the mossy cells and GAB- physiological evidence that CA3 directly inner-
Aergic neurons, and the onset of hilar neuron vates the granule cells of the dentate gyrus, but
bursts immediately preceded granule cell depolari- absence of evidence is not a proof of absence.
zation, suggesting single synaptic delays of a CA3- At the present time, the evidence suggests that
hilus-granule cell pathway (Scharfman, 1994a, c). the granule cells, even if they receive a direct
633

Fig. 4. The CA3 excitatory backprojection is dependent on hilar mossy cells. (A) Schematic illustrating the experimental approach for
results shown in B–C. A site in CA3c was used for extracellular recording while recording intracellularly from a granule cell. Pressure
application of CNQX was used at two distinct sites in the hilus or in the inner molecular layer. CNQX was applied in microdrops that
were barely detectable by eye, allowing preferential application to select areas of the slice. (B) In the presence of bicuculline, CA3
epileptiform discharges were evoked by fimbria stimulation, and following the onset of the burst discharge, a granule cell depolarized
and discharged two action potentials. After CNQX was pressure-applied to the hilus, the granule cell response decreased but the CA3
discharge remained unaffected. (C) In a different slice, CNQX pressure-application to the inner molecular layer, near the recorded
granule cell, reversibly decreased the EPSP of the granule cell, suggesting an inner molecular layer glutamatergic synapse was necessary
for the EPSP. From Scharfman (1994a). Reprinted with permission from the Society for Neuroscience.

projection from CA3 neurons, are primarily in- involve layer II of entorhinal cortex specifically, but
hibited by the backprojection, not activated. How- it may be rare that layer II is ever activated in iso-
ever, once they are relieved of GABAA receptor- lation, given its broad input from other layers
mediated inhibition, a strong excitatory circuit is within the entorhinal cortex, and the diverse inputs
unmasked. The excitatory circuit appears to be to entorhinal cortex. In actuality, it is highly likely
primarily derived from the mossy cell projection to that the sensory and other cortical input that comes
the inner molecular layer. to entorhinal cortex would also involve subcortical
structures that at the same time might reach CA3
through the fimbria. Furthermore, the tempo-
Functional implications of the CA3 backprojection roammonic pathway may activate CA3 neurons at
lower threshold than granule cells (see Chapter by
Hippocampal information processing Derrick, this volume), and this further supports the
tenet that incoming information might first activate
Is CA3 the point of entry to hippocampus? the CA3 region of hippocampus, rather than do so
via dentate granule cells. Given the data that mossy
Given the projection of CA3 pyramidal cells inner- fibers primarily inhibit CA3 pyramidal cells, the
vates CA1, the dentate gyrus, other CA3 neurons dentate may actually function to keep CA3 from
by recurrent collaterals, and sends information both being activated excessively by the concomitant in-
ipsilaterally and contralaterally, the argument can puts from the cortex and subcortical zones.
be made that it is a cell type that may be central to
hippocampal information processing (Fig. 1B). In
light of this, it is interesting that the general as- The dentate gyrus has two ‘‘gates’’ rather than one
sumption is ‘‘trisynapto-centric,’’ that granule cells
receive the primary input, and the mossy fiber con- Another implication of the discussion above, par-
trol how the hippocampus processes the incoming ticularly the physiological studies’ is that the dent-
input. This view might be correct in cases that ate gyrus actually has two ‘‘gates’’ (Fig. 5). The
634

If this occurs, it would be predicted that asso-


ciative processes would be enhanced. In other
words, not only could there be recurrent excitatory
circuits among CA3 pyramidal cells that perform
autoassociative functions, but polysynaptic auto-
association would also potentially develop be-
tween CA3 and the granule cells when GABAergic
inhibition is weak. One argument that makes this
even more compelling is that granule cells do not
necessarily need to reach threshold in order for
mossy fiber transmission to occur (Alle and Gei-
ger, 2006). Subthreshold depolarizations in gran-
Fig. 5. The trisynaptic circuit and CA3 backprojection sup-
ports a bi-directional gate to the dentate gyrus granule cells.
ule cells can lead to effects in granule cell targets.
An important implication is that autoassociative
networks, particularly the complex CA3–dentate
first is a gate that prevents activation by the ent- network discussed here, is likely to be controlled
orhinal cortex (see Chapter by Hsu, this volume). by GABAergic hilar neurons and mossy cells.
This may not be an all-or-none function, so the
word ‘‘filter’’ may be better (see Chapter by Dudek
and Sutula, this volume). However, the concept is Pathological conditions
appropriate regardless of the semantics: cortical
input is not very effective in activating granule The CA3 backprojection may subserve an impor-
cells above their threshold. tant role in conditions like ischemia or epilepsy,
It appears that a second ‘‘gate’’ or ‘‘filter’’ also when hilar GABAergic neurons are injured or killed
exists, from CA3 pyramidal cells back to the dent- (Johansen et al., 1987; Sloviter, 1987; see Chapter by
ate gyrus. Thus, the backprojection does not ap- Tallent, this volume). In both conditions, the som-
pear to be effective in depolarizing granule cells, atostatin-immunoreactive hilar neurons that project
and it appears that the reason is the divergence of preferentially to the outer molecular layer are lost.
CA3 hilar collaterals to both excitatory hilar In animal models of epilepsy that use status epilep-
mossy cells, and inhibitory GABAergic interneu- ticus as the initial experimental manipulation to in-
rons. Evidently, the inhibitory neurons exert a duce an epileptic state, there are many types of
stronger net influence, and it is important to note GABAergic neurons that may be affected, but most
that in vivo it may be different, because mossy cell laboratories suggest a relative preservation of parv-
axons are truncated in the slice much more than albumin-immunoreactive basket cells. The animal
GABAergic neurons. In vivo, one would predict models of epilepsy are also interesting because both
that CA3 would inhibit granule cells within the the granule cell axons and the pyramidal cell axons
same lamella, but activate them in distal sections may sprout new collaterals in response to injury
of hippocampus because of the long translamellar adjacent to them. Granule cell axons sprout into the
mossy cell axon projection. inner molecular layer and hilus, as well as stratum
When inhibition is reduced experimentally, CA3 oriens in CA3. It would be of interest to determine
discharge is extremely robust in depolarizing gran- whether the backprojection increases in response to
ule cells, and it appears to be primarily due to injury.
inner molecular synapses of mossy cells. Thus, the Regardless of these changes, it does appear that
hilar neurons modulate the CA3 backprojection the backprojection is robust under epileptic condi-
under normal conditions, and any condition that tions. The data are from hippocampal slices from
would reduce GABAergic inhibition would be ex- animals that had status epilepticus and became ep-
pected to facilitate CA3 activation of dentate ileptic, i.e., they developed recurrent spontaneous
granule cells. seizures for the rest of their lifespan. In most of
635

Fig. 6. Evidence for associative networks and reverberatory circuits in the dentate gyrus under control of GABAergic inhibition. (A) A
schematic illustrates electrode locations for the recordings shown in B. (B) Stimulation of specific sites in the slice evoked bursts of
action potentials in CA3 pyramidal cells, and simultaneous intracellular recordings from a granule cell show bi-directional activation
of CA3 and the granule cell. (C) In a different slice from B where CA3 epileptiform bursts were followed by numerous after discharges,
CA3 appeared to precede activation of the simultaneously-recorded granule cell, but during the after discharges, the converse appeared
to develop. From Scharfman (1994a). Reprinted with permission from the Society for Neuroscience.

these animals, CA3 pyramidal cell burst discharges (Fig. 6). During burst discharges in CA3, we have
occur when slices are prepared, even when bicucul- found that the onset of the burst discharge occurs
line is absent (Scharfman et al., 2001). This affords in pyramidal cells first, and then in granule cells,
an opportunity to examine the backprojection by but during after discharges that follow the primary
simply examining concurrent activity in hilar neu- burst, granule cell discharge may precede the py-
rons and granule cells. When this has been done, it ramidal cells. These data suggest CA3 may be a
is clear that the backprojection to mossy cells still generator of activity initially, but granule cells can
exists, because mossy cells exhibit synchronized initiate further excitation. In the whole animal,
burst discharges with CA3 pyramidal cells in slices such reverberatory excitation could lead to sub-
from rats with recurrent seizures (Scharfman et al., stantial network activity of CA3/dentate neurons,
2001). Interneurons in the hilus do as well, although and possibly trigger seizures after a certain thresh-
the population of both mossy cells and interneurons old of network activity is reached. One would pre-
is reduced relative to the normal rat, and it is not dict that this might occur in more ventral
yet clear which GABAergic neurons are involved of hippocampus than dorsal, given that the backpro-
the many types that exist. Granule cells usually ap- jection is more prevalent in ventral hippocampus.
pear to be relatively quiet when CA3 burst dis- This prediction is supported by findings in slices of
charges occur (Scharfman and McCloskey, 2007) temporal hippocampus, where seizure activity is
but this is not always the case: some granule cells most robust (Bragdon et al., 1986; Borck and
have depolarizations during the CA3 burst dis- Jefferys, 1999; Scharfman et al., 2002).
charges. In this case, bicuculline may not be re-
quired because of a loss of GABAergic inhibition
when hilar inhibitory neurons are killed after status Conclusions
epilepticus.
This pathway may be relevant to the seizures The CA3 region is often considered to be the sec-
that occur in the whole animal, because the CA3 ond subfield that receives information from ex-
discharges appear to generate reverberatory activ- trinsic afferents to hippocampus, while the dentate
ity between the dentate and CA3 pyramidal cells gyrus is the initial entry point. Here we suggest
636

that CA3 may be a major gateway to hippocam- Bilkey, D.K. and Schwartzkroin, P.A. (1990) Variation in
pus, particularly with respect to subcortical input. electrophysiology and morphology of hippocampal CA3 py-
ramidal cells. Brain Res., 514: 77–83.
One reason to think of CA3 as a central access
Borck, C. and Jefferys, J.G. (1999) Seizure-like events in dis-
point is that it has a projection that reaches the inhibited ventral slices of adult rat hippocampus. J. Ne-
vast majority of hippocampal neurons, both ipsi- urophysiol., 82: 2130–2142.
laterally and contralaterally, including both CA1 Bragdon, A.C., Taylor, D.M. and Wilson, W.A. (1986) Potas-
and the dentate gyrus. The projection to the dent- sium-induced epileptiform activity in area CA3 varies mark-
ate gyrus, a ‘backprojection,’ has not been studied edly along the septotemporal axis of the rat hippocampus.
Brain Res., 378: 169–173.
in as much detail as the Schaffer collateral system, Chang, L.W. and Dyer, R.S. (1985) Septotemporal gradients of
leading to more appreciation of CA1–CA3 inter- trimethyltin-induced hippocampal lesions. Neurobehav. Tox-
actions when CA3–dentate networks may actually icol. Teratol., 7: 43–49.
be just as important to hippocampal function. Colom, L.V. and Saggau, P. (1994) Spontaneous interictal-like
activity originates in multiple areas of the CA2-CA3 region
Physiological data suggest that the CA3 backpro-
of hippocampal slices. J. Neurophysiol., 71: 1574–1585.
jection engages intermediary hilar neurons to exert Freund, T.F., Ylinen, A., Miettinen, R., Pitkanen, A., Lah-
its primary effect on granule cells, although direct tinen, H., Baimbridge, K.G. and Riekkinen, P.J. (1992) Pat-
projections to granule cells in a subset of CA3 tern of neuronal death in the rat hippocampus after status
neurons in ventral hippocampus have been shown epilepticus. Relationship to calcium binding protein content
by anatomical methods. GABAergic inhibition and ischemic vulnerability. Brain Res. Bull., 28: 27–38.
Haas, K.Z., Sperber, E.F., Opanashuk, L.A., Stanton, P.K. and
appears to regulate the ability of CA3 pyramidal Moshe, S.L. (2001) Resistance of immature hippocampus to
cells to exert a primarily inhibitory or excitatory morphologic and physiologic alterations following status ep-
effect on granule cells, and to control the possible ilepticus or kindling. Hippocampus, 11: 615–625.
reverberatory activity between the two subfields. Ishizuka, N., Weber, J. and Amaral, D.G. (1990) Organi-
CA3–dentate interactions may subserve an impor- zation of intrahippocampal projections originating from
CA3 pyramidal cells in the rat. J. Comp. Neurol., 295:
tant associative function in the behaving animal, 580–623.
influenced by factors that depress GABAergic Johansen, F.F., Zimmer, J. and Diemer, N.H. (1987) Early loss
neuron activity or GABAergic transmission; this of somatostatin neurons in dentate hilus after cerebral is-
may also apply under pathological conditions chemia in the rat precedes CA-1 pyramidal cell loss. Acta
Neuropathol. (Berl.), 73: 110–114.
when GABAergic neurons are damaged or lost,
Kneisler, T.B. and Dingledine, R. (1995) Synaptic input from
such as temporal lobe epilepsy. CA3 pyramidal cells to dentate basket cells in rat hippocam-
pus. J. Physiol., 487: 125–146.
Knowles, W.D., Traub, R.D. and Strowbridge, B.W. (1987)
Acknowledgments The initiation and spread of epileptiform bursts in the in vitro
hippocampal slice. Neuroscience, 21: 441–455.
Li, X.G., Somogyi, P., Ylinen, A. and Buzsaki, G. (1994) The
NIH (NINDS) and the New York State Depart- hippocampal CA3 network: an in vivo intracellular labeling
ment of Health. study. J. Comp. Neurol., 339: 181–208.
Miettinen, R., Kotti, T., Tuunanen, J., Toppinen, A., Ri-
ekkinen Sr., P. and Halonen, T. (1998) Hippocampal damage
after injection of kainic acid into the rat entorhinal cortex.
References Brain Res., 813: 9–17.
Scharfman, H.E. (1993a) Activation of dentate hilar neurons by
Acsady, L., Kamondi, A., Sik, A., Freund, T. and Buzsaki, G. stimulation of the fimbria in rat hippocampal slices. Neurosci
(1998) GABAergic cells are the major postsynaptic targets of Lett., 156: 61–66.
mossy fibers in the rat hippocampus. J. Neurosci., 18: Scharfman, H.E. (1993b) Spiny neurons of area CA3c in rat
3386–3403. hippocampal slices have similar electrophysiological charac-
Alle, H. and Geiger, J.R. (2006) Combined analog and action teristics and synaptic responses despite morphological vari-
potential coding in hippocampal mossy fibers. Science, 311: ation. Hippocampus, 3: 9–28.
1290–1293. Scharfman, H.E. (1994a) EPSPs of dentate gyrus granule cells
Amaral, D.G. and Witter, M.P. (1989) The three-dimensional during epileptiform bursts of dentate hilar ‘‘mossy’’ cells and
organization of the hippocampal formation: a review of an- area CA3 pyramidal cells in disinhibited rat hippocampal
atomical data. Neuroscience, 31: 571–591. slices. J. Neurosci., 14: 6041–6057.
637

Scharfman, H.E. (1994b) Evidence from simultaneous intracellu- in the normal and epileptic rat dentate gyrus. J. Comp. Ne-
lar recordings in rat hippocampal slices that area CA3 pyram- urol., 454: 424–439.
idal cells innervate dentate hilar mossy cells. J. Neurophysiol., Schwartzkroin, P.A. and Prince, D.A. (1977) Penicillin-induced
72: 2167–2180. epileptiform activity in the hippocampal in vitro preparation.
Scharfman, H.E. (1994c) Synchronization of area CA3 hippo- Ann. Neurol., 1: 463–469.
campal pyramidal cells and non-granule cells of the dentate Schwerdtfeger, W.K. and Sarvey, J.M. (1983) Connectivity of
gyrus in bicuculline-treated rat hippocampal slices. Neuro- the hilar region of the hippocampal formation in the rat. J.
science, 59: 245–257. Hirnforschung, 24: 201–207.
Scharfman, H.E. (1995) Electrophysiological evidence that Sloviter, R.S. (1987) Decreased hippocampal inhibition and a
dentate hilar mossy cells are excitatory and innervate both selective loss of interneurons in experimental epilepsy. Sci-
granule cells and interneurons. J. Neurophysiol., 74: 179–194. ence, 235: 73–76.
Scharfman, H.E., Kunkel, D.D. and Schwartzkroin, P.A. (1990) Turner, D.A., Li, X.G., Pyapali, G.K., Ylinen, A. and Buzsaki,
Synaptic connections of dentate granule cells and hilar neu- G. (1995) Morphometric and electrical properties of recon-
rons: results of paired intracellular recordings and intracellular structed hippocampal CA3 neurons recorded in vivo. J.
horseradish peroxidase injections. Neuroscience, 37: 693–707. Comp. Neurol., 356: 580–594.
Scharfman, H.E., Goodman, J.H. and McCloskey, D.P. (2007) Urban, N.N., Henze, D.A. and Barrionuevo, G. (2001) Revis-
Ectopic granule cells of the rat dentate gyrus. Dev. Neurosci., iting the role of the hippocampal mossy fiber synapse. Hip-
29: 14–27. pocampus, 11: 408–417.
Scharfman, H.E., Smith, K.L., Goodman, J.H. and Sollas, A.L. Wu, K., Canning, K.J. and Leung, L.S. (1998) Functional in-
(2001) Survival of dentate hilar mossy cells after pilocarpine- terconnections between CA3 and the dentate gyrus revealed
induced seizures and their synchronized burst discharges with by current source density analysis. Hippocampus, 8: 217–230.
area CA3 pyramidal cells. Neuroscience, 104: 741–759. Zimmer, J. (1971) Ipsilateral afferents to the commissural zone
Scharfman, H.E., Sollas, A.L., Smith, K.L., Jackson, M.B. and of the fascia dentata, demonstrated in decommissurated rats
Goodman, J.H. (2002) Structural and functional asymmetry by silver impregnation. J. Comp. Neurol., 142: 393–416.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 35

Modeling the dentate gyrus

Robert J. Morgan1,, Vijayalakshmi Santhakumar2 and Ivan Soltesz1

1
Department of Anatomy and Neurobiology, 193 Irvine Hall, University of California, Irvine, CA 92697, USA
2
Department of Neurology, David Geffen School of Medicine, University of California, Los Angeles, CA, USA

Abstract: Computational modeling has become an increasingly useful tool for studying complex neuronal
circuits such as the dentate gyrus. In order to effectively apply computational techniques and theories to
answer pressing biological questions, however, it is necessary to develop detailed, data-driven models.
Development of such models is a complicated process, akin to putting together a jigsaw puzzle with the
pieces being such things as cell types, cell numbers, and specific connectivity. This chapter provides a
walkthrough for the development of a very large-scale, biophysically realistic model of the dentate gyrus.
Subsequently, it demonstrates the utility of a modeling approach in asking and answering questions about
both healthy and pathological states involving the modeled brain region. Finally, this chapter discusses
some predictions that come directly from the model that can be tested in future experimental approaches.

Keywords: computational model; epilepsy; granule cell; interneuron; mossy cell; sclerosis

The dentate gyrus in the mammalian brain is a (Bartos et al., 2001; Vida et al., 2006), homeostatic
complex neuronal circuit that is thought to serve plasticity (Howard et al., 2007), and topological al-
as a gate for activity propagation throughout the terations during epileptogenesis (Dyhrfjeld-Johnsen
limbic system (Heinemann et al., 1992; Lothman et al., 2007). These studies demonstrate the impor-
et al., 1992). Scientific understanding of such com- tance and utility of computational modeling in con-
plex systems has traditionally been obtained tributing to our understanding of the dentate gyrus
through use of animal models, culture systems, both in healthy animals and in pathological states
and numerous other experimental approaches. In such as epilepsy.
the past several decades, however, computational In this chapter, we will use the analogy of a
modeling has become an increasingly useful tool for jigsaw puzzle to walk through the process of
studying and understanding neuronal network modeling the dentate gyrus neural network. In the
structure and function. In fact, the dentate gyrus first and second parts, ‘‘Finding the right pieces’’
has been the subject of a number of recent compu- and ‘‘Putting the pieces together,’’ we will discuss
tational models looking at such issues as mossy fiber what types of considerations are necessary when
sprouting (Lytton et al., 1998; Santhakumar et al., building a large-scale, data-driven model of a nor-
2005; Dyhrfield-Johnsen et al., 2007) and hilar cell mal, healthy rat dentate gyrus (all data presented
loss (Santhakumar et al., 2005; Dyhrfjeld-Johnsen in this chapter comes from the rat dentate). Then,
et al., 2007), interneurons and shunting inhibition in ‘‘The big picture’’ we will examine how such a
model is useful for studying the effects of struc-
Corresponding author. Tel.: +1 (949) 824-3967; tural and functional network alterations that occur
E-mail: rjmorgan@uci.edu during epileptogenesis. Finally, we will conclude

DOI: 10.1016/S0079-6123(07)63035-0 639


640

with ‘‘Tackling future puzzles and problems’’ by cells), hilar cells with axonal projections to the
discussing future directions for dentate modeling commissural-associational pathway (HICAP cells),
and some predictions that can be tested experi- and interneuron-specific (IS) cells (Fig. 1) (San-
mentally. thakumar and Soltesz, 2004). The second piece is
the number of cells of each type, data that can be
estimated from published data (Table 1). The third
Finding the right pieces piece is the specific connectivity between cell types.
This quantifies precisely how many postsynaptic
Assembling a computational model is much like cells among each of the eight cell types a single
putting together a puzzle. The pieces include cell presynaptic neuron of a given type innervates (e.g.,
types, cell numbers, synapses, and other cellular from the third row, second column in Table 1; a
interactions such as gap junctions, and distributions single basket cell innervates approximately 1,250
of axons and dendrites. These are assembled in a GCs; mean and ranges are indicated, with refer-
specific fashion upon a backdrop that includes the ences). The final piece required for constructing the
software necessary for implementing the model and connection matrix of the dentate is the spatial con-
the hardware, the physical computer and memory straint in cell-to-cell connectivity. For each cell
used for the model calculations. Just as the number type, the extent of the axons of single cells along the
and size of pieces in a puzzle varies, the number of septo-temporal axis of the dentate gyrus can be de-
components and the details of those components termined from in vivo single cell fills published in
vary dramatically in computational models. The the literature (Fig. 2).
precise amount of detail necessary in a model de-
pends directly on the question the model is designed
The first piece: cell types
to answer. Examining the effects of the loss of spe-
cific subgroups of cell types, for example, requires
Determination of cell type is based on several fac-
the model to capture the relative changes in cell
tors including: (1) the nature of the neurotrans-
densities, necessitating the development of large-
mitter released; (2) location in the laminar
scale models. The first two sections of this chapter
structure of the dentate gyrus; (3) morphological
will provide an overview of the general approach
features such as shape of the soma, dendritic ar-
for developing such a large-scale model to study
bor, and axonal distribution pattern; and (4) pres-
both the healthy rat dentate gyrus and the effects of
ence of specific markers such as calcium-binding
epileptogenesis with emphasis on mossy fiber
proteins and neuropeptides (for detailed descrip-
sprouting and hilar cell loss.
tion of cell type discrimination see Freund and
With over a million cells and a billion synapses
Buzsaki, 1996). According to these criteria, two
amongst at least eight distinct cell types in the
excitatory cell types (GCs and MCs) and at least
dentate gyrus, the first step in developing a model is
six distinct inhibitory interneurons can be identi-
to determine the anatomy of the dentate network.
fied in the dentate gyrus (Fig. 1).
This can be accomplished by creating a connection
matrix of the dentate gyrus (Patton and Mcnaugh-
ton, 1995), which can be thought of as the border or The second piece: cell numbers
backbone of our puzzle. The matrix itself contains a
number of distinct pieces that are outlined here and A practical approach for determining cell numbers
described in detail below. The first piece is the cell for various cell types involves first identifying the
type. Eight types of dentate cells can be identified as cell layers and the total number of neurons in each
anatomically well-described: granule cells (GCs), layer. Then, it is possible to estimate subtype num-
mossy cells (MCs), basket cells (BCs), axo-axonic bers by estimating the percentage of neurons that
cells, molecular layer cells with axonal projections express specific markers (Buckmaster et al., 1996,
to the perforant path (MOPP cells), hilar cells with 2002a; Nomura et al., 1997a, b). The dentate gyrus
axonal projections to the perforant path (HIPP is composed of three cell layers: the molecular layer,
641

Fig. 1. Schematic of the basic circuitry of the dentate gyrus. Relational representation of the healthy dentate gyrus illustrating the
network connections between the eight major cell types (GC, granule cell; BC, basket cell; MC, mossy cell; AAC, axo-axonic cells;
MOPP, molecular layer interneurons with axons in perforant-path termination zone; HIPP, hilar interneurons with axons in perforant-
path termination zone; HICAP, hilar interneurons with axons in the commissural/associational pathway termination zone; IS, in-
terneuron selective cells). The schematic shows the characteristic location of the various cell types within the three layers of the dentate
gyrus. Note, however, that this diagram does not indicate the topography of axonal connectivity or the somato-dendritic location of
the synapses. Reproduced with permission from Dyhrfjeld-Johnsen et al. (2007).

granule cell layer, and the hilus (Fig. 1). The GCs (presumed MCs) as approximately 30,000. The
are the primary projection cells of the dentate GABAergic interneurons are located in all three
gyrus. These are small, densely packed cells, typi- layers of the dentate gyrus. These include the BCs,
cally located in the granule cell layer of the dentate parvalbumin-positive (PV) axo-axonic cells (whose
gyrus. The number of GCs in the rat dentate gyrus axon terminals project exclusively to the axon initial
is approximately 1,000,000 (Gaarskjaer, 1978; Boss segment of excitatory cells), somatostatin-positive
et al., 1985; West, 1990; Patton and Mcnaughton, HIPP cells (Freund and Buzsaki, 1996; Katona et
1995; Freund and Buzsaki, 1996). The other major al., 1999), nitric oxide synthase-positive HICAP
excitatory cell type, the MC, is located in the hilus cells (Freund and Buzsaki, 1996), and aspiny and
and does not express the GABA synthetic enzyme calretinin-positive hilar IS cells (Gulyas et al., 1996).
glutamic acid decarboxylase (GAD). Buckmaster The number of GABAergic cell types along the
and Jongen-Relo (1999) estimated the number of granule cell-hilar border and in the hilus can be
GAD-mRNA negative neurons in the dentate hilus determined based on published histochemical data
642

Table 1. Connectivity matrix for the neuronal network of the control dentate gyrus

Granule cells Mossy cells Basket cells Axo-axonic ells MOPP cells HIPP cells HICAP cells IS cells

Granule cells X 9.5 15 3 X 110 40 20


(1,000,000) X 7–12 10–20 1–5 X 100–120 30–50 10–30
Ref. [1–5] Ref. [6] Ref. [7] Ref. [6–9] Ref. [6,7,9] Ref. [6] Ref. [4,10,11] Ref. [4,7,10,11] Ref. [7]
Mossy cells 32,500 350 7.5 7.5 5 600 200 X
(30,000) 30,000–35,000 200–500 5–10 5–10 5 600 200 X
Ref. [11] Ref. [4,11–13] Ref. [12,13] Ref. [13] Ref. [13] Ref. [14] Ref. [12,13] Ref. [12,13] Ref. [15]
Basket cells 1250 75 35 X X 0.5 X X
(10,000) 1000–1500 50–100 20–50 X X 0–1 X X
Ref. [16,17] Ref. [4,16–19] Ref. Ref. Ref. [18] Ref. [18] Ref. [18] Ref. [18] Ref. [10,20]
[11,16,17,19] [16,17,20,21]
Axo-axonic 3000 150 X X X X X X
cells
(2000) 2000–4000 100–200 X X X X X X
Ref. [4,22] Ref. [4,18,22] Ref. Ref. [5,18] Ref. [5,18] Ref. [5,18] Ref. [5,18] Ref. [5,18] Ref. [5,18,19]
[4,5,11,14,23]
MOPP cells 7500 X 40 1.5 7.5 X 7.5 X
(4000) 5000–10,000 X 30–50 1–2 5–10 X 5–10 X
Ref. [11,14] Ref. [14] Ref. [14,24] Ref. [14,25] Ref. [14,26] Ref. [14,25] Ref. [14,20,25] Ref. [14,25] Ref. [14,15]
HIPP cells 1550 35 450 30 15 X 15 X
(12,000) 1500–1600 20–50 400–500 20–40 10–20 X 10–20 X
Ref. [11] Ref. [4,11,20] Ref. Ref. [4,11,20] Ref. [20,25] Ref. [25] Ref. [14,20,25] Ref. [25] Ref. [15,20]
[4,11,12,27,28]
HICAP cells 700 35 175 X 15 50 50 X
(3000) 700 30–40 150–200 X 10–20 50 50 X
Ref. [5,29,30] Ref. [4,11,20] Ref. [20] Ref. [4,11,20] Ref. [20] Ref. [14,20] Ref. [20] Ref. [20]
IS cells X X 7.5 X X 7.5 7.5 450
(3000) X X 5–10 X X 5–10 5–10 100–800
Ref. [15,29,30] Ref. [15] Ref. [15] Ref. [15,19] Ref. [15] Ref. [19] Ref. [19] Ref. [15]

Notes: Cell numbers and connectivity values were estimated from published data for granule cells, mossy cells, basket cells, axo-axonic cells, molecular
layer interneurons with axonal projections to the perforant path (MOPP), hilar interneurons with axonal projections to the perforant path (HIPP), hilar
interneurons with axonal projections to the commissural/associational pathway (HICAP) and interneuron specific cells (IS). Connectivity is given as
number of connections to a postsynaptic population (row 1) from a single presynaptic neuron (column 1). The average number of connections used in the
1:1 structural model is given in bold. Note, however, that the small world structure (see Section ‘‘Structural alterations in the dentate during epile-
ptogenesis’’) was preserved even if only the extreme low or the extreme high estimates were used for the calculation of L and C. References given in table
correspond to: 1Gaarskjaer (1978), 2Boss et al. (1985), 3West (1990), 4Patton and McNaughton (1995), 5Freund and Buzsaki (1996), 6Buckmaster and
Dudek (1999), 7Acsady et al. (1998), 8Geiger et al. (1997), 9Blasco-Ibanez et al. (2000), 10Gulyas et al. (1992), 11Buckmaster and Jongen-Relo (1999),
12
Buckmaster et al. (1996), 13Wenzel et al. (1997), 14Han et al. (1993), 15Gulyas et al. (1996), 16Babb et al. (1988), 17Woodson et al. (1989), 18Halasy and
Somogyi (1993), 19Acsady et al. (2000), 20Sik et al. (1997), 21Bartos et al. (2001), 22Li et al. (1992), 23Ribak et al. (1985), 24Frotscher et al. (1991), 25Katona
et al. (1999), 26Soriano et al. (1990), 27Claiborne et al. (1990), 28Buckmaster et al. (2002a), 29Nomura et al. (1997a), 30Nomura et al. (1997b). Reproduced
with permission from Dyhrfjeld-Johnsen et al. (2007).

(Buckmaster et al., 1996, 2002a; Nomura et al., The third piece: connectivity between cell types
1997a, b). Buckmaster and Jongen-Relo (1999) es-
timated the total number of GABAergic neurons in In the past decade, a vast amount of high-quality
the molecular layer of the dentate gyrus as approx- data about the connectivity of the dentate gyrus
imately 10,000. An even distribution between inner, has been collected, and this data serves as an in-
medial, and outer molecular layers is assumed, valuable resource in constructing the cell-type spe-
yielding 4,000 MOPP cells with somata located in cific connectivity matrix for the dentate gyrus. The
the inner molecular layer (Han et al., 1993). Note connectivity for each cell type is summarized in
that this estimate excludes the molecular layer Table 1 and described below. Each value was de-
interneurons that project primarily outside of the termined by a detailed survey of the anatomical
dentate gyrus such as the outer molecular layer literature and was based on an assumption of uni-
interneurons projecting to the subiculum (Ceranik form bouton density along the axon of the presy-
et al., 1997). naptic cell. This assumption is in agreement with
643

Fig. 2. Gaussian fits to experimentally determined distributions of axonal branch length used in construction of the models of the
dentate gyrus. (A) Plot shows the averaged axonal distribution of 13 granule cells (Buckmaster and Dudek, 1999) and the corre-
sponding Gaussian fit. (B) Fit to the septo-temporal distribution of axonal lengths of a filled and reconstructed basket cell (Sik et al.,
1997). (C) Fit to the axonal distribution of a CA1 axo-axonic cell (Li et al., 1992). (D) Gaussian fit to the averaged axonal distributions
of three HIPP cells from gerbil (Buckmaster et al., 2002a, b). (E) Fit to averaged axonal distributions of three mossy cells illustrates the
characteristic bimodal pattern of distribution (Buckmaster et al., 1996). (F) Histogram of the axonal lengths of a HICAP cell along the
long axis of the dentate gyrus (Sik et al., 1997) and the Gaussian fit to the distribution. All distributions were based on axonal
reconstruction of cells filled in vivo. In all plots, the septal end of the dentate gyrus is on the left (indicated by negative coordinates) and
the soma is located at zero. Reproduced with permission from Dyhrfjeld-Johnsen et al. (2007).

the in vivo data of Sik et al. (1997), and it greatly with few collaterals (3%) in the GC layer (Buck-
simplifies the estimation of connectivity from an- master and Dudek, 1999). In addition to MCs
atomical data on axonal length and synapse den- (Acsady et al., 1998), mossy fibers have also been
sity per unit length of axon. shown to contact BCs (Buckmaster and Schwa-
rtzkroin, 1994; Geiger et al., 1997) and PV inter-
Granule cells neurons (Blasco-Ibanez et al., 2000). With a total
of 400–500 synaptic contacts made by a single
Mossy fibers (GC axons) in a healthy rat dentate mossy fiber (Acsady et al., 1998), the 3% of each
gyrus are primarily restricted to the hilus (97%), axon located in the GC layer (Buckmaster and
644

Dudek, 1999) is estimated to contact 15 BCs and 3 postsynaptic hilar interneuron receives two synap-
axo-axonic cells. In the hilus, a single GC forms tic contacts from a single mossy cell axon (Buck-
large, complex mossy fiber boutons that innervate master et al., 1996), each mossy cell is estimated to
7–12 MCs (Acsady et al., 1998), while an estimated contact 600 HIPP and 200 HICAP cells. There is
100–150 mossy fiber terminals target hilar inter- very low mossy cell to interneuron connectivity in
neurons with approximately one synapse per the inner molecular layer (Wenzel et al., 1997);
postsynaptic interneuron (Acsady et al., 1998). MCs could contact 5–10 basket and axo-axonic
Gulyas et al. (1992) estimated that a single spiny cells and approximately 5 MOPP cells with somata
calretinin-positive cell (presumed HIPP cell) is in the inner molecular layer (Han et al., 1993).
contacted by approximately 9,000 GCs. With
12,000 HIPP cells and 1,000,000 GCs, each GC
Basket cells
can be estimated to contact approximately 110
HIPP cells and 40 HICAP cells. Additionally, in
In the CA3 region of the rat hippocampus, each
agreement with the presence of mossy fiber termi-
principal cell is contacted by approximately 200
nals on aspiny calretinin-positive interneurons
BCs (Halasy and Somogyi, 1993), but a GC in the
(Acsady et al., 1998), 15 mossy fibers are expected
dentate could be contacted by as few as 30 BCs.
to synapse onto IS cells. Since mossy fibers avoid
Assuming that each of the 1,000,000 GCs is con-
the molecular layer (Buckmaster and Dudek,
tacted by 115 BCs each making 1–20 synaptic
1999) in the healthy dentate gyrus, it is assumed
connections (Halasy and Somogyi, 1993; Acsady
that they do not contact MOPP cells. However, in
et al., 2000), it can be estimated that each BC
animal models of temporal lobe epilepsy, sprouted
contacts approximately 1,250 GCs. MCs receive
mossy fibers have been shown to contact up to 500
10–15 basket cell synapses (Acsady et al., 2000),
postsynaptic GCs (Buckmaster et al., 2002b).
resulting in an estimate of 75 BC to mossy cell
synapses per BC. Approximately 1% of the 11,000
synapses made by a single basket cell axon in the
Mossy cells
GC layer are onto other BCs (Sik et al., 1997) with
3–7 synapses per postsynaptic cell (Bartos et al.,
A single filled mossy cell axon has been reported to
2001). Consequently, each BC in the dentate gyrus
make 35,000 synapses in the inner molecular layer
contacts approximately 35 other BCs. Since hilar
(Buckmaster et al., 1996; Wenzel et al., 1997). As-
and molecular layer interneurons are not a major
suming a single synapse per postsynaptic cell, a
target of basket cells (Halasy and Somogyi, 1993),
single mossy cell is estimated to contact
a single BC may contact 0–1 HIPP cells. Similarly,
30,000–35,000 GCs. Of the 2700 synapses made
the BC synapses onto axo-axonic cells, HICAP,
by a single MC axon in the hilus, approximately
and MOPP cells are assumed to be negligible. As
40% target GABA-negative neurons (Wenzel et
PV cells preferentially contact other PV-positive
al., 1997). As each mossy cell is estimated to make
cells in the hilus (Acsady et al., 2000), we assume
1–5 synaptic contacts on a single postsynaptic
that BCs do not contact calretinin-positive IS cells
mossy cell (Buckmaster et al., 1996), it is estimated
(Gulyas et al., 1992).
that each mossy cell contacts approximately 350
other MCs (this is likely to be a generous estimate
since it is based on the assumption that all GAD Axo-axonic cells
negative somata in the hilus represent MCs). The
remaining 60% of the hilar mossy cell axons target Most synapses made by axo-axonic cell axons are
GABA-positive cells (Buckmaster et al., 1996; thought to target GC axon initial segments
Wenzel et al., 1997), with no reports demonstrat- (Halasy and Somogyi, 1993). However, a small
ing mossy cell synapses onto IS cells. Assuming fraction of axon collaterals also descend into the
that there is no preferential target selectivity be- superficial and deep hilus (Han et al., 1993; Freund
tween HIPP and HICAP cells and that each and Buzsaki, 1996). In neocortex, an axo-axonic
645

cell makes 4–10 synapses on the axon intial seg- MOPP cells
ment of a postsynaptic cell (Li et al., 1992). With
22,000,000 estimated axon initial segment synapses MOPP cells target an estimated 7,500 GCs in the
in the GC layer (Halasy and Somogyi, 1993) and 4 rat dentate gyrus. While MOPP cell axons project
synapses per postsynaptic cell (based on the data in the horizontal axis to a similar extent as HIPP
from the neocortex from Li et al., 1992), each of cells, they show considerably less collateralization
the 2,000 axo-axonic cells are estimated to target (Han et al., 1993), resulting in an estimate of half
approximately 3,000 GCs. MCs receive axo-axonic as many synapses onto MOPP and HICAP cells as
cell input (Ribak et al., 1985), and with the com- HIPP cells make. As MOPP cell axons are re-
paratively small fraction of axons from axo-axonic stricted to the molecular layer (Han et al., 1993)
cells in the hilus (Han et al., 1993; Freund and and do not target the basal dendrites of BCs, they
Buzsaki, 1996), it can be estimated that axo-axonic are assumed to contact less than 1/10 the number
cells target a number of MCs equal to approxi- of BCs targeted by HIPP cells. Likewise, MOPP
mately 5% of their GC targets, which results cells with axons restricted to the outer and middle
in approximately 150 MCs. Since axo-axonic molecular layers (Han et al., 1993) would not tar-
cells primarily target the axon initial segment get the hilar dendrites of axo-axonic cells (Soriano
of non-GABAergic cells (Halasy and Somogyi, et al., 1990) or the axo-axonic cells with somata
1993; Freund and Buzsaki, 1996), it may be and proximal dendrites in the hilus (Han et al.,
assumed that these cells do not project to inter- 1993). It is estimated that MOPP cells only contact
neurons. 1–2 axo-axonic cells. As the MOPP cell axonal
arbors in the molecular layer (Han et al., 1993) do
not overlap with major parts of the dendritic ar-
HIPP cells borizations of MCs (Frotscher et al., 1991), HIPP
cells (Han et al., 1993; Sik et al., 1997; Katona et
HIPP cells have been estimated to contact about al., 1999) or IS cells (Gulyas et al., 1996), the con-
1,600 GCs and 450 BCs with 1–5 synapses per nectivity to these cells is negligible.
postsynaptic cell (Sik et al., 1997). MCs can have
one dendrite in the molecular layer (Amaral, 1978;
Ribak et al., 1985; Scharfman, 1991) which can be HICAP cells
targeted by HIPP cell axons, whereas GCs have two
primary dendrites (Claiborne et al., 1990; Lubke Sik et al. (1997) estimated that the septo-temporal
et al., 1998). Since the mossy cell to GC ratio is extent and bouton density of HICAP cell axons is
approximately 1:30, the mossy cell dendrites con- similar to the HIPP cell axons, whereas the esti-
stitute only approximately 1/60 of the targets for mated axonal length of HICAP cells is approxi-
HIPP cells. Increasing this fraction to approxi- mately half of the HIPP cell axonal length. Thus, it
mately 1/45 to account for the presence of a few is estimated that HICAP cells contact about half
HIPP cell contacts on MC in the hilus (Buckmaster the number of GCs contacted by HIPP cells.
et al., 2002a) results in an estimate that each HIPP However, since HICAP cells have an additional
cell contacts approximately 35 MCs. HIPP cell 3% of axon collaterals in the hilus (Sik et al.,
axonal divergence onto HICAP and MOPP cells in 1997), their number of postsynaptic MCs can be
the molecular layer can be assumed to be similar to assumed to be the same as that of the HIPP cells.
that of somatostatin-positive cells in CA1 (Katona HICAP cells are assumed to contact less than half
et al., 1999) and estimated to be 15 connections the number of BCs targeted by HIPP cells (175)
onto each population. The HIPP cell axonal diver- and a negligible number of axo-axonic cells. With
gence to axo-axonic cells is estimated to be between a total of 26,000 synapses from a single HICAP
the divergence to basket and HICAP cells; therefore cell axon (Sik et al., 1997), approximately 700
the HIPP cell axon likely contacts 30 axo-axonic synapses should be present in the hilus. Assuming
cells. 2–5 synapses per postsynaptic cell, each HICAP
646

cell could contact 100–300 hilar cells. Each HI- of modeling, this corresponds to a number of dis-
CAP cell is assumed to target 50 HIPP and HI- tinct obstacles. Single cell models must be created
CAP cells, which, along with 35 synapses on MCs, with sufficient detail to function appropriately in
is in the estimated range. Although the total ax- the network; network scaling must be implemented
onal length of HICAP cells is only about half of (i.e., what percentage of the actual cell numbers,
that of HIPP cells, the number of MOPP cells connections, etc., will be represented in the net-
targeted is assumed to be the same (10–20), as work); receptor types and synaptic data must be
the HICAP cell axons primarily project to the in- included; and network topology must be deter-
ner molecular layer where both cell bodies and mined (will the network be a strip or a ring, e.g.,
proximal dendrites of MOPP cells are located and how will connection probabilities be imple-
(Han et al., 1993). mented to account for the shape and size of the
network). Finally, the stimulation paradigm must
be determined so that the fully connected network
IS cells will actually start to produce data. These obsta-
cles, and ways to overcome them, are discussed in
IS cells contact an estimated 100–800 other IS cells detail in this section.
and 5–10 (presumably CCK positive) BCs (Gulyas
et al., 1996). Acsady et al. (2000) suggested that
CCK cells would include both basket cell and HI- Multicompartmental single cell models
CAP morphologies and that IS cells also project to
somatostatin-positive HIPP cells. These data result In a network model that proposes to simulate both
in an estimate that IS cells project to 5–10 HICAP somatic and dendritic inputs and model synaptic
cells and HIPP cells. contacts on specific dendritic compartments, it is
necessary to develop multicompartmental, biophys-
The fourth piece: spatial constraints and axonal ically realistic models of individual cell types. It is
distributions not always necessary to develop these models from
scratch, however, as ModelDB (http://sense-
In addition to the number of connections made by lab.med.yale.edu/senselab/modeldb/) is a search-
any given cell, the distribution of these connections able resource that makes published models of
along the septo-temporal axis of the dentate gyrus several cell types including dentate GCs available
is very important. Fortunately, in vivo single cell for download (Aradi and Holmes, 1999a; San-
fills have been performed which provide a very thakumar et al., 2005). These models can be readily
complete description of the axonal extent of the adapted to a wide variety of networks. Morpho-
eight cell types discussed above. Averages of these logical properties for multicompartmental models
in vivo axonal distributions can be fit with either a of other cell types not available on ModelDB, such
single or double Gaussian for each cell type, which as some of the other dentate cell types, can be
then defines the model axonal distribution for each developed based on the data reported in the liter-
cell of that type (Fig. 2). These Gaussians then ature (Buckmaster et al., 1993, 2002b; Freund and
describe the probability of any given cell connect- Buzsaki, 1996; Geiger et al., 1997; Bartos et al.,
ing to another cell based on the distance between 2001).
the two. The intrinsic properties of cell types can be
modeled based on data from whole cell physio-
logical experiments obtained in the presence of
Putting the pieces together blockers of synaptic activity (Lubke et al., 1998).
Detailed descriptions of how to model passive and
Once the border of the puzzle is complete (the active properties of individual cells can be found in
connection matrix in our analogy), it is possible to The Neuron Book (Carnevale and Hines, 2006), but
start assembling the bulk of the picture. In the case we will discuss a couple of relevant issues here
647

briefly. First, since GCs (Desmond and Levy, by BCs is considered crucial in maintaining weak
1985) and MCs (Amaral, 1978) are rich in dendri- GC activity in response to afferent input. The
tic spines, it is necessary to account for the mem- HIPP cells, on the other hand, synapse on the
brane area contribution of the spines (Rall et al., dendritic region near the afferent inputs, and they
1992). This can be accomplished by decreasing are likely to modulate integration of dendritic in-
membrane resistivity (increasing leak conduct- puts. The loss of this dendritic inhibition with an
ance) and increasing membrane capacitance essentially intact or increased somatic inhibition
(Aradi and Holmes, 1999b). An additional con- has been proposed to be particularly relevant to
sideration is the presence of spontaneous activity epileptogenesis (Cossart et al., 2001). Further-
in some cell types such as MCs (Ishizuka et al., more, BCs are resistant to cell death whereas HIPP
1995; Ratzliff et al., 2004), as well as their lower cells are extremely vulnerable in epileptic tissue
input resistance in the presence of background (Buckmaster et al., 2002a). These features make
synaptic activity (Ratzliff et al., 2004), compared the basket and HIPP cells essential constituents of
to the input resistance in ionotropic glutamate and the dentate model.
GABA receptor antagonists (Lubke et al., 1998). The remaining four interneurons also possess
Spontaneous firing rate can be simulated via a unique features of inhibition. The MOPP cells op-
constant direct current injection (Santhakumar erate by providing GCs with feed-forward dendri-
et al., 2005), and the synaptic background activity tic inhibition (Halasy and Somogyi, 1993). The
can be simulated as a fluctuating point conduct- HICAP cells synapse on the proximal dendrites of
ance with balanced excitation and inhibition, as GCs, near where mossy cell axons terminate and
described in Destexhe et al. (2001). provide feedback inhibition (Freund and Buzsaki,
Both excitatory cell types (GCs and MCs) are 1996). The axo-axonic cells provide GABAergic
well defined and have a vast amount of electro- input at the axon initial segments of GCs and MCs
physiological data available for the construction of and potentially play powerful roles in modifying
single cell models. However, the wealth of infor- spike initiation (Freund and Buzsaki, 1996; Ho-
mation available for the interneurons is substan- ward et al., 2005). The IS cells connect exclusively
tially less for many cell types. Additionally, the to other interneurons and form a well connected
role of interneurons in the network must be con- network that could modulate excitability and
sidered with regard to the question being asked. synchrony of the network.
For example, in a model network that is pursuing Unfortunately, detailed physiological data for
the question of the role of mossy fiber sprouting the development of realistic models of MOPP,
and hilar cell loss in epileptogenesis (Santhakumar HICAP, axo-axonic, and IS cells are not readily
et al., 2005; Dyhrfjeld-Johnsen et al., 2007), several available. Therefore, in an approach that avoids
factors must be considered. First, inhibition can errors of commission one could choose to include
operate in two principal modes. One mode is feed- only basket and HIPP cells in the network. How-
forward inhibition, where an interneuron is acti- ever, this makes it essential to perform control
vated by the same afferent input as the excitatory simulations to examine whether augmenting inhi-
neuron that it contacts. The basket cell is the pri- bition (in order to compensate for excluded cell
mary source of feed-forward inhibition in the types) influences the overall outcome of the net-
dentate gurus (Freund and Buzsaki, 1996). The work simulations.
second mode is feedback inhibition, where inter-
neurons are activated by excitatory cells which
they subsequently inhibit. The BCs and HIPP cells Network scaling
are the two key feedback inhibitory cells in the
dentate (Freund and Buzsaki, 1996; Buckmaster et Once the single cell models are complete and ready
al., 2002a). Another important aspect of inhibition to go, it is time to start thinking about how big the
is the location of synaptic contact between inter- network is going to be. With the immense in-
neuron and principal cell. Perisomatic inhibition creases in computational power over the last
648

couple of decades, very large networks are quite the size of a realistic dentate model can be scaled up
feasible. However, some questions may not require further to a 1:20 scale model with 50,000 GCs
such large networks, and often computational re- (Dyhrfjeld-Johnsen et al., 2007). The specific con-
sources will be limited. Also important in deter- nection parameters that can be used for a model of
mining what scale a network should be is whether this size are shown in Table 2. In Section ‘‘The big
the network includes functional model cells that picture,’’ we will discuss results obtained from this
require a great deal of computational power to very large-scale network with direct implications for
process. Purely structural network models (that do epileptogenesis.
not use multicompartmental functional single cell
models), also called network graphs, can provide a
great deal of useful information about neuronal Receptor types and synapses
circuits. Indeed, Dyhrfjeld-Johnsen et al. (2007)
created an extraordinarily large 1:1 scale structural Building a model of a dentate network with hun-
model of the dentate gyrus that yielded a great dreds or thousands of biophysically realistic multi-
deal of new data and will be discussed briefly in compartmental model cells is only a meaningful
Section ‘‘The big picture’’. venture if those cells have a way to talk to one an-
Putting purely structural graphs aside for now other. For this reason, we must consider neuro-
though, and focusing on the functional model that transmitter-receptor types that must be included in
we are building in this chapter, we can determine the network. Again, the types of receptors, synapses,
how to scale the cell numbers. We know that the and other cellular interaction devices (such as gap
GC:MC:BC:HIPP ratio is 1000:30:10:12. In agree- junctions) to include in the network will depend on
ment with these ratios, the Lytton et al. (1998) the particular question the network is designed to
model that focused on the network effects of answer. For the example network we have been
sprouting included 50 GCs, 2 MCs, and 2 inter- building throughout this chapter, the goal is to de-
neurons. Using this relatively small network repre- termine the effects of mossy fiber sprouting and hilar
sentation of the dentate they demonstrated that the cell loss on excitability. To examine these effects, the
effect of sprouting on enhancing network excitabil- minimal requirement is to incorporate ionotropic
ity was strongly limited by the intrinsic firing prop- glutamatergic AMPA synapses and GABAA
erties of GCs (e.g., strong spike frequency synapses. Experiments from both normal and epi-
adaptation). Although this model made the predic- leptic animals are constantly ongoing, and as the
tion that mossy fiber sprouting could lead to hyper- data become available it will be possible to add
excitability, the network was highly interconnected other receptors such as NMDA and metabotropic
as a consequence of the small size. This network receptors. It would also be interesting to incorporate
architecture, combined with a very strong perforant greater diversity at individual synapses through in-
path synaptic conductance, could have resulted in corporation of mechanisms such as short-term plas-
artificial synchronous activity in the network. In- ticity. Each addition to the network must be
creasing the size of the Lytton et al. (1998) model by carefully considered, however, as greater complex-
10-fold entails building a 1:2000 scale dentate with ity necessarily increases the load on the computa-
500 GCs, 15 MCs, 6 BCs and 6 HIPP cells. This tional resources. Since the basic circuit effects of
500+ cell model can be readily simulated on a sprouting and cell loss can be simulated by including
laptop and allows for manipulating hilar cell num- AMPA and GABAA synapses, these will be the
bers and connectivity patterns (Santhakumar et al., types included in our example model.
2005). However, even in a network of this size, ar- Postsynaptic conductances can be represented as
tificially large conductances as well as other factors a sum of two exponentials (Bartos et al., 2001).
(e.g., strong edge effects and unrealistic distance The peak conductance (gmax), rise and decay time
dependent cell-type specific connectivity) could in- constants, and the synaptic delays (distinct from
fluence the results. With the current computational the axonal conduction delay described later) for
power of state-of-the-art multiprocessor computers, each network connection can be estimated from
649

Table 2. Parameters of the 50,000+ neuron functional model network

Functional model network parameters

From To –> GC MC BC HC

Granule cellsa Convergence 68.03 78.05 370.95 2266.64


(50,000) Divergence 68.03 2.34 3.71 27.19
Synapse weight (nS) 1.00 0.20 0.94 0.10
Mossy cells Convergence 243.62 87.23 5.59 375.53
(1500) Divergence 8120.82 87.23 1.86 150.21
Synapse weight (nS) 0.30 0.50 0.30 0.20
Basket cells Convergence 3.11 6.31 8.98 n/a
(500) Divergence 313.22 18.93 8.98 n/a
Synapse weight (nS) 1.60 1.50 7.60 n/a
HIPP cells Convergence 4.82 3.76 140.13 n/a
(600) Divergence 401.86 9.39 116.77 n/a
Synapse weight (nS) 0.50 1.00 0.50 n/a

Perforant pathb Synapse weight (nS) 20 17.5 10 n/a

Notes: The cell numbers (column 1) and synaptic connectivity values and strengths in the functional model network used for the activity simulations in
Fig. 4. Note that this network is smaller (50,000+ cells) than the full-scale dentate gyrus (over 1 million cells) model described in Section ‘‘Structural
alterations in the dentate during epileptogenesis’’. Therefore, the connectivity had to be adjusted from what is shown in Table 1. Convergence is given as
number of connections onto a single postsynaptic neuron (row 1) from a presynaptic neuronal population (column 1). For example, 243 mossy cells
converge on a single granule cell in this network. Divergence is given as the number of connections to a postsynaptic population (row 1) from a single
presynaptic neuron (column 1). For example, a single mossy cell makes synapses on 8120 postsynaptic granule cells in this network. The strengths of the
connections are given in nS. For example, the strength of the excitatory synapse formed by a single mossy cell on a single granule cell is 0.3 nS.
a
Granule cell to granule cell connections represent values at 100% sprouting.
b
Perforant path input to 5000 granule cells (2 synapses each), 50 basket cells (2 synapses each) and 10 mossy cells (1 synapse each) in the central 10 bins of
the network model. Reproduced with permission from Dyhrfjeld-Johnsen et al. (2007).

experimental data (Kneisler and Dingledine, 1995; Network topology


Geiger et al., 1997; Bartos et al., 2001; Santhaku-
mar et al., 2005) and details of all of the specific The shape of the network and the way in which
cellular and synaptic properties can be found in connections are made depend strongly on the size
Santhakumar et al. (2005). When making such es- of the network. For instance, a network of ap-
timates, care must be taken to ensure that compa- proximately 50 or 500 cells (Lytton et al., 1998;
rable conditions were used in the experiments that Santhakumar et al., 2005, respectively) may re-
provide the data. For example, since the data for quire a ring structure to avoid edge effects (i.e.,
synaptic conductances is typically obtained by cells on the edge of a network have as few as half
paired recording experiments from several differ- as many pre- and postsynaptic targets compared
ent groups, the temperature at which the record- to cells in the middle). A network of 50,000+
ings were performed must be accounted for. Q10 cells (Dyhrfjeld-Johnsen et al., 2007), on the other
estimates can be used to convert kinetic data from hand can be set up as a linear strip, more similar
room temperature recordings to the appropriate to the actual topology of the biological dentate.
values at physiological temperature. Axons can be Similarly, small networks may require non-topo-
modeled implicitly by activating a given synapse graphic connectivity. That is, the postsynaptic
after either a static or distance-dependent delay targets of each cell are selected at random from
once a presynaptic cell crosses a preset membrane the pool of potential target neurons while main-
potential threshold (Bartos et al., 2001). taining the cell type specific divergence and
650

convergence. Larger networks can incorporate The big picture


the spatial rules discussed above in Section 7 The
fourth piece: Spatial constraints and axonal dis- Following essentially the strategy outlined in the
tributions; the connectivity between cells can be previous two sections, dentate gyrus modelers
distributed according to the experimentally de- have assembled a number of different puzzles,
rived axonal distributions of the various cell types producing dentate models of a wide range of sizes
(Fig. 2). aimed at answering a large number of important
questions. We have walked through the creation of
a very large-scale dentate model capable of han-
Network stimulation dling functional network simulations with greater
than 50,000 cells. However, we have not yet dis-
A network such as the dentate gyrus, in which cussed how a model of this type can help us gain a
most cells are not designed to be spontaneously better understanding of disease processes and
active, requires some afferent input to initiate net- hopefully how to treat them. Without the ability
work activity. This input can come in a number of to provide information of this nature, models are
forms, again dependent on the specific question as simply incomplete representations of reality and of
well as the particular network being modeled. For little use. Fortunately, models can provide a great
the dentate gyrus as a whole, it makes sense to deal of information because they give scientists
simulate a perforant path input from the ent- complete control over variables that cannot be
orhinal cortex. This input to the network is located isolated in typical experimental situations. In this
on both dendrites of all GCs and the apical dend- section, we will briefly describe an example in
rites of all BCs. Since 15% of MCs have also been which using large scale dentate gyrus modeling al-
shown to receive direct perforant path input lowed Dyhrfjeld-Johnsen et al. (2007) to isolate
(Scharfman, 1991; Buckmaster et al., 1992), the structural changes in the dentate during epilepto-
number of MCs that should receive input in a genesis and conclude that structural changes alone
given model can be easily calculated. Mass stim- can lead to hyperexcitability.
ulation of the perforant path can be simulated by
activating a maximum peak AMPA conductance
in cells postsynaptic to the stimulation (Santhaku- Structural alterations in the dentate during
mar et al., 2000). The strength of the perforant epileptogenesis
path input in the 50,000+ cell functional model
from Dyhrfjeld-Johnsen et al. (2007) is given in As mentioned throughout this chapter, two pri-
Table 2 (note the very large synaptic weight for mary structural alterations occur in the dentate
perforant path input compared to the small gyrus during epileptogenesis: mossy fiber sprout-
weights for cell to cell synapses). In the biological ing and hilar cell death (including hilar interneu-
network and in physiological studies it is unlikely rons and MC). These two processes (which we will
that the entire network is activated simultane- refer to simply as ‘‘sclerosis,’’ defined here as the
ously. Rather, focal activation can be simulated by two common structural changes in the dentate as-
activating the input to GCs and BCs located in a sociated with mesial temporal lobe sclerosis) can
model hippocampal ‘‘lamella’’. If focal network be implemented in a dentate gyrus model accord-
activation is undesirable, spontaneous network ac- ing to the data driven approach given above. De-
tivity can be simulated instead. This can be ac- tailed descriptions of how to implement both of
complished by providing uncorrelated activation these processes can be found in Santhakumar et al.
(perforant path inputs with Poisson distributed (2005) and Dyhrfjeld-Johnsen et al. (2007). Briefly,
interspike intervals) to each granule cell, basket a maximal level of sclerosis can be simulated by
cell, and 15% of the MCs. Other methods for in- adding an average of approximately 275 granule
itiation of network activity can be devised as nec- cell to granule cell connections (Buckmaster et al.,
essary. 2002b) and by removing all hilar cells from the
651

network. Using this maximum value as a point of that utilizes this parameter is that anyone on the
reference, the severity of sclerosis can be approx- planet is separated from anyone else by up to six
imated by simulating percentages of the maximum degrees of separation, or six people. While this
(e.g., at 50% sclerosis, simulating epileptogenesis theory has never been satisfactorily proven, it has
following a moderate head injury (Howard et al., led to interesting experiments and even a party
2007), each granule cell would contact approxi- game called the Six Degrees of Kevin Bacon, in
mately 137 other GCs, and 50% of hilar cells which the participants have to connect any actor
would be lost). or actress to Kevin Bacon through their film roles
A 1:1 scale structural model of the dentate gyrus in as few steps as possible.
can be constructed where each cell is simply a node Based on these two topological measures, net-
in the computer (nodes have no functional prop- works can be divided into three classes: (1) Reg-
erties and are therefore much less computationally ular, with a long path length and large clustering
demanding than the multicompartmental cell coefficient; (2) Random, with a short path length
models described previously) and synapses be- and small clustering coefficient; and (3) Small
tween cells are represented as non-functioning di- World, with a short path length and large cluster-
rected links (Dyhrfjeld-Johnsen et al., 2007). This ing coefficient. Representative graphs of these
structural network can then be utilized to under- types of networks are displayed in Fig. 3.
stand the basic structure of the healthy dentate As shown, the regular network (Fig. 3A) is
gyrus and how that structure changes during epile- characterized by a high local connectivity, but the
ptogenesis. distance (via connected nodes) between any two
In order to understand how network structure random nodes, especially nodes at the ends of the
can be measured and later how structural changes graph, can be quite large. On the other hand, the
have functional implications, it is necessary to ex- random network (Fig. 3B) contains many long
plain the parameters that define network topology. distance connections, contributing to a very short
Two measures that Watts and Strogatz (1998) average path length. There is little local connec-
originally employed to assess the structure of the tivity however, as nodes are just as likely to con-
nervous system of the worm C. elegans can be used nect to far away nodes as they are to their
to analyze the structure of the dentate network: the neighbors. The small world network (Fig. 3C) acts
average path length, L, and the average clustering as a sort of compromise between the random and
coefficient, C. The average path length is defined regular graphs. Indeed, there is a high degree of
as the average number of steps required to move local connectivity combined with a number of
from any node to any other node in the network, long-range connections, yielding a both locally
and it is therefore a measure of how well connected and globally well-connected network. Small world
the network is globally. The average clustering networks can also be thought of in terms of social
coefficient, on the other hand, is a measure of local networking. Humans often form a cohesive group
connectivity. It is defined as the fraction of all of friends who are well connected locally due to
possible connections between ‘‘postsynaptic’’ various constraints such as geography, common
nodes of a given node that are actually formed. activities and so on. Additionally, any given per-
The average clustering coefficient for an entire son in the group may know some people who are
network is simply the average of the clustering totally unrelated to the local group, but who each
coefficients for each node in the network. have cohesive local groups of their own, thus cre-
The topological measures, C and L, can be best ating numerous long distance connections between
thought of in terms of a social network, or a net- the locally well-connected networks.
work of friends. C measures the probability that So how do these L and C values play a role in
any two friends of a given person also know each understanding the dentate gyrus? Analysis of the
other. L measures how many steps in a chain of structural model of the dentate gyrus in the healthy
friends or acquaintances would be necessary to state, without any mossy fiber sprouting or hilar cell
connect two random people. A common theory loss, indicates that the dentate gyrus is a small
652

Fig. 3. Schematics of three basic network topologies. (A) Regular network topology. The nodes in a regular network are connected to
their nearest neighbors, resulting in a high degree of local interconnectedness (high clustering coefficient C), but also requiring a large
number of steps to reach other nodes in the network from a given starting point (high average path length L). (B) Random network
topology. In a random network, there is no spatial restriction on the connectivity of the individual nodes, resulting in a network with a
low average path length L but also a low clustering coefficient C. (C) Small world network topology. Reconnection of even a few of the
local connections in a regular network to distal nodes in a random manner results in the emergence of a small world network with a
high clustering coefficient C but a low average path length L. Reproduced with permission from Dyhrfjeld-Johnsen et al. (2007).

world network. It has a very high clustering coeffi- of nodes lost due to hilar cell death is only 4.5% at
cient and a path length that is only 2.68, approx- maximal sclerosis, yet the network becomes seem-
imately the same as the path length of the C. ingly more locally and globally well connected. This
elegans nervous system, a nervous system with only finding predicts that during epileptogenesis, the
282 cells! This finding means that any cell is con- dentate gyrus may become more readily able to
nected to any of the other one million cells in the transmit information throughout the network and
dentate network by less than three synapses, and likely increase synchronous firing as well, phenom-
cells are highly connected locally, resulting in fast ena that could very well contribute to a seizure
local computations and the ability to efficiently re- phenotype. It is only at nearly 100% sclerosis that
lay signals to distant parts of the network. the dentate gyrus network transforms into a more
While the analysis of the healthy dentate gyrus regular network structure, predicting a counterin-
provides useful and novel information about its tuitive decrease in hyperexcitability at maximal
structure, perhaps the most interesting and coun- sclerosis.
terintuitive structural information comes from the
analysis of networks undergoing sclerosis. These
networks demonstrate that the small world charac- Functional implications of structural alterations
teristics of the dentate gyrus are actually enhanced
up to approximately 80% sclerosis, despite a mas- Now that we have discussed the structural analysis
sive loss of connections (see Section ‘‘Mossy cells’’ of the healthy and sclerotic dentate gyrus, it re-
for MC connectivity; note the massive divergence mains to be seen whether network activity is ac-
onto GCs that is lost as sclerosis progresses and tually influenced by the structural changes in the
MCs die). In fact, the total number of links in the ways we might expect based on the predictions in
network decreases by 74% while the total number the preceding section. As in the 1:1 scale structural
653

model, sclerosis can be implemented in a func- transforms into a pattern with distinct waves of ac-
tional model of the dentate gyrus. This model tivity (from 80 to 100% sclerosis; Fig. 4E, F) that
contains approximately 50,000 cells, owing to the can collide and mutually annihilate (Netoff et al.,
increased need for computational resources to 2004; Roxin et al., 2004).
perform the calculations necessary with multicom- These findings not only support the important
partmental biophysically realistic cells (see Section role of structural alterations in determining func-
‘‘Multicompartmental single cell models’’). Simu- tional network activity, but they are in agreement
lations can be performed as sclerosis progresses in with experimental observations in both epileptic
steps (e.g., perform simulations at 20 then 40%, animals and humans. Indeed, no studies that have
and so on), and activity in the network can be quantified hilar cell loss in animal models of epi-
quantified by a number of different measures. lepsy ever reported 100% cell loss (Cavazos and
Four different measures were used in Dyhrfjeld- Sutula, 1990; Cavazos et al., 1994; Leite et al., 1996;
Johnsen et al. (2007): (1) duration of granule cell Buckmaster and Dudek, 1997; Mathern et al., 1997;
firing (the time of the last granule cell action po- Buckmaster and Jongen-Relo, 1999; Gorter et al.,
tential in the network minus the time of the first 2001; van Vliet et al., 2004; Zappone and Sloviter,
granule cell action potential); (2) average number 2004). Additionally, in surgically removed speci-
of action potentials per granule cell; (3) time until mens from pharmacologically intractable human
activity of the most distant GCs from the stimu- temporal lobe epilepsy patients (Gabriel et al.,
lation point (i.e., latency to full network activity); 2004), cell counts showed that only approximately
and (4) synchrony of granule cell action potentials. 80% of hilar cells were lost, even in patients with
Results of the functional simulations reveal that severe sclerosis (Blumcke et al., 2000). This finding
network activity very closely parallels structural al- coincides perfectly with the period of maximal
terations in the dentate gyrus. The healthy non-scle- hyperexcitability in the functional model networks,
rotic dentate model shows very limited firing in when sclerosis reaches 80%.
response to a single perforant path input (Fig. 4A),
in accordance with the biological data (Santhaku-
mar et al., 2001). Additionally, as sclerosis pro- Understanding the limitations of models
gresses, the functional network becomes increasingly
hyperexcitable up to approximately 80% sclerosis It is important to consider that no model can ever
(Figs 4B–E), again agreeing with in vitro measures fully replicate the system that it is modeling. As a
of epileptiform activity (Rafiq et al., 1995) and in result, it is crucial to be certain that results and
accordance with the enhanced small world features conclusions derived from models are robust under
revealed by the structural analysis. A very interest- a number of potentially confounding situations.
ing phenomenon then occurs at levels of sclerosis The modeling strategy presented in this chapter
exceeding 80%. Hyperexcitability of the functional attempts to ensure that available biological data is
model network actually decreases (Fig. 4F)! The represented and accounted for. However, a num-
level of sclerosis at which hyperexcitability dimin- ber of specific components of the dentate circuit
ishes corresponds perfectly to the point at which the and connectivity often cannot be modeled due to
network transforms from a small world structure to lack of precise data. Additionally, depending on
a more regular network. Thus, the functional net- the particular question the model is designed to
work follows the same pattern that is seen in the answer, several factors may be purposely omitted
structural network, and the functional effects are in order to best focus on the relevant issues. For
solely due to the structural network alterations, as the model presented in this section, for example,
no intrinsic cellular or synaptic properties are altered changes to intrinsic cellular and synaptic proper-
in the model as sclerosis progresses. Interestingly, ties during sclerosis were omitted since the exper-
network dynamics are also affected by structural iments were designed to study structural
changes; a relatively uniform pattern of granule cell alterations in isolation. For these reasons, an ex-
activity (from 40 to 60% sclerosis; Fig. 4C, D) tensive series of control simulations must be
654

Fig. 4. Effects of the sclerosis-related topological changes on granule cell activity in functional model networks. A–F. Raster plots of
the first 300 ms of action potential discharges of granule cells in the 50,000+ neuron functional model network (granule cells number
1–50,000, plotted on the y-axis; each dot corresponds to an action potential fired by a granule cell) at increasing degrees of sclerosis.
Network activity was initiated by a single stimulation of the perforant path input to granule cells number 22,500–27,499 and to 10
mossy cells and 50 basket cells (distributed in the same area as the stimulated granule cells) at t ¼ 5 ms (as in Santhakumar et al., 2005).
Note that the most pronounced hyperactivity was observed at sub-maximal (80%) sclerosis. Reproduced with permission from
Dyhrfjeld-Johnsen et al. (2007).

performed to test the effects of a wide variety of 7) Implementation of a bilateral model of the
conditions that are not well constrained by the dentate gyrus including commissural projec-
available experimental data. Controls will vary tions.
from model to model, but many are needed in any
situation. Those performed for the structural All of the controls employed for the structural
model presented here were the following: model were also used as controls for the functional
model. In addition, the functional model had its
1) Variations in cell numbers. own set of controls, including:
2) Variations in connectivity estimates.
3) Inhomogeneous distribution of neuron den- 1) Double inhibitory synaptic strengths.
sities along the septo-temporal axis. 2) Axonal conduction delays.
4) Inhomogeneity in connectivity along the 3) Spontaneous instead of stimulation-evoked
transverse axis. activity.
5) Altered axonal distributions at the septal and
temporal poles (the anatomical boundaries of All of the results from these control simulations
the dentate gyrus). supported the basic findings of the original mod-
6) Offset degrees of sprouting and hilar neuron els, thus greatly strengthening the primary conclu-
loss. sions.
655

Tackling future puzzles and problems (on average, 5–6 times the average number of con-
nections) could serve as hubs and strongly promote
Modeling has come a long way in the last several hyperexcitability within the sclerotic dentate gyrus
decades. Even 10 years ago, large-scale realistic network. Interestingly, this finding is supported by
modeling was nearly impossible due to computa- the presence, in epileptic rats, of a small percentage
tional restrictions and the requirement of vast su- of GCs that have basal dendrites (Spigelman et al.,
percomputers to perform necessary calculations. 1998) and which receive a vast number of addi-
Today, however, it is possible to perform quite in- tional mossy fiber contacts (Buckmaster and Thind,
depth modeling studies on a home desktop. Models 2005). Computational modeling studies provide the
such as the 50,000 cell network in Dyhrfjeld-John- best way of isolating structural changes from in-
sen et al. (2007) can be realistically run on a dual trinsic cellular and synaptic changes that occur si-
processor system (albeit with 16–32 GB of RAM). multaneously in virtually all pathological states.
Smaller networks such as the 500 cell model in Large-scale simulations will also make it possi-
Santhakumar et al. (2005) can be run in a few min- ble to hold structure constant and test the circuit
utes on a typical laptop. These are wonderful ad- effects of specific therapeutic interventions,
vancements, as the development of large-scale thereby allowing fast, cheap, efficient, and specific
models that can simulate brain circuits with greater analysis of treatments for a variety of pathological
realism is critical to our understanding of the ex- processes. The fast pace at which scientists are
perimentally determined cellular and molecular un- gleaning knowledge about the dentate from animal
derpinnings of diseases like epilepsy. Additionally, models will allow for inclusion of a complete com-
large-scale realistic networks are necessary for our plement of interneuronal subtypes, various recep-
understanding of structure–function relationships, tor and modulatory systems, realistic channel
especially in a highly plastic system like the dentate distributions, dendritic integration processes, and
gyrus. synaptic plasticity into computational models.
The exact structural changes that occur in many These additions will greatly broaden the scope of
pathological states such as epilepsy are as yet un- questions that a model can address and the
known. For example, recent evidence indicates that model’s predictive power. In the coming years, re-
local connection probability in various brain areas alistic modeling is only going to become more im-
may be modified by intra-class correlations (Yos- portant in bridging the gaps between animals and
himura and Callaway, 2005; Yoshimura et al., humans, molecules and behavior.
2005) and over-representation of small network
motifs (Milo et al., 2002; Reigl et al., 2004; Sporns
and Kotter, 2004; Song et al., 2005). While no di- Acknowledgements
rect evidence is present to implicate such factors in
the dentate gyrus, it is likely that unknown struc- The authors would like to thank Dr. Jonas
tural alterations have a significant impact on net- Dyhrfjeld-Johnsen for his contributions to the re-
work activity. Computational modeling studies search discussed in this chapter and his helpful
provide an excellent tool for determining what the comments on the chapter. This work was funded by
probable result of a wide variety of structural al- NIH grant NS35915 to IS and UCI MSTP to RM.
terations could be. Subsequent animal model ex-
periments can then attempt to visualize the
structural alterations predicted by the computation References
models. In an example of this type of approach,
Morgan and Soltesz (2006) have studied probable Acsady, L., Kamondi, A., Sik, A., Freund, T. and Buzsaki, G.
(1998) GABAergic cells are the major postsynaptic targets of
connectivity patterns within the recurrent GC net-
mossy fibers in the rat hippocampus. J. Neurosci., 18:
work that results from mossy fiber sprouting. Their 3386–3403.
results indicate that the presence of a small number Acsady, L., Katona, I., Martinez-Guijarro, F.J., Buzsaki, G.
(at most 5%) of GCs with increased connectivity and Freund, T.F. (2000) Unusual target selectivity of
656

perisomatic inhibitory cells in the hilar region of the rat hip- Buckmaster, P.S., Yamawaki, R. and Zhang, G.F. (2002a)
pocampus. J. Neurosci., 20: 6907–6919. Axon arbors and synaptic connections of a vulnerable pop-
Amaral, D.G. (1978) A Golgi study of cell types in the hilar ulation of interneurons in the dentate gyrus in vivo. J. Comp.
region of the hippocampus in the rat. J. Comp. Neurol., 182: Neurol., 445: 360–373.
851–914. Buckmaster, P.S., Zhang, G.F. and Yamawaki, R. (2002b)
Aradi, I. and Holmes, W.R. (1999a) Role of multiple calcium Axon sprouting in a model of temporal lobe epilepsy creates
and calcium-dependent conductances in regulation of hippo- a predominantly excitatory feedback circuit. J. Neurosci., 22:
campal dentate granule cell excitability. J. Comput. Neuro- 6650–6658.
sci., 6: 215–235. Carnevale, N.T. and Hines, M.L. (2006) The Neuron Book.
Aradi, I. and Holmes, W.R. (1999b) Active dendrites regulate Cambridge University Press, Cambridge, UK.
spatio-temporal synaptic integration in hippocampal dentate Cavazos, J.E., Das, I. and Sutula, T.P. (1994) Neuronal loss
granule cells. Neurocomputing, 26–27: 45–51. induced in limbic pathways by kindling: evidence for induc-
Babb, T.L., Pretorius, J.K., Kupfer, W.R. and Brown, W.J. tion of hippocampal sclerosis by repeated brief seizures. J.
(1988) Distribution of glutamate-decarboxylase-immunore- Neurosci., 14: 3106–3121.
active neurons and synapses in the rat and monkey hippo- Cavazos, J.E. and Sutula, T.P. (1990) Progressive neuronal loss
campus: light and electron-microscopy. J. Comp. Neurol., induced by kindling: a possible mechanism for mossy fiber
278: 121–138. synaptic reorganization and hippocampal sclerosis. Brain
Bartos, M., Vida, I., Frotscher, M., Geiger, J.R.P. and Jonas, P. Res., 527: 1–6.
(2001) Rapid signaling at inhibitory synapses in a dentate Ceranik, K., Bender, R., Geiger, J.R.P., Monyer, H., Jonas, P.,
gyrus interneuron network. J. Neurosci., 21: 2687–2698. Frotscher, M. and Lubke, J. (1997) A novel type of GAB-
Blasco-Ibanez, J.M., Martinez-Guijarro, F.J. and Freund, T.F. Aergic interneuron connecting the input and the output re-
(2000) Recurrent mossy fibers preferentially innervate parv- gions of the hippocampus. J. Neurosci., 17: 5380–5394.
albumin-immunoreactive interneurons in the granule cell Claiborne, B.J., Amaral, D.G. and Cowan, W.M. (1990) Quan-
layer of the rat dentate gyrus. Neuroreport, 11: 3219–3225. titative, three-dimensional analysis of granule cell dendrites
Blumcke, I., Suter, B., Behle, K., Kuhn, R., Schramm, J., Elger, in the rat dentate gyrus. J. Comp. Neurol., 302: 206–219.
C.E. and Wiestler, O.D. (2000) Loss of hilar mossy cells in Cossart, R., Dinocourt, C., Hirsch, J.C., Merchan-Perez, A.,
Ammon’s horn sclerosis. Epilepsia 41 Suppl 6: S174–180. De Felipe, J., Ben-Ari, Y., Esclapez, M. and Bernard, C.
Boss, B.D., Peterson, G.M. and Cowan, W.M. (1985) On the (2001) Dendritic but not somatic GABAergic inhibition is
number of neurons in the dentate gyrus of the rat. Brain Res., decreased in experimental epilepsy. Nat. Neurosci., 4: 52–62.
338: 144–150. Desmond, N.L. and Levy, W.B. (1985) Granule cell dendritic
Buckmaster, P.S. and Dudek, F.E. (1997) Network properties spine density in the rat hippocampus varies with spine shape
of the dentate gyrus in epileptic rats with hilar neuron loss and location. Neurosci. Lett., 54: 219–224.
and granule cell axon reorganization. J. Neurophysiol., 77: Destexhe, A., Rudolph, M., Fellous, J.M. and Sejnowski, T.J.
2685–2696. (2001) Fluctuating synaptic conductances recreate in vivo-
Buckmaster, P.S. and Dudek, F.E. (1999) In vivo intracellular like activity in neocortical neurons. Neuroscience, 107:
analysis of granule cell axon reorganization in epileptic rats. 13–24.
J. Neurophysiol., 81: 712–721. Dyhrfield-Johnsen, J., Santhakumar, V., Morgan, R.J., Huerta,
Buckmaster, P.S. and Jongen-Relo, A.L. (1999) Highly specific R., Tsimring, L. and Soltesz, I. (2007) Topological determi-
neuron loss preserves lateral inhibitory circuits in the dentate nants of epileptogenesis in large-scale structural and func-
gyrus of kainate-induced epileptic rats. J. Neurosci., 19: tional models of the dentate gyrus derived from experimental
9519–9529. data. J. Neurophysiol., 97(2): 1566–1587.
Buckmaster, P.S., Strowbridge, B.W., Kunkel, D.D., Schmiege, Freund, T.F. and Buzsaki, G. (1996) Interneurons of the hip-
D.L. and Schwartzkroin, P.A. (1992) Mossy cell axonal pro- pocampus. Hippocampus, 6: 347–470.
jections to the dentate gyrus molecular layer in the rat hip- Frotscher, M., Seress, L., Schwerdtfeger, W.K. and Buhl, E.
pocampal slice. Hippocampus, 2: 349–362. (1991) The mossy cells of the fascia dentata: a comparative
Buckmaster, P.S., Strowbridge, B.W. and Schwartzkroin, P.A. study of their fine structure and synaptic connections in ro-
(1993) A comparison of rat hippocampal mossy cells and dents and primates. J. Comp. Neurol., 312: 145–163.
CA3c pyramidal cells. J. Neurophysiol., 70: 1281–1299. Gaarskjaer, F.B. (1978) Organization of mossy fiber system of
Buckmaster, P.S. and Schwartzkroin, P.A. (1994) Hippocampal rat studied in extended hippocampi.1. Terminal area related
mossy cell function: a speculative view. Hippocampus, 4: to number of granule and pyramidal cells. J. Comp. Neurol.,
393–402. 178: 49–71.
Buckmaster, P.S. and Thind, K. (2005) American epilepsy so- Gabriel, S., Njunting, M., Pomper, J.K., Merschhemke, M.,
ciety meeting abstract. Epilepsia, 46: 91–131. Sanabria, E.R.G., Eilers, A., Kivi, A., Zeller, M., Meencke,
Buckmaster, P.S., Wenzel, H.J., Kunkel, D.D. and Schwa- H.-J., Cavalheiro, E.A., Heinemann, U. and Lehmann, T.-N.
rtzkroin, P.A. (1996) Axon arbors and synaptic connections of (2004) Stimulus and potassium-induced epileptiform activity
hippocampal mossy cells in the rat in vivo. J. Comp. Neurol., in the human dentate gyrus from patients with and without
366: 270–292. hippocampal sclerosis. J. Neurosci., 24: 10416–10430.
657

Gorter, J.A., van Vliet, E.A., Aronica, E. and Lopes da Silva, Lubke, J., Frotscher, M. and Spruston, N. (1998) Specialized
F.H. (2001) Progression of spontaneous seizures after electrophysiological properties of anatomically identified
status epilepticus is associated with mossy fibre sprouting neurons in the hilar region of the rat fascia dentata. J. Ne-
and extensive bilateral loss of hilar parvalbumin and som- urophysiol., 79: 1518–1534.
atostatin-immunoreactive neurons. Eur. J. Neurosci., 13: Lytton, W.W., Hellman, K.M. and Sutula, T.P. (1998) Com-
657–669. puter models of hippocampal circuit changes of the kindling
Geiger, J.R.P., Lubke, J., Roth, A., Frotscher, M. and Jonas, P. model of epilepsy. Artif. Intell. Med., 13: 81–97.
(1997) Submillisecond AMPA receptor-mediated signaling at Mathern, G.W., Bertram III, E.H., Babb, T.L., Pretorius, J.K.,
a principal neuron-interneuron synapse. Neuron, 18: Kuhlman, P.A., Spradlin, S. and Mendoza, D. (1997) In
1009–1023. contrast to kindled seizures, the frequency of spontaneous
Gulyas, A.I., Hajos, N. and Freund, T.F. (1996) Interneurons epilepsy in the limbic status model correlates with greater
containing calretinin are specialized to control other inter- aberrant fascia dentata excitatory and inhibitory axon
neurons in the rat hippocampus. J. Neurosci., 16: 3397–3411. sprouting, and increased staining for N-methyl-D-aspartate,
Gulyas, A.I., Miettinen, R., Jacobowitz, D.M. and Freund, AMPA and GABA(A) receptors. Neuroscience, 77:
T.F. (1992) Calretinin is present in nonpyramidal cells of the 1003–1019.
rat hippocampus .1. A new type of neuron specifically asso- Milo, R., Shen-Orr, S., Itzkovitz, S., Kashtan, N., Chklovskii,
ciated with the mossy fiber system. Neuroscience, 48: 1–27. D. and Alon, U. (2002) Network motifs: simple building
Halasy, K. and Somogyi, P. (1993) Subdivisions in the multiple blocks of complex networks. Science, 298: 824–827.
gabaergic innervation of granule cells in the dentate gyrus of Morgan, R.J. and Soltesz, I. (2006) American epilepsy society
the rat hippocampus. Eur. J. Neurosci., 5: 411–429. meeting abstract. Epilepsia, 47: 18–28.
Han, Z.S., Buhl, E.H., Lorinczi, Z. and Somogyi, P. (1993) A Netoff, T.I., Clewley, R., Arno, S., Keck, T. and White, J.A.
high-degree of spatial selectivity in the axonal and dendritic (2004) Epilepsy in small-world networks. J. Neurosci., 24:
domains of physiologically identified local-circuit neurons in 8075–8083.
the dentate gyrus of the rat hippocampus. Eur. J. Neurosci., Nomura, T., Fukuda, T., Aika, Y., Heizmann, C.W., Emson,
5: 395–410. P.C., Kobayashi, T. and Kosaka, T. (1997a) Laminar distri-
Heinemann, U., Beck, H., Dreier, J.P., Ficker, E., Stabel, J. and bution of non-principal neurons in the rat hippocampus, with
Zhang, C.L. (1992) The dentate gyrus as a regulated gate for special reference to their compositional difference among
the propagation of epileptiform activity. Epilepsy Res. Sup- layers. Brain Res., 764: 197–204.
pl., 7: 273–280. Nomura, T., Fukuda, T., Aika, Y., Heizmann, C.W., Emson,
Howard, A., Tamas, G. and Soltesz, I. (2005) Lighting the P.C., Kobayashi, T. and Kosaka, T. (1997b) Distribution of
chandelier: new vistas for axo-axonic cells. Trends Neurosci., nonprincipal neurons in the rat hippocampus, with special
28: 310–316. reference to their dorsoventral difference. Brain Res., 751:
Howard, A.L., Neu, A., Morgan, R.J., Echegoyen, J.C. and 64–80.
Soltesz, I. (2007) Opposing Modifications in Intrinsic Cur- Patton, P.E. and Mcnaughton, B. (1995) Connection matrix of
rents and Synaptic Inputs in Post-Traumatic Mossy Cells: the hippocampal-formation .1. The dentate gyrus. Hippo-
Evidence for Single-Cell Homeostasis in a Hyperexcitable campus, 5: 245–286.
Network. J. Neurophysiol., 97(3): 2394–2409. Rafiq, A., Zhang, Y.F., DeLorenzo, R.J. and Coulter, D.A.
Ishizuka, N., Cowan, W.M. and Amaral, D.G. (1995) A quan- (1995) Long-duration self-sustained epileptiform activity in
titative analysis of the dendritic organization of pyramidal the hippocampal-parahippocampal slice: a model of status
cells in the rat hippocampus. J. Comp. Neurol., 362: 17–45. epilepticus. J. Neurophysiol., 74: 2028–2042.
Katona, I., Acsady, L. and Freund, T.F. (1999) Postsynaptic Rall, W., Burke, R.E., Holmes, W.R., Jack, J.J., Redman, S.J.
targets of somatostatin-immunoreactive interneurons in the and Segev, I. (1992) Matching dendritic neuron models to
rat hippocampus. Neuroscience, 88: 37–55. experimental data. Physiol. Rev., 72: S159–S186.
Kneisler, T.B. and Dingledine, R. (1995) Synaptic input from Ratzliff, A.H., Howard, A.L., Santhakumar, V., Osapay, I. and
CA3 pyramidal cells to dentate basket cells in rat hippocam- Soltesz, I. (2004) Rapid deletion of mossy cells does not result
pus. J. Physiol., 487(Pt 1): 125–146. in a hyperexcitable dentate gyrus: implications for epilepto-
Leite, J.P., Babb, T.L., Pretorius, J.K., Kuhlman, P.A., Yeo- genesis. J. Neurosci., 24: 2259–2269.
man, K.M. and Mathern, G.W. (1996) Neuron loss, mossy Reigl, M., Alon, U. and Chklovskii, D.B. (2004) Search for com-
fiber sprouting, and interictal spikes after intrahippocampal putational modules in the C. elegans brain. BMC Biol., 2: 25.
kainate in developing rats. Epilepsy Res., 26: 219–231. Ribak, C.E., Seress, L. and Amaral, D.G. (1985) The develop-
Li, X.G., Somogyi, P., Tepper, J.M. and Buzsaki, G. (1992) ment, ultrastructure and synaptic connections of the mossy
Axonal and dendritic arborization of an intracellularly labe- cells of the dentate gyrus. J. Neurocytol., 14: 835–857.
led chandelier cell in the ca1 region of rat hippocampus. Exp. Roxin, A., Riecke, H. and Solla, S.A. (2004) Self-sustained ac-
Brain Res., 90: 519–525. tivity in a small-world network of excitable neurons. Phys.
Lothman, E.W., Stringer, J.L. and Bertram, E.H. (1992) The Rev. Lett., 92: 198101.
dentate gyrus as a control point for seizures in the hippo- Santhakumar, V., Aradi, I. and Soltesz, I. (2005) Role of mossy
campus and beyond. Epilepsy Res. Suppl., 7: 301–313. fiber sprouting and mossy cell loss in hyperexcitability: a
658

network model of the dentate gyrus incorporating cell types Vida, I., Bartos, M. and Jonas, P. (2006) Shunting inhibition
and axonal topography. J. Neurophysiol., 93: 437–453. improves robustness of gamma oscillations in hippocampal
Santhakumar, V., Bender, R., Frotscher, M., Ross, S.T., Ho- interneuron networks by homogenizing firing rates. Neuron,
llrigel, G.S., Toth, Z. and Soltesz, I. (2000) Granule cell 49: 107–117.
hyperexcitability in the early post-traumatic rat dentate van Vliet, E.A., Aronica, E., Tolner, E.A., Lopes da Silva, F.H.
gyrus: the ‘irritable mossy cell’ hypothesis. J. Physiol., and Gorter, J.A. (2004) Progression of temporal lobe epilepsy
524(Pt 1): 117–134. in the rat is associated with immunocytochemical changes in
Santhakumar, V., Ratzliff, A.D., Jeng, J., Toth, Z. and inhibitory interneurons in specific regions of the hippocampal
Soltesz, I. (2001) Long-term hyperexcitability in the hippo- formation. Exp. Neurol., 187: 367–379.
campus after experimental head trauma. Ann. Neurol., 50: Watts, D.J. and Strogatz, S.H. (1998) Collective dynamics of
708–717. ‘small-world’ networks. Nature, 393: 440–442.
Santhakumar, V. and Soltesz, I. (2004) Plasticity of intern- Wenzel, H.J., Buckmaster, P.S., Anderson, N.L., Wenzel, M.E.
euronal species diversity and parameter variance in neuro- and Schwartzkroin, P.A. (1997) Ultrastructural localization
logical diseases. Trends Neurosci., 27: 504–510. of neurotransmitter immunoreactivity in mossy cell axons
Scharfman, H.E. (1991) Dentate hilar cells with dendrites in the and their synaptic targets in the rat dentate gyrus. Hippo-
molecular layer have lower thresholds for synaptic activation campus, 7: 559–570.
by perforant path than granule cells. J. Neurosci., 11: West, M.J. (1990) Stereological studies of the hippocampus: a
1660–1673. comparison of the hippocampal subdivisions of diverse spe-
Sik, A., Penttonen, M. and Buzsaki, G. (1997) Interneurons in cies including hedgehogs, laboratory rodents, wild mice and
the hippocampal dentate gyrus: an in vivo intracellular study. men. Prog. Brain Res., 83: 13–36.
Eur. J. Neurosci., 9: 573–588. Woodson, W., Nitecka, L. and Benari, Y. (1989) Organization
Song, S., Sjostrom, P.J., Reigl, M., Nelson, S. and Chklovskii, of the gabaergic system in the rat hippocampal-formation: a
D.B. (2005) Highly nonrandom features of synaptic connec- quantitative immunocytochemical study. J. Comp. Neurol.,
tivity in local cortical circuits. PLoS Biol., 3: e68. 280: 254–271.
Soriano, E., Nitsch, R. and Frotscher, M. (1990) Axo-axonic Yoshimura, Y. and Callaway, E.M. (2005) Fine-scale specificity
chandelier cells in the rat fascia dentata: Golgi-electron mi- of cortical networks depends on inhibitory cell type and
croscopy and immunocytochemical studies. J. Comp. Ne- connectivity. Nat. Neurosci., 8: 1552–1559.
urol., 293: 1–25. Yoshimura, Y., Dantzker, J.L. and Callaway, E.M. (2005) Ex-
Spigelman, I., Yan, X.X., Obenaus, A., Lee, E.Y., Wasterlain, citatory cortical neurons form fine-scale functional networks.
C.G. and Ribak, C.E. (1998) Dentate granule cells form Nature, 433: 868–873.
novel basal dendrites in a rat model of temporal lobe epi- Zappone, C.A. and Sloviter, R.S. (2004) Translamellar disin-
lepsy. Neuroscience, 86: 109–120. hibition in the rat hippocampal dentate gyrus after seizure-
Sporns, O. and Kotter, R. (2004) Motifs in brain networks. induced degeneration of vulnerable hilar neurons. J. Neuro-
PLoS Biol., 2: e369. sci., 24: 853–864.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 36

Hippocampal granule cells in normal aging: insights


from electrophysiological and functional imaging
experiments

Monica K. Chawla1,2 and Carol A. Barnes1,2,3,

1
Arizona Research Laboratories Division of Neural Systems, Memory and Aging, University of Arizona, Life Sciences
North, Room 384, P.O. Box 245115, Tucson, AZ 85724, USA
2
Evelyn F. McKnight Brain Institute, University of Arizona, Tucson, AZ 85724, USA
3
Departments of Psychology and Neurology, University of Arizona, Tucson, AZ 85724, USA

Abstract: Normal aging, in the absence of neurodegenerative disease, can provide important insights into
the mechanisms by which the brain can maintain cognitive abilities across the lifespan. Hippocampal-
dependent memory processes can become vulnerable as age advances. The focus of this chapter is the
contribution of hippocampal granule cells to cognitive impairments that are observed during aging. A
number of alterations in structure, function, and gene expression have been observed in aged granule cells,
any of which may lead to adaptive, compensatory or detrimental consequences to hippocampal function.
As the average life span of humans continues to increase, those who reach 100 years or beyond is more
common. Individuals that have aged successfully, and exhibit high levels of cognitive ability can provide
useful clues into the enormous potential possessed by the mammalian brain.

Keywords: fascia dentata; dentate gyrus; LTP; fluorescence in situ hybridization; FMRI

Introduction appear to be involved in cell maintenance include


cytokine molecules, expression of antiapoptotic
Aging is accompanied by numerous molecular, cel- proteins, and generation of various survival-related
lular, and structural changes that have functional proteins like chaperones and antioxidants. As
consequences. Successful aging depends upon mul- human populations continue to shift towards older
tiple factors including genetic background, and average ages, it is becoming increasingly important
positive or negative environmental or social condi- to understand the factors that can ‘‘tip the balance’’
tions. The observation that the brain is able to so that more individuals age successfully and fewer
adapt to a variety of insults, suggests a large degree develop pathological conditions associated with ag-
of heterogeneity in the expression of processes that ing, including Alzheimer’s and Parkinson’s diseases.
can be employed by neurons or glia to optimize Normal aging in the absence of disease can provide
function. Among the physiological factors that insight into the stable, plastic, and adaptive strat-
egies adopted by the brain in these high functioning
Corresponding author. Tel.: +1 520-626-2616; aged individuals. Indeed, an important goal for re-
Fax: +1 520-626-2618; E-mail: carol@nsma.arizona.edu search in the neurobiology of aging is to identify

DOI: 10.1016/S0079-6123(07)63036-2 661


662

mechanisms by which some aged individuals show Granule cell numbers in aging
minimal decrement in cognitive, motor and sensory
skills across their lifespan. Early ideas suggested that cognitive decline ob-
There is now a general consensus that some as- served during aging is due to global loss of neurons
pects of learning and memory functions decline in rats, monkeys, and humans is now known to be
with advancing age (for reviews see, Landfield, incorrect. Brody (1955) first reported that the re-
1988; Rosenzweig and Barnes, 2003; Driscoll and duction in brain weight observed in older individ-
Sutherland, 2005; Burke and Barnes, 2006). Mem- uals was due to massive neuronal loss across the
ory processes that rely on hippocampal integrity cortex. While these observations were later repli-
are clearly vulnerable to advancing age in humans cated in rats and humans in other laboratories
(e.g., Cabeza et al., 2005), monkeys (e.g., Rapp (Ball, 1977; Coleman and Flood, 1987; Coleman et
and Amaral, 1991; Keuker et al., 2000) and in ro- al., 1987), these early studies used cell counting
dent models of aging (e.g., Barnes, 1979; Geinis- methods without stereological techniques, and were
man et al., 1986; Barnes, 1988, 1994; Kadar et al., found to be incorrect in later experiments that used
1994; Gallagher and Rapp, 1997; Bach et al., 1999; unbiased cell counting methods (Sterio, 1984; West,
Rapp et al., 1999; Driscoll et al., 2006). 1993a; West et al., 1994; Morrison and Hof, 1997).
This chapter focuses on the contribution of hip- It is now generally accepted that age-related cog-
pocampal granule cells to cognitive impairments of nitive decline is not due to massive neuronal loss.
aged animals. First, the basic anatomical, bio- Using very detailed electron microscopic meth-
chemical and biophysical properties of young ods and serial reconstructions Geinisman and
granule cells will be compared with older granule Bondareff (1976) first reported that granule cell
cells. Whether these modifications in aging are numbers are not decreased in aged rat, but rather,
adaptive, compensatory or detrimental to the they lose about a third of their medial entorhinal
hippocampus will be discussed. Next, functional synaptic input. A number of more recent investi-
connectivity, inferred by extracellular electrophys- gations have replicated these results using modern
iological recording methods, is examined in an at- counting methods and have found no age-related
tempt to relate functional synaptic changes to neuronal loss of CA pyramidal cells or dentate
cognitive decline in aged rats. Additionally, since gyrus granule cells (Rapp and Gallagher, 1996; Ra-
aging is a multidimensional process that can affect smussen et al., 1996) in rats. Similarly, in aged
many complex cellular mechanisms, transcrip- monkeys (Peters et al., 1996), and in humans (West,
tional regulation of immediate early genomic re- 1993b), there appears to be no hippocampal gran-
sponses are also examined here. The triggers and ule or CA pyramidal cell loss in aging. Moreover,
consequences of altered gene responses are poorly entorhinal cortical layers II and III cell numbers do
understood at present. Recent advances in imaging not change during aging in rats or monkeys (Merrill
technologies have enabled both the identification et al., 2000, 2001; Rapp et al., 2002).
of regions of brain that are activated by specific
behavioral experiences in monkeys and humans
(Small et al., 2002, 2004) as well as specific cells Granule cell neurogenesis and aging
that are actively engaged in a particular behavioral
experience (Guzowski et al., 1999). The examples Granule cells are unusual, in that they are among
that follow will illustrate the complex pattern of the few neuronal cells that undergo neurogenesis
changes that occur during aging with a focus in adulthood (Altman and Das, 1965; Gould and
on granule cells. These patterns are not only Cameron, 1996; Eriksson, 2003). Neurogenesis has
subtle but highly sub-region specific and point also been observed in olfactory bulb neurons
towards compensatory mechanisms engaged by across the lifespan of mammals. The newborn
the brain to adapt to the complexities of the mul- granule cells originate from the subgranular zone
tidimensional, enigmatic process called normal of the dentate gyrus and migrate into the granular
aging. layer (Kuhn et al., 1996), develop dendritic
663

processes (Ribak et al., 2004) and mature into neurons remain stable from middle age into older
functional neurons displaying some of the same ages (Rao et al., 2005). Moreover, the increase in
synaptic responsiveness and other electrophysio- neurogenesis that is induced by specific types of
logical properties as do existing granule cells (van learning (contextual fear conditioning) is reduced in
Praag et al., 2002; Song et al., 2005). Continous aged rats, as well as the survival of new granule cells
neurogenesis in the dentate gyrus throughout the (Wati et al., 2006). The reduction in integration of
life span of mammals may be of great functional newborn granule cells into hippocampal circuitry
significance. For example, Shors et al. (2001) have with age may contribute to the cognitive impair-
shown that numbers of newborn granule cells cor- ment in aged animals.
relate with effective hippocampal-dependent mem-
ory function. Whether, newly born granule cells
are integrated into functional networks mediating Age-related changes in granule cell synaptic
spatial information processing was recently inves- contacts
tigated by Ramirez-Amaya et al. (2006). Using
triple immunohistochemistry and confocal micros- The constancy in granule cell numbers (Geinisman
copy, they were able to show that granule cells and Bondareff, 1976) and the volume of the gran-
born in the mature dentate gyrus (5-month old), ule cell layer (Coleman and Flood, 1987) with in-
newborn granule cells demonstrated by the mitotic creasing age in rodents stands in contrast to
marker-BrdU express the plasticity-related imme- reports of a decrease in synapse numbers. As was
diate-early gene Arc protein in response to spatial the case with cell counts in early literature, synapse
exploration. Additionally, the proportion of Arc counts conducted across age groups produced var-
protein-containing cells was significantly higher in iable results, possibly due to methods that were
the newly born granule cell population compared less accurate than current techniques (Geinisman
to the preexisting population of granule cells and Bondareff, 1976; Geinisman et al., 1977;
(2.8% vs. 1.6%, respectively). Geinisman, 1979; Hoff et al., 1982; McWilliams
Recently it was shown that exercise-induced in- and Lynch, 1984; Bertoni-Freddari et al., 1986;
creases in neurogenesis are associated with better Badiali de Giorgi et al., 1987). Using unbiased
performance on the Morris water maze (van Praag stereological techniques, Geinisman et al. (1992)
et al., 2005), and a correlation between higher lev- demonstrated a loss of axospinous synapse num-
els of neurogenesis and preserved spatial memory bers in the middle third of the molecular layer of
in old rats has been observed by Drapeau et al. the dentate gyrus in memory-impaired old rats.
(2003). Other studies, however, have not found When the entire hippocampus was homogenized
correlations between neurogenesis and behavior in preparation for protein analysis to examine pos-
(Bizon and Gallagher, 2003), or in some instances sible age-related differences in levels of synapse-
a negative correlation (Bizon et al., 2004) in the related proteins (synaptogamin-, synaptophysin-, or
aged rodent. An interesting hypothesis recently synaptosomal-associated protein 25), no age effects
proposed for the role of adult neurogenesis, sug- were observed (Nicolle et al., 1999). However, with
gests that the newborn granule cells participate in the use of methods that can detect circuit-specific
forming time-tagged associations for laying down differences in synaptic protein levels, age-related
new memories (Aimone et al., 2006). differences are revealed. Smith et al. (2000) exam-
Aged rats do show granule cell neurogenesis and ined the layer II entorhinal cortical projection that
old granule cell survival rates can be modified by inervates both CA3 stratum lacunosum moleculare
enriched environments (Kempermann and Gage, and dentate gyrus molecular layer. There was a
1999; Kempermann et al., 2002). Generation of significant loss of synaptophysin in both areas in
newly born granule cells from progenitor cells, memory-impaired old rats. Studies measuring
however, show a marked decline beginning in mid- synaptic markers, however, cannot elucidate the
dle age (Kuhn et al., 1996; Rao et al., 2005), but the possible functional changes that result from alter-
numbers of newly born granule cells that become ations in these proteins, such as synaptic strength.
664

The answer to these kinds of functional questions, Electrophysiological comparisons of granule cells
require electrophysiological investigation (see of young and old rats have been made using both in
Synaptic Response section below). vivo and in vitro preparations. These studies reveal
that most of the basic granule cell electrical prop-
erties remain constant over the lifespan of rodents,
Granule cell dendrites in aging
mice, and rabbits (Barnes et al., 1994). For exam-
ple, studies investigating resting membrane poten-
As discussed above, since numbers of granule cells
tial, membrane time constant, the relation between
do not change with age, it follows that the num-
applied current and input resistance, the amplitude
bers of primary dendritic shafts should also remain
and duration of the Na+-dependent action poten-
unchanged. One study, however, did report a re-
tial, threshold of action potential following depo-
duction in numbers of granule cell shafts in old
larizing current injection, and rise time and half
rats (Geinisman et al., 1978). Because nonstereo-
width of the intracellular EPSP (Barnes and McN-
logical methods were used in that study, the results
aughton, 1979, 1980a; Barnes et al., 1987; Baskys et
may most easily be explained by sampling bias. An
al., 1987; Niesen et al., 1988; Foster et al., 1991)
investigation in young and old rhesus monkeys,
have not revealed age-associated changes (for re-
using stereological techniques, revealed no signifi-
view, see Barnes, 1994; Rosenzweig and Barnes,
cant difference in total dendritic length, number of
2003).
dendritic branch points and total segment number
Interestingly, the action potential threshold to
between young and old granule cells (Luebke and
orthodromic stimulation is decreased in aged gran-
Rosene, 2003). They did, however, find that the
ule cells (Barnes and McNaughton, 1980a). The
molecular layer was decreased in width (between
observation that threshold to action potential dis-
the granule cell layer of the upper blade and the
charge to direct current does not change during
hippocampal fissure) in old (>24 years) vs. young
aging, while the action potential threshold to or-
(o11 years) monkeys.
thodromic stimulation does change, might be ex-
Can spatial memory deficits with advancing age
plained in several ways. Among these was the
be explained in part by a decrease in spine numbers?
hypothesis that there is increased electrotonic cou-
von Bohlen und Halbach et al., 2006 recently re-
pling between granule cells with age (Barnes et al.,
ported no significant reduction in spine density in
1987). If there were more gap junctional connec-
the apical dendrites of the upper blade DG and area
tions between granule cells, current would flow
CA1 of old mice, although the mean spine length
from the granule cell with the lowest threshold into
did show a reduction in both areas in young (6–7
adjacent cells that had not reached threshold.
months) compared to old (21–22 months) mice. In-
Overall, this would result in a lowering of the
terestingly, Samsonovich and Ascoli (2006) have
threshold for the entire population resulting from
recently concluded, using statistical analysis of dig-
stimulation of the perforant path. Previously, it
itally reconstructed neurons (from several different
had been shown by MacVicar and Dudek (1982)
brain regions and different experimental manipula-
that 60% of granule cells were electrotonically
tions) that there is morphological homeostasis in
coupled, using Lucifer yellow injections and freeze
apical and basal dendritic processes of neurons.
fracture methods. In the Barnes et al. (1987), study
Therefore, dendritic changes in one area may be
a low molecular weight fluorescence dye, 5,6-car-
compensated by changes in other areas in order to
boxyfluorescein, was used to explore intercellular
preserve the dendritic tree, spine numbers, and
connections via gap junctions in in vitro slice
other characteristics of granule cells.
preparations. Increased dye coupling was observed
in all three hippocampal subregions with age. In
Biophysical properties of aged granule cells fact, granule cells showed the most extensive cou-
pling (i.e., 64% of old granule cells were coupled
Does the absence of gross granule cell loss in to at least one other cell compared to 49% of
aging mean a general preservation of function? young granule cells). Accompanying this increased
665

Fig. 1. (A) Two representative granule cells filled with 5,6 carboxyfluorescein from the dentate gyrus of a 24-month-old rat. (B)
Histograms showing the numbers of carboxyfluorescein injections that resulted in single, double, triple, or greater numbers of granule
cells filled with dye. Aged rats showed significantly increased electrotonic coupling compared to young animals. This may account for
the increased excitability of old granule cells. Adapted with modifications from Barnes et al. (1987). (See Color Plate 36.1 in color plate
section.)

electrotonic coupling, old granule cells also exhib- condition in the dentate gyrus (Jung and McN-
ited an apparent increase in postsynaptic excita- aughton, 1993; Shen et al., 1998; Gothard et al.,
bility, as assessed by the ratio of the population 2001). Additionally, granule cells were originally
spike area to EPSP amplitude (26%). Figure 1A misclassified as having ‘‘theta-cell-like’’ firing char-
shows an example of two granule cells filled with acteristics (Deadwyler et al., 1976; Rose et al.,
5,6-carboxyfluorescein taken from the dentate 1983). It is now known, however, that granule cells
gyrus of a 24-month-old rat. The bar graph shown have low firing rates and selective place fields in
in Fig. 1B shows that older animals have signifi- environments, similar to hippocampal pyramidal
cantly higher electrotonic coupling probabilities cell place fields. For example, Jung and McNaugh-
compared to young granule cells. The increase in ton (1993) recorded single unit activity from gran-
electrotonic coupling and excitability may contrib- ule cells while rats performed a spatial task on an
ute to compensatory processes acting to keep the eight-arm radial maze. It was found that granule
probability of discharge of individual granule cells cells showed both spatial and direction specificity in
constant in spite of a considerable loss (20–30%) their firing properties, although their place fields
of entorhinal cortical afferents to the hippocampal were smaller and more ‘‘discontigous’’ than pyram-
dentate gyrus. idal cell place fields. Additionally, single granule
cells have multiple place fields that were stable over
at least several days. It was also observed that fewer
Single unit recordings in granule cells granule cells, compared with hippocampal pyram-
idal cells, showed place-specific firing within any
Very few studies have investigated the spontaneous given environment (Barnes et al., 1990).
firing characteristics of granule cells in freely be- These electrophysiological data are consistent
having animals. This is partially due to the fact that with theoretical ideas suggesting that it is advan-
few granule cells fire during any given behavioral tageous for hippocampal autoassociative memory
666

processing for information to be coded sparsely current transients at the granule cell bodies. Thus,
(Marr, 1971; McNaughton and Morris, 1987; the waveform of the population synaptic responses
Rolls, 1990). If the dentate gyrus performs the recorded below the granule cell layer, should reflect
function of a pattern separator, the formation of the passive electrical properties of the dendrites,
distinct representations would be facilitated by and location of the active synapses as in an equiv-
amplification of small differences in the input to alent cylinder model. Because a square root rela-
the granule cell network. Although few granule tionship between synaptic distance and the rise time
cells may fire in any given situation, consistent is empirically robust (Jefferys, 1975), the square
with a pattern separation mechanism, granule cell root of the rise time of the extracellular synaptic
firing patterns undergo substantial modification response recorded from the hilus was calculated
after only minimal changes in an environment and used as an electrotonic location parameter for
(Leutgeb et al., 2007). comparison across the lifespan of the rat (Barnes
As discussed above, earlier studies mistakenly and McNaughton, 1979). By selectively stimulating
identified granule cells to possess firing character- small subsets of perforant path fibers along its
istics similar to inhibitory interneurons. This may medio-lateral axis, synapses at increasing distances
be explained in part, because of the ease with from the soma were activated and the response
which higher firing rate cells are encountered along waveform assessed (as illustrated in Fig. 2). When
a recording tract trajectory. Mizumori et al. (1989) the mean electrotonic location parameter was com-
were able to identify characteristics that could pared across age groups, the rise time of the extra-
differentiate between interneurons (basket cells) cellular synaptic response was similar across the
and granule cells. For example, interneurons have lifespan. When the medians of the frequency distri-
high spontaneous discharge rates and a relative butions (as shown in Fig. 3) of the electrotonic lo-
absence of spike accommodation, while granule cation parameter were compared in three age
cells have lower firing rates. There have been no groups of rats, there was a small shift towards
systematic age comparisons of the firing charac- lower values with advancing age. This small shift in
teristics of granule cells; however, a radial maze the distribution of extracellularly recorded synaptic
spatial working memory task was used to assess response rise times may suggest atrophy of finer
the possible contribution of age-related changes in dendritic branches, which was, in fact an observa-
the firing characteristics of theta cells (putative in- tion made by Geinisman et al. (1978) in the supra-
terneurons) (Mizumori et al., 1992). The mean granular zone of old rats.
discharge rates of theta cells were significantly Barnes (1979) first noted, using extracellular
higher in the stratum granulosum in old rats com- field potential recordings in young and old rats,
pared to young animals. This region-specific that, for a given stimulus intensity delivered to the
change in the firing characteristics of theta cells perforant path afferent fibers, old Long Evans rats
may have important consequences for information showed reduced field EPSP amplitudes compared
processing with increasing age. with young Long Evans rats. Using both in vivo
and in vitro preparations to record extracellularly
in young and old rats, Barnes and McNaughton
Perforant path granule cell synaptic responses in (1980a) found that while there was no difference in
vivo and in vitro the EPSP size at threshold stimulation intensities,
at higher stimulus strengths, the EPSP was con-
Depending on the location of termination of acti- sistently smaller in old rats, over a wide range of
vated perforant path synaptic contacts onto the intensities (see Fig. 4).
granule cell dendritic tree, the resulting field poten- It is also possible to record the compound action
tial responses have distinctive rise time and shape potentials of incoming perforant path fibers (i.e.,
characteristics. Because of the curved geometry of the presynaptic fiber potential) to estimate the
the dentate gyrus, field potentials recorded within numbers of perforant path axon collaterals that
the hilar region are nearly proportional to the terminate in the granule cell molecular layer.
667

Fig. 2. (A) Schematic diagram of a granule cell in the dentate gyrus showing the extracellular recording electrode placed slightly below
the cell body layer (in the hilus). (B) Representative extracellularly recorded synaptic responses evoked from stimulation of different
subpopulations of perforant path fibers are shown adjacent to their approximately corresponding input location on the granule cell
dendrites. Note that the rise time characteristics of the synaptic responses vary as a function of the dendritic location of the activated
synapses. Adapted with modifications from Barnes and McNaughton (1979).

When the presynaptic fiber potential was meas- granule cell, i.e., have stronger weights. The later
ured in young and old rats, again there was no observation is presumed to be due to axon collat-
change in the response amplitude at threshold eral pruning in the perforant path projection from
stimulus intensity levels, however, above thresh- the entorhinal cortex. This hypothesis was tested
old, younger animals showed larger fiber potential in vitro, using minimal stimulation methods and
response amplitudes. This finding was consistent quantal analysis techniques while recording in-
with the reported decline in anatomical synaptic tracellularly from granule cells of aged rats (Foster
contacts, as shown by Geinisman (1979). et al., 1991). Foster et al. tested rats of three
A smaller presynaptic fiber potential implies different ages (5, 9, and 24 months) and found that
that fewer axon collaterals are activated by the the increased EPSP amplitude of granule cells in
stimulus, and is therefore a possible explanation old rats was due to an increase in quantal size,
for the smaller extracellular field EPSP observed in which implicates an increase in postsynaptic sen-
old rats. When field EPSP amplitudes were plotted sitivity. Since the magnitude of paired-pulse facil-
against fiber potential amplitudes, however, the itation was unchanged between age groups, the
synaptic response was larger for a given input size. increase in synaptic strength could not be ex-
Thus, smaller fiber potentials gave rise to larger plained by an increase in the probability or
synaptic responses in old rats, inspite of the fact number of transmitter quanta released. These post
that at stimulus intensities above threshold fiber synaptic changes in synaptic strength have impor-
potential amplitudes were smaller. tant functional implications for maintaining gran-
One explanation for this result would be that ule cell activity within a functional range and may
individual synapses that survive in the old rat suggest a compensatory mechanism employed by
would be more effective in depolarizing the the aging dentate gyrus.
668

Fig. 3. (A) The frequency distributions of the electrotonic location parameter from adolescent (2 months), middle-aged (12 months), or
old (24 months) rats as percentage of number of observations for each age group. (B) Bar graph shows the medians of the frequency
distributions of the electrotonic parameter in adolescent, middle-aged, or old rats. Note the small shift towards lower values with
advancing age. This small shift in the distribution of extracellularly recorded synaptic response rise times towards lower values with
advancing age may suggest atrophy of the finer dendritic branches. Adapted and modified from Barnes and McNaughton (1979).

Cholinergic responses in vitro in aging granule cells excitability of granule cells, which could signifi-
cantly influence hippocampal network behavior
When cholinergic afferent fibers are stimulated at (Huerta and Lisman, 1993; Hasselmo et al., 1996;
40 Hz, a slow cholinergic (atropine-sensitive) Hasselmo, 1999; Barnes et al., 2000b; Ikonen et al.,
EPSP can be elicited in all three hippocampal sub- 2002).
regions. When the response to this type of stim-
ulus, delivered to stratum granulosum, was
evaluated in animals from three different age Modification of synaptic weights at the perforant
groups (3, 9, or 24 months) there was a signifi- path-granule cell synapse
cant age-related decline (50%) in the amplitude
of the slow cholinergic EPSP (Shen and Barnes, The classic studies published in 1973 by Bliss and
1996). There was also a similar decine in choliner- Gardner-Medwin (Bliss and Gardner-Medwin,
gic transmission in the CA3 and CA1 regions of 1973; Bliss and Lomo, 1973) demonstrate that
old rats. Therefore, the neuromodulatory role of synaptic projections from the entorhinal cortex
the cholinergic system in aging may alter could be modified, both in vivo (Bliss and
669

Fig. 4. The relation between field EPSP amplitude and perforant path stimulus intensity is shown in young (K) and old (J) rats for in
vivo (A) and in vitro (D) preparations. The response waveforms (averaged across age groups) are shown for young (B) and old (C)
animals from an in vivo experiment. Superimposed traces at various stimulus levels recorded simultaneously from the granule cell layer
(E) and the molecular layer (F) are shown for an in vitro slice experiment. The dashed lines in B and C and E indicate the time of
measurement of the field EPSP (2 ms after stimulus onset). The sharp negative deflections in B, C, and E are population spikes
(asterisks) whereas the early negative deflection in F represents the presynaptic fiber response (indicated by arrow). Adapted from
Barnes and McNaughton (1980a).

Gardner-Medwin, 1973) and in vitro (Bliss and The relationship between the durability of LTP
Lomo, 1973) by patterned activation of perforant and spatial memory was first reported in 1979
path afferents to hippocampal granule cells. This (Barnes, 1979), with the demonstration that spatial
phenomenon, now referred to as long-term po- performance accuracy on the circular platform
tentiation or LTP (Douglas and Goddard, 1975), task was correlated with the durablility of LTP, as
required convergence of inputs onto granule cells assessed over a number of days. This experiment
or ‘‘cooperativity’’ (McNaughton et al., 1978), was conducted in young and old Long Evans rats,
consistent with Hebb’s neural postulate for asso- trained on the circular platform spatial memory
ciation (Hebb, 1949). It was subsequently found task, where the goal was to escape from a brightly
that this electrophysiological demonstration of illuminated open surface by finding the one hole
coincidence detection worked through a novel (identified by the distal visual cues in the room)
glutamatergic receptor, the NMDA receptor (Col- that would allow the rat to escape into a dark box.
lingridge, 1985). The NMDA receptor was found Following behavioral testing, 32 young and 32 old
to be voltage-dependent and to allow influx of animals were implanted bilaterally with indwelling
calcium, presumed at that time to be necessary for electrodes that could monitor field potentials in the
initiating durable forms of plasticity. In fact, Mor- dentate gyrus elicited by stimulation of the medial
ris et al. (1986) showed that NMDA-receptor an- perforant path to granule cells. The results of this
tagonism (via APV administration) was able to experiment were able to show, for the first time,
disrupt performance of the spatial, hippocampal- that within each age group there was a significant
dependent version of the Morris swim task, while correlation between the decay time constant of late
leaving the nonspatial version intact. phase LTP and spatial memory. The same
670

relationship between spatial behavior and LTP dentate gyrus are complex during aging (Magnus-
maintenance has been found in young and old son et al., 2002, 2003), and it is not clear whether
mice at the Schaffer collateral — CA1 synapse in changes in receptor-related functions would result
vitro (Bach et al., 1999). In these experiments a in the changes in plasticity (e.g., Wenk and Barnes,
nonspatial version of the circular platform task 2000).
was also administered, and the performance on the While, the examination of spatial memory and
cued task did not correlate with LTP durability. LTP in the hippocampus of old rats has provided
To examine the LTP induction process over important insights into age-related functional de-
days, as well as its maintenance after induction cline, it has also raised a number of questions.
levels were saturated, young and old rats were Among these is, what causes the faster decay in
prepared for long-term field potential recordings LTP in old animals? If LTP decay could be ma-
from the dentate gyrus. After baseline evoked re- nipulated, memory function might be facilitated in
sponse stability to low frequency test pulses had older animals. The mechanisms involved in the
been established, bursts of high frequency stimu- maintenance of LTP are, unfortunately, less well
lation were delivered once per day for 7 consec- understood than those involved in induction of
utive days. Although the old rats reached LTP. As discussed in the following section, some
asymptotic levels of LTP following this treatment insight may be gained through an examination of
more slowly than did the younger rats, they did the role played by immediate early genes in dura-
achieve the same proportional levels of synaptic ble forms of synaptic changes that may underlie
strengthening. Decay of LTP was then monitored memory changes in old mammals.
for several weeks. Old rats showed faster decay
over days than did young rats (time constants of
17 and 37 days, respectively); Barnes and McN- Immediate early genes
aughton (1980a).
As discussed above, if LTP is induced using ro- The induction and translation of immediate early
bust stimulus parameters at the perforant path- genes, lead to changes in multiple intracellular
granule cell synapse, there the degree of LTP is signaling cascades, and may play an important
equivalent between age groups (Barnes, 1979; role in changes observed in age-related plasticity
Barnes and McNaughton, 1980a). One question mechanisms. A number of IEGs are powerfully
is whether there is an altered threshold for induc- activated by stimulation parameters that lead to
tion of LTP at this synapse. Certainly there are LTP induction. As discussed in the previous sec-
fewer synaptic contacts in the termination zone of tion, old animals, in addition to changes in LTP
the medial perforant path synapses (Geinisman et induction thresholds, also exhibit decreases in LTP
al., 1992b) and thus, less synaptic convergence maintenance that are not detectable until several
from perforant path stimulation would be days after robust induction protocols have been
achieved in older rats. applied. Studies examining the basis of the main-
To address the possibility that the threshold for tenance phase of long-term plasticity have dem-
LTP was altered in old rats, Barnes et al. (2000a) onstrated the requirement for RNA and protein
directly depolarized the granule cells with intra- synthesis following the conditioning stimulus
cellular current injection paired with orthodromic (Montarolo et al., 1986; Nguyen et al., 1994; La-
stimulation, thereby eliminating the necessity for nahan et al., 1997). Therefore, it is plausible that
synaptic convergence, which is typically required age-dependent memory decline could result from
to induce LTP (McNaughton et al., 1978). Using alterations in pathways subsequent to transcrip-
this paradigm more current was required to induce tional responses mediated by IEGs.
LTP in aged animals. One possible mechanism for Cole et al. (1989) established a relationship be-
this apparent deficit in LTP induction in the aged tween LTP induction and activation of several
dentate gyrus could be alterations in NMDA re- IEGs (including, zif268, c-fos, c-jun, jun-B, and
ceptors. The biochemical data reported for the Arc). In that study, when LTP at the perforant
671

path-granule cell synapse was induced by high fre- As the studies of specific IEGs in activity-depend-
quency stimulation, levels of zif 268 mRNA in- ent synaptic plasticity continued, an increase in a
creased significantly. This increase in zif 268 unique gene, activity-regulated cytoskeleton-associ-
mRNA could be blocked both by NMDA-recep- ated gene (Arc, also known as Arg 3.1) was iden-
tor antagonists and by convergent synaptic inhib- tified, following seizures, or LTP induced at the
itory inputs known to block LTP, thus establishing perforant path granule cell synapse (Link et al.,
a correlation between IEG expression and LTP. 1995; Lyford et al., 1995). It seemed logical to in-
Similar findings were reported in un-anesthetized vestigate whether Arc had a role in the maintenance
rats by Dragunow et al. (1989). The early studies phase of LTP. Guzowski et al. (2000) used intra-
that linked IEG expression in granule cells with hippocampal infusions of antisense oligodeoxynuc-
LTP, were however, called into question by sug- leotides to block the synthesis of Arc protein and
gestions that the thresholds for LTP induction and monitored the effect on LTP and spatial memory.
gene activation are different. To address this issue, Blocking the synthesis of Arc protein in the dorsal
Worley et al. (1993) used a chronic in vivo record- hippocampus impaired the maintenance phase of
ing approach to monitor two different patterns of LTP induced at the perforant path-granule cell
LTP-inducing stimuli to perforant path afferents. synapse, and the long-term consolidation of memory
The RNA expression pattern and the protein for a spatial location (Morris water maze) without
products of several IEGs (zif 268, jun-B, c-jun, and interfering with either the induction phase of LTP or
c-fos) were monitored in young and old animals, water maze acquisition. Thus, a role for Arc in hip-
following the two different electrical stimulation pocampal-dependent plasticity was established.
protocols (10 pulse train or 50 pulse train, deliv- The IEG Arc proved to be a useful tool to clarify
ered at 400 Hz). Although, different patterns of behavior-related synaptic plasticity as well, as it is
IEG expression emerged, only zif 268 was detect- rapidly induced when animals explore their envi-
able at thresholds that induced LTP. Additionally, ronment. Guzowski et al. (1999) showed that Arc
when the in situ autoradiograms were quantified can be induced in hippocampal CA1 neurons fol-
using densitometry, there were no significant age- lowing two episodes of spatial experience and that
related differences in zif 268 mRNA levels between the proportions of cells expressing Arc matched
groups. predictions derived from electrophysiological re-
Using the more sensitive reverse Northern cordings. In that investigation it was also shown
method, Lanahan et al. (1997), examined a panel that Arc RNA is transcribed within 1–2 min fol-
of several genes (c-fos, jun B, c-jun, fos B, fra1, and lowing behavior, and that it appears as discrete
zif 268) known to be induced by LTP. In that transcriptional foci in the nucleus (for 2–15 min).
study, an LTP induction protocol (high frequency Within 15–30 min the RNA translocates to the cy-
stimulation of the perforant fibers) known to result toplasm, and can be subsequently found in the
in LTP maintenance deficits in old rats (Barnes dendrites (45 min). This unique property of Arc
and McNaughton, 1980b) was used. The dentate formed the basis for the development of a powerful
gyrus was extracted from young (9 month) and old new technique called compartment analysis of tem-
(26 month) animals and reverse Northern blots poral activity by fluorescence in situ hybridization
were quantified using a phosphoimager. Of all the or catFISH. With this technique, the behavior-in-
genes examined (including zif 268) only c-fos RNA duced activity history of neuronal ensembles could
level differed between young and old rats. Inter- be inferred at two different time points, throughout
estingly, while both young and old rats showed the brain (e.g., Guzowski et al., 1999; Chawla et al.,
elevated c-fos expression following LTP induction, 2004). Although, many IEGs can be used in the
the older rats showed greater levels of expression catFISH technique, Arc provides distinct advan-
than did the young animals. This indicated that tages in that it is dynamically regulated in many
changes in c-fos, may contribute to altered signa- brain regions.
ling mechanisms that are involved in age-related When the pattern of spatial exploration-induced
hippocampal plasticity deficits observed in aging. Arc expression was examined in hippocampal
672

granule cells of young rats, it was found to be


predominantly located in the upper blade of the
dentate gyrus (Chawla et al., 2005). No major an-
atomical differences in afferents to the upper vs.
lower blade easily account for this interesting ob-
servation, since only minor differences in anatom-
ical connectivity exist between these regions. The
proportion of active cells following spatial experi-
ence is low for granule cells compared to CA1 py-
ramidal cells (4% vs. 40%), and is consistent
with the sparse firing pattern observed in the dent-
ate gyrus in electrophysiological studies (Barnes et
al., 1990; Jung and McNaughton, 1993; Shen et
al., 1997; Gothard et al., 2001). Figure 5 illustrates
the expression of behaviorally induced Arc in
young rats (Fig. 5A), which is significantly ele-
vated compared to baseline levels of expression in
young animals that have not performed the spatial
foraging task (Fig. 5B). The difference in upper
and lower blade Arc expression cannot be ex-
plained by the inability of the lower blade to ex-
press Arc, because as early as 5 min following
maximal electroconvulsive shock (MECS) treat-
ment, both the upper and lower blades show ro-
bust Arc expression. It is possible that granule cells
in the lower blade are largely inactive for a variety
of experiences.

Vulnerability of dentate gyrus in aging

The information gathered over the past several Fig. 5. Image of granule cells (blue) following in situ hybrid-
ization using a c-RNA probe for the IEG Arc. (A) Arc positive
decades from electrophysiological recording and cells (shown in red) in the granule cell layer from the upper
anatomical studies, has been critical in furthering blade of a young rat following two 5 min (separated by a 20 min
our understanding of the kinds of learning and rest period in the home cage) episodes of spatial exploratory
memory deficits that should be expected in normal behavior. (B) Granule cells from a caged control animal of the
aging. Anatomical and physiological studies can same age, that did not perform the exploratory behavior, are
shown, illustrating baseline or negligible expression of Arc
be complimented by functional imaging ap- mRNA. (See Color Plate 36.5 in color plate section.)
proaches that can directly identify behavioral cor-
relates of neuronal function. Several imaging
techniques are now available that provide cellular values in 14 young and old rhesus monkeys (7–31
and temporal resolution in this regard. Among years) that also performed a memory assessment
them is the functional imaging method based on task (delayed nonmatching-to-sample, DNMS).
measurement of cerebral blood flow volumes using Representative CBV meaures from a young and
magnetic resonance imaging (MRI) that have been old monkey is shown in Fig. 6A. The CBV meas-
shown to correlate with regional energy metabo- ures indicated a significant inverse correlation with
lism (e.g., Gonzalez et al., 1995; Hyder et al., age only in the dentate gyrus (Fig. 6B). In addi-
2001). Recently, Small et al. (2004) measured CBV tion, when CBV measurements were plotted
673

Fig. 6. (A) Examples of gadolinium-induced change in MRI signal, (a measure of cerebral blood flow volume, CBV, that is a correlate
of brain metabolism) from a young and old rhesus monkey. Warm colors indicate higher CBV values. The white circle, identified in
precontrast images, overlies the dentate gyrus. (B) CBV measurements plotted vs. age in the dentate gyrus indicates a significant decline
with advancing age. (C) CBV measurements in the dentate gyrus of old monkeys plotted against memory performance on the delayed
nonmatching-to-sample test. A significant relationship between CBV and memory performance is observed, selectively in the old
animals. Adapted with modifications from Small et al. (2004). (See Color Plate 36.6 in color plate section.)

against memory performance on a delayed non- subiculum and dentate gyrus show selective
matching-to-sample task; again, only the dentate changes in normal aging that could be correlated
gyrus in old monkeys showed a significant corre- with memory decline in these individuals. Thus, a
lation with CBV activity (Fig. 6C). As discussed cross species consensus is now emerging in that the
previously, catFISH can provide both temporal dentate gyrus in rats, monkeys, and humans is the
and cellular resolution in determining the activity hippocampal subregion that is particularly vulner-
history of individual neurons. When cellular Arc able to advancing age.
RNA was determined using catFISH combined
with high-resolution confocal microscopy in
young, middle-aged, and old rats, only the dent- Conclusions and summary
ate gyrus showed age-related changes. As shown in
Fig. 7B, a significant age-related decline in Arc- Aging is accompanied by selective memory
RNA-containing granule cells was observed in changes that depend on proper hippocampal func-
aged animals. Importantly, as was seen in mon- tion. Investigation of hippocampal anatomy, phys-
keys with MRI imaging, the pyramidal subregion iology, and gene expression in normal aging
in rats was unaffected by advancing age, i.e., there provides a foundation for understanding age-re-
was no change in Arc-expressing pyramidal cell lated cognitive decline. A common theme that
numbers. emerges from examination of the neurobiological
Furthermore, using MRI studies in humans, changes in aging granule cells, is that cell loss can-
Small et al. (2002) have also reported that not account for the cognitive decline, as granule
674

Fig. 7. Behavior-induced Arc mRNA expression pattern in young and old rat granule cells of the dentate gyrus. (A) Individual
examples of behaviorally induced Arc RNA. Fluorescence-tagged CY3 Arc RNA containing cells are green, and granule cell nuclei are
shown in red (counterstained with propydium iodide, a nucleic acid stain). (B) The average proportion of Arc-positive neurons (% of
total cells) measured from the three different age groups (9, 15, and 24 months) is shown for the dentate gyrus. Note that the dentate
gyrus of old rats contains a significantly lower proportion of behaviorally induced Arc-positive granule cells compared to young or
middle-aged animals. (See Color Plate 36.7 in color plate section.)

cell numbers are preserved during aging. Although With the human population living well into their
there is an absence of neuronal degeneration and a eighties or nineties, it is becoming increasingly im-
general preservation of basic biophysical proper- portant to find ways to preserve cognitive func-
ties, there is a significant loss of axodentritic tion, and maintain an optimal quality of life
synapses onto granule cells from entorhinal affer- during the last decades. New imaging techniques,
ents. In addition to the synapse loss, there are a pharmacological interventions or genetic manipu-
number of other changes that occur in aged ani- lations, can advance our knowledge regarding
mals. Among them, is an altered threshold for normal aging. A multidisciplinary effort will be
LTP induction, an increase in strength of the sur- essential for identifying early interventions to
viving synapses, and increased electrotonic cou- ameliorate declining cognition with advancing age.
pling between the old granule cells. Together, these
factors may result in an attenuation of the dele-
terious impact of aging on cognition. Acknowledgements
In spite of the compensatory changes men-
tioned above, granule cells appear to be partic- We thank Michelle Carroll for her excellent techni-
ularly vulnerable to the effects of aging compared cal assistance in preparing the figures. The authors
to hippocampal pyramidal cells. For example, and some of this work were supported by grants
there is a reduction in the numbers of granule from the National Institute of Health AG009219
cells that show behavior-induced gene expression and AG003376, the state of Arizona and ADHS
in old rats. Furthermore, older monkeys that and the McKnight Brain Research Foundation.
have difficulty performing a hippocampal-de-
pendent learning task also have lower functional
activity in the dentate gyrus as assessed by cer- References
ebral blood volume measures using MRI tech-
niques. Additionally, MRI studies in humans also Aimone, J.B., Wiles, J. and Gage, F.H. (2006) Potential role for
indicate that dentate gyrus is also a hippocampal adult neurogenesis in the encoding of time in new memories.
subregion with selective changes in normal aging Nat. Neurosci., 9: 723–727.
Altman, J. and Das, G.D. (1965) Autoradiographic and histo-
that could be correlated with memory decline in logical evidence of postnatal hippocampal neurogenesis in
these individuals. rats. J. Comp. Neurol., 124: 319–335.
675

Bach, M.E., Barad, M., Son, H., Zhuo, M., Lu, Y.F., Shih, R., synaptic junctions in rat dentate gyrus during aging. Brain
Mansuy, I., Hawkins, R.D. and Kandel, E.R. (1999) Age- Res., 366: 187–192.
related defects in spatial memory are correlated with defects in Bizon, J.L. and Gallagher, M. (2003) Production of new cells in
the late phase of hippocampal long-term potentiation in vitro the rat dentate gyrus over the lifespan: relation to cognitive
and are attenuated by drugs that enhance the cAMP signaling decline. Eur. J. Neurosci., 18: 215–219.
pathway. Proc. Natl. Acad. Sci. U.S.A., 96: 5280–5285. Bizon, J.L., Lee, H.J. and Gallagher, M. (2004) Neurogenesis in
Badiali de Giorgi, L., Bonvicini, F., Bianchi, D., Bossoni, G. a rat model of age-related cognitive decline. Aging Cell, 3:
and Laschi, R. (1987) Ultrastructural aspects of ageing rat 227–234.
hippocampus and effects of L-acetyl-carnitine treatment. Bliss, T.V.P. and Gardner-Medwin, A.R. (1973) Long-lasting
Drugs Exp. Clin. Res., 13: 185–189. potentiation of synaptic transmission in the dentate area of
Ball, M.J. (1977) Neuronal loss, neurofibrillary tangles and the unanaesthetised rabbit following stimulation of the per-
granulovacuolar degeneration in the hippocampus with age- forant path. J. Physiol., 232: 357–374.
ing and dementia. A quantitative study. Acta Neuropathol. Bliss, T.V.P. and Lomo, T. (1973) Long-lasting potentiation of
(Berl.), 37: 111–118. synaptic transmission in the dentate area of the anesthetized
Barnes, C.A. (1979) Memory deficits associated with senes- rabbit following stimulus of perforant path. J. Physiol., 232:
cence: a neurophysiological and behavioral study in the rat. J. 331–356.
Comp. Physiol. Psychol., 93: 74–104. von Bohlen und Halbach, O., Zacher, C., Gass, P. and Un-
Barnes, C.A. (1988) Spatial learning and memory processes: the sicker, K. (2006) Age-related alterations in hippocampal
search for their neurobiological mechanisms in the rat. spines and deficiencies in spatial memory in mice. J. Neuro-
Trends Neurosci., 11: 163–169. sci. Res., 83: 525–531.
Barnes, C.A. (1994) Normal aging: regionally specific changes Brody, H. (1955) Organization of the cerebral cortex. III. A
in hippocampal synaptic transmission. Trends Neurosci., 17: study of aging in the human cerebral cortex. J. Comp. Ne-
13–18. urol., 102: 511–556.
Barnes, C.A. and McNaughton, B.L. (1979) Neurophysiolog- Burke, S.N. and Barnes, C.A. (2006) Neural plasticity in the
ical comparison of dendritic cable properties in adolescent, ageing brain. Nat. Rev. Neurosci., 7: 30–40.
middle-aged, and senescent rats. Exp. Aging Res., 5: 195–206. Cabeza, R., Nyberg, L. and Park, D. (2005) Cognitive Neuro-
Barnes, C.A. and McNaughton, B.L. (1980a) Physiological science of Aging. Oxford University Press, New York.
compensation for loss of afferent synapses in rat hippocam- Chawla, M.K., Guzowski, J.F., Ramirez-Amaya, V., Lipa, P.,
pal granule cells during senescence. J. Physiol., 309: 473–485. Hoffman, K.L., Marriott, L.K., Worley, P.F., McNaughton,
Barnes, C.A. and McNaughton, B.L. (1980b) Spatial memory B.L. and Barnes, C.A. (2005) Sparse, environmentally selec-
and hippocampal synaptic plasticity in senescent and middle- tive expression of Arc RNA in the upper blade of the rodent
aged rats. In: Stein D. (Ed.), The Psychobiology of Aging: fascia dentata by brief spatial experience. Hippocampus, 15:
Problems and Perspectives. Elsevier Press, New York, 579–586.
pp. 253–272. Chawla, M.K., Lin, G., Olson, K., Vazdarjanova, A., Burke,
Barnes, C.A., McNaughton, B.L., Mizumori, S.J., Leonard, S.N., McNaughton, B.L., Worley, P.F., Guzowski, J.F.,
B.W. and Lin, L.H. (1990) Comparison of spatial and tem- Roysam, B. and Barnes, C.A. (2004) 3D-catFISH: a system
poral characteristics of neuronal activity in sequential stages for automated quantitative three-dimensional compartmental
of hippocampal processing. Prog. Brain Res., 83: 287–300. analysis of temporal gene transcription activity imaged by flu-
Barnes, C.A., Meltzer, J., Houston, F., Orr, G., McGann, K. orescence in situ hybridization. J. Neurosci. Methods, 139:
and Wenk, G.L. (2000b) Chronic treatment of old rats with 13–24.
donepezil or galantamine: effects on memory, hippocampal Cole, A.J., Saffen, D.W., Baraban, J.M. and Worley, P.F.
plasticity and nicotinic receptors. Neuroscience, 99: 17–23. (1989) Rapid increase of an immediate early gene messenger
Barnes, C.A., Rao, G. and Houston, F.P. (2000a) LTP induc- RNA in hippocampal neurons by synaptic NMDA receptor
tion threshold change in old rats at the perforant path — activation. Nature, 340: 474–476.
granule cell synapse. Neurobiol. Aging, 21: 613–620. Coleman, P.D. and Flood, D.G. (1987) Neuron numbers and
Barnes, C.A., Rao, G. and McNaughton, B.L. (1987) Increased dendritic extent in normal aging and Alzheimer’s disease.
electrotonic coupling in aged rat hippocampus: a possible Neurobiol. Aging, 8: 521–545.
mechanism for cellular excitability changes. J. Comp. Ne- Coleman, P.D., Flood, D.G. and West, M.J. (1987) Volumes of
urol., 259: 549–558. the components of the hippocampus in the aging F344 rat. J.
Barnes, C.A., Treves, A., Rao, G. and Shen, J. (1994) Electro- Comp. Neurol., 266: 300–306.
physiological markers of cognitive aging: region specificity and Collingridge, G.L. (1985) Long term potentiation in the hip-
computational consequences. Semin. Neurosci., 6: 359–367. pocampus: Mechanisms of initiation and modulation by ne-
Baskys, A., Niesen, C.E. and Carlen, P.L. (1987) Altered mod- urotransmitters. Trends Pharmacol. Sci., 6: 407–411.
ulatory actions of serotonin on dentate granule cells of aged Deadwyler, S.A., Gribkoff, V., Cotman, C. and Lynch, G.
rats. Brain Res., 419: 112–118. (1976) Long lasting changes in the spontaneous activity of
Bertoni-Freddari, C., Giuli, C., Pieri, C. and Paci, D. (1986) hippocampal neurons following stimulation of the entorhinal
Quantitative investigation of the morphological plasticity of cortex. Brain. Res. Bull., 1: 1–7.
676

Douglas, R.M. and Goddard, G.V. (1975) Long-term potent- Gothard, K.M., Hoffman, K.L., Battaglia, F.P. and McN-
iation of the perforant path-granule cell synapse in the rat aughton, B.L. (2001) Dentate gyrus and ca1 ensemble activity
hippocampus. Brain Res., 86: 205–215. during spatial reference frame shifts in the presence and ab-
Dragunow, M., Abraham, W.C., Goulding, M., Mason, S.E., sence of visual input. J. Neurosci., 21: 7284–7292.
Robertson, H.A. and Faull, R.L. (1989) Long-term potent- Gould, E. and Cameron, H.A. (1996) Regulation of neuronal
iation and the induction of c-fos mRNA and proteins in the birth, migration and death in the rat dentate gyrus. Dev.
dentate gyrus of unanesthetized rats. Neurosci. Lett., 101: Neurosci., 18: 22–35.
274–280. Guzowski, J.F., Lyford, G.L., Stevenson, G.D., Houston,
Drapeau, E., Mayo, W., Aurousseau, C., Le Moal, M., Piazza, F.P., McGaugh, J.L., Worley, P.F. and Barnes, C.A. (2000)
P.V. and Abrous, D.N. (2003) Spatial memory performances Inhibition of activity-dependent arc protein expression in the
of aged rats in the water maze predict levels of hippocampal rat hippocampus impairs the maintenance of long-term po-
neurogenesis. Proc. Natl. Acad. Sci. U.S.A., 100: 14385–14390. tentiation and the consolidation of long-term memory.
Driscoll, I., Howard, S.R., Stone, J.C., Monfils, M.H., To- J. Neurosci., 20: 3993–4001.
manek, B., Brooks, W.M. and Sutherland, R.J. (2006) The Guzowski, J.F., McNaughton, B.L., Barnes, C.A. and Worley,
aging hippocampus: a multi-level analysis in the rat. Neuro- P.F. (1999) Environment-specific expression of the immedi-
science, 139: 1173–1185. ate-early gene Arc in hippocampal neuronal ensembles. Nat.
Driscoll, I. and Sutherland, R.J. (2005) The aging hippocam- Neurosci., 2: 1120–1124.
pus: navigating between rat and human experiments. Rev. Hasselmo, M.E. (1999) Neuromodulation: acetylcholine and
Neurosci., 16: 87–121. memory consolidation. Trends Cogn. Sci., 3: 351–359.
Eriksson, P.S. (2003) Neurogenesis and its implications for re- Hasselmo, M.E., Wyble, B.P. and Wallenstein, G.V. (1996)
generation in the adult brain. J. Rehabil. Med., 41(Suppl.): Encoding and retrieval of episodic memories: role of
17–19. cholinergic and GABAergic modulation in the hippocam-
Foster, T.C., Barnes, C.A., Rao, G. and McNaughton, B.L. pus. Hippocampus, 6: 693–708.
(1991) Increase in perforant path quantal size in aged F-344 Hebb, D.O. (1949) The Organization of Behavior. Wiley, New
rats. Neurobiol. Aging, 12: 441–448. York.
Gallagher, M. and Rapp, P.R. (1997) The use of animal models Hoff, S.F., Scheff, S.W., Bernardo, L.S. and Cotman, C.W.
to study the effects of aging on cognition. Annu. Rev. Psy- (1982) Lesion-induced synaptogenesis in the dentate gyrus of
chol., 48: 339–370. aged rats: I. Loss and reaquisition of normal synaptic den-
Geinisman, Y. (1979) Loss of axosomatic synapses in the dent- sity. J. Comp. Neurol., 205: 246–252.
ate gyrus of aged rats. Brain Res., 168: 485–492. Huerta, P.T. and Lisman, J.E. (1993) Heightened synaptic
Geinisman, Y. and Bondareff, W. (1976) Decrease in the plasticity of hippocampal CA1 neurons during a cholinergic-
number of synapses in the senescent brain: A quantitative ally induced rhythmic state. Nature, 364: 723–725.
electron microscopic analysis of the dentate gyrus molecular Hyder, F., Kida, I., Behar, K.L., Kennan, R.P., Maciejewski,
layer in the rat. Mech. Aging Dev., 5: 11–23. P.K. and Rothman, D.L. (2001) Quantitative functional im-
Geinisman, Y., Bondareff, W. and Dodge, J.T. (1977) Partial aging of the brain: towards mapping neuronal activity by
deafferntation of neurons in the dentate gyrus of the senes- BOLD fMRI. NMR Biomed., 14: 413–431.
cent rat. Brain Res., 134: 541–545. Ikonen, S., McMahan, R., Gallagher, M., Eichenbaum, H. and
Geinisman, Y., Bondareff, W. and Dodge, J.T. (1978) Dendritic Tanila, H. (2002) Cholinergic system regulation of spatial
atrophy in the dentate gyrus of the senescent rat. Am. J. representation by the hippocampus. Hippocampus, 12:
Anat., 152: 321–329. 386–397.
Geinisman, Y., de Toledo-Morrell, L. and Morrell, F. (1986) Jefferys, J.G.R. (1975) Propagation of action potentials into the
Loss of perforated synapses in the dentate gyrus: morpho- dendrite of hippocampal granule cells in vitro. J. Physiol.,
logical substrate of memory deficit in aged rats. Proc. Natl. 249: 16–18.
Acad. Sci. U.S.A., 83: 3027–3031. Jung, M.W. and McNaughton, B.L. (1993) Spatial selectivity of
Geinisman, Y., de Toledo-Morrell, L., Morrell, F., Persina, I.S. unit activity in the hippocampal granular layer. Hippocam-
and Rossi, M. (1992) Age-related loss of axospinous synapses pus, 3:165–182.
formed by two afferent systems in the rat dentate gyrus as Kadar, T., Arbel, I., Silbermann, M. and Levy, A. (1994) Mor-
revealed by the unbiased stereological dissector technique. phological hippocampal changes during normal aging and
Hippocampus, 2: 437–444. their relation to cognitive deterioration. J. Neural. Transm.
Gonzalez, R.G., Fischman, A.J., Guimaraes, A.R., Carr, C.A., Suppl., 44: 133–143.
Stern, C.E., Halpern, E.F., Growdon, J.H. and Rosen, B.R. Kempermann, G. and Gage, F.H. (1999) Experience-dependent
(1995) Functional MR in the evaluation of dementia: corre- regulation of adult hippocampal neurogenesis: effects of
lation of abnormal dynamic cerebral blood volume measure- long-term stimulation and stimulus withdrawal. Hippocam-
ments with changes in cerebral metabolism on positron pus, 9: 321–332.
emission tomography with fludeoxyglucose F 18. AJNR Am. Kempermann, G., Gast, D. and Gage, F.H. (2002) Neuroplas-
J. Neuroradiol., 16: 1763–1770. ticity in old age: sustained fivefold induction of hippocampal
677

neurogenesis by long-term environmental enrichment. Ann. behaviorally characterized aged rats. J. Comp. Neurol., 438:
Neurol., 52: 135–143. 445–456.
Keuker, J.I., Luiten, P.G. and Fuchs, E. (2000) Capillary Merrill, D.A., Roberts, J.A. and Tuszynski, M.H. (2000) Con-
changes in hippocampal CA1 and CA3 areas of the aging servation of neuron number and size in entorhinal cortex
rhesus monkey. Acta Neuropathol. (Berl.), 100: 665–672. layers, I.I., III, and V/VI of aged primates. J. Comp. Neurol.,
Kuhn, H.G., Dickinson-Anson, H. and Gage, F.H. (1996) Ne- 422: 396–401.
urogenesis in the dentate gyrus of the adult rat: age-related Mizumori, S.J., Barnes, C.A. and McNaughton, B.L. (1989)
decrease of neuronal progenitor proliferation. J. Neurosci., Reversible inactivation of the medial septum: selective effects
16: 2027–2033. on the spontaneous unit activity of different hippocampal cell
Lanahan, A., Lyford, G., Stevenson, G.S., Worley, P.F. and types. Brain Res., 500: 99–106.
Barnes, C.A. (1997) Selective alteration of long-term potent- Mizumori, S.J., Barnes, C.A. and McNaughton, B.L. (1992)
iation-induced transcriptional response in hippocampus of Differential effects of age on subpopulations of hippocampal
aged, memory-impaired rats. J. Neurosci., 17: 2876–2885. theta cells. Neurobiol. Aging, 13: 673–679.
Landfield, P.W. (1988) Hippocampal neurobiological mecha- Montarolo, P.G., Goelet, P., Castellucci, V.F., Morgan, J.,
nisms of age-related memory dysfunction. Neurobiol. Aging, Kandel, E.R. and Schacher, S. (1986) A critical period for
9: 571–579. macromolecular synthesis in long-term heterosynaptic facil-
Leutgeb, J.K., Leutgeb, S., Moser, M.B. and Moser, E.I. (2007) itation in Aplysia. Science, 234: 1249–1254.
Pattern separation in the dentate gyrus and CA3 of the Morris, R.G.M., Anderson, E., Lynch, G.S. and Baudry, M.
hippocampus. Science, 315: 961–966. (1986) Selective impairment of learning and blockade of long-
Link, W., Konietzko, U., Kauselmann, G., Krug, M., Schwa- term potentiation by an N-methyl-D-aspartate receptor an-
nke, B., Frey, U. and Kuhl, D. (1995) Somatodendritic ex- tagonist, AP5. Nature, 319: 774–776.
pression of an immediate early gene is regulated by synaptic Morrison, J.H. and Hof, P.R. (1997) Life and death of neurons
activity. Proc. Natl. Acad. Sci. U.S.A., 92: 5734–5738. in the aging brain. Science, 278: 412–419.
Luebke, J.I. and Rosene, D.L. (2003) Aging alters dendritic Nguyen, P.V., Abel, T. and Kandel, E.R. (1994) Requirement
morphology, input resistance, and inhibitory signaling in of a critical period of transcription for induction of a late
dentate granule cells of the rhesus monkey. J. Comp. Neurol., phase of LTP. Science, 265: 1104–1107.
460: 573–584. Nicolle, M.M., Gallagher, M. and McKinney, M. (1999) No
Lyford, G.L., Yamagata, K., Kaufmann, W.E., Barnes, C.A., loss of synaptic proteins in the hippocampus of aged, be-
Sanders, L.K., Copeland, N.G., Gilbert, D.J., Jenkins, N.A., haviorally impaired rats. Neurobiol. Aging, 20: 343–348.
Lanahan, A.A. and Worley, P.F. (1995) Arc, a growth factor Niesen, C.E., Baskys, A. and Carlen, P.L. (1988) Reversed et-
and activity-regulated gene, encodes a novel cytoskeleton- hanol effects on potassium conductances in aged hippocam-
associated protein that is enriched in neuronal dendrites. pal dentate granule neurons. Brain Res., 445: 137–141.
Neuron, 14: 433–445. Peters, A., Rosene, D.L., Moss, M.B., Kemper, T.L., Abraham,
MacVicar, B.A. and Dudek, F.E. (1982) Electrotonic coupling C.R., Tigges, J. and Albert, M.S. (1996) Neurobiological
between granule cells of rat dentate gyrus: Physiological and bases of age-related cognitive decline in the rhesus monkey. J.
anatomical evidence. J. Neurophysiol., 47: 579–592. Neuropathol. Exp. Neurol., 55: 861–874.
Magnusson, K.R., Nelson, S.E. and Young, A.B. (2002) Age- van Praag, H., Schinder, A.F., Christie, B.R., Toni, N., Palmer,
related changes in the protein expression of subunits of the T.D. and Gage, F.H. (2002) Functional neurogenesis in the
NMDA receptor. Brain Res. Mol. Brain Res., 99: 40–45. adult hippocampus. Nature, 415: 1030–1034.
Magnusson, K.R., Scruggs, B., Aniya, J., Wright, K.C., Ontl, van Praag, H., Shubert, T., Zhao, C. and Gage, F.H. (2005)
T., Xing, Y. and Bai, L. (2003) Age-related deficits in mice Exercise enhances learning and hippocampal neurogenesis in
performing working memory tasks in a water maze. Behav. aged mice. J. Neurosci., 25: 8680–8685.
Neurosci., 117: 485–495. Ramirez-Amaya, V., Gage, F.H., Worley, P.F. and Barnes,
Marr, D. (1971) Simple memory: a theory for archicortex. Phi- C.A. (2006) Integration of new neurons into functional neu-
los. Trans. R. Soc. B, 262: 23–81. ral networks. J. Neurosci., 26: 12237–12241.
McNaughton, B.L., Douglas, R.M. and Goddard, G.V. (1978) Rao, M.S., Hattiangady, B., Abdel-Rahman, A., Stanley, D.P.
Synaptic enhancement in fascia dentata: cooperativity among and Shetty, A.K. (2005) Newly born cells in the ageing dent-
coactive afferents. Brain Res., 157: 277–293. ate gyrus display normal migration, survival and neuronal
McNaughton, B.L. and Morris, R.G.M. (1987) Hippocampal fate choice but endure retarded early maturation. Eur. J.
synaptic enhancement and information storage within a dis- Neurosci., 21: 464–476.
tributed memory system. Trends Neurosci., 10: 408–415. Rapp, P.R. and Amaral, D.G. (1991) Recognition memory
McWilliams, J.R. and Lynch, G. (1984) Synaptic density and deficits in a subpopulation of aged monkeys resemble the
axonal aprouting in rat hippocampus: stability in adulthood effects of medial temporal lobe damage. Neurobiol. Aging,
abd decline in late adulthood. Brain Res., 294: 152–156. 12: 481–486.
Merrill, D.A., Chiba, A.A. and Tuszynski, M.H. (2001) Con- Rapp, P.R., Deroche, P.S., Mao, Y. and Burwell, R.D. (2002)
servation of neuronal number and size in entorhinal cortex of Neuron number in the parahippocampal region is preserved
678

in aged rats with spatial learning deficits. Cereb. Cortex, 12: Shors, T.J., Miesegaes, G., Beylin, A., Zhao, M., Rydel, T. and
1171–1179. Gould, E. (2001) Neurogenesis in the adult is involved in the
Rapp, P.R. and Gallagher, M. (1996) Preserved neuron number formation of trace memories. Nature, 410: 372–376.
in the hippocampus of aged rats with spatial learning deficits. Small, S.A., Chawla, M.K., Buonocore, M., Rapp, P.R. and
Proc. Natl. Acad. Sci. U.S.A., 93: 9926–9930. Barnes, C.A. (2004) Imaging correlates of brain function in
Rapp, P.R., Stack, E.C. and Gallagher, M. (1999) Morpho- monkeys and rats isolates a hippocampal subregion differ-
metric studies of the aged hippocampus: I. Volumetric anal- entially vulnerable to aging. Proc. Natl. Acad. Sci. U.S.A.,
ysis in behaviorally characterized rats. J. Comp. Neurol., 403: 101: 7181–7186.
459–470. Small, S.A., Tsai, W.Y., DeLaPaz, R., Mayeux, R. and Stern,
Rasmussen, T., Schliemann, T., Sorensen, J.C., Zimmer, J. and Y. (2002) Imaging hippocampal function across the human
West, M.J. (1996) Memory impaired aged rats: no loss of life span: is memory decline normal or not? Ann. Neurol., 51:
principal hippocampal and subicular neurons. Neurobiol. 290–295.
Aging, 17: 143–147. Smith, T.D., Adams, M.M., Gallagher, M., Morrison, J.H. and
Ribak, C.E., Korn, M.J., Shan, Z. and Obenaus, A. Rapp, P.R. (2000) Circuit-specific alterations in hippocampal
(2004) Dendritic growth cones and recurrent basal dend- synaptophysin immunoreactivity predict spatial learning im-
rites are typical features of newly generated dentate gran- pairment in aged rats. J. Neurosci., 20: 6587–6593.
ule cells in the adult hippocampus. Brain Res., 1000: Song, H., Kempermann, G., Overstreet Wadiche, L., Zhao, C.,
195–199. Schinder, A.F. and Bischofberger, J. (2005) New neurons in
Rolls, E.T. (1990) Functions of the primate hippocampus in the adult mammalian brain: synaptogenesis and functional
spatial processing and memory. In: Kesner R.P. and Olton integration. J. Neurosci., 25: 10366–10368.
D.S. (Eds.), Neurobiology of Comparative Cognition. Com- Sterio, D.C. (1984) The unbiased estimation of number and
parative Cognition and Neuroscience. Lawrence Erlbaum sizes of arbitrary particles using the disector. J. Microsc., 134:
Associates, Inc., Hillsdale, MI, pp. 339–362. 127–136.
Rose, G., Diamond, D. and Lynch, G.S. (1983) Dentate gran- Wati, H., Kudo, K., Qiao, C., Kuroki, T. and Kanba, S. (2006)
ule cells have the behavioral characteristics of theta neurons. A decreased survival of proliferated cells in the hippocampus
Brain Res., 266: 29–37. is associated with a decline in spatial memory in aged rats.
Rosenzweig, E.S. and Barnes, C.A. (2003) Impact of aging on Neurosci Lett., 399: 171–174.
hippocampal function: plasticity, network dynamics, and Wenk, G.L. and Barnes, C.A. (2000) Regional changes in the
cognition. Prog. Neurobiol., 69: 143–179. hippocampal density of AMPA and NMDA receptors across
Samsonovich, A.V. and Ascoli, G.A. (2006) Morphological the lifespan of the rat. Brain Res., 885: 1–5.
homeostasis in cortical dendrites. Proc. Natl. Acad. Sci. West, M.J. (1993a) New stereological methods for counting
U.S.A., 103: 1569–1574. neurons. Neurobiol. Aging, 14: 275–285.
Shen, J. and Barnes, C.A. (1996) Age-related decrease in West, M.J. (1993b) Regionally specific loss of neurons in the
cholinergic synaptic transmission in three hippocampal sub- aging human hippocampus. Neurobiol. Aging, 14: 287–293.
fields. Neurobiol. Aging, 17: 439–451. West, M.J., Coleman, P.D., Flood, D.G. and Troncoso, J.C.
Shen, J., Barnes, C.A., McNaughton, B.L., Skaggs, W.E. and (1994) Differences in the pattern of hippocampal neuronal
Weaver, K.L. (1997) The effect of aging on experience-de- loss in normal ageing and Alzheimer’s disease. Lancet, 344:
pendent plasticity of hippocampal place cells. J. Neurosci., 769–772.
17: 6769–6782. Worley, P.F., Bhat, R.V., Baraban, J.M., Erickson, C.A.,
Shen, J., Kudrimoti, H.S., McNaughton, B.L. and Barnes, C.A. McNaughton, B.L. and Barnes, C.A. (1993) Thresholds for
(1998) Reactivation of neuronal ensembles in hippocampal synaptic activation of transcription factors in hippocampus:
dentate gyrus during sleep after spatial experience. J. Sleep correlation with long-term enhancement. J. Neurosci., 13:
Res., 7(Suppl 1): 6–16. 4776–4786.
Plate 33.1. Wiring diagram of the hippocampus (interneurons excluded) showing dual inputs from the lateral and medial entorhinal
cortex. The layer 2 cortical inputs to the dentate and CA3 diverge fan out (f) widely over these networks and then provide convergent
input to individual dentate granule cells. In contrast, the layer 3 cortical inputs are specialized for individual subregions of CA1. This is
an example of a point-to-point (p-p) connection. Within the hippocampus, granule cells provide input, via mossy fibers to the mossy
cells of the dentate and to CA3 cells. CA3 cells make feedback connections to themselves and to the mossy cells of the dentate. These,
in turn, provide excitatory input to granule cells in the inner third of the granule cell dendritic tree. (For B/W version, see page 616 in
the volume.)

Plate 36.1. (A) Two representative granule cells filled with 5,6 carboxyfluorescein from the dentate gyrus of a 24-month-old rat. (B)
Histograms showing the numbers of carboxyfluorescein injections that resulted in single, double, triple, or greater numbers of granule
cells filled with dye. Aged rats showed significantly increased electrotonic coupling compared to young animals. This may account for
the increased excitability of old granule cells. Adapted with modifications from Barnes et al. (1987). (For B/W version, see page 665 in
the volume.)
Plate 36.5. Image of granule cells (blue) following in situ hybridization using a c-RNA probe for the IEG Arc. (A) Arc positive cells
(shown in red) in the granule cell layer from the upper blade of a young rat following two 5 min (separated by a 20 min rest period in
the home cage) episodes of spatial exploratory behavior. (B) Granule cells from a caged control animal of the same age, that did not
perform the exploratory behavior, are shown, illustrating baseline or negligible expression of Arc mRNA. (For B/W version, see page
672 in the volume.)
Plate 36.6. (A) Examples of gadolinium-induced change in MRI signal, (a measure of cerebral blood flow volume, CBV, that is a
correlate of brain metabolism) from a young and old rhesus monkey. Warm colors indicate higher CBV values. The white circle,
identified in precontrast images, overlies the dentate gyrus. (B) CBV measurements plotted vs. age in the dentate gyrus indicates a
significant decline with advancing age. (C) CBV measurements in the dentate gyrus of old monkeys plotted against memory per-
formance on the delayed nonmatching-to-sample test. A significant relationship between CBV and memory performance is observed,
selectively in the old animals. Adapted with modifications from Small et al. (2004). (For B/W version, see page 673 in the volume.)

Plate 36.7. Behavior-induced Arc mRNA expression pattern in young and old rat granule cells of the dentate gyrus. (A) Individual
examples of behaviorally induced Arc RNA. Fluorescence-tagged CY3 Arc RNA containing cells are green, and granule cell nuclei are
shown in red (counterstained with propydium iodide, a nucleic acid stain). (B) The average proportion of Arc-positive neurons (% of
total cells) measured from the three different age groups (9, 15, and 24 months) is shown for the dentate gyrus. Note that the dentate
gyrus of old rats contains a significantly lower proportion of behaviorally induced Arc-positive granule cells compared to young or
middle-aged animals. (For B/W version, see page 674 in the volume.)
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 37

The effects of aging on dentate circuitry and function

Peter R. Patrylo1, and Anne Williamson2

1
Department of Physiology, Southern Illinois University School of Medicine Carbondale, IL 62901, USA
2
Department of Neurosurgery, Yale University School of Medicine New Haven, CT 06520, USA

Abstract: The central nervous system (CNS) undergoes a variety of anatomic, physiologic, and behavioral
changes during aging. One region that has received a great deal of attention is the hippocampal formation
due to the increased incidence of impaired spatial learning and memory with age. The hippocampal
formation is also highly susceptible to Alzheimer’s disease, ischemia/hypoxia, and seizure generation, the
three most common aging-related neurological disorders. While data reveal that the dentate gyrus plays a
key role in hippocampal function and dysfunction, the majority of electrophysiological studies that have
examined the effects of age on the hippocampal formation have focused on CA3 and CA1. We perceive this
to be an oversight and consequently will highlight data in this review which demonstrate an age-related
disruption in dentate circuitry and function, and propose that these changes contribute to the decline in
hippocampal-dependent behavior seen with ‘‘normal’’ aging.

Keywords: dentate gating; mossy fibers; recurrent excitation; IPSP; epileptogenicity; spatial learning and
memory; aging

Dentate filter function function comes from several experimental findings.


First, 2-deoxyglucose studies and electrophysiolog-
‘‘Normal’’ and pathologic hippocampal function ical recordings have shown that seizures are unlikely
(e.g., spatial learning and memory, and epileptiform to invade the hippocampus unless epileptiform ac-
activity, respectively) are both associated with an tivity is generated within the dentate gyrus itself
increase in excitatory activity between principal cells. (Heinemann et al., 1992; Lothman et al., 1992).
Determining whether ‘‘normal’’ or pathological Second, low frequency orthodromic activation of
function occurs has been proposed to depend, in dentate granule cells normally suppresses CA3 py-
part, on the dentate gyrus due to it’s strategic po- ramidal cell activity (Bragin et al., 1995; Penttonen
sition within the tri-synaptic circuit of the hippo- et al., 1997) since granule cell axons preferentially
campal formation that allows it to ‘‘filter/gate’’ the form synapses with inhibitory interneurons in CA3
majority of excitatory input entering hippocampal (Acsady et al., 1998). If granule cell activity is in-
region CA3, a region highly susceptible to epilepti- creased, however, glutamatergic excitation can be
form activity (Johnson and Brown, 1984; Miles and evoked in CA3 due to greater frequency facilitation
Wong, 1987; Christian and Dudek, 1988). Evidence at synapses between granule cell axons (mossy
that the dentate gyrus subserves a gating/filter fibers) and pyramidal neurons (Henze et al., 2002;
Lawrence and McBain, 2003). Third, several groups
Corresponding author. Tel.: +1618 453 6743; have shown that, if the intrinsic properties or
Fax: +1618 453 1517; E-mail: ppatrylo@siumed.edu connectivity of the dentate gyrus are altered

DOI: 10.1016/S0079-6123(07)63037-4 679


680

experimentally or naturally (e.g., insult or genetic 344 (21–32 months) rats. The Fischer 344 rat is one
variability), hippocampal-dependent function is of the most commonly used strain of rodents in
affected adversely including, learning and memory neurobiology of aging research, and the ages used
(Lipp et al., 1984; Crusio et al., 1987; McNaughton reflects the convention of classifying rodents as
et al., 1989; Nanry et al., 1989; Bernasconi-Guastalla aged when approximately 50% of their cohorts
et al., 1994; Schuster et al., 1997), synaptic plasticity have perished. To ensure that the activity elicited
(Beck et al., 2000), and seizure susceptibility (Grimes in CA3 was disynaptic in nature, and thus a re-
et al., 1990; Cronin et al., 1992; Grecksch et al., flection of dentate filtering, only responses with a
1995; Wuarin and Dudek, 1996; Patrylo and Dudek, delay Z2 ms were examined (monosynaptic inter-
1998). connections normally exhibit a delay r2 ms).
The cellular and circuit properties that bestow These experiments revealed that during a train of
the dentate gyrus with the capacity to ‘‘filter/gate’’ 5 Hz molecular layer stimulation (theta frequency)
excitatory activity are unknown. However this re- hyperexcitable activity (multiple population
gion does exhibit several unique physiological spikes) occurs in area CA3 in approximately
properties that are likely to contribute to this 30% of aged slices (Fig. 1C). These slices also fre-
function. Among these properties are: (1) the prin- quently exhibited spontaneous interictal-like ac-
cipal cells (granule cells) having a relatively hyper- tivity in CA3 following termination of the train. In
polarized resting membrane potential and strong contrast, 100% of adult slices and the remaining
spike frequency adaptation (Fricke and Prince, 70% of aged slices exhibited either suppression or
1984; Staley et al., 1992); (2) the capacity of dent- no change in the amplitude of the activity evoked
ate granule cells to fire normally in single spike during the train and no subsequent epileptiform
mode following disinhibition versus a graded burst activity. Further, this compromise in dentate filter
discharge (Schwartzkroin and Prince, 1978; Fricke function appeared to be associated with an im-
and Prince, 1984; Lynch et al., 2000); (3) the pres- paired capacity of the dentate gyrus to respond to
ence of strong feedfoward and feedback inhibitory afferent input, since 5 Hz molecular layer stimula-
circuits (Andersen, 1975; Freund and Buzsáki, tion evoked hyperexcitable activity in the dentate
1996); (4) a paucity of recurrent excitatory inter- gyrus in approximately the same percentage of
connections (Schwartzkroin and Prince, 1978; aged slices as were noted to generate hyperexcit-
Fricke and Prince, 1984; Wuarin and Dudek, able activity in CA3 (Patrylo et al., 2007). These
1996; Nadler, 2003); and (5) the dynamic proper- data were also proposed to reflect an aging-related
ties of mossy fiber — CA3 synaptic transmission compromise in inhibitory activity in CA3 since
(e.g., Lawrence and McBain, 2003). modeling studies indicate that epileptiform activity
is generated in hippocampal region CA3 if granule
cell activity is increased and inhibitory tone is
Aging and dentate filter function compromised (Lawrence and McBain, 2003).
This aging-related breakdown of dentate filter
Spatial learning and memory, synaptic plasticity, function poses two questions. (1) What are the un-
and seizure susceptibility are all affected adversely derlying mechanisms for this decline in function? (2)
with age. Moreover, available data suggest that Does this disruption in dentate function contribute
several of the cellular and circuit properties of the to aging-related hippocampal dysfunction (e.g., im-
dentate gyrus are altered during aging (see below) paired spatial learning and memory)? While, the
and consequently it raises the question whether answers to these questions are currently unclear, in
dentate filter function is also comprised. To begin the subsequent sections we will highlight aging-re-
to determine the effects of age on dentate filter lated changes that are likely to contribute to this
function, Patrylo et al. (2007) examined field po- decline in dentate filter function, and propose that
tential activity elicited in hippocampal region CA3 this change in function contributes to the decline in
with dentate molecular layer stimulation in brain spatial learning and memory and enhanced suscep-
slices from adult (2–11 months) and aged Fisher tibility for complex partial seizures during aging.
681

Fig. 1. Dentate filter function is impaired in a subpopulation of aged rodents. (A) Schematic diagram illustrating the position of the
recording and stimulating electrodes used to assess dentate filter function. Note that the recordings were performed in CA3. (B) The
experimental paradigm illustrated schematically. (C) In adult slices, 5 Hz molecular layer stimulation elicited minimal change in the
evoked response, recorded from CA3. In contrast, similar stimulation could elicit hyperexcitability in approximately 30% of aged
slices. (D) The majority of aged slices exhibiting multiple population spikes during 5 Hz molecular layer stimulation developed
spontaneous epileptiform activity subsequently.

Potential mechanisms contributing to the aging- is illustrated in Table 1. To summarize briefly, ag-
related decline in dentate filter function ing is associated with: (1) a change in synaptic
connectivity (e.g., Geinisman et al., 1992; Patrylo
Numerous changes have been reported to occur in et al., 2007); (2) an increase in dye coupling be-
the dentate gyrus during aging and a partial listing tween granule cells (Barnes et al., 1987); (3) a
682

change in receptor and channel properties (e.g., During aging, the inhibitory system in the dent-
Gutierrez et al., 1996; Magnusson et al., 2006); (4) ate gyrus exhibits a variety of anatomic and bio-
a compromise in synaptic plasticity (e.g., Barnes chemical changes that suggest a compromise in
and McNaughton, 1980b; de Toledo-Morrell function. Among these changes are: (1) a decrease
et al., 1988; Barnes, 2000; Barnes et al., 2000); in glutamic acid decarboxylase (GAD; Shetty and
(5) a decrease in extracellular volume (Sykova Turner, 1998; Vela et al., 2003; Ling et al., 2005)
et al., 1998); and (6) a change in the levels of and extracellular GABA levels (Almaguer-Melian
growth factors that can affect neuronal survival, et al., 2005); (2) an increase in GABA transami-
metabolism, and activity (e.g., Katoh-Semba et al., nase (Lumeng and Li, 1974; Kugler and Baier,
1998; Sonntag et al., 1999). While all of these 1992; Hwang et al., 2004); and (3) a decrease in the
changes would be expected to alter dentate net- number of neurons immunopositive for GAD,
work activity and thus filter function, it is beyond and/or other peptides/neuromodulators that co-
the scope of this article to provide a detailed re- localize in inhibitory neurons (Shetty and Turner,
view of each change. Consequently, we will focus 1998; Cadacio et al., 2003; Vela et al., 2003;
on data demonstrating an aging-related change Stanley and Shetty, 2004). It is important to note
in local inhibitory and excitatory circuitry (GAB- however that several other changes occur during
Aergic and glutamatergic respectively) since sim- aging that have been suggested to be compensa-
ilar alterations have been reported in models of tory. For example, an increase in the levels of the
epilepsy and have been associated with a break- GABA a1 subunit has been noted in the hippo-
down of dentate filter function (e.g., Behr et al., campus and dentate gyrus of aged rodents and has
1998; Finnerty et al., 2001; Nadler, 2003). been proposed to be an adaptive response to the
decrease in GAD levels (e.g., Gutierrez et al., 1996)
since incorporation of the a1 subunit into a GABA
GABAergic circuitry receptor/channel complex results in a current that
has a prolonged decay constant (e.g., Fisher,
Interneurons help regulate principal cell excitabil- 2004). Examination of the anatomical data on
ity, receptive field size, and plasticity, and conse- the distribution and number of GAD immunore-
quently play a critical role in information active neurons suggest that not all inhibitory cir-
processing as well as pathophysiology (Freund cuits are affected equally with age. Specifically,
and Buzsáki, 1996; Paulsen and Moser, 1998). In Shetty and Turner (1998) reported that the
the dentate gyrus a tremendous diversity of inter- number of GAD-immunoreactive neurons within
neurons are found (Freund and Buzsáki, 1996) and subjacent to the granule cell layer is preserved
that can be divided into two main categories based with age, while GAD immunopositive neuron
on the distribution of their axonal arbor and sub- number in the hilus is decreased. Based on the
sequently physiological function. The first group characteristic laminar location of specific types of
exhibits an axonal arbor that forms synapses with interneurons (e.g., Freund and Buzsaki, 1996),
the somata, initial dendritic tree, and axon hillock these findings suggest that the basket and chande-
of granule cells and are involved with regulating lier cells, those primarily responsible for somatic
the capacity of granule cells to discharge. The sec- and axonal inhibition, are preserved with age and
ond group exhibits axonal projections that form subsequently suggest that feedback inhibition
synapses with the distal segments of granule cell should be preserved during aging. In agreement
dendrites and subsequently help regulate afferent with this suggestion, the number of parvalbumin-
input (i.e., perforant path input). These two types immunoreactive neurons within and adjacent to
of interneurons play distinct roles in ‘‘normal’’ the granule cell layer is preserved with age (Shetty
hippocampal function (Austin et al., 1989; Moser, and Turner, 1998) and basket and chandelier cells
1996) and are affected differentially in specific ne- co-localize this calcium binding protein (e.g.,
uropathological conditions (e.g., Kobayashi and Soriano et al., 1990; Ribak, 1992; Ribak et al.,
Buckmaster, 2003). 1993; Freund and Buzsáki, 1996). Further,
Table 1. Aging-related changes in the dentate gyrus

Findings Species Exemplary reference

Glutamatergic NT
Entorhinal Decreased axospinous synapse number in the MML Rat – F344 Geinisman et al. (1992)
Decreased ratio of Timm staining volume in the MML/ Rat – Long Evans Rapp et al. (1999)
ML
Decreased pre-synaptic volley Rat – F344 Barnes and McNaughton (1980a)
Assoc/Com Decreased axospinous synapse number in the IML Rat – F344 Geinisman et al. (1992)
Expanded zone of Timm staining in the IML Rat – Long Evans Rapp et al. (1999)
Mossy fibers Elaboration of Timm stained mossy fibers into the GCL Guinea pig Wolfer and Lipp (1995)
and IML — can decrease slightly in very old subjects
Elaboration of Timm stained mossy fibers into the IML Human Cassell and Brown (1984)
AMPA receptors Decreased GluR1 and GluR2 mRNA levels Rat – Wistar Pagliusi et al. (1994)
No significant change in [3H] AMPA binding density Rat – Long Evans Nicolle et al. (1996)
Decreased [3H] AMPA binding density Rat – SD Wenk and Barnes (2000)
KA receptors Decreased [3H]KA binding density in the hilus Rat – Long Evans Nagahara et al. (1993)
Expanded zone, but not density, of [3H]KA binding in the Rat – Long Evans Nagahara et al. (1993); Nicolle et
IML al. (1996)
NMDA receptors Non-significant trend for decreased [3H]NMDA receptor Rat – Long Evans Nicolle et al. (1996)
binding in ML
No significant change in [3H] glutamate binding (AP-5 Rat – SD Wenk and Barnes (2000)
subtraction)
Non-significant trend for decreased NR2B mRNA Rhesus monkey Bai et al. (2004)
Decreased NR2B mRNA; decreased NR1 mRNA in the Mice – C57Bl/6 Magnusson (2000)
ventral DG
Decreased [3H] glutamate and [3H] CPP binding density Mice – C57Bl/6 Magnusson (2000)
Decreased NR2B, but not NR2A and NR1, binding Mice – C57BL/6J Magnusson et al. (2006)
density; NR2B decreased preferentially in the intermediate
DG
Decreased NR1-IR Rhesus monkey Gazzaley et al. (1996)
Miscellaneous Increased probability for prolonged EPSPs in BIC Rat – F344 Patrylo et al. (2007)

683
684
Table 1 (continued )

Findings Species Exemplary reference

GABAergic NT
Cell number Decreased number of GAD-immunopositive neurons Rat – F344 Shetty and Turner, (1998)
Synapse number Decreased axodendritic synapse number in OML (on Rhesus monkey Tigges et al. (1995)
shaft)
Receptors Preserved a1 receptor subunit IR Rhesus monkey Rissman et al. (2006)
Binding Increased a1 (but not b2, b3, and g2) mRNA levels and IR Rat – F344 & SD Gutierrez et al. (1996)
Increased [3H] zolpidem binding Rat – Wistar Ruano et al. (1995)
Electrophysiology Preserved paired-pulse inhibition and polysynaptic evoked Rat – F344 Barnes (1979); Patrylo et al. (2007)
IPSPs (conductances and reversal potentials)
Decreased sIPSP frequency Rat – F344 Patrylo et al. (2007)
Increased mIPSC decay time constant Rhesus monkey Luebke and Rosene (2003)
Increased benzodiazepine potentiation of mIPSC decay Rhesus monkey Luebke and Rosene (2003)
constant
Plasticity
Decreased LTP maintenance Rat – F344 Barnes (1979); Barnes and
McNaughton (1980b)
Decreased capacity for LTP induction at low, but not Rat – F344 Barnes (1979); Barnes et al. (2000);
high, stimulus intensities Orr et al. (2001)
Decreased capacity for perforant path — kindling Rat – F344 de Toledo-Morell et al. (1984)
Decreased BDNF-induced LTP Rat – Wistar Gooney et al. (2004)
Ion channels and intrinsic
properties
Ca2+ Decreased calcium currents (e.g., L-type) Rat – F344 Reynolds and Carlen (1989)
(however see Herman et al., 1998;
Thibault et al., 2001, regarding
CA1)
Intrinsic membrane No change in RMP, Rin, spike width, AP amplitude, and Rat – F344 Barnes and McNaughton (1980a);
properties time constant Niesen et al. (1988); Patrylo et al.
(2007)
No change in AHP and spike frequency adaptation Rat – F344 Niesen et al. (1988)
Other NTs and neuropeptides
NPY Decreased number of NPY immunopositive neurons Rat – F344 Hattiangady et al. (2005); Cadacio
et al. (2003)
Somatostatin Decreased somatostatin mRNA levels and IR Rat – Wistar Vela et al. (2003)
ACh Decreased vesicular ACh transporter IR Rhesus monkey Calhoun et al. (2004)
Decreased ChAT IR Rat – Wistar Lukoyanov et al. (1999)
Decreased amplitude of the slow cholinergic EPSP Rat – F344 Shen and Barnes (1996)
Decreased M2 and M1 receptor binding; trend for Rat – F344 Tayebati et al. (2002)
increased M3 receptor binding
Increased [3H]AF-DX 384 binding (increased M2 levels) Rat – Long Evans Aubert et al. (1995)
Decreased nAChRa4 IR in polymorphic region Mouse – CBA/J Rogers et al. (1998)
Growth factors
BDNF Decreased BDNF activity and IR Rat – F344 Hattiangady et al. (2005)
Decreased BDNF IR Rhesus monkey Hayashi et al. (2001)
Increased BDNF IR in the hilus Rat – SD Katoh-Semba et al. (1998)
Insulin and IGF No change in binding density for the IGF1, IGF2, and Rat – Long Evans Dore et al. (1997)
insulin receptors
NGF and NT3 Unknown whether changes are specifically seen within the
dentate gyrus
Neurogenesis
Decreased rate of basal neurogenesis Rat – F344 Kuhn et al. (1996); Merrill et al.
(2003)
Maturation of newly generated neurons retarded Rat – F344 Rao et al. (2005)
Miscellaneous
Decreased extracellular volume Rat – Wistar Sykova et al. (1998)
Increased dye coupling Rat – F344 Barnes et al. (1987)

ACh, acetylcholine; AF-DX, 5,11-dihydro-11-[[(2-(2-[(dipropylamino) methyl]-1-piperidinyl) ethyl) amino] carbonyl]-6H-pyrido [2, 3-b] [1,4]-benzodiazepin-6-one; AHP, after
hyperpolarization; AMPA, alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid; AP, action potential; AP-5, amino phosphonovalerate; Assoc/Com, associational/commissural; BDNF,
brain-derived neurotrophic factor; BIC, bicuculline methiodide; CPP, [(7)-2-carboxypiperazin-4-yl] propyl-1-phosphonic acid; DG, dentate gyrus; EPSP, excitatory post-synaptic potential;
F344, Fischer 344; GAD, glutamic acid decarboxylase; GCL, granule cell layer; IGF1, insulin-like growth factor-1; IGF2, insulin-like growth factor-2; IML, inner molecular layer; IPSP,
inhibitory post-synaptic potential; IR, immunoreactivity; KA, kainate; LTP, long term potentiation; mIPSC, miniature inhibitory post-synaptic current; ML, molecular layer; MML, middle
molecular layer; NGF, nerve growth factor; NMDA, n-methyl-d-aspartate; NPY, neuropeptide Y; NT, neurotransmission; NT3, neurotrophin-3; OML, outer molecular layer; Rin, input
resistance; RMP, resting membrane potential; SD, Sprague-Dawleys; IPSP, Spontaneous IPSP.

685
686

electrophysiological data from hippocampal slices dendrites and presynaptic afferent fibers play a
from aged rodents have shown that paired-pulse critical role in dentate filter function and encoding
inhibition is preserved in the aged dentate gyrus of new spatial information due to increasing the
(Barnes 1979; Patrylo et al., 2007), and whole cell signal to noise ratio of incoming cortical input
current clamp data reveal no significant change in (Paulsen and Moser, 1998). These findings also
the reversal potentials and conductances of fast suggest that if such dendritic/afferent inhibition is
and slow inhibitory post-synaptic potentials compromised, both functions would be affected
(IPSPs) evoked in granule cells from aged versus adversely. Based on the data presented above, we
adult rodents (Patrylo et al., 2007). propose that inhibitory input onto the distal dend-
Regarding the aging-related decrease in number rites of granule cells, and potentially afferent ter-
of GAD immunopositive neurons within the hilus, minals, is decreased during aging and results in
anatomic and electrophysiological data suggest that abnormal processing of afferent input during spa-
these interneurons are likely to be those involved tial exploration, and subsequently decreases the
with inhibiting the distal dendrites of granule signal to noise ratio below the level required for
cells and subsequently modulating afferent input. transmitting new spatial information. Experiments
Specifically, anatomic studies have shown that the examining dendritic inhibition in behaviorally
number of somatostatin (SOM) and neuropeptide-Y characterized aged animals are required to address
(NPY) immunoreactive interneurons are decreased this possibility, however.
within the hilus with age (Cadacio et al., 2003; Vela Whether the decrease in the number of GAD-IR
et al., 2003) and these types of interneurons are neurons within the hilus of the dentate gyrus of aged
known to have an axonal arborization that targets rodents reflects a loss of SOM and NPY containing
the distal dendrites of granule cells (e.g., Leranth interneurons, a change in their properties (decreased
et al., 1990; Freund and Buzsáki, 1996; Katona protein levels), or both, is unclear with currently
et al., 1999; Maccaferri et al., 2000). Further, whole available data supporting either possibility. For
cell patch recordings from hippocampal slices from example, studies examining total hilar neuron
adult and aged Fisher 344 rats have shown that the number reveal that there is no aging-related change
overall frequency of spontaneous IPSPs (sIPSPs) is (Rasmussen et al., 1996), while studies examining
decreased in aged granule cells, although the fre- the number of NPY immunoreactive neurons sug-
quency of sIPSPs with a large amplitude and pro- gest that a relatively small subpopulation of neurons
longed duration is preserved (Patrylo et al., 2007). may be lost (Cadacio et al., 2003; Vela et al., 2003).
Since large amplitude composite inhibitory events It should be noted, however, that attention to ac-
are believed to arise from input impinging on the curate immunohistochemical cell quantification is
soma and proximal dendrites of granule cells, while critical, since the level of an immunohistochemical
smaller unitary events are believed to result from marker may simply decrease below the level re-
more distal input (Soltèsz et al., 1995; Williams et quired for visual detection without a loss of the cell
al., 1998), this aging-related decrease in sIPSP fre- expressing the protein. In this regard, it is interesting
quency is likely to reflect a reduction of inhibitory to note that Stanley and Shetty (2004) have reported
input onto the distal dendrites of granule cells (Pat- that the loss of GAD-IR neurons in hippocampal
rylo et al., 2007). In vitro studies on hippocampal regions CA3 and CA1 does not correspond with a
slices from aged primates have also noted a trend for decrease in total neuron number, and that several
reduced IPSC frequency, although this change did groups have shown that GAD, SOM, and NPY
not reach significance (Luebke and Rosene, 2003). protein expression can be regulated by afferent
It is interesting to note that afferent input is input and neurotrophic levels (Akhtar and Land,
suppressed during theta wave activity (Moser, 1991; Hyman et al., 1994; Carnahan and Nawa,
1996), while feedback inhibition onto the soma/ 1995; Patel et al., 1995; Aamodt et al., 2000;
axon of granule cells is decreased (Austin et al., Marty, 2000; Reibel et al., 2000; de Almeida et al.,
1989; Moser, 1996). Consequently, it has been 2002) which may be altered with age (e.g., Geinis-
proposed that inhibitory input onto distal man et al., 1992; Katoh-Semba et al., 1998; Patrylo
687

et al., 2007). Consequently, additional studies are unlikely to contribute to this change since neuron
required to determine whether aging is associated number in the entorhinal cortex is preserved in
with a loss of interneurons in the dentate gyrus, or a aged animals (Merrill et al., 2001; Rapp et al.,
change in protein expression and/or function. 2002). Subsequent studies using synaptophysin
Regardless, these aging-related changes in in- immunohistochemistry (Smith et al., 2000) have
hibitory efficacy are likely to affect granule cell provided corroboratory findings. Specifically,
activity and synchronization, and consequently Rapp and colleagues noted that there is a decrease
their capacity to respond to afferent input. If this in the level of synaptophysin immunostaining in
decline in inhibitory efficacy is further coupled the molecular layer of the dentate gyrus, as well as
with an enhancement of NMDA-receptor medi- stratum lacunosum moleculare of CA3 in aged
ated activity, or an increase in recurrent excitatory rodents with cognitive impairment. Although this
interconnections, one would predict a breakdown study did not determine whether this change in
of the dentate gate (e.g., Lynch et al., 2000; Finn- synaptophysin immunohistochemistry reflects a
erty et al., 2001; Nadler, 2003). During aging it is loss or modification of synapses, the results do re-
unlikely that an enhancement of NMDA receptor- veal an aging-related change in a pre-synaptic
mediated activity in the dentate gyrus contributes marker that is found preferentially in cognitively
to the decline in filter function, since decreases in impaired rodents. Based on these results, Rapp
the protein levels of certain NMDA receptor sub- and colleagues proposed that the decline in spatial
units (Gazzaley et al., 1996; Nicolle et al., 1996; learning and memory during aging results from the
Magnusson, 2000; Bai et al., 2004; Magnusson et disruption in information flow from the entorhinal
al., 2006), the strength of NMDA-receptor medi- cortex to the hippocampus. It should be noted
ated activity (Rao et al., 1994), and NMDA-re- however, that the decrease in axospinous synapse
ceptor mediated plasticity have been reported with number in the molecular layer of aged rodents
age (Barnes and McNaughton, 1980b; de Toledo- could also reflect a decrease in excitatory input
Morrell et al., 1988; Barnes et al., 2000). As for a onto inhibitory interneurons with dendritic spines
change (increase) in local excitatory interconnec- (Soriano and Frotscher, 1993; Freund and Buz-
tivity contributing to the compromise in dentate sáki, 1996; Mott et al., 1997), although one could
filter function, we will review data in the next sec- argue that these are few in number compared to
tion that address this possibility. granule cell dendrites. Further, a decrease of in-
hibitory input onto granule cell dendritic spines is
unlikely to be a significant contributor because
Glutamatergic circuitry GABAergic synapses are not typically axospinous
(Halasy and Somogyi, 1993; Soriano and Frotsc-
In the early 1990s, Geinisman et al. (1992) re- her, 1993; Freund and Buzsáki, 1996; Katona et
ported that axospinous synapses number is de- al., 1999).
creased in the molecular layer (middle and inner In the adult CNS, a loss of entorhinal input into
molecular layers) of the dentate gyrus in aged ro- the dentate gyrus (e.g., unilateral lesion) results in
dents and is associated with the cognitive status of a transient compromise in cognitive function and
the animal (loss occurred in cognitively impaired an increased propensity for epileptiform activity
subjects). This loss of axospinous synapses is be- (e.g., Steward, 1982; Kelley and Steward, 1996).
lieved to reflect a decrease in entorhinal and asso- These effects eventually reverse, however, due to
ciational/commissural input onto granule cell reorganization of the remaining excitatory circuits,
dendrites, because electrophysiological studies in- in particular the temporodentate pathway arising
dicate that the amplitude of the pre-synaptic volley from the contralateral entorhinal cortex (e.g.,
and resulting EPSP are decreased in the dentate Steward, 1982; Reeves and Smith, 1987). Asso-
gyrus of aged rats (e.g., Barnes and McNaughton, ciational/commissural fibers and the mossy fibers
1980a, b). Further, a loss of presynaptic neurons have also been shown to reorganize following a
(layer II neurons in the entorhinal cortex) is loss of entorhinal input (e.g., Lynch et al., 1976;
688

Steward et al., 1976; Steward, 1982, 1992; West With regard to the CSD studies (Barnes and
and Dewey, 1984; Reeves and Smith, 1987). Dur- McNaughton, 1983) however, this expansion is
ing aging, the loss of entorhinal input is bilateral in unlikely to produce the novel net inward current in
nature, and therefore suggests that the capacity for the granule cell and inner molecular layer of the
restorative reorganization is likely to be reduced. aged dentate gyrus since even though granule
This raises the question whether there is evidence cells form a feedback circuit with hilar mossy
for reorganization of the associational/commis- cells and CA3 pyramidal neurons in the in vivo
sural fibers and/or mossy fibers in the aged dentate dentate gyrus (Ishizuka et al., 1990; Li et al., 1994;
gyrus. Reorganization of these pathways would Buckmaster et al., 1996; Wenzel et al., 1997) the
increase recurrent excitatory interconnections majority of these interconnections are severed dur-
within the hippocampus and thus potentially affect ing the preparation of transverse hippocampal
dentate filter function as has been suggested in slices (Ishizuka et al., 1990; however see Scharf-
models of temporal lobe epilepsy (e.g., Behr et al., man, 1994; Buckmaster et al., 1996; Wenzel et al.,
1998; Finnerty et al., 2001; Nadler, 2003). 1997). Thus, unless the distribution of the associ-
Electrophysiological studies suggest that local ation/commissural fibers is altered along the lon-
excitatory circuits do reorganize within the dentate gitudinal axis of the hippocampus during aging, it
gyrus with age. Current source density (CSD) stud- is unlikely that this reorganization accounts for the
ies have shown that antidromic stimulation elicits increase in excitatory activity in hippocampal
an additional peak of net inward current in the slices from aged rodents. Moreover, Geinisman
granule cell and inner molecular layer of the dentate et al. (1992) suggest that there is a decrease in
gyrus in slices from aged, but not adult, rodents associational/commissural — granule cell axospi-
(Barnes and McNaughton, 1983). While the reason nous synapses within the inner molecular layer. An
for this additional current sink is unknown, it is alternate explanation for the net inward current in
consistent with the presence of additional synaptic the granule cell and inner molecular layer is that
input onto the soma and proximal dendrites of the distribution of mossy fibers changes with age.
granule cells. Furthermore, whole cell patch clamp In this regard, studies that have labeled the mossy
recordings from transverse hippocampal slices have fibers selectively in tissue from humans and ro-
demonstrated that granule cells from aged rodents dents using the Timm stain have shown that mossy
have an increased propensity for generating spon- fibers can extend up into the granule cell and inner
taneous, prolonged, large amplitude EPSPs follow- molecular layers in the transverse plane of the
ing exposure to the GABAA receptor antagonist hippocampus in aged subjects (Cassell and Brown,
bicuculline methiodide (Patrylo et al., 2007). These 1984; Wolfer and Lipp, 1995). While this distri-
events resemble the giant synaptic potentials that bution pattern of mossy fibers corresponds with
underlie paroxysmal depolarizing shifts, and sug- the location of the novel net current flux in the
gest that aberrant excitatory circuits are present in aged dentate gyrus (Barnes and McNaughton,
the aged dentate gyrus, since granule cells do not 1983), no data are currently available to verify that
typically exhibit reverberatory activity following they lead to increased granule cell to granule cell
disinhibition (Schwartzkroin and Prince, 1978; synapses within the aged dentate gyrus.
Fricke and Prince, 1984; Wuarin and Dudek, 1996). Recent data suggest an additional mechanism
Although the cellular player(s) that contribute that could contribute to the aberrant excitatory
to this aberrant excitatory circuit are unclear, an- interconnections in the aged dentate gyrus (Rao et
atomic data suggest several possibilities. First, the al., 2005). Specifically, it was demonstrated that
volume of [3H] kainate binding (Nicolle et al., newly generated granule cells in middle-aged and
1996) and Timm staining (Rapp et al., 1999) aged rodents exhibit a slowing of developmental
within the inner molecular layer relative to the maturation and an increased incidence of basal
middle/outer molecular layers is increased in the dendrites relative to granule cells born in younger
aged dentate gyrus, and thus it could reflect an animals. While it is unclear whether these basal
expansion of the associational/commissural fibers. dendrites persist in the newly generated granule
689

cells, it is interesting to note that GABAA recep- is not only associated with a decline in memory re-
tor-mediated inhibition plays a critical role in in- tention, but also a reduced capacity to encode new
tegrating newly generated granule cells into information rapidly (Tanila et al., 1997; Rosenzweig
dentate circuitry (Ge et al., 2006) and inhibitory and Barnes, 2003; Wilson et al., 2005), and encod-
tone is disrupted during aging (see above). Several ing depends on the entorhinal–dentate gyrus–CA3
groups have reported basal dendrites in animal pathway (e.g., Lee and Kesner, 2004; Jerman et al.,
models of temporal lobe epilepsy (Spigelman et al., 2006; Rolls and Kesner, 2006). Recent data reveal
1998; Buckmaster and Dudek, 1999; Ribak et al., that aged rodents with a decreased ability to encode
2000) and have provided evidence that these ab- new spatial information exhibit an increase in spon-
errant processes can create anomalous excitatory taneous activity in CA3 (Wilson et al., 2005). It has
circuits capable of positive feedback (e.g., Ribak been suggested that this increase in basal activity in
et al., 2000). CA3 decreases the signal-to-noise ratio by increas-
In summary, the efficacy of the inhibitory sys- ing the ‘‘noise’’, and thus obscures the influx of new
tem is compromised within the dentate gyrus dur- pertinent spatial information. The hyperexcitability
ing aging, and excitatory interconnections appear observed in CA3 in aged rodents during and after
to be increased. This disruption in the balance of 5 Hz perforant path stimulation (Patrylo et al.,
inhibition relative to excitation could impair dent- 2007) could be an in vitro equivalent of the increase
ate filter function, as has been shown in animal in basal CA3 activity reported in vivo by Wilson et
models of epilepsy that exhibit similar, although al. (2005). Consequently, we postulate that the ag-
more severe, changes in dentate circuitry. ing-related disruption in dentate gating, an effect
that occurs in a fraction of aged subjects, contrib-
utes to a decreased ability to encode new informa-
The dentate gyrus, spatial learning and memory, tion and thus the compromise in spatial learning
and aging and memory seen in a subpopulation of aged sub-
jects. This increase in basal activity could also affect
One of the most common aging-related neurological cognitive function negatively by changing the ca-
dysfunctions is a decline in spatial learning and pacity for synaptic plasticity, due to a metaplastic
memory, which has been reported to occur in shift in network dynamics (Abraham and Bear,
humans, non-human primates, and rodents (Barnes 1996).
1979; Gage et al., 1984; de Toledo-Morrell et al.,
1984; Barnes and McNaughton, 1985; Gallagher
and Pelleymonter, 1988; Uttl and Graf, 1993; Rapp The dentate gyrus, seizure susceptibility, and aging
and Gallagher, 1996; Wilkniss et al., 1997). While
the dentate gyrus exhibits several physiological Epidemiologic studies indicate that aging is also
characteristics that are believed to be critical for associated with an increase in the incidence and
spatial learning and memory, including burst dis- prevalence for seizure disorders (Loiseau et al.,
charges during theta (Rose et al., 1983; Buzsaki and 1990; Hauser et al., 1991, 1996; Hauser, 1992). In-
Czeh, 1992), phase precession (Skaggs et al., 1996), deed, seizure disorders are the third most common
and reactivation during sleep (Shen et al., 1998), and neurological disorder seen in the elderly, with up to
data indicate that alternations of dentate intrinsic 5% of individuals Z65 years of age affected (Tallis
properties or connectivity can have detrimental et al., 1991). The most common type of seizure
effects on learning and memory, it is not entirely observed in this age group is complex-partial in
clear whether aging-related changes in dentate cir- nature (Hauser, 1997; Rowan et al., 2005), and it is
cuitry and function contribute to the decline in cog- well established that limbic structures (i.e., hippo-
nitive function seen with age. campus) are frequently involved. Furthermore,
There is good reason to define the potential role 50% of newly identified seizure disorders in the
that the dentate gyrus plays in aging-related cogni- elderly have no identifiable antecedent (i.e., are
tive decline, because recent data suggest that aging cryptogenic), which has led to the suggestion that
690

aging itself may be epileptogenic (Ng et al., 1985; kindling and kainic acid models of epilepsy suggest
Hauser et al., 1996; Hauser, 1997). that aging-related changes in dentate circuitry and
Data from laboratory animals support and ex- function are likely to contribute to the increased
tend these clinical observations with an enhanced incidence for complex-partial seizures with age.
susceptibility for hippocampal-dependent seizures
seen in aged rodents. For example, aged rodents
exhibit longer duration afterdischarges during Conclusions
kindling, and faster seizure propagation to the cont-
ralateral hemisphere (Chiba et al., 1992). The rate In conclusion, extensive changes occur in the dent-
of kindling is slower in aged rodents however ate gyrus at the structural and functional levels
(de Toledo-Morrell et al., 1984; Chiba et al., during aging, and we suggest that one critical con-
1992), although this is likely to reflect a decreased sequence is the breakdown in filter function.
capacity for plasticity during aging, rather than a Changes in dentate gyrus gating could explain the
decreased susceptibility to seizures, since the pro- increase in hippocampal-dependent seizure suscep-
gression of kindling is dependent on NMDA recep- tibility that occurs during the aging process. These,
tor-mediated plasticity (Mody et al., 1988; Sayin and additional effects in associated structures such
et al., 1999) and decreases in NMDA receptor- as area CA3, may also contribute to the decline in
mediated properties occur with age (e.g., Rao et al., cognitive function with age. However, the precise
1994; Barnes et al., 2000; Magnusson et al., 2006). role of the dentate gyrus in age-related functional
Aged rodents also demonstrate increased seizure changes (i.e., whether it is the cause or the effect of
susceptibility when tested with the convulsant the increase in seizures susceptibility or cognitive
kainic acid. Systemic treatment with kainic acid decline) remains to be addressed completely.
results in exacerbated pre-seizure behavioral man-
ifestations in aged rodents, compared to adults
Acknowledgments
(e.g., wet-dog shakes), and a decreased latency to
onset for limbic seizures and sustained seizures
Supported by a grant from the NIH (AG00795)
(Darbin et al., 2004). Since the dentate gyrus plays
and the Epilepsy Foundation through the gener-
a role in generating wet-dog shakes (Damiano and
ous support of the CDC (PRP).
Connor, 1984; Frush and McNamara, 1986; Barn-
es and Mitchell, 1990; Grimes et al., 1990; Mitchell
et al., 1990), and regulating limbic seizures (Dash- References
eiff and McNamara, 1982; Lothman et al., 1992;
Grecksch et al., 1995), it is plausible that a break- Aamodt, S.M., Shi, J., Colonnese, M.T., Veras, W. and Cons-
down of dentate filter function results in an en- tantine-Paton, M. (2000) Chronic NMDA exposure acceler-
ates development of GABAergic inhibition in the superior
hanced propensity for dentate–CA3 activation and
colliculus. J. Neurophysiol., 83: 1580–1591.
thus an increased susceptibility to seizures. Sup- Abraham, W.C. and Bear, M.F. (1996) Metaplasticity: the plas-
port for this hypothesis comes from two sources. ticity of synaptic plasticity. Trends Neurosci., 19: 126–130.
First, the mossy fiber–CA3 synapse is preferen- Acsady, L., Kamondi, A., Sik, A., Freund, T. and Buzsaki, G.
tially susceptible to kainic acid (Ben-Ari and Coss- (1998) GABAergic cells are the major postsynaptic targets of
art, 2000), and preliminary data suggest that aged mossy fibers in the rat hippocampus. J. Neurosci., 18:
3386–3403.
hippocampal slices exhibit a lower threshold for Akhtar, N.D. and Land, P.W. (1991) Activity-dependent reg-
kainic acid-induced epileptiform activity (Willing- ulation of glutamic acid decarboxylase in the rat barrel cor-
ham et al., 2006). Second, a similar, although more tex: effects of neonatal versus adult sensory deprivation. J.
severe, disruption of dentate circuitry and gating Comp. Neurol., 307: 200–213.
Almaguer-Melian, W., Cruz-Aguado, R., Riva Cde, L., Kend-
has been observed in the dentate gyrus in models
rick, K.M., Frey, J.U. and Bergado, J. (2005) Effect of LTP-
of epilepsy that are associated with hippocampal reinforcing paradigms on neurotransmitter release in the
activation (e.g., Behr et al., 1998; Finnerty et al., dentate gyrus of young and aged rats. Biochem. Biophys.
2001; Nadler, 2003). Taken together, the data from Res. Commun., 327: 877–883.
691

de Almeida, O.M., Gardino, P.F., Loureiro dos Santos, N.E., Behr, J., Lyson, K.J. and Mody, I. (1998) Enhanced propaga-
Yamasaki, E.N., de Mello, M.C., Hokoc, J.N. and de Mello, tion of epileptiform activity through the kindled dentate
F.G. (2002) Opposite roles of GABA and excitatory amino gyrus. J. Neurophysiol., 79: 1726–1732.
acids on the control of GAD expression in cultured retina Ben-Ari, Y. and Cossart, R. (2000) Kainate, a double agent that
cells. Brain Res., 925: 89–99. generates seizures: two decades of progress. Trends Neuro-
Andersen, P. (1975) Organization of hippocampal neurons and sci., 23: 580–587.
their interconnections. In: Isaacson R.L. and Pribram K.H. Bernasconi-Guastalla, S., Wolfer, D.P. and Lipp, H.P. (1994)
(Eds.), The Hippocampus, Vol. 1. Plenum Press, New York, Hippocampal mossy fibers and swimming navigation in mice:
pp. 155–175. correlations with size and left-right asymmetries. Hippocam-
Aubert, I., Rowe, W., Meaney, M.J., Gauthier, S. and Quirion, pus, 4: 53–63.
R. (1995) Cholinergic markers in aged cognitively impaired Bragin, A., Jando, G., Nadasdy, Z., van Landeghem, M. and
Long-Evans rats. Neuroscience, 67: 277–292. Buzsaki, G. (1995) Dentate EEG spikes and associated in-
Austin, K.B., Bronzino, J.D. and Morgane, P.J. (1989) Paired- terneuronal population bursts in the hippocampal hilar re-
pulse facilitation and inhibition in the dentate gyrus is de- gion of the rat. J. Neurophysiol., 73: 1691–1705.
pendent on behavioral state. Exp. Brain Res., 77(3): 594–604. Buckmaster, P.S. and Dudek, F.E. (1999) In vivo intracellular
Bai, L., Hof, P.R., Standaert, D.G., Xing, Y., Nelson, S.E., analysis of granule cell axon reorganization in epileptic rats.
Young, A.B. and Magnusson, K.R. (2004) Changes in the J. Neurophysiol., 81: 712–721.
expression of the NR2B subunit during aging in macaque Buckmaster, P.S., Wenzel, H.J., Kunkel, D.D. and Schwa-
monkeys. Neurobiol. Aging, 25: 201–208. rtzkroin, PA. (1996) Axon arbors and synaptic connections
Barnes, C.A. (1979) Memory deficits associated with senes- of hippocampal mossy cells in the rat in vivo. J. Comp.
cence: a neurophysiological and behavioral study in the rat. J. Neurol., 366: 271–292.
Comp. Psychol., 93: 74–104. Buzsaki, G. and Czeh, G. (1992) Physiological function
Barnes, C.A. (2000) Plasticity in the aging central nervous sys- of granule cells: a hypothesis. Epilepsy Res. Suppl., 7:
tem. Int. Rev. Neurobiol., 45: 339–354. 281–290.
Barnes, C.A. and McNaughton, B.L. (1980a) Physiological Cadacio, C.L., Milner, T.A., Gallagher, M. and Pierce, J.P.
compensation for loss of afferent synapses in rat hippocam- (2003) Hilar neuropeptide Y interneuron loss in the aged rat
pal granule cells during senescence. J. Physiol. Lond., 309: hippocampal formation. Exp. Neurol., 183: 147–158.
473–485. Calhoun, M.E., Mao, Y., Roberts, J.A. and Rapp, P.R. (2004)
Barnes, C.A. and McNaughton, B.L. (1980b) Spatial memory Reduction in hippocampal cholinergic innervation is unre-
and hippocampal synaptic plasticity in middle-aged and se- lated to recognition memory impairment in aged rhesus
nescent rats. In: Stein D. (Ed.), Psychobiology of Aging: monkeys. J. Comp. Neurol., 475: 238–246.
Problems and Perspectives. Elsevier/North-Holland, New Carnahan, J. and Nawa, H. (1995) Regulation of neuropeptide
York, pp. 253–272. expression in the brain by neurotrophins. Potential role in
Barnes, C.A. and McNaughton, B.L. (1983) Excitability vivo. Mol. Neurobiol., 10: 135–149.
changes in hippocampal granule cells of senescent rat. In: Cassell, M.D. and Brown, M.W. (1984) The distribution of
Changeux J.P., Glowinski J., Imbert M. and Bloom F.E. Timm’s stain in the nonsulphide-perfused human hippocam-
(Eds.), Molecular and Cellular Interactions Underlying pal formation. J. Comp. Neurol., 222: 461–471.
Higher Brain Function, Progress in Brain Research. Elsevi- Chiba, S., Muneoka, Y., Sato, Y. and Miyagishi, T. (1992)
er, New York, pp. 445–451. Seizure susceptibility in aged rats — pentylenetetrazol-in-
Barnes, C.A. and McNaughton, B.L. (1985) An age-compar- duced seizures and amygdaloid kindling seizures. No To
ison of the rates of acquisition and forgetting of spatial in- Shinkei [Brain Nerve], 44: 559–564.
formation in relation to long-term enhancement of Christian, E.P. and Dudek, F.E. (1988) Characteristics of local
hippocampal synapses. Behav. Neurosci., 6: 563–571. excitatory circuits studied with glutamate microapplication in
Barnes, C.A., Rao, G. and Houston, F.P. (2000) LTP induction the CA3 area of rat hippocampal slices. J. Neurophysiol., 59:
threshold change in old rats at the perforant path — granule 90–109.
cell synapse. Neurobiol. Aging, 21: 613–620. Cronin, J., Obenaus, A., Houser, C.R. and Dudek, F.E. (1992)
Barnes, C.A., Rao, G. and McNaughton, B.L. (1987) Increased Electrophysiology of dentate granule cells after kainate-in-
electrotonic coupling in aged rat hippocampus: a possible duced synaptic reorganization of the mossy fibers. Brain
mechanism for cellular excitability changes. J. Comp. Ne- Res., 573: 305–310.
urol., 259: 549–558. Crusio, W.E., Schwegler, H. and Lipp, H.P. (1987) Radial-
Barnes, M.I. and Mitchell, C.L. (1990) Differential effects of maze performance and structural variation of the hippocam-
colchicine lesions of dentate granule cells on wet dog shakes pus in mice: a correlation with mossy fibre distribution. Brain
and seizures elicited by direct hippocampal stimulation. Res., 425: 182–185.
Physiol. Behav., 48: 131–138. Damiano, B.P. and Connor, J.D. (1984) Hippocampal
Beck, H., Goussakov, I.V., Lie, A., Helmstaedter, C. and Elger, mediation of shaking behavior induced by electrical stimu-
C.E. (2000) Synaptic plasticity in the human dentate gyrus. J. lation of the perforant path in the rat. Brain Res., 308:
Neurosci., 20: 7080–7086. 383–386.
692

Darbin, O., Naritoku, D. and Patrylo, P.R. (2004) Aging alters changes of the GABAA receptor in the rat hippocampus.
electroencephalographic and clinical manifestations of kain- Neuroscience, 74: 341–348.
ate-induced status epilepticus. Epilepsia, 45: 1219–1227. Halasy, K. and Somogyi, P. (1993) Distribution of GABAergic
Dasheiff, R.M. and McNamara, J.O. (1982) Intradentate col- synapses and their targets in the dentate gyrus of rat: a
chicine retards the developments of amygdala kindling. Ann. quantitative immunoelectron microscopic analysis. J. Hi-
Neurol., 11: 347–352. rnforsch., 34: 299–308.
Dore, S., Kar, S., Rowe, W. and Quirion, R. (1997) Distribu- Hattiangady, B., Rao, M.S., Shetty, G.A. and Shetty, A.K.
tion and levels of [125I]IGF-I, [125I]IGF-II and [125I]insulin (2005) Brain-derived neurotrophic factor, phosphorylated
receptor binding sites in the hippocampus of aged memory- cyclic AMP response element binding protein and neuropep-
unimpaired and -impaired rats. Neuroscience, 80: 1033–1040. tide Y decline as early as middle age in the dentate gyrus and
Finnerty, G.T., Whittington, M.A. and Jefferys, J.G. (2001) CA1 and CA3 subfields of the hippocampus. Exp. Neurol.,
Altered dentate filtering during the transition to seizure in the 195: 353–371.
rat tetanus toxin model of epilepsy. J. Neurophysiol., 86: Hauser, W.A. (1992) Seizure disorders: the changes with age.
2748–2753. Epilepsia, 33: S6–S14.
Fisher, J.L. (2004) The alpha 1 and alpha 6 subunit subtypes of Hauser, W.A. (1997) Epidemiology of seizures and epilepsy in
the mammalian GABA(A) receptor confer distinct channel the elderly. In: Rowan A.J. and Ramsay R.E. (Eds.), Seizures
gating kinetics. J. Physiol., 561: 433–448. and Epilepsy in the Elderly. Butterworth-Heinemann, Bos-
Freund, T.F. and Buzsáki, G. (1996) Interneurons of the hip- ton, MA, pp. 7–18.
pocampus. Hippocampus, 6: 347–470. Hauser, W.A., Annegers, J.F. and Kurland, L.T. (1991) Prev-
Fricke, R.A. and Prince, D.A. (1984) Electrophysiology of alence of epilepsy in Rochester, Minnesota. Epilepsia, 32:
dentate gyrus granule cells. J. Neurophysiol., 51: 195–209. 429–445.
Frush, D.P. and McNamara, J.O. (1986) Evidence implicating Hauser, W.A., Annegers, J.F. and Rocca, W.A. (1996) De-
dentate granule cells in wet dog shakes produced by kindling scriptive epidemiology of epilepsy: contributions of popula-
stimulations of entorhinal cortex. Exp. Neurol., 92: 102–113. tion-based studies from Rochester, Minnesota. Mayo Clin.
Gage, F.H., Bjorklund, A., Stenevi, U., Dunnett, S.B. and Proc., 71: 576–586.
Kelly, P.A.T. (1984) Intrahippocampal septal grafts amelio- Hayashi, M., Mistunaga, F., Ohira, K. and Shimizu, K. (2001)
rate learning impairments in aged rats. Science, 225: 533–536. Changes in BDNF-immunoreactive structures in the hippo-
Gallagher, M. and Pelleymonter, M.A. (1988) Spatial learning campal formation of the aged macaque monkey. Brain Res.,
deficits in old rats: a model for memory decline in the aged. 918: 191–196.
Neurobiol. Aging, 9: 549–556. Heinemann, U., Beck, H., Dreier, J.P., Ficker, E., Stabel, J. and
Gazzaley, A.H., Siegel, S.J., Kordower, J.H., Mufson, E.J. and Zhang, C.L. (1992) The dentate gyrus as a regulated gate for
Morrison, J.H. (1996) Circuit-specific alterations of N-me- the propagation of epileptiform activity. Epilepsy Res. Sup-
thyl-D-aspartate receptor subunit 1 in the dentate gyrus of pl., 7: 273–280.
aged monkeys. Proc. Natl. Acad. Sci. U.S.A., 93: 3121–3125. Henze, D.A., Wittner, L. and Buzsaki, G. (2002) Single granule
Ge, S., Goh, E.L., Sailor, K.A., Kitabatake, Y., Ming, G.L. cells reliably discharge targets in the hippocampal CA3 net-
and Song, H. (2006) GABA regulates synaptic integration of work in vivo. Nat. Neurosci., 5: 790–795.
newly generated neurons in the adult brain. Nature, 439: Herman, J.P., Chen, K.C., Booze, R. and Landfield, P.W.
589–593. (1998) Up-regulation of alpha1D Ca2+ channel subunit
Geinisman, Y., de Toledo-Morrell, L., Morrell, F., Persina, I.S. mRNA expression in the hippocampus of aged F344 rats.
and Rossi, M. (1992) Age-related loss of axospinous synapses Neurobiol. Aging, 19: 581–587.
formed by two afferent systems in the rat dentate gyrus as Hwang, I.K., Kim, D.W., Yoo, K.Y., Kim, D.S., Kim, K.S.,
revealed by the unbiased stereological dissector technique. Kang, J.H., Choi, S.Y., Kim, Y.S., Kang, T.C. and Won,
Hippocampus, 2: 437–444. M.H. (2004) Age-related changes of the gamma-amino-
Gooney, M., Messaoudi, E., Maher, F.O., Bramham, C.R. and butyric acid transaminase immunoreactivity in the hippo-
Lynch, M.A. (2004) BDNF-induced LTP in dentate gyrus is campus and dentate gyrus of the Mongolian gerbil. Brain
impaired with age: analysis of changes in cell signaling events. Res., 1017: 77–84.
Neurobiol. Aging, 25: 1323–1331. Hyman, C., Juhasz, M., Jackson, C. and Wright, P. (1994) Ip
Grecksch, G., Ruethrich, H., Bernstein, H.G. and Becker, A. NY, Lindsay RM. Overlapping and distinct actions of the
(1995) PTZ-kindling after colchicine lesion in the dentate neurotrophins BDNF, NT-3, and NT-4/5 on cultured do-
gyrus of the rat hippocampus. Physiol. Behav., 58: 695–698. paminergic and GABAergic neurons of the ventral me-
Grimes, L.M., Earnhardt, T.S., Mitchell, C.L., Tilson, H.A. sencephalon. J. Neurosci., 14: 335–347.
and Hong, J.S. (1990) Granule cells in the ventral, but not Ishizuka, N., Weber, J. and Amaral, D.G. (1990) Organization
dorsal, dentate gyrus are essential for kainic acid-induced wet of intrahippocampal projections originating from CA3 py-
dog shakes. Brain Res., 514: 167–170. ramidal cells in the rat. J. Comp. Neurol., 295: 580–623.
Gutierrez, A., Khan, Z.U., Ruano, D., Miralles, C.P., Vitorica, Jerman, T., Kesner, R.P. and Hunsaker, M.R. (2006) Discon-
J. and De Blas, A.L. (1996) Aging-related subunit expression nection analysis of CA3 and DG in mediating encoding but
693

not retrieval in a spatial maze learning task. Learn. Mem., 13: Lothman, E.W., Stringer, J.L. and Bertram, E.H. (1992) The
458–464. dentate gyrus as a control point for seizures in the hippo-
Johnson, D. and Brown, T.H. (1984) Mechanisms of neuronal campus and beyond. Epilepsy Res. Suppl., 7: 301–313.
burst generation. In: Schwartzkroin P.A. and Wheal H.V. Luebke, J.I. and Rosene, D.L. (2003) Aging alters dendritic
(Eds.), Electrophysiology of Epilepsy. Academic Press, New morphology, input resistance, and inhibitory signaling in
York, pp. 277–301. dentate granule cells of the rhesus monkey. J. Comp. Neurol.,
Katoh-Semba, R., Semba, R., Takeuchi, I.K. and Kato, K. 460: 573–584.
(1998) Age-related changes in levels of brain-derived ne- Lukoyanov, N.V., Andrade, J.P., Dulce Madeira, M. and
urotrophic factor in selected brain regions of rats, normal Paula-Barbosa, M.M. (1999) Effects of age and sex on the
mice and senescence-accelerated mice: a comparison to those water maze performance and hippocampal cholinergic fibers
of nerve growth factor and neurotrophin-3. Neurosci. Res., in rats. Neurosci. Lett., 269: 141–144.
31: 227–234. Lumeng, L. and Li, T.K. (1974) Vitamin B6 metabolism in
Katona, I., Acsady, L. and Freund, T.F. (1999) Postsynaptic chronic alcohol abuse. Pyridoxal phosphate levels in plasma
targets of somatostatin-immunoreactive interneurons in the and the effects of acetaldehyde on pyridoxal phosphate syn-
rat hippocampus. Neuroscience, 88: 37–55. thesis and degradation in human erythrocytes. J. Clin. In-
Kelley, M.S. and Steward, O. (1996) The process of reinner- vest., 53: 693–704.
vation in the dentate gyrus of adult rats: physiological events Lynch, G., Gall, C., Rose, G. and Cotman, C. (1976) Changes
at the time of the lesion and during the early postlesion in the distribution of the dentate gyrus associational system
period. Exp. Neurol., 139: 73–82. following unilateral or bilateral entorhinal lesions in the adult
Kobayashi, M. and Buckmaster, P.S. (2003) Reduced inhibition rat. Brain Res., 110: 57–71.
of dentate granule cells in a model of temporal lob epilepsy. J. Lynch, M., Sayin, U., Golarai, G. and Sutula, T. (2000)
Neurosci., 23: 2440–2452. NMDA receptor-dependent plasticity of granule cell spiking
Kugler, P. and Baier, G. (1992) Mitochondrial enzymes related in the dentate gyrus of normal and epileptic rats. J. Ne-
to glutamate and GABA metabolism in the hippocampus urophysiol., 84: 2868–2879.
of young and aged rats: a quantitative histochemical study. Maccaferri, G., Roberts, J.D., Szucs, P., Cottingham, C.A. and
Neurochem. Res., 17: 179–185. Somogyi, P. (2000) Cell surface domain specific postsynaptic
Kuhn, H.G., Dickinson-Anson, H. and Gage, F.H. (1996) currents evoked by identified GABAergic neurons in rat hip-
Neurogenesis in the dentate gyrus of the adult rat: age-related pocampus in vitro. J. Physiol., 524: 91–116.
decrease of neuronal progenitor proliferation. J. Neurosci., Magnusson, K.R. (2000) Declines in mRNA expression of
16: 2027–2033. different subunits may account for differential effects of ag-
Lawrence, J.J. and McBain, C.J. (2003) Interneuron diver- ing on agonist and antagonist binding to the NMDA recep-
sity series: containing the detonation — feedforward in- tor. J. Neurosci., 20: 1666–1674.
hibition in the CA3 hippocampus. Trends Neurosci., 26: Magnusson, K.R., Kresge, D. and Supon, J. (2006) Differential
631–640. effects of aging on NMDA receptors in the intermediate versus
Lee, I. and Kesner, R.P. (2004) Encoding versus retrieval of the dorsal hippocampus. Neurobiol. Aging, 27: 324–333.
spatial memory: double dissociation between the dentate Marty, S. (2000) Differences in the regulation of neuropeptide
gyrus and the perforant path inputs into CA3 in the dorsal Y, somatostatin and parvalbumin levels in hippocampal in-
hippocampus. Hippocampus, 14: 66–76. terneurons by neuronal activity and BDNF. Prog. Brain
Leranth, C., Malcolm, A.J. and Frotscher, M. (1990) Afferent Res., 128: 1.
and efferent synaptic connections of somatostatin-immuno- McNaughton, B.L., Barnes, C.A., Meltzer, J. and Sutherland,
reactive neurons in the rat fascia dentate. J. Comp. Neurol., R.J. (1989) Hippocampal granule cells are necessary for nor-
295: 111–122. mal spatial learning but not for spatially-selective pyramidal
Li, X.G., Somogyi, P., Ylinen, A. and Buzsaki, G. (1994) The cell discharge. Exp. Brain Res., 76: 485–496.
hippocampal CA3 network: an in vivo intracellular labeling Merrill, D.A., Chiba, A.A. and Tuszynski, M.H. (2001) Con-
study. J. Comp. Neurol., 339: 181–208. servation of neuronal number and size in the entorhinal cor-
Ling, L.L., Hughes, L.F. and Caspary, D.M. (2005) Age-related tex of behaviorally characterized aged rats. J. Comp. Neurol.,
loss of the GABA synthetic enzyme glutamic acid dec- 438: 445–456.
arboxylase in rat primary auditory cortex. Neuroscience, 132: Merrill, D.A., Karim, R., Darraq, M., Chiba, A.A. and
1103–1113. Tuszynski, M.H. (2003) Hippocampal cell genesis does not
Lipp, H.-P., Schwegler, H. and Driscoll, P. (1984) Postnatal correlate with spatial learning ability in aged rats. J. Comp.
modification of hippocampal circuitry alters avoidance learn- Neurol., 459: 201–207.
ing in adult rats. Science, 225: 80–82. Miles, R. and Wong, R.K.S. (1987) Inhibitory control of local
Loiseau, J., Louiseau, P., Duche, B., Guyot, M., Dartigues, J.F. excitatory circuits in the guinea pig hippocampus. J. Physiol.,
and Aublet, B. (1990) A survey of epileptic disorders in 388: 611–629.
southwest France: seizures in elderly patients. Ann. Neurol., Mitchell, C.L., Grimes, L.M. and Hong, J.S. (1990) Granule
27: 232–237. cells in the ventral dentate gyrus are essential for kainic
694

acid-induced wet dog shakes but not those induced by pre- Paulsen, O. and Moser, E.I. (1998) A model of hippocampal
cipitated abstinence in morphine-dependent rats. Brain Res., memory encoding and retrieval: GABAergic control of
511(2): 338–340. synaptic plasticity. Trends Neurosci., 21: 273–278.
Mody, I., Stanton, P.K. and Heinemann, U. (1988) Activation Penttonen, M., Kamondi, A., Sik, A., Acsady, L. and Buzsaki,
of N-methyl-D-aspartate receptors parallels changes in cellu- G. (1997) Feed-forward and feed-back activation of the
lar and synaptic properties of dentate gyrus granule cells after dentate gyrus in vivo during dentate spikes and sharp wave
kindling. J. Neurophysiol., 59: 1033–1054. bursts. Hippocampus, 7: 437–450.
Moser, E.I. (1996) Altered inhibition of dentate granule cells Rao, G., Barnes, C.A., McNaughton, B.L. and Shen, J. (1994)
during spatial learning in an exploration task. J. Neurosci., Age-related decrease in the NMDA-mediated EPSP in FD.
16: 1247–1259. Soc. Neurosci. Abstr., 20: 1207.
Mott, D.D., Turner, D.A., Okazaki, M.M. and Lewis, D.V. Rao, M.S., Hattiangady, B., Abdel-Rahman, A., Stanley, D.P.
(1997) Interneurons of the dentate-hilus border of the rat and Shetty, A.K. (2005) Newly born cells in the ageing dent-
dentate gyrus: morphological and electrophysiological heter- ate gyrus display normal migration, survival and neuronal
ogeneity. J. Neurosci., 17: 3990–4005. fate choice but endure retarded early maturation. Eur. J.
Nadler, J.V. (2003) The recurrent mossy fiber pathway of the Neurosci., 21: 464–476.
epileptic brain. Neurochem. Res., 28: 1649–1658. Rapp, P.R., Deroche, P.S., Mao, Y. and Burwell, R.D. (2002)
Nagahara, A.H., Nicolle, M.M. and Gallagher, M. (1993) Neuron number in the parahippocampal region is preserved
Alterations in [3H]-kainate receptor binding in the hippocam- in aged rats with spatial learning deficits. Cereb. Cortex, 12:
pal formation of aged Long-Evans rats. Hippocampus, 3: 1171–1179.
269–277. Rapp, P.R. and Gallagher, M. (1996) Preserved neuron number
Nanry, K.P., Mundy, W.R. and Tilson, H.A. (1989) Colchicine- in the hippocampus of aged rats with spatial learning deficits.
induced alterations of reference memory in rats: role of spa- Proc. Natl. Acad. Sci., 93: 9926–9930.
tial versus non-spatial task components. Behav. Brain Res., Rapp, P.R., Stack, E.C. and Gallagher, M. (1999) Morpho-
35: 45–53. metric studies of the aged hippocampus: I. Volumetric anal-
Ng, S.K., Hauser, W.A., Brust, J.C.M., Healton, E.B. and ysis in behaviorally characterized rats. J. Comp. Neurol., 403:
Susser, M.W. (1985) Risk factors for adult-onset first sei- 459–470.
zures. Ann. Neurol., 18: 153. Rasmussen, T., Schliemann, T., Sorensen, J.C., Zimmer, J. and
Nicolle, M.M., Bizon, J.L. and Gallagher, M. (1996) In vitro West, M.J. (1996) Memory impaired aged rats: no loss of
autoradiography of ionotropic glutamate receptors in hip- principal hippocampal and subicular neurons. Neurobiol.
pocampus and striatum of aged Long-Evans rats: relation- Aging, 17: 143–147.
ship to spatial learning. Neuroscience, 74: 741–756. Reeves, T.M. and Smith, D.C. (1987) Reinnervation of the
Niesen, C.E., Baskys, A. and Carlen, P.L. (1988) Reversed et- dentate gyrus and recovery of alternation behavior following
hanol effects on potassium conductances in aged hippocam- entorhinal cortex lesions. Behav. Neurosci., 101: 179–186.
pal dentate granule neurons. Brain Res., 445: 137–141. Reibel, S., Vivien-Roels, B., Le, B.T., Larmet, Y., Carnahan, J.,
Orr, G., Rao, G., Houston, F.P., McNaughton, B.L. and Marescaux, C. and Depaulis, A. (2000) Overexpression of
Barnes, C.A. (2001) Hippocampal synaptic plasticity is mod- neuropeptide Y induced by brain-derived neurotrophic factor
ulated by theta rhythm in the fascia dentata of adult and aged in the rat hippocampus is long lasting. Eur. J. Neurosci., 12:
freely behaving rats. Hippocampus, 11: 647–654. 595–605.
Pagliusi, S.R., Gerrard, P., Abdallah, M., Talabot, D. and Reynolds, J.N. and Carlen, P.L. (1989) Diminished calcium
Catsicas, S. (1994) Age-related changes in expression of currents in aged hippocampal dentate gyrus granule neu-
AMPA-selective glutamate receptor subunits: is calcium-per- rones. Brain Res., 479: 384–390.
meability altered in hippocampal neurons? Neuroscience, 61: Ribak, C.E. (1992) Local circuitry of GABAergic basket cells in
429–433. the dentate gyrus. Epilepsy Res. Suppl., 7: 29–47.
Patel, Y.C., Liu, J.L., Warszynska, A., Kent, G., Papachristou, Ribak, C.E., Seress, L. and Leranth, C. (1993) Electron micro-
D.N. and Patel, S.C. (1995) Differential stimulation of som- scopic immunocytochemical study of the distribution of
atostatin but not neuropeptide Y gene expression by parvalbumin-containing neurons and axon terminals in the
quinolinic acid in cultured cortical neurons. J. Neurochem., primate dentate gyrus and Ammon’s horn. J. Comp. Neurol.,
65: 998–1006. 327: 298–321.
Patrylo, P.R. and Dudek, F.E. (1998) Physiological unmasking Ribak, C.E., Tran, P.H., Spigelman, I., Okazaki, M.M. and
of new glutamatergic pathways in the dentate gyrus of hip- Nadler, J.V. (2000) Status epilepticus-induced hilar basal
pocampal slices from kainate-induced epileptic rats. J. Ne- dendrites on rodent granule cells contribute to recurrent ex-
urophysiol., 79: 418–429. citatory circuitry. J. Comp. Neurol., 428: 240–253.
Patrylo, P.R., Tyagi, I., Willingham, A.L., Lee, S. and Will- Rissman, R.A., Nocera, R., Fuller, L.M., Kordower, J.H. and
iamson, A. (2007) Dentate filter function is altered in a pro- Armstrong, D.M. (2006) Age-related alterations in
epileptic fashion during aging. Epilepsia (in press) doi: GABA(A) receptor subunits in the nonhuman primate hip-
10.1111/j.1528-1167.2007.01139.X pocampus. Brain Res., 1073-1074: 120–130.
695

Rogers, S.W., Gahring, L.C., Collins, A.C. and Marks, M. synaptophysin immunoreactivity predict spatial learning im-
(1998) Age-related changes in neuronal nicotinic acetylcho- pairment in aged rats. J. Neurosci., 20: 6587–6593.
line receptor subunit alpha4 expression are modified by long- Soltèsz, I., Smetters, D.K. and Mody, I. (1995) Tonic inhibition
term nicotine administration. J. Neurosci., 18: 4825–4832. originates from synapses close to the soma. Neuron, 14:
Rolls, E.T. and Kesner, R.P. (2006) A computational theory of 1273–1283.
hippocampal function, and empirical tests of the theory. Sonntag, W.E., Lynch, C.D., Bennett, S.A., Khan, A.S.,
Prog. Neurobiol., 79: 1–48. Thornton, P.L., Cooney, P.T., Ingram, R.L., McShane, T.
Rose, G., Diamond, D. and Lynch, G.S. (1983) Dentate gran- and Brunso-Bechtold, J.K. (1999) Alterations in insulin-like
ule cells in the rat hippocampal formation have the behavi- growth factor-1 gene and protein expression and type 1 in-
oral characteristics of theta neurons. Brain Res., 266: 29–37. sulin-like growth factor receptors in the brains of ageing rats.
Rosenzweig, E.S. and Barnes, C.A. (2003) Impact of aging on Neuroscience, 88: 269–279.
hippocampal function: plasticity, network dynamics, and Soriano, E. and Frotscher, M. (1993) GABAergic innervation
cognition. Prog. Neurobiol., 69: 143–179. of the rat fascia dentata: a novel type of interneuron in the
Rowan, A.J., Ramsay, R.E., Collins, J.F., Pryor, F., Board- granule cell layer with extensive axonal arborization in the
man, K.D., Uthman, B.M., Spitz, M., Frederick, T., Towne, molecular layer. J. Comp. Neurol., 334: 385–396.
A., Carter, G.S., Marks, W., Felicetta, J., Tomyanovich, Soriano, E., Nitsch, R. and Frotscher, M. (1990) Axo-axonic
M.L. and VA Cooperative Study 428 Group. (2005) chandelier cells in the rat fascia dentata: golgi-electron mi-
New onset geriatric epilepsy: a randomized study of gaba- croscopy and immunocytochemical studies. J. Comp. Ne-
pentin, lamotrigine, and carbamazepine. Neurology, 64: urol., 293: 1–25.
1868–1873. Spigelman, I., Yan, X.X., Obenaus, A., Lee, E.Y., Wasterlain,
Ruano, D., Benavides, J., Machado, A. and Vitorica, J. (1995) C.G. and Ribak, C.E. (1998) Dentate granule cells form
Aging-associated changes in the pharmacological properties novel basal dendrites in a rat model of temporal lobe epi-
of the benzodiazepine (omega) receptor isotypes in the rat lepsy. Neuroscience, 86: 109–120.
hippocampus. J. Neurochem., 64: 867–873. Staley, K.J., Otis, T.S. and Mody, I. (1992) Membrane prop-
Sayin, Ü., Rutecki, P. and Sutula, T. (1999) NMDA-dependent erties of dentate gyrus granule cells: comparison of sharp
currents in granule cells of the dentate gyrus contribute to microelectrode and whole-cell recordings. J. Neurophysiol.,
induction but not permanence of kindling. J. Neurophysiol., 67: 1346–1358.
81: 564–574. Stanley, D.P. and Shetty, A.K. (2004) Aging in the rat hippo-
Scharfman, H.E. (1994) Synchronization of area CA3 hippo- campus is associated with widespread reductions in the
campal pyramidal cells and non-granule cells of the dentate number of glutamate decarboxylase-67 positive interneurons
gyrus in bicuculline-treated rat hippocampal slices. Neuro- but not interneuron degeneration. J. Neurochem., 89:
science, 59: 245–257. 204–216.
Schuster, G., Cassel, J.C. and Will, B. (1997) Comparison of the Steward, O. (1982) Assessing the functional significance of le-
behavioral and morphological effects of colchicine- or neu- sion-induced neuronal plasticity. Int. Rev. Neurobiol., 23:
tral fluid-induced destruction of granule cells in the dentate 197–254.
gyrus of the rat. Neurobiol. Learn. Mem., 68: 86–91. Steward, O. (1992) Lesion-induced synapse reorganization in
Schwartzkroin, P.A. and Prince, D.A. (1978) Cellular and field the hippocampus of cats: sprouting of entorhinal, commis-
potential properties of epileptogenic hippocampal slices. sural/associational, and mossy fiber projections after unilat-
Brain Res., 147: 117–130. eral entorhinal cortex lesions, with comments on the normal
Shen, J. and Barnes, C.A. (1996) Age-related decrease in organization of these pathways. Hippocampus, 2: 247–268.
cholinergic synaptic transmission in three hippocampal sub- Steward, O., Loesche, J. and Horton, W.C. (1976) Behavioral
fields. Neurobiol. Aging, 17: 439–451. correlates of denervation and reinnervation of the hippo-
Shen, J., Kudrimoti, H.S., McNaughton, B.L. and Barnes, C.A. campal formation of the rat: open field activity and cue uti-
(1998) Reactivation of neuronal ensembles in hippocampal lization following bilateral entorhinal cortex lesions. Brain
dentate gyrus during sleep after spatial experience. J. Sleep Res. Bull., 2: 41–48.
Res., 7: 6–16. Sykova, E., Mazel, T. and Simonova, Z. (1998) Diffusion con-
Shetty, A.K. and Turner, D.A. (1998) Hippocampal interneu- straints and neuron-glia interaction during aging. Exp. Ger-
rons expressing glutamic acid decarboxylase and calcium- ontol., 33: 837–851.
binding proteins decrease with aging in Fischer 344 rats. J. Tallis, R., Hall, G., Craig, I. and Dean, A. (1991) How common
Comp. Neurol., 394: 252–269. are epileptic seizures in old age? Age Ageing, 20: 442–448.
Skaggs, W.E., McNaughton, B.L., Wilson, M.A. and Barnes, Tanila, H., Shapiro, M., Gallagher, M. and Eichenbaum, H.
C.A. (1996) Theta phase precession in hippocampal neuronal (1997) Brain aging: changes in the nature of information
populations and the compression of temporal sequences. coding by the hippocampus. J. Neurosci., 17: 5155–5166.
Hippocampus, 6: 149–172. Tayebati, S.K., Amenta, F., El-Assouad, D. and Zaccheo, D.
Smith, T.D., Adams, M.M., Gallagher, M., Morrison, J.H. and (2002) Muscarinic cholinergic receptor subtypes in the hip-
Rapp, P.R. (2000) Circuit-specific alterations in hippocampal pocampus of aged rats. Mech. Ageing Dev., 123: 521–528.
696

Thibault, O., Hadley, R. and Landfield, P.W. (2001) Elevated West, J.R. and Dewey, S.L. (1984) Mossy fiber sprouting in the
postsynaptic [Ca2+]i and L-type calcium channel activity in fascia dentata after unilateral entorhinal lesions: quantitative
aged hippocampal neurons: relationship to impaired synaptic analysis using computer-assisted image processing. Neuro-
plasticity. J. Neurosci., 21: 9744–9756. science, 13: 377–384.
Tigges, J., Herndon, J.G. and Rosene, D.L. (1995) Mild age- Wilkniss, S.M., Jones, M.G., Korol, D.L., Gold, P.E. and
related changes in the dentate gyrus of adult rhesus monkeys. Manning, C.A. (1997) Age-related differences in an ecolog-
Acta Anat. (Basel), 153: 39–48. ically based study of route learning. Psychol. Aging, 12:
de Toledo-Morrell, L., Geinisman, Y. and Morrell, F. (1988) 372–375.
Age-dependent alterations in hippocampal synaptic plastic- Williams, S.R., Buhl, E.H. and Mody, I. (1998) The dynamics
ity: relation to memory disorders. Neurobiol. Aging, 9: of synchronized neurotransmitter release determined from
581–590. compound spontaneous IPSCs in rat dentate granule neu-
de Toledo-Morrell, L., Morrell, F. and Fleming, S. (1984) Age- rones in vitro. J. Physiol., 510: 477–497.
dependent deficits in spatial memory are related to impaired Willingham, A., Tokhi, A., Jones, R. and Patrylo, P.R. (2006)
hippocampal kindling. Behav. Neurosci., 98: 902–907. Hippocampal slices from aged rats have an enhanced suscep-
Uttl, B. and Graf, P. (1993) Episodic spatial memory in adult- tibility to kainate-induced epileptiform activity. Abstract #
hood. Psychol. Aging, 8: 257–273. 3.102, 60th annual meeting of the American Epilspy Society.
Vela, J., Gutierrez, A., Vitorica, J. and Ruano, D. (2003) Rat Wilson, I.A., Ikonen, S., Gallagher, M., Eichenbaum, H. and
hippocampal GABAergic molecular markers are differen- Tanila, H. (2005) Age-associated alterations of hippocampal
tially affected by ageing. J. Neurochem., 85: 368–377. place cells are subregion specific. J. Neurosci., 25: 6877–6886.
Wenk, G.L. and Barnes, C.A. (2000) Regional changes in the Wolfer, D.P. and Lipp, H.P. (1995) Evidence for physiological
hippocampal density of AMPA and NMDA receptors across growth of hippocampal mossy fiber collaterals in the guinea
the lifespan of the rat. Brain Res., 885: 1–5. pig during puberty and adulthood. Hippocampus, 5:
Wenzel, H.J., Buckmaster, P.S., Anderson, N.L., Wenzel, M.E. 329–340.
and Schwartzkroin, P.A. (1997) Ultrastructural localization Wuarin, J.P. and Dudek, F.E. (1996) Electrographic seizures
of neurotransmitter immunoreactivity in mossy cell axons and new recurrent excitatory circuits in the dentate gyrus of
and their synaptic targets in the rat dentate gyrus. Hippo- hippocampal slices from kaintae-treated epileptic rats. J. Ne-
campus, 7: 559–570. urosci., 16: 4438–4448.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 38

Dentate gyrus neurogenesis and depression

Amar Sahay1,2,, Michael R. Drew1,2 and Rene Hen1,2,3,

1
Departments of Neuroscience and Psychiatry, Columbia University, New York, NY 10032, USA
2
Division of Integrative Neuroscience, Columbia University, New York, NY 10032, USA
3
Department of Pharmacology, Columbia University, New York, NY 10032, USA

Abstract: Major depressive disorder (MDD) is a debilitating and complex psychiatric disorder that in-
volves multiple neural circuits and genetic and non-genetic risk factors. In the quest for elucidating the
neurobiological basis of MDD, hippocampal neurogenesis has emerged as a candidate substrate, both for
the etiology as well as treatment of MDD. This chapter critiques the advances made in the study of
hippocampal neurogenesis as they relate to the neurogenic hypothesis of MDD. While an involvement of
neurogenesis in the etiology of depression remains highly speculative, preclinical studies have revealed a
novel and previously unrecognized role for hippocampal neurogenesis in mediating some of the behavioral
effects of antidepressants. The implications of these findings are discussed to reevaluate the role of
hippocampal neurogenesis in MDD.

Keywords: dentate gyrus; depression; neurogenesis; serotonin; antidepressants; hippocampus

Introduction ability to concentrate, diminished interest in pleas-


urable activities, daily insomnia or hypersomnia,
Understanding the neurobiological basis of major weight loss or gain, and recurrent suicidal ideation.
depressive disorder (MDD) is one of the most The diagnostic criteria for MDD convey the com-
pressing challenges for today’s society. Severe plexity of the disease and suggest that multiple
forms of depression affect 2–5% of the U.S. pop- neural circuits subserving distinct cognitive and
ulation, and mood disorders impact 7% of the affective processes are likely to be involved.
world’s population and rank among the top ten Our comprehension of the mechanisms under-
causes of disability (Murray and Lopez, 1996). The lying the pathogenesis of MDD has evolved con-
diagnosis of MDD based on the criteria established siderably since the formulation of the monoamine
by the Diagnostics and Statistical Manual of Men- hypothesis (Bunney and Davis, 1965; Schildkraut,
tal Disorders (American Psychological Association, 1965; Nestler et al., 2002). The recent emphasis on
2000) includes the persistence of depressed mood, neural circuits as opposed to a chemical imbalance
low self esteem, feelings of hopelessness, decreased catalyzed a fundamental shift in our conceptual-
ization of MDD and psychiatric disorders. It pro-
Corresponding author: Tel.: +1 212-543-5477; vided a framework to understand how genes,
Fax: +1 212-543-5074; through their effects on neural circuits, influence
E-mail: as2619@columbia.edu (A. Sahay) our ability to encode experience and adapt to en-
Corresponding author: Tel: +1 212-543-5328; vironmental stimuli and stressors. Implicit in this
Fax: +1 212-543-5074; E-mail: rh95@columbia.edu (R. Hen) idea is that genes moderate vulnerability to the

DOI: 10.1016/S0079-6123(07)63038-6 697


698

effects of environmental stress during particularly periods in brain development could, therefore, be
sensitive or critical periods in brain development pathogenic in that they profoundly impact the
by determining the optimal range of neuronal trajectory of emotional development. Deficits in
circuit function for the organism. Indeed, the ne- adult hippocampal neurogenesis could compro-
urotrophic, neuroplasticity and network hypothe- mise hippocampal-dependent functions and con-
ses of MDD all reflect the biology of gene products tribute to the pathophysiology of MDD.
in the context of synaptic and structural plasticity In this chapter, we focus on hippocampal ne-
of neural circuitry (Duman et al., 1997; Duman, urogenesis as it relates to MDD. Our aim is to
2002; Nestler et al., 2002; Castren, 2005). distill the observations made in the rapidly grow-
Dentate gyrus neurogenesis has gained consid- ing field of hippocampal neurogenesis and to crit-
erable attention as both a form of structural plas- ically assess the putative role of neurogenesis in the
ticity and as a neural substrate for the patho- etiology and treatment of MDD. We begin by de-
physiology of MDD. The neurogenic hypothesis fining a framework for the reader to understand
posits that a decrease in the production of new- how neurogenesis can contribute to dentate gyrus
born dentate granule cells in the hippocampus function. Within this framework, we will first eval-
causally relates to the pathogenesis and patho- uate evidence for deficits in hippocampal neuro-
physiology of MDD and that enhanced neurogen- genesis in patients with MDD. We will then
esis is necessary for treatment of depression examine the role of the serotonergic system in
(Duman et al., 2000; Jacobs et al., 2000). The hy- hippocampal neurogenesis because the best char-
pothesis, when first proposed, was predicated on acterized genetic risk alleles for MDD encode
the following observations, which are reviewed in components of this system. Because susceptibility
greater detail in subsequent sections. First, stress, to MDD conferred by genes is likely to be revealed
which is widely recognized as a major causal factor by environmental risk factors such as stress, we
in MDD, is known to suppress neurogenesis. Sec- will discuss the relationship between stress and
ond, most antidepressant (AD) treatments increase neurogenesis. We then review the considerable ev-
hippocampal neurogenesis. Third, imbalance in idence linking the effects of ADs with increased
the serotonin system influences hippocampal ne- hippocampal neurogenesis. Finally, we will turn to
urogenesis. Fourth, the induction of neurogenesis evidence provided by studies using preclinical
is contingent upon chronic but not subchronic models that attempt to establish a causal link be-
(acute) selective serotonin reuptake inhibitor tween hippocampal neurogenesis and the etiology,
(SSRI) treatment, paralleling the time course for and pathophysiology of MDD and the require-
therapeutic actions of ADs. Finally, the therapeu- ment for neurogenesis in mediating the behavioral
tic lag in the response to SSRIs in patients with effects of ADs.
MDD mirrors the timeline of maturation and
integration of newborn dentate granule cells.
Consequently, the dentate gyrus and neurogenesis Neurogenesis and MDD
therein are potential substrates for the AD
response. A general framework for neurogenesis and dentate
Central to the neurogenic hypothesis is the as- gyrus function
sumption that the dentate gyrus plays an impor-
tant role in mediating cognitive and affective Since the seminal findings of Altman and Das in
processes. Moreover, since levels of neurogenesis 1965, it is now well accepted that the adult hip-
change during the lifetime of the organism, pocampus is host to the birth and integration of
changes in dentate neurogenesis may contribute newborn dentate granule cells in the dentate gyrus
to dentate gyrus function in different ways. An- (Altman and Das, 1965). In the rat, the species for
other assumption is that neurogenesis represents which the best data are available, it is estimated
a potentially adaptive mechanism or form of plas- that 9000 new cells are born each day in the DG,
ticity. Deficits in neurogenesis during critical and, of these, approximately 50% go on to express
699

neuron-specific markers. At this rate, the number 1. Increase the number of mature dentate gran-
of new granule neurons born each month is equal ule cells
to 6% of the mature granule cell population The integration of newborn neurons can re-
(Cameron and McKay, 2001). In non-human pri- sult in an increase in the granule cell layer of
mates, the rate of neurogenesis may be lower than the dentate gyrus. It is conceivable that a net
the rate documented in rodents (Kornack and increase in size is possible only within a cer-
Rakic, 1999; Gould et al., 1999b). One should bear tain period in an animal’s life. An increase in
in mind that these data reflect neurogenesis under cell number can result from an enhancement
laboratory housing conditions, and given the in- in the rate of proliferation or the percentage
crease in neurogenesis with environmental enrich- of newborn neurons that survive.
ment (Kempermann et al., 1997; Gould et al., 2. Provide a reservoir of highly plastic immature
1999a), could underestimate the rates of neuro- neurons in the adult dentate gyrus
genesis in the normal habitat. Likewise, rates of Newly generated dentate granule cells also
neurogenesis in man maybe underestimated by exhibit forms of synaptic plasticity distinct
available data, because human data were based on from those of mature cells in the adult hip-
a single study in which tissue samples were taken pocampus. Newborn dentate granule cells
from cancer patients injected with a mitotic show unique physiological properties such as
marker, bromo-deoxyuridine (BrdU) before death lower thresholds for induction of long-term
(Eriksson et al., 1998). The number of BrdU-labe- potentiation and long-term depression than
led neurons entering the neuronal lineage was do mature neurons (Schmidt-Hieber et al.,
lower than that reported for marmosets and ro- 2004; Song et al., 2005). Moreover, newborn
dents, but the age of subjects could explain the dentate granule cells, unlike mature granule
difference because they were old, and neurogenesis cells, are able to undergo LTP under condi-
declines with age (Seki and Arai, 1995; Kuhn et al., tions of increased GABAergic inhibition
1996; Rao and Shetty, 2004). Therefore, it is un- (Wang et al., 2000; Snyder et al., 2001; Saxe
clear to what extent the relatively low level of ne- et al., 2006). Thus, in addition to conferring
urogenesis observed in the human subjects was due structural plasticity to the dentate gyrus, ne-
to real species difference. urogenesis also creates a transient reservoir of
The study of adult hippocampal neurogenesis excitable, highly plastic cells that may serve a
has revealed it to be a robust phenomenon that is unique biological function, distinct from that
capable of conferring previously unrecognized of mature granule neurons. The size of such a
forms of plasticity to the dentate gyrus. For ex- reservoir can by influenced by numerous
ample, it is clear that both net addition of newly physiological and environmental factors and
generated neurons and replacement of mature cells contingencies.
occur in the adult dentate gyrus and that the extent 3. Generate multiple cell types in the dentate
to which these processes occur may vary with the gyrus
animals age, and environmental and physiological While it is widely agreed that neurogenesis in
parameters (Bayer et al., 1982; West, 1993; Kem- the subgranular zone (SGZ) results in gener-
permann et al., 1998; Nottebohm, 2002; Amrein ation of dentate granule cells, there is one
et al., 2004; Wiskott et al., 2006). Modeling and report showing that GABAergic basket cells
computational approaches have revealed merits of in the dentate gyrus incorporate BrdU and
both net addition and replacement in optimizing form functional inhibitory synapses with
hippocampal network function (Chambers et al., dentate granule cells (Liu et al., 2003). Thus,
2004; Becker, 2005; Meltzer et al., 2005; Wiskott it is plausible that the generation of interneu-
et al., 2006). Figure 1 illustrates the distinct, but rons may occur under certain conditions to
potentially interrelated, ways by which neurogen- influence network activity. Clearly more
esis can modify the cellular composition of the evidence is needed to support this possibility.
dentate gyrus. In addition to the generation of neurons in
700

Fig. 1. A schematic of the dentate gyrus granule cell layer (GCL) illustrating the different ways by which neurogenesis can influence its
structure and function. Boxed panel reveals a cross section of the dentate GCL with the different populations that reside within it:
mature granule cells born during development (light blue), adult-generated mature granule cells (dark blue), adult-born immature
neurons (red) and interneurons (green). Over the lifespan, the GCL may increase in size due to a net addition of new neurons (A) or
may remain unchanged due to a net replacement of developmentally generated granule cells (B). Changes in neurogenesis can result in
increased representation of interneurons (C), a larger pool of adult generated immature neurons (D) or the generation of mature
neurons with distinct physiological and biochemical properties (E). Conceivably, neurogenesis may be altered in any one of these ways
in MDD. Conversely, AD drugs may influence DG function in more than one way to exert their behavioral effects. (See Color Plate
38.1 in color plate section.)

the dentate gyrus, proliferation in the SGZ Such a mechanism, when predominant,
also generates glial cells. The emerging role would not increase the size of the dentate
for glial cells in modulating synaptic function gyrus but ensure replacement of cells, whose
in health and disease underscores the need to functions are impaired, and rejuvenate the
understand how newly generated glial cells network with new cells (Nottebohm, 2002).
contribute to hippocampal physiology and There is some evidence to suggest that adult-
function (Ma et al., 2005; Haydon and generated mature dentate granule cells, while
Carmignoto, 2006). sharing electrophysiological properties with
4. Drive turnover and replacement of mature their early-development-born counterparts,
dentate granule cells exhibit greater plasticity in response to be-
The integration of newborn neurons can oc- haviorally relevant stimuli (Laplagne et al.,
cur to replace the death of mature neurons. 2006; Ramirez-Amaya et al., 2006).
701

Independent of the balance between the inte- the depressed brain are made using neuroimaging
gration of new neurons and the death of mature techniques such as positron emission tomography
neurons, it is conceivable that a specific form of (PET), a powerful way of identifying neural struc-
experience can result in a larger representation tures with altered metabolic activity. Studies on
of specific granule cells selected for by that kind patients with MDD have revealed alterations, but
of experience. Such a representation may manifest only in a very small number of studies and with
in a distinct pattern of biochemical and electro- conflicting results. This is partly due to the limited
physiological properties found in one cohort of resolution of PET. Using cerebral blood flow PET,
newborn cells versus another. Functional hetero- one group reported increased blood flow in the
geneity within the hippocampus and dentate gyrus hippocampus of acutely depressed patients with a
supports the possibility that subsets of neurons short duration of illness (Videbech et al., 2001,
within different regions of the dentate gyrus could 2002; Videbech and Ravnkilde, 2004). By contrast,
reflect distinct experiences (Moser and Moser, two other studies have shown either a decrease or
1998; Scharfman et al., 2002; Silva et al., 2006). no change in metabolism in the hippocampus of
The aforementioned ways by which neurogene- patients with MDD using fluorodeoxyglucose
sis contributes to the structure and function of the (FDG)–PET imaging (Saxena et al., 2001; Dre-
dentate gyrus convey the complexity of the phe- vets et al., 2002; Kimbrell et al., 2002). Differences
nomenon of hippocampal neurogenesis. They also in patient profile with regards to severity and du-
remind us of the many ways by which neurogenesis ration of illness and treatment could explain these
may be altered in pathological conditions. differences. Alternatively, it has been suggested
that increased activity, if untreated, may result in
hippocampal atrophy and decreased metabolism.
Hippocampal dysfunction and atrophy in MDD Atrophy would occlude detection of changes in
activity.
Hippocampal dysfunction in MDD is well sup- A consistent finding that has emerged from
ported by clinical studies which have shown that magnetic resonance imaging (MRI) studies on pa-
MDD is often accompanied by deficits in declar- tients with MDD is a reduction in hippocampal
ative learning and memory and diminished cogni- volume. Despite a few studies that failed to report
tive flexibility that are dissociable from changes in any differences between patients with MDD and
motivation (Austin et al., 2001; Fossati et al., control groups using MRI, there is consensus for
2002). The anatomical and functional segregation reduced hippocampal volume in MDD (Sheline,
of the hippocampus along its septotemporal axis 1996; Sheline et al., 1999; Bremner et al., 2000; von
suggests roles for the hippocampus in both cogni- Gunten et al., 2000; Vakili et al., 2000; Neumeister
tive and emotional processes (Moser and Moser, et al., 2005). Two recent meta-analyses of studies
1998; Strange and Dolan, 1999; Strange et al., measuring temporal lobe structures in MDD com-
1999; Bannerman et al., 2003). As a first step to pellingly demonstrate a reduction in hippocampus
thinking about the contribution of the dentate in people with recurrent depression relative to age-
gyrus and dentate neurogenesis to the pathophys- and sex-matched controls (Campbell et al., 2004;
iology of MDD, one must consider the direct ev- Videbech and Ravnkilde, 2004). Interestingly,
idence for hippocampal dysfunction and atrophy frequency of depressive episodes and the dura-
in MDD. tion for which depression is untreated corre-
While longitudinal studies linking changes in late with magnitude of reduction in hippocampal
hippocampal structure and function with the etio- volume (MacQueen et al., 2003; Sheline et al.,
logy of depression are lacking, we have made 2003). Taken together, the evidence argues for
progress in identifying changes in hippocampal reduced hippocampal volume in MDD and that
function, cellular structure and volume that are such changes are likely to be a result of depres-
associated with pathophysiology of depression. sion rather than a cause. However, it is worth
Direct measurements of hippocampal function in mentioning here that a smaller hippocampus is
702

thought to be a predisposing factor for, rather dentate gyrus and pyramidal neurons and glial
than a consequence of, post-traumatic stress dis- cells in the CA fields (Stockmeier et al., 2004). In
order (Gilbertson et al., 2002). addition, the authors reported a reduction in soma
The significance of hippocampal volume change size of pyramidal neurons and a trend towards the
in the context of cognitive deficits is conveyed by a same in dentate granule cells. Finally, one group
recent study that shows that healthy individuals directly examined the proliferation of cells in the
who complained of memory impairments had adult dentate gyrus of MDD patients using an
smaller hippocampal volumes than non-impaired M-phase marker, Ki-67 (Reif et al., 2006). Their
controls (van der Flier et al., 2004). Another study results showed no changes in Ki-67 immunopos-
showed a correlation between deficits in recollec- itive cells in hippocampus of depressed patients.
tion memory performance and reduced hippocam- While this study is informative and is the first to
pal volume in elderly depressed patients estimate levels of proliferation in the MDD brain,
(von Gunten and Ron, 2004). However, these it must be interpreted with several caveats in mind.
studies are limited in number and comprised First, a reduction in neurogenesis in patients with
elderly individuals who are likely to have other MDD could be masked by AD-mediated increase
brain changes that may contribute to the memory in cell proliferation. Second, and for obvious rea-
impairments. sons, the study could not measure changes in sur-
Changes in hippocampal volume can be ex- vival of newborn cells or examine the kinetics
plained by several different mechanisms that may of turnover and maturation of newborn dentate
operate in concert at the cellular and circuit levels granule cells.
including: (i) Increased apoptosis of mature neu- The pathohistological analyses suggest that
rons or glial cells. (ii) A loss of neuropil which may changes in neuropil, rather than neurogenesis,
involve changes in dendritic complexity, spine may account for reductions in hippocampal vol-
density, and number and size of afferent and effer- ume. Indeed, the effects of stress on hippocampal
ent axonal projections. (iii) Reduced neurogenesis white matter are well documented. Preclinical
or gliogenesis in the SGZ of dentate gyrus. The studies have shown that volumetric changes result
link between neurogenesis and hippocampal vol- from reduced dendritic complexity and not abla-
ume has been addressed in histological analyses tion of hippocampal neurogenesis (Santarelli et al.,
of postmortem tissue obtained from brains of 2003; McEwen, 2005). It should be noted that
patients with MDD, and will be discussed next. while these data argue against the possibility that
reduced cell number owing to extensive cell death
or decreased neurogenesis is a primary mediator of
Neurogenesis and cell death in MDD hippocampal volume change, they do not directly
address the possibility that neurogenesis is altered
Pathohistological studies of postmortem tissue of in MDD. Since most of the patients were on med-
patients, while small in number, have provided ication at the time of death, and since AD drugs
some clues about the nature of cellular changes in potently upregulate hippocampal neurogenesis, it
the hippocampus of a depressed individual. One is possible that depression-related alterations in
study examined synaptic density and glial cell neurogenesis could have been masked in these
number using synaptophysin and GFAP-immuno- studies. Moreover, since hippocampal neurogene-
reactivity, respectively, and found no differences in sis in humans is likely to change with age, small
the hippocampus of medicated patients with MDD differences in proliferation or survival of newborn
relative to controls (Muller et al., 2001). Another neurons in postmortem analyses of older patients
study revealed low levels of apoptosis in the dent- could be difficult to detect. Further studies on
ate gyrus, CA1, CA4, subiculum and entorhinal postmortem tissue of medicated and non-medi-
cortex of patients with MDD (Lucassen et al., cated individuals are needed to identify specific
2001). A third study showed a significant increase changes in hippocampal neurogenesis associated
in cell density of granule and glial cells in the with MDD.
703

Genes, environment and MDD Importantly, the association between the 5-HTT
polymorphism and depression is only observed
Depression is a complex and multifactorial illness in individuals who had experienced significant
with genetic and non-genetic underpinnings. The stressful life events. These findings argue that the
heritability of MDD is likely to be in the range of serotonergic system has a critical influence on ne-
40–50% and there is substantial evidence to sug- urodevelopmental processes that lead to MDD. If
gest that the phenotypic expression of MDD is hippocampal neurogenesis is to be a considered as
contingent upon interactions between the genetic a candidate neural substrate for depression, then
make-up of the individual and environmental fac- the effects of serotonergic dysregulation on it war-
tors, an interaction that has a dramatic effect on rants comment. The following section addresses
the formation and functioning of neural circuitry the role of the serotonergic system in modulating
(Sullivan et al., 2000; Kendler et al., 2001; Caspi dentate gyrus structure and function.
and Moffitt, 2006; Leonardo and Hen, 2006;
Levinson, 2006). Here, it must be emphasized that
human susceptibility to MDD as revealed by en- The 5-HT system and hippocampal neurogenesis
vironmental factors is tremendously magnified in
early life. For example, adults who had experi- Serotonergic terminals originating from the dorsal
enced four out of seven traumatic events in early raphe nucleus (DRN) and median raphe nuclei
life had a 4.6-fold increased risk of developing de- (MRN) diffusely innervate multiple structures in
pressive symptoms later in life and were 12.2-fold the vertebrate forebrain and reach the ventricles
more likely to commit suicide (Felitti et al., 1998; via the supra-ependymal plexus (Azmitia and
Chapman et al., 2004). These studies underscore Segal, 1978; Freund et al., 1990). The serotonergic
the idea of a critical period in ‘‘emotional devel- innervation of the hilus, molecular layers of the
opment’’ when mechanisms mediating neural cir- dentate gyrus and the SGZ supports the possibility
cuit synaptic- and structural plasticity are that 5-HT signaling may influence adult neuro-
particularly susceptible to environmental input, genesis (Oleskevich et al., 1991). The idea that
and if compromised by a vulnerability conferred serotonin can influence neurogenesis was first pro-
by genes, can result in maladaptive alterations in posed three decades ago (Lauder and Krebs,
neural circuit function and pathological behavior. 1978). However, the specific ways by which the
While the search for candidate genes for MDD has 5-HT system influences adult dentate neurogenesis
yielded some convergence from linkage studies was established only relatively recently. The first
that certain genetic loci are involved, more data studies to address the role of 5-HT in adult hip-
are clearly needed to conclusively implicate a spe- pocampal neurogenesis used serotonin depletion
cific gene in the etiology of MDD. A notable analyses. Injection of the serotonin neurotoxin
exception is the serotonin transporter (5-HTT) 5,7-dihydroneurotoxin (5,7-DHT) or 5-HT syn-
polymorphism. In a landmark study, Caspi and thesis inhibitor parachlorophenylalanine (PCPA)
colleagues found that a functional polymorphism into the raphe of young female rats resulted in a
in the 5-HTT gene moderated the sensitivity of reduction of dentate granule cell proliferation and
individuals to the depressogenic effects of early life the number of immature neurons as assessed by
stress, a finding recently replicated by Kendler and BrdU uptake and PSA-NCAM immunostaining,
colleagues (Caspi et al., 2003; Kendler et al., 2005). respectively (Brezun and Daszuta, 1999, 2000).
Caspi and colleagues found that people who car- Moreover, the same group showed that they could
ried one or two copies of the ‘‘short’’ allele of rescue the deficit in hippocampal proliferation in
5-HTT, associated with lower levels of 5-HTT these rats following intrahippocampal grafts of
and impaired reuptake of serotonin (5-HT) at embryonic 5-HT neurons (Brezun and Daszuta,
synapses, had more depressive symptoms and su- 2000). A potential role for 5-HT in influencing
icidal behavior in relation to stressful life events the maturation of dentate granule cells comes
than did people who had the ‘‘long allele’’. from studies in which rat pups were treated
704

with phenylchloromethamphetamine (PCA) or dentate gyrus (Radley and Jacobs, 2002). Consist-
5,7-DHT to reduce serotonergic innervation of ent with these findings, acute or chronic treatment
the forebrain. Analysis of dentate granule cells in with the 5-HT1A agonist 8-OH DPAT increases
rodents with reduced serotonergic innervation re- proliferation in the SGZ and the number of adult
vealed fewer dendritic spines and synapses, but born neurons (Santarelli et al., 2003; Banasr et al.,
otherwise normal dendritic complexity (Yan et al., 2004). The effects of activating 5-HT1AR appear
1997; Faber and Haring, 1999), suggesting that to be restricted to proliferation and do not affect
serotonergic signaling is important for selected, the differentiation of newborn progenitors into
and not all, aspects of the neuronal maturation neurons or glial cells. The increase in proliferation
process. While the limitations inherent to these could reflect a change in rate of progression
pharmacological lesion studies must be acknowl- through the cell cycle or an increase in the size of
edged, these studies illustrate how deficits in the the proliferative pool in the SGZ. It is unclear,
5-HT system can have consequences for the mat- given the experimental design employed in these
uration of dentate granule cells. Given the striking studies, whether activation of 5-HT1AR also
recapitulation in adult neurogenesis of the early- influences the survival of newborn neurons. Inter-
developmental neuronal maturation process, it is estingly, mice lacking the 5-HT1AR fail to re-
likely that changes in 5-HT signaling have similar spond to the neurogenic effects of chronic
consequences on neurons born in adulthood fluoxetine (Santarelli et al., 2003).
(Esposito et al., 2005; Laplagne et al., 2006; Over- Two other 5-HT receptors, the 5-HT2A and
street-Wadiche and Westbrook, 2006). Indeed, the 5-HT1B receptors, have also been implicated in
well-characterized effects of 5-HT receptor agonist cell proliferation in the adult SGZ. While neither
and antagonists and SSRIs on adult neurogenesis 5-HT1B agonists nor antagonists affect baseline
(Malberg et al., 2000) solidify the link between cell proliferation, the former can, however, restore
5-HT and adult hippocampal neurogenesis. Im- normal levels of proliferation in PCPA pretreated
portantly, studies on 5-HT receptors, which are rats. These data suggest that effects of the 5-HT1B
discussed next, offer a glimpse into the ways by receptor on cell proliferation are small under
which altered serotonin levels as a consequence of physiological conditions, but can become impor-
a genetic polymorphism, such as the 5-HTT poly- tant when 5-HT levels are decreased. The phar-
morphism, can influence the birth and maturation macology of the 5-HT2A receptor is also complex.
of newborn dentate granule cells. 5-HT2A antagonists decrease cell proliferation
The effects of 5-HT levels on neurogenesis re- but agonists have no effect (Banasr et al., 2004),
flect the sum of interactions between the synthesis suggesting that under physiological conditions,
of 5-HT, its release and its actions at different the 5-HT2A-dependent signaling pathways
5-HT postsynaptic receptors acting in both a cell that modulate neurogenesis may be saturated.
autonomous and non-cell autonomous manner. Taken together, these observations reveal the
The effects of 5-HT on a newborn neuron depend differential effects of recruiting different post-
on the repertoire of 5-HT receptors that it ex- synaptic 5-HT receptors on hippocampal neuro-
presses. Since the maturation of newborn neurons genesis.
is intimately connected with the activity of the Central to understanding how changes in 5-HT
network, it is also influenced by the actions of levels influence neurogenesis is knowledge of the
different 5-HT receptors expressed within the hip- expression of different 5-HT receptors in neural
pocampal formation in interneurons, mature dent- progenitors and at different stages of their matu-
ate granule cells and afferent projections arising ration. Conspicuously absent from the pharmaco-
in the entorhinal cortex. The 5-HT1A receptor logical studies is precise information for 5-HT
(5-HT1AR) is the best-studied 5-HT receptor in receptor distribution in the adult SGZ. The
the context of adult hippocampal neurogenesis. 5-HT1AR is an exception to the rule. We know
Acute administration of 5-HT1A antagonists re- that the 5-HT1AR is expressed at very low levels,
sults in decreased cell proliferation in the adult if any, in the SGZ of the rodent hippocampus.
705

In both rat and mouse, 5-HT1AR expression is the glucocorticoids, which mobilize energy, in-
restricted to the mature dentate granule cells crease cardiovascular tone and influence immune
rather than the immature population of cells dur- and nervous system functions. In response to a
ing development of the dentate gyrus (Patel and stressful event, for example, a neuroendocrine cas-
Zhou, 2005; Sahay and Hen, unpublished data). cade is initiated in which corticotropin releasing
The effect of 5-HT1AR agonists on cell prolifer- hormone (CRH) is released from neurons in the
ation is, therefore, likely to be non-cell autono- paraventricular nucleus (PVN) in the hypo-
mous. It is possible that the 5-HT1AR is required thalamus, which then triggers release of corticotro-
in hilar interneurons or mature dentate granule pin by the anterior pituitary to stimulate
cells to mediate the effects of 5-HT on cell prolif- glucorticoid secretion by the adrenal cortex. A hy-
eration. Cell-type specific ablation and overexpres- peractive HPA, on the other hand, results in the
sion of the 5-HT1AR will reveal its precise oversecretion of glucocorticoids with deleterious
contribution in different cell types to adult hippo- consequences for the physiology of the viscera and
campal neurogenesis. brain (Sapolsky, 2000). In addition, increased
The specific role of different 5-HT receptors in glucocorticoid levels can impair serotonergic
the maturation and integration of newborn neu- signaling (Joels and van Riel, 2004). It is there-
rons is still to be elucidated. It is also unclear how fore not surprising that the stress response is
5-HT may impact the turnover of mature dentate tightly regulated by efferents from multiple brain
granule cells or influence the survival of newborn regions that converge onto the PVN, where
dentate granule cells. The role for distinct 5-HT incoming information is integrated to elicit an
receptors in regulation of developmental processes adaptive response. Afferent inputs of the PVN are
such as dendritic development, synaptogenesis and inhibitory or facilitative in nature and arise in
glutamate receptor trafficking (Kondoh et al., brain stem nuclei, amygdala, cortex, septum and
2004; Kvachnina et al., 2005; Yuen et al., 2005a, the hippocampus, and are in turn regulated by a
b) suggests that 5-HT receptor-dependent mecha- negative feedback system. The hippocampus, for
nisms during development may be conserved in example, exercises a powerful inhibitory influence
neurogenesis in the adult brain depending on on HPA function to terminate the stress response,
which 5-HT receptors are expressed and when and is in turn regulated by glucocorticoids acting
during neuronal maturation. In addition, 5-HT on cognate receptors expressed within the hippo-
signaling can induce the production of ne- campal formation. Dysregulation of such control
urotrophins and growth factors known to regu- can, therefore, result in hypersecretion of gluco-
late hippocampal neurogenesis. corticoids and an exaggerated stress response
that can severely impact neural circuit function.
Consistent with preclinical findings, about half of
The role of stress in MDD and its effects on all depressed patients show a blunted response to
neurogenesis the dexamethasone suppression test (Carroll et al.,
1968b; Holsboer et al., 1982), which measures the
Stressful life experiences play a pivotal role in de- ability of an exogenous glucocorticoid receptor
velopment of MDD in individuals with a genetic agonist to suppress endogenous stress hormone
vulnerability (Holsboer and Barden, 1996; Gold release. Moreover, patients with MDD show ele-
and Chrousos, 2002). Major stressors precede the vated levels of CRH in the cerebrospinal fluid,
appearance of the first symptoms, and dysregula- increased numbers of CRH containing cells in the
tion of the hypothalamic-pituitary-adenocortical PVN, and decreased CRH binding in the prefron-
(HPA) system is often observed in patients with tal cortex. Thus, the optimal function of the
MDD (Carroll et al., 1968a). An optimally func- hippocampal formation is a critical factor in mod-
tional HPA system enables the organism to re- ulation of the stress response, and an exaggerated
spond appropriately to stressful stimuli by stress response can, in turn, negatively impact hip-
controlling the production of adrenal steroids, pocampal function.
706

One effect of stress on neural circuitry within the mind. For example, glucocorticoids have deleteri-
hippocampal formation is the suppression of ne- ous effects on neurogenesis during stressful epi-
urogenesis by glucocorticoids. Stress-induced sup- sodes but not during physical activity. This may be
pression of cell proliferation in the DG has been due to the fact that physical activity, unlike stress,
reported in several different mammalian species elicits the production of growth factors that may
and there is considerable evidence arguing for a buffer the effects of corticosteroids on neurons.
role for glucocorticoids as the mediators of the Likewise, given the differences in the developing
stress response. Elimination of circulating adrenal and adult brain, an increase in glucocorticoids
steroids by adrenalectomy, for example, increases during early postnatal life may have profoundly
cell proliferation and neurogenesis in the adult different effects from those in adulthood. A recent
dentate gyrus (Cameron and McKay, 1999). study modeling early life stress using a maternal
Exogenous administration of corticosterone, on separation paradigm in rodents supports this idea
the other hand, suppresses proliferation (Cameron (Mirescu et al., 2004). Maternal separation during
and Gould, 1994). Glucocorticoids have been the early postnatal period in rodents leads to per-
shown to inhibit the proliferation and differenti- sistent changes in the HPA axis, protracted release
ation of neural progenitors, and also the survival of corticosterone in response to mild stressors,
of young neurons (Wong and Herbert, 2004, increased anxiety behavior (Huot et al., 2001),
2006). These effects are likely to be mediated di- impaired maternal care (Lovic et al., 2001) and
rectly through high affinity mineralocorticoid re- impaired spatial navigation (Huot et al., 2002).
ceptors (MR) and low affinity glucocorticoid Rats subjected to prolonged maternal deprivation
receptors (GR) that are expressed at various stages during the early postnatal period also showed re-
of maturation and also indirectly through changes duced proliferation in the dentate gyrus of neural
in network activity of the hippocampal formation progenitors in adulthood. Interestingly, the early
(see chapter by Joels in this volume). Analysis of life stressor did not affect the number of mature
receptor distribution in the dentate gyrus of ro- neurons, suggesting a compensatory increase in
dents reveals that GRs are expressed in both neu- survival of newly generated neurons. The authors
ral progenitors and mature dentate granule cells, in this study showed that the change in prolifer-
while MRs are expressed only in the latter. ation could be rescued in adulthood by adrenal-
Throughout most of adulthood, neither GRs ectomy and by reducing corticosteroid production.
nor MRs are expressed in immature neurons Without adrenalectomy, the corticosterone levels
(Cameron et al., 1993; Garcia et al., 2004a). How- were normal in stressed animals, not only under
ever, in aged rodents GR and MR expression is baseline conditions but also in response to a
seen in immature neurons, suggesting that imma- stressor, suggesting that the suppressed prolifera-
ture neurons at this stage, but not earlier in the tion was likely to be result of increased sensitivity
animal’s life, may show increased sensitivity to to corticosteroids rather than increased levels
corticosterone action. (Mirescu et al., 2004). It is tempting to speculate
The dynamic expression of GR and MR during that a shift in the temporal pattern of MR and GR
neurogenesis and the ability of corticosteroids to distribution during neurogenesis may contribute to
regulate the expression of growth factors such as this increased sensitivity.
IGF-1, BDNF and EGF (Islam et al., 1998; Schaaf It is possible that changes in neurogenesis as a
et al., 2000), which have distinct effects on prolif- result of HPA axis dysregulation contribute to the
eration and survival (Kuhn et al., 1997; Aberg pathophysiology of MDD by affecting the role of
et al., 2000; Sairanen et al., 2005), illustrate the cell the hippocampus in learning and other cognitive-
autonomous and non-cell autonomous ways by emotional processes. Preclinical studies (discussed
which corticosteroids can affect neurogenesis. It later) have proven tremendously informative in
should be noted that the sensitivity of neurons or defining the physiological relevance of adult hip-
neural progenitors to corticosteroids must be stud- pocampal neurogenesis to these processes, and
ied with the state of the neuron or network in have greatly facilitated our understanding of how
707

stress-mediated suppression of neurogenesis may onset. This delay is likely to reflect changes in
be relevant to the pathophysiology of MDD. structural and synaptic plasticity in the brain
A second possible consequence of the stress-medi- mediated by multiple mechanisms involving
ated suppression of neurogenesis is that the ability monoaminergic signaling and neurotrophins.
of the hippocampus to regulate HPA activity be- PET imaging studies on MDD patients treated
comes compromised. However, evidence directly with SSRIs such as paroxetine and fluoxetine have
linking changes in dentate gyrus function with helped define a neuroanatomical basis comprising
HPA activity is scarce. Lesions of the hippocam- corticolimbic circuits (Seminowicz et al., 2004).
pus and fimbria-fornix transections reduce the Structures that showed changes in metabolic ac-
ability of dexamethasone to inhibit stress induced tivity included the subgenual cingulate, hippocam-
adrenocortical responses and results in hyperse- pus and prefrontal cortex (Mayberg et al., 2000;
cretion of glucocorticoids and ACTH following Kennedy et al., 2001). One form of structural
stressful stimuli (Knigge, 1961; Sapolsky et al., plasticity within the hippocampus that is consist-
1989; Herman et al., 1992; Feldman and ent with the delayed onset of ADs is the birth and
Weidenfeld, 1993; Bratt et al., 2001; Goursaud subsequent integration of newborn dentate gran-
et al., 2006). Antagonism of GR in the rat hippo- ule cells in the adult dentate gyrus. Moreover,
campus results in hypersecretion of ACTH almost all ADs known to date increase adult
and corticosterone following stressful stimuli neurogenesis. Therefore, the idea that ADs may
(Sapolsky, 1994; Feldman and Weidenfeld, 1999). work through enhancing neurogenesis has received
Analysis of subfields within the hippocampal for- abundant attention and is now considered central
mation confirms the differential contributions to the neurogenic hypothesis of MDD (Malberg
of CA fields and the dentate gyrus in the ventral and Schechter, 2005).
hippocampus in regulating HPA reactivity. In-
triguingly, lesion studies indicate that damage of
ventral subiculum but not ventral CA1 or dentate Neurogenic effects of antidepressant treatments
gyrus results in prolonged glucocorticoid re-
sponses (Herman et al., 1992, 1995). By contrast, Numerous groups have shown that different
electrical stimulation of DG in anesthesized ani- classes of ADs including 5-HT and norepinephrine
mals inhibits corticosteroid secretion (Dunn and selective reuptake inhibitors, tricyclics, monoa-
Orr, 1984). These studies suggest that lesions mine oxidase inhibitors, phosphodiesterase inhib-
of DG in adulthood may be compensated for by itors and electroconvulsive shock therapy increase
activity in structures downstream such as the sub- neurogenesis (Madsen et al., 2000; Malberg et al.,
iculum. Genetic manipulations that specifically 2000; Manev et al., 2001; Nakagawa et al., 2002b).
impair glucocorticoid signaling in the DG both in AD treatment does not appear to affect the ratio
adulthood and in early postnatal life will prove of newly generated neurons to glial cells, with the
critical in establishing the link between hippocam- majority of newborn cells adopting the neuronal
pal neurogenesis and HPA reactivity. fate. The neurogenic effects of ADs are specific to
In sum, the well-documented effects of stress on the SGZ, and are not observed in other compo-
neurogenesis suggest that patients with MDD are nents of the ventricular system such as the lateral
likely to show reductions in neuronal proliferation. ventricles or the subventricular zone. Moreover,
While it is plausible that a reduction in neurogen- administration of non-AD psychotropic drugs
esis in turn contributes to HPA dysregulation, such as haloperidol does not increase hippocam-
there is as yet no evidence for this. pal neurogenesis (Eisch, 2002). Other treatments
reported to have AD effects, including exercise
Neurogenesis and antidepressants (Babyak et al., 2000; Singh et al., 2001; Motl et al.,
2004), environmental enrichment and estrogen,
It is well recognized that all of the major classes of have also been shown to increase neurogenesis
ADs are associated with a several week delay in (van Praag et al., 1999; Tanapat et al., 1999;
708

Rhodes et al., 2003; Meshi et al., 2006). In addi- SSRIs act directly on progenitors or immature
tion, also lithium, which is used in the treatment of neurons to influence processes such as prolifera-
bipolar disorder increases neurogenesis (Chen tion, differentiation, maturation and survival. In
et al., 2000). The one AD treatment that has not addition, SSRIs are also likely to modulate net-
been shown to enhance neurogenesis is repetitive work activity within the dentate gyrus and as a
transcranial magnetic stimulation or rTMS, which result, regulate neurogenesis indirectly. By virtue
is still awaiting FDA approval (Loo and Mitchell, of their effects on neurogenesis, SSRIs may be ca-
2005). While rTMS has been shown to reverse the pable of driving replacement or turnover within
effects of chronic psychosocial stress on stress the dentate gyrus and catalyzing the insertion of
hormone levels, it does not upregulate neurogen- newly generated neurons with distinct electrophys-
esis (Czeh et al., 2002). iological and biochemical properties. The best-
That ADs block the behavioral effects of stress characterized effect of SSRIs to date is the increase
and restore normal levels of neurogenesis in the in proliferation of neural progenitors in the SGZ.
adult hippocampus lends further credence to the Studies in rodents indicate that a 14-day, but not
possibility that ADs may work by increasing ne- shorter-term, administration of fluoxetine (1–5
urogenesis to exert their behavioral effects. In tree days) is sufficient to upregulate cell proliferation
shrews, chronic exposure to psychosocial conflict (Malberg et al., 2000). By contrast, a longer treat-
results in a decrease in cell proliferation, which is ment regimen is required to enhance survival
blocked by treatment with the atypical AD tianep- of newly generated neuroblasts. Fluoxetine treat-
tine (Czeh et al., 2001). In another model of de- ment for 28 days, but not 14 days, following
pression, the learned helplessness (LH) paradigm, BrdU injection resulted in an increase in cell sur-
exposure to inescapable shock engenders prode- vival (Malberg et al., 2000; Nakagawa et al.,
pressive behavior and a reduction in hippocampal 2002a).
cell proliferation, both of which, are reversed by The delay with which the neurogenic effects
AD treatment (Cryan et al., 2002; Malberg and emerge after the initiation of SSRI treatment
Duman, 2003). In addition, ECS enhances cell provides a potential mechanism to explain the
proliferation after chronic corticosterone treat- therapeutic lag in the effects of these drugs.
ment (Hellsten et al., 2002). Namely, it could be that the therapeutic effects
It is well known that ADs have pleiotropic depend on the increase in proliferation, which it-
effects on neuronal circuits. That a diverse range self requires several weeks of treatment. However,
of ADs appears to enhance neurogenesis indicates closer inspection of this hypothesis reveals several
that the dentate gyrus may be a neuroanatomical shortcomings. It is unlikely that a boost in prolif-
substrate to target for the development of novel eration would produce an immediate psychologi-
AD treatments. In the next section, we describe cal effect. Rodent studies indicate that newborn
recent preclinical studies elucidating how SSRIs neurons do not become functionally integrated
modulate adult hippocampal neurogenesis, and until approximately 2–4 weeks after exiting the cell
then in the following sections we describe work cycle, so it would seem that any functional effects
from animal models aimed at identifying whether elicited by the increase in proliferation would not
neurogenesis is necessary for the behavioral effects appear until that time (Esposito et al., 2005). This
of AD treatments. means that behavioral effects of SSRIs mediated
by increased proliferation should not manifest
until about 4 weeks after treatment initiation. In
Serotonin-dependent ADs and hippocampal contrast, some behavioral effects of AD drugs in
neurogenesis animal models begin immediately following acute
treatment, and virtually all the behavioral effects
SSRIs represent the most successful class of ADs manifest within 4 weeks of SSRI treatment. In
identified to date. Based on our understanding of primates the time required for maturation of new
neurogenesis in the SGZ, it is conceivable that neurons is greater than in rodents (Kohler et al.,
709

2006), so the predicted therapeutic lag would be the increase in ANPs translates into a net increase
even longer. A recent meta-analysis of clinical data in the number of new neurons.
indicates that some of the therapeutic effects of A net increase in the number of mature neurons
SSRIs may commence very rapidly after treatment implies that SSRIs also increases the population of
initiation (within 1–2 weeks) (Taylor et al., 2006). immature neurons. It is presently not clear how
Thus, increases in proliferation are unlikely to un- SSRIs influence the maturation of immature neu-
derlie the early onset effects of these drugs, but it rons with regards to their physiological properties,
remains plausible that neurogenesis contributes to synaptic connectivity and dendritic complexity.
the more long-term effects. What is well appreciated, however, is that SSRI
Consistent with this interpretation, remarkable treatment results in the induction of growth fac-
preliminary data from Meltzer, Deisseroth and tors and neurotrophins whose effects on matura-
colleagues suggest that mature adult-born neurons tion and survival are well understood (Carlezon
contribute to the behavioral effects of SSRIs et al., 2005; Duman and Monteggia, 2006) and,
(Meltzer et al., 2006). In this study, rats were importantly, whose receptors are expressed in im-
given 7 days of fluoxetine treatment and then mature neurons. It is also possible that SSRIs may
tested behaviorally 1 month later. At that time induce the secretion of growth factors from neural
point, the previously treated rats showed enhanced progenitors, which then influence the function of
performance in the forced swim test, and histology neighboring mature granule cells in a paracrine
revealed an increase in the number of newly gen- manner. There is no evidence for this as yet.
erated neurons. The results suggest that the pro- The consequences of increased proliferation and
liferative effects of SSRIs may appear sooner than survival of newly generated cells following fluoxe-
once thought (after 1 week rather than 2 weeks tine treatment for the net size of the granule cell
of treatment), and that some behavioral effects of population of the dentate gyrus have not been as-
these drugs depend not on the acute presence of certained. One study suggested that there may not
the drug but rather on slow, time-dependent proc- be a net change in the size of the dentate gyrus
esses initiated by drug treatment, such as neuro- because the increase in newly generated neurons is
genesis and circuit reorganization. offset by death of previously born mature granule
The studies that examined the effects of fluoxe- cells (Sairanen et al., 2005). Based on their data
tine on proliferation do not identify the specific showing that chronic fluoxetine treatment in-
types of neural progenitors that respond to creases not only proliferation and survival of
changes in 5-HT levels. A recent study addressed newly generated neurons but also the rate of
this question using transgenic mice in which ex- apoptosis, the authors argue that the net size of the
pression of a fluorescent reporter gene was regu- dentate gyrus does not change with chronic AD
lated by a nestin promoter fragment. Because treatment.
nestin is expressed in multiple progenitor cell Finally, it is plausible that SSRI treatment alters
types, this approach allowed the visualization the physiological properties of newly generated
and quantification of the distinct sub types of neurons, generating a cohort of cells within the
neural progenitors that reside within the SGZ dentate that are unique. These unique cells may
(Encinas et al., 2006). The results of this study confer greater adaptive potential to the dentate
showed that only a specific proliferative cell type, gyrus than naı̈ve newly generated neurons. Much
the transiently amplifying neural progenitor work remains to be done in this area to address the
(ANP) that exists as an intermediate between the different ways by which SSRIs influence neuro-
type I radial glial-like neural progenitor and the genesis.
type II cell (Filippov et al., 2003; Tozuka et al., Delineating the pattern of expression of distinct
2005; Encinas et al., 2006), directly responds to 5-HT receptor subtypes within different cell types
fluoxetine. Moreover, the study confirmed that in the SGZ and the DG will shed light on how
once exposure to fluoxetine ends, the rate of pro- changes in 5-HT can elicit the diverse range of
genitor cell division is restored to baseline and that effects discussed here. It follows that the
710

identification of genes downstream of 5-HT recep- one side of the chamber to the other cancels or
tors that mediate the behavioral effects of ADs will terminates the shock). Exposure to inescapable
pave the way for developing neurogenic non- shock impairs subsequent acquisition of the es-
monoamine based therapeutics. Preclinical studies cape/avoidance task (relative to naı̈ve subjects),
have proven invaluable in defining the contribu- arguably because the animal has learned it is help-
tion of neurogenesis to the etiology and treatment less (Willner and Mitchell, 2002). LH training also
of MDD as they allow for discernment between reduces cell proliferation in the DG. Treatment
correlation and causality. with AD drugs alleviates both the behavioral help-
lessness (Malberg and Duman, 2003; Chourbaji et
al., 2005) and the reduction in cell proliferation
Preclinical studies: in the search for causality
(Malberg and Duman, 2003). However, there are
several reasons why reductions in proliferation are
Ultimately, whether neurogenesis is causally re-
not a likely mechanism of the behavioral helpless-
lated to the etiology or treatment of depression
ness. First, behavioral helplessness manifests im-
requires the use of animal models in which neuro-
mediately with exposure to inescapable shock, and
genesis and emotional state can be experimentally
it is unlikely that a reduction in proliferation could
manipulated. If reduced neurogenesis contributes
so rapidly give rise to a behavioral effect. If the
to depression, it should be possible to produce a
reduction in proliferation were to impact behavior,
depressive phenotype by experimentally reducing
the effects would more likely manifest 1–3 weeks
neurogenesis. Conversely, behavioral manipula-
after the onset of the reduction in proliferation, at
tions that produce a depressive phenotype should
the time when the newborn cells would be becom-
reduce neurogenesis and do so before the behavi-
ing functionally integrated into DG circuits.
oral manifestations of depression develop. Of
Similarly, acute dosing with AD drugs is sufficient
course, the predictive value of such experiments
to alleviate behavioral helplessness (Malberg and
will depend on the specificity of the methods for
Duman, 2003; Chourbaji et al., 2005), but the
manipulating neurogenesis and the validity of the
neurogenic effect of these drugs requires chronic
animal models of depression. This section will re-
treatment (Malberg et al., 2000). Finally, a recent
view recent work that addresses the role of neuro-
study has demonstrated that LH training in rats
genesis in the etiology and treatment of MDD
reduces cell proliferation in all subjects, but only a
assessed in preclinical models.
subset of subjects display behavioral helplessness
(Vollmayr et al., 2003). One unexplored possibility
Lessons from stress based depression paradigms and invoked by these studies is that the extant popu-
neurogenesis lation of immature neurons, rather than the pro-
liferative pool, is a substrate for the depressogenic
One prediction drawn from the neurogenic hy- effects of stress. Studies are underway to test this
pothesis of MDD is that behavioral manipulations hypothesis.
that produce depressed behavior in animals should Neurogenesis can more plausibly be linked to
produce a concomitant decrease in neurogenesis. the effects of chronic exposure to stress, which
Several methods have been used to produce de- have been studied extensively in animals. This
pression-like behavior in rodents, all involving the work typically involves exposing animals to vari-
exposure to inescapable stress. One such procedure ety of mild stressors over a period of several weeks.
is LH, originally developed by Seligman and col- Stressors include food and water deprivation, tem-
leagues (Overmier and Seligman, 1967; Seligman perature changes, restraint and tail suspension
and Beagley, 1975). LH is a relatively short-term (Strekalova et al., 2004; Willner, 2005; Mineur
procedure in which animals are exposed to ines- et al., 2006). There are variations in the methods
capable shock inside a conditioning chamber. used for chronic stress in rats (Willner et al., 1987,
Subjects are then given an escape/avoidance task 1992; Willner, 2005) and mice (Mineur et al., 2003,
in which shock is controllable (e.g., shuttling from 2006; Strekalova et al., 2004). The effects of
711

chronic stress include a reduction in sucrose pref- used to target X-rays specifically over the hippo-
erence, which has been interpreted as anhedonia campus, while protecting regions anterior and
(Willner et al., 1987; Strekalova et al., 2004), al- posterior, including the subventricular zone.
terations in sleep–wake cycle, reduced sexual and Blocking neurogenesis with this procedure has no
self-care behavior, and increases in anxiety-like effect in a number of relevant behavioral tasks,
behavior in traditional tests such as the elevated- including the novelty-suppressed feeding test
plus maze (Willner, 2005). Unlike the effects of (Santarelli et al., 2003) and the novelty-induced
LH, the behavioral effects of chronic stress are hypophagia test (our unpublished data). Both of
ameliorated by chronic but not acute treatment these tests are AD screens that measure the latency
with ADs (Willner et al., 1987; Yalcin et al., 2005), of a mouse to venture into the center of an open
suggesting that chronic stress may be a better field or novel context to obtain food. Latency-to-
model of depression because it captures the ther- feed is decreased by chronic AD drug treatment
apeutic lag seen in human patients. An abundance and acute anxiolytic treatment, but not by acute
of research demonstrates that hippocampal neuro- AD drug treatment (Bodnoff et al., 1989; Dulawa
genesis is reduced by a variety of chronic stress et al., 2004). Irradiation does not affect behavior in
procedures (for review, see Duman, 2004), and this two traditional anxiety tests, the elevated-plus
reduction in neurogenesis is blocked by chronic maze and light-dark choice test (our unpublished
treatment with AD drugs (Alonso et al., 2004). data), nor does it increase the susceptibility of mice
Moreover, a recent study has shown that among to the effects of chronic stress (Santarelli et al.,
rats exposed to chronic stress, only a subset re- 2003).
spond behaviorally to SSRI treatment (Jayatissa We have also examined some of these behaviors
et al., 2006). Interestingly, neurogenesis was re- in a transgenic mouse line in which neuronal pro-
stored to normal levels only in the behaviorally liferation can be blocked conditionally. The mouse
identified responders. line expresses herpes-simplex virus thymidine kin-
ase (HSV-TK) under control of the GFAP pro-
moter (GFAP-TK) (Garcia et al., 2004b). Mitotic
Does blockade of neurogenesis produce symptoms of cells expressing HSV-TK are killed by the antiviral
depression? drug ganciclovir. Thus, in this mouse, the dividing
neuronal progenitors, which express GFAP, are
The animal research on chronic stress evidences an killed after ganciclovir is administered, and as
intriguing correlation between the rate of neuro- a result, neurogenesis is reduced to low levels
genesis and emotional state. Chronic stress pro- (Garcia et al., 2004b; Saxe et al., 2006). As with
duces a behavioral phenotype analogous to aspects irradiation, blocking neurogenesis in this mouse
of depression while reducing neurogenesis. AD line had no effect on anxiety-like behavior in sev-
drugs restore neurogenesis and ameliorate the be- eral tests, including the open field, NSF test, or
havioral phenotype. Does this correlation reflect light-dark choice test (Saxe et al., 2006). Thus, we
that neurogenesis is a causal mechanism for these have not found any evidence that blocking neuro-
behavioral changes? Or is neurogenesis simply a genesis in adult mice produces a depression-like
marker for changes in other biological pathways phenotype.
or network activity?
We have begun to address this question exper-
imentally by blocking hippocampal neurogenesis A role for hippocampal neurogenesis in learning
and then examining the behavioral consequences.
One method of arresting neurogenesis is to subject The only domain in which direct behavioral effects
the brain to low doses of X-irradiation. Irradiation of arresting neurogenesis have been reported is in
kills mitotic cells such that neurogenesis is arrested learning and memory. Indeed, the very first evi-
virtually completely (Monje et al., 2002, 2003; dence for a behavioral function of neurogenesis in
Wojtowicz, 2006). In our laboratory, a shield is mammals came from studies of learning and
712

memory. These studies showed that participating requirement of neurogenesis for the encoding of
in a hippocampus-dependent classical condition- novel contexts and/or for assigning emotional va-
ing procedure (trace eyeblink conditioning) en- lence to contexts.
hances the survival of newborn neurons in the How (and whether) these learning impairments
SGZ (Gould et al., 1999a) and that reducing ne- relate to depression is unclear. Cognitive impair-
urogenesis with the anti-mitotic compound me- ments have been reported in depression, but cog-
thylazoxymethanol acetate (MAM) impairs nitive impairment is not a cardinal feature of the
acquisition of the same task (Shors et al., 2001). disease, in contrast to some other psychiatric ill-
The use of MAM is encumbered by deleterious nesses, namely schizophrenia. Still, it is certainly
side effects and moreover, it does not completely the case that depressed patients have an impaired
block neurogenesis (Dupret et al., 2005). Subse- ability to ‘‘contextualize’’ negative emotions, in
quent experiments using more targeted methods that these emotions are overgeneralized across ex-
for arresting neurogenesis have confirmed that ne- periences and situations. Much more research into
urogenesis is required for some hippocampus- the putative role of neurogenesis in contextual
dependent learning tasks (Snyder et al., 2005; learning will be needed to determine whether re-
Saxe et al., 2006; Winocur et al., 2006). A recent duced neurogenesis could give rise to these features
study from our laboratory has also revealed a of depression.
paradoxical role for hippocampal neurogenesis
in a hippocampus-dependent working memory
paradigm, where ablation of newly generated Neurogenesis is required for some behavioral
neurons results in improved performance (Saxe effects of AD drugs
et al., 2007).
Perhaps the most compelling of these findings is Although animal models have not provided evi-
the requirement of neurogenesis for contextual dence that reduced neurogenesis causes depres-
fear conditioning, which has been demonstrated in sion-like symptoms, these models have produced
two species (rats and mice) using two different evidence that neurogenesis is involved in the ther-
methods for arresting neurogenesis (Saxe et al., apeutic effects of AD drugs. Three recent studies
2006; Winocur et al., 2006). Contextual fear con- have used the targeted irradiation procedure de-
ditioning is a form of Pavlovian conditioning pro- scribed above to test whether DG neurogenesis is
duced by pairing a distinctive context (spatial required for the behavioral effects of AD drugs in
location) with footshock. As a result of the pair- rodent models. A study conducted in our labora-
ing, animals exhibit characteristic fear responses tory demonstrated that neurogenesis is required
(e.g., freezing, defecation, potentiation of the star- for the effects of both imipramine, a classic tricy-
tle response) when re-exposed to the context. clic AD, and fluoxetine in two mouse behavioral
Lesions to the hippocampus often impair this screens for AD activity (Santarelli et al., 2003), the
form of learning (Phillips and LeDoux, 1992; NSF test and a chronic stress procedure. Impor-
Matus-Amat et al., 2004), presumably because the tantly, this study has been replicated in our lab-
hippocampus participates in mnemonic encoding oratory using the GFAP-TK mice (unpublished
of the spatial context. Blocking neurogenesis prior results). In a separate series of experiments con-
to training in this procedure reduces the amount of ducted by another laboratory, the synthetic can-
the contextual fear expressed when rodents are re- nabinoid HU210 was shown to have AD-like
exposed to the training context. Importantly, effects in the NSF paradigm following 10 days of
blocking neurogenesis does not impair fear condi- treatment, and, interestingly, this effect was
tioning to a discrete tone stimulus, indicating that blocked by X-ray irradiation (Jiang et al., 2005).
shock sensitivity and motor control of the fear In addition, there is a recent preliminary report
response are not impaired. The impairment of using rats (Meltzer et al., 2006) that irradiation
contextual fear conditioning may thus reflect a blocks the behavioral effects of fluoxetine in the
713

forced swim test. Thus, a neurogenic dependence Summary


for the behavioral effects of ADs has been revealed
for three different drugs in three different AD General insights: is neurogenesis a missing link or
screens and using two different ways to ablate ne- is the link still missing?
urogenesis.
The above work suggests that neurogenesis may Research on MDD in the last decade has led to
be a critical substrate for AD efficacy. However, considerable maturation of our understanding of
an important limitation of this work is the reliance how different neural circuits function in a normal
on a very limited number of animal models and brain and in the context of pathology. Several no-
AD treatments. It is unlikely that the three be- table findings have emerged from studies in hu-
havioral assays used in these experiments capture mans with MDD and preclinical models of MDD.
all the clinically relevant features of AD treat- First, the AD response and the pathogenesis
ments, and consequently it remains possible that of MDD may have different neural substrates.
some clinically important features of these treat- Second, the pathogenic mechanisms may differ
ments are neurogenesis-independent. Indeed, two from those that underlie the pathophysiology of
more recent studies from our lab confirm that this MDD. Third, any model explaining the etiology of
caveat is valid. MDD must incorporate genetic vulnerability,
One study from our laboratory (Holick et al., stressors, critical periods and multiple neural cir-
2007) examined the effects of AD drugs on be- cuits. In other words, MDD is more likely than
havior and DG neurogenesis in the Balb-c mouse not, a result of multiple ‘‘hits’’ to the brain (deficits
strain, a strain that exhibits high anxiety in be- in multiple neural substrates). This last finding
havioral tests. In this strain, chronic fluoxetine underscores the need for more refined preclinical
treatment reduced anxiety-like and depressive be- models for depression. In this section, we revisit
havior in the novelty-induced hypophagia and the neurogenic hypothesis of MDD with attention
forced-swim paradigms but failed to increase neu- to the evidence discussed thus far in the context of
ronal proliferation. Not surprisingly, the behavi- etiology, pathophysiology and treatment.
oral effects of these drugs are not blocked by
irradiation in this strain. A second study found
that the anxiolytic effects of environmental en- The neurogenic hypothesis and the etiology of MDD
richment do not require neurogenesis (Meshi et al.,
2006). Only recently the neurogenic hypothesis of MDD
In sum, these studies suggest that AD-like joined the neurotrophic and network hypotheses
effects can be achieved through at least two differ- as a candidate to explain the neurobiological basis
ent pathways, one that is neurogenesis-dependent of MDD.
and one that is neurogenesis-independent. AD Unlike the neurotrophic and network hypothe-
drugs and cannabinoids appear to use a neuro- sis, however, the neurogenic hypothesis implicates
genesis-dependent pathway in some circumstances only one brain region, the dentate gyrus, as the
but not in others; chronic stress may be one factor primary neural substrate for the etiology of MDD
that governs this dichotomy. Enrichment appears and the AD response.
to use a neurogenesis-independent pathway either As this review makes clear, the evidence for ne-
alone or in combination with the neurogenesis- urogenesis as an etiological factor in MDD is scarce
dependent pathway. An important outstanding at best. Pathological analyses of postmortem tissue
question not addressed by these experiments is obtained from patients with MDD have not re-
whether the upregulation of neurogenesis is suffi- vealed alterations in the size of the dentate gyrus,
cient to produce AD effects. Testing this hypoth- decreased number of proliferative cells, or changes
esis will require new methods for very specifically in the degree of ongoing apoptosis. The data on
upregulating aspects of neurogenesis. apoptosis do not indicate the age of neurons that
714

Fig. 2. A model to evaluate the role of neurogenesis as a substrate in the etiology, pathophysiology and treatment of MDD. MDD is
likely to arise from the synergistic effects of stress, biological vulnerability conferred by risk alleles and deficits in multiple neural
circuits. Whether hippocampal neurogenesis is involved in the etiology of MDD is at present unclear. Altered hippocampal neuro-
genesis may result as a consequence of pathogenic mechanisms and contribute to the pathophysiology of MDD. The treatment of
some, but not all, symptoms of MDD may rely on neurogenesis.

are dying and therefore, preclude an assessment of directing the emotional trajectory. Finally, longi-
changes in rates of turnover or survival. As noted tudinal neuroimaging studies in humans are
earlier in this chapter, more studies on postmortem needed to reveal how the hippocampal landscape
tissue obtained from non-medicated patients are changes with the appearance of the first symptoms
required to conclusively characterize dentate gyrus of MDD and over time.
neurogenesis in depressed individuals. Neverthe-
less, these data establish that changes in hippo-
campal volume are unlikely to result from changes The neurogenic hypothesis and the pathophysiology
in hippocampal neurogenesis. Preclinical studies of MDD
have yielded data more directly controverting the
role of neurogenesis in the etiology of MDD: ab- The increasingly appreciated role for neurogenesis
lation of neurogenesis in adult otherwise normal in hippocampus-dependent learning as defined by
animals does not engender a depression-like phe- work from several different laboratories using ro-
notype. However, there are some important cave- dents provides evidence that neurogenesis makes a
ats to these preclinical studies. It is almost functionally significant contribution to hippocam-
certainly the case that MDD involves the simul- pal circuits. It is thus plausible that the putative
taneous presence of multiple risk factors or ‘‘hits’’ reductions in neurogenesis associated with MDD
(Fig. 2). If this were the case, ablation of neuro- have important psychological implications. These
genesis in wild-type adult animals would not be could include cognitive deficits, which are a pos-
sufficient to elicit a depression-like phenotype. The sible mechanism of the emotional symptoms of the
ablation of neurogenesis might create a diathesis disease (Beck, 2005). In addition, it is now appre-
for depression that would only be revealed in the ciated that the hippocampus, particularly its
presence of other genetic or environmental insults. ventral (anterior, in humans) extent, has a central
Moreover, the timing with respect to when the role in emotional regulation. The development of
brain experiences an insult, whether it be genetic higher resolution neuroimaging techniques will
or environmental, is also critical. Clearly, more enable us to visualize changes in dentate gyrus ac-
studies are needed that incorporate the effects of tivity in patients with MDD and during learning.
stressors and assess the consequences of altering The use of novel genetic approaches to selectively
neurogenesis during the early postnatal period in manipulate the maturation of newborn neurons
rodents with genetic backgrounds that harbor influence their survival, and drive turnover of ma-
different risk alleles. These studies will reveal ture dentate granule cells will undoubtedly en-
whether changes in neurogenesis lend a diathesis hance our understanding of how neurogenesis
for MDD, and inform us about the complex in- contributes to dentate function and to the deficits
terplay between multiple neural circuits in seen in MDD.
715

The neurogenic hypothesis and the treatment of potentially different ways by which ADs may rely
MDD on neurogenesis for their behavioral effects (Fig. 1,
Drew and Hen, in press). It is plausible that
The finding that neurogenesis is one mechanism the short-term effects of ADs, for example, are
used by ADs to exert their behavioral effects has mediated by ADs acting on the extant reservoir
now been repeated by several different laborato- of adult-generated immature neurons or by accel-
ries using rodents. This is in striking contrast to erating the maturation process of an extant pop-
the absence of data implicating neurogenesis in the ulation of adult-generated immature neurons with
pathogenesis of MDD. One interpretation of this concomitant replacement/addition to the dentate
apparent paradox (besides the limitations of cur- gyrus. Likewise, the rapid effects of ADs in re-
rent preclinical models highlighted in section ‘‘The versing behavioral changes induced by stress in
neurogenic hypothesis and the etiology of MDD’’) paradigms such as LH discussed earlier could also
is that the etiology and treatment of MDD may be mediated through the extant population of im-
have different neural substrates. Such a dissocia- mature neurons or through adult-generated ma-
tion is suggested by a recent association study that ture dentate granule cells. In this regard, it is
looked at 21 candidate polymorphisms and worthy to note that inducing BDNF and CREB
showed that the genetic basis for the capacity to expression in the DG but not other hippocampal
respond to monoamine-based ADs differed from subfields (or other brain regions) is sufficient
that of susceptibility to MDD (Garriock et al., to elicit an AD response (Chen et al., 2001;
2006). An interesting parallel was demonstrated by Shirayama et al., 2002; Duman and Monteggia,
preclinical studies on BDNF and depression, 2006). Thus, the DG is an attractive neural
which showed that increased BDNF signaling in substrate for the AD response and these studies
the hippocampus is sufficient to induce AD-like highlight a previously unrecognized function for
effects, but genetic ablation of BDNF on its own the dentate gyrus. Given our present understand-
does not elicit a depression-like phenotype ing of DG function (see chapter by Kesner in
(Duman and Monteggia, 2006). While the data this volume), it is unclear how enhancement in
from preclinical studies and ADs are encouraging, processes such as pattern separation and conjunc-
more work is needed to solidify the requirement tive encoding contribute to AD-like behavior
for neurogenesis in mediating the behavioral assessed in these depression paradigms. Whether
effects of ADs. Conspicuously absent from the cognitive behavior therapy, an endorsed line of
roster of experiments are those that show that in- treatment often used in parallel with AD drugs,
creased neurogenesis is sufficient for the behavioral increases activity in the DG is yet to be deter-
effects of ADs. Experiments along these lines using mined.
genetic approaches to specifically increase the In conclusion, there is much work to be done on
number of newly generated neurons are currently hippocampal neurogenesis to ascertain its role in
underway in our laboratory. Moreover, given the the brain and still much more to establish a role
pleiotropic effects of ADs on neurogenesis, it at for it in MDD.
present unclear whether immature neurons or the The convergence of insights from preclinical
adult-generated mature dentate granule cells are studies and neuroimaging studies and identifi
required to induce behavioral change. Inducible cation of novel genetic risk alleles will undoubt-
genetic approaches using promoters specific to edly help establish whether the neurogenic
these different cell types will allow for cell type hypothesis can explain facets of this complex
specific manipulations to unequivocally identify and debilitating psychiatric disorder. At the
the cellular substrates and mechanisms underlying very least, the neurogenic hypothesis, like any
the neurogenic-dependent AD response. elegant hypothesis, has succeeded in generating
In interpreting the data on the neurogenic de- the momentum needed to rigorously test its
pendence of ADs, we must remind ourselves of the tenets.
716

Acknowledgments Becker, S. (2005) A computational principle for hippocampal


learning and neurogenesis. Hippocampus, 15: 722–738.
Bodnoff, S.R., Suranyi-Cadotte, B., Quirion, R. and Meaney,
The authors would like to thank members of the
M.J. (1989) A comparison of the effects of diazepam versus
Hen laboratory for helpful discussions. Funding several typical and atypical anti-depressant drugs in an an-
support was provided by NARSAD (A.S, M.R.D imal model of anxiety. Psychopharmacology (Berl.), 97:
and R.H) and by Charles H. Revson Foundation 277–279.
Senior Fellowship in Biomedical Science (M.R.D). Bratt, A.M., Kelley, S.P., Knowles, J.P., Barrett, J., Davis, K.,
Davis, M. and Mittleman, G. (2001) Long term modulation
of the HPA axis by the hippocampus. Behavioral, biochem-
ical and immunological endpoints in rats exposed to chronic
mild stress. Psychoneuroendocrinology, 26: 121–145.
References Bremner, J.D., Narayan, M., Anderson, E.R., Staib, L.H.,
Miller, H.L. and Charney, D.S. (2000) Hippocampal volume
Aberg, M.A., Aberg, N.D., Hedbacker, H., Oscarsson, J. and reduction in major depression. Am. J. Psychiatry, 157:
Eriksson, P.S. (2000) Peripheral infusion of IGF-I selectively 115–118.
induces neurogenesis in the adult rat hippocampus. J. Ne- Brezun, J.M. and Daszuta, A. (1999) Depletion in serotonin
urosci., 20: 2896–2903. decreases neurogenesis in the dentate gyrus and the subven-
Alonso, R., Griebel, G., Pavone, G., Stemmelin, J., Le Fur, G. tricular zone of adult rats. Neuroscience, 89: 999–1002.
and Soubrie, P. (2004) Blockade of CRF(1) or V(1b) recep- Brezun, J.M. and Daszuta, A. (2000) Serotonin may stimulate
tors reverses stress-induced suppression of neurogenesis in a granule cell proliferation in the adult hippocampus, as ob-
mouse model of depression. Mol. Psychiatry, 9: 278–286 224. served in rats grafted with foetal raphe neurons. Eur. J. Ne-
Altman, J. and Das, G.D. (1965) Autoradiographic and histo- urosci., 12: 391–396.
logical evidence of postnatal hippocampal neurogenesis in Bunney Jr., W.E. and Davis, J.M. (1965) Norepinephrine in
rats. J. Comp. Neurol., 124: 319–335. depressive reactions. A review. Arch. Gen. Psychiatry, 13:
Amrein, I., Slomianka, L. and Lipp, H.P. (2004) Granule cell 483–494.
number, cell death and cell proliferation in the dentate gyrus Cameron, H.A. and Gould, E. (1994) Adult neurogenesis is
of wild-living rodents. Eur. J. Neurosci., 20: 3342–3350. regulated by adrenal steroids in the dentate gyrus. Neurosci-
American Psychiatric Association (2000) Diagnostic and Sta- ence, 61: 203–209.
tistical Manual of Mental Disorders. Fourth Edition—Text Cameron, H.A. and McKay, R.D. (1999) Restoring production
Revision (DSMIV-TR). of hippocampal neurons in old age. Nat. Neurosci., 2:
Austin, M.P., Mitchell, P. and Goodwin, G.M. (2001) Cogni- 894–897.
tive deficits in depression: possible implications for functional Cameron, H.A. and McKay, R.D. (2001) Adult neurogenesis
neuropathology. Br. J. Psychiatry, 178: 200–206. produces a large pool of new granule cells in the dentate
Azmitia, E.C. and Segal, M. (1978) An autoradiographic anal- gyrus. J. Comp. Neurol., 435: 406–417.
ysis of the differential ascending projections of the dorsal and Cameron, H.A., Woolley, C.S. and Gould, E. (1993) Adrenal
median raphe nuclei in the rat. J. Comp. Neurol., 179: steroid receptor immunoreactivity in cells born in the adult
641–667. rat dentate gyrus. Brain Res., 611: 342–346.
Babyak, M., Blumenthal, J.A., Herman, S., Khatri, P., Dorai- Campbell, S., Marriott, M., Nahmias, C. and Macqueen, G.M.
swamy, M., Moore, K., Craighead, W.E., Baldewicz, T.T. (2004) Lower hippocampal volume in patients suffering
and Krishnan, K.R. (2000) Exercise treatment for major de- from depression: a meta-analysis. Am. J. Psychiatry, 161:
pression: maintenance of therapeutic benefit at 10 months. 598–607.
Psychosom. Med., 62: 633–638. Carlezon Jr., W.A., Duman, R.S. and Nestler, E.J. (2005) The
Banasr, M., Hery, M., Printemps, R. and Daszuta, A. (2004) many faces of CREB. Trends Neurosci., 28: 436–445.
Serotonin-induced increases in adult cell proliferation and Carroll, B.J., Martin, F.I. and Davies, B. (1968a) Pituitary-
neurogenesis are mediated through different and common 5- adrenal function in depression. Lancet, 1: 1373–1374.
HT receptor subtypes in the dentate gyrus and the subven- Carroll, B.J., Martin, F.I. and Davies, B. (1968b) Resistance to
tricular zone. Neuropsychopharmacology, 29: 450–460. suppression by dexamethasone of plasma 11-O. H.C.S. levels
Bannerman, D.M., Grubb, M., Deacon, R.M., Yee, B.K., Fel- in severe depressive illness. Br. Med. J., 3: 285–287.
don, J. and Rawlins, J.N. (2003) Ventral hippocampal lesions Caspi, A. and Moffitt, T.E. (2006) Gene-environment interac-
affect anxiety but not spatial learning. Behav. Brain Res., tions in psychiatry: joining forces with neuroscience. Nat.
139: 197–213. Rev. Neurosci., 7: 583–590.
Bayer, S.A., Yackel, J.W. and Puri, P.S. (1982) Neurons in the Caspi, A., Sugden, K., Moffitt, T.E., Taylor, A., Craig, I.W.,
rat dentate gyrus granular layer substantially increase during Harrington, H., Mcclay, J., Mill, J., Martin, J., Braithwaite,
juvenile and adult life. Science, 216: 890–892. A. and Poulton, R. (2003) Influence of life stress on depres-
Beck, A.T. (2005) The current state of cognitive therapy: a sion: moderation by a polymorphism in the 5-HTT gene.
40-year retrospective. Arch. Gen. Psychiatry, 62: 953–959. Science, 301: 386–389.
717

Castren, E. (2005) Is mood chemistry? Nat. Rev. Neurosci., 6: Duman, R.S. and Monteggia, L.M. (2006) A neurotrophic
241–246. model for stress-related mood disorders. Biol. Psychiatry, 59:
Chambers, R.A., Potenza, M.N., Hoffman, R.E. and Miranker, 1116–1127.
W. (2004) Simulated apoptosis/neurogenesis regulates learn- Dunn, J.D. and Orr, S.E. (1984) Differential plasma corticos-
ing and memory capabilities of adaptive neural networks. terone responses to hippocampal stimulation. Exp. Brain
Neuropsychopharmacology, 29: 747–758. Res., 54: 1–6.
Chapman, D.P., Whitfield, C.L., Felitti, V.J., Dube, S.R., Ed- Dupret, D., Montaron, M.F., Drapeau, E., Aurousseau, C., Le
wards, V.J. and Anda, R.F. (2004) Adverse childhood expe- Moal, M., Piazza, P.V. and Abrous, D.N. (2005) Methyl-
riences and the risk of depressive disorders in adulthood. J. azoxymethanol acetate does not fully block cell genesis in the
Affect. Disord., 82: 217–225. young and aged dentate gyrus. Eur. J. Neurosci., 22:
Chen, A.C., Shirayama, Y., Shin, K.H., Neve, R.L. and 778–783.
Duman, R.S. (2001) Expression of the cAMP response ele- Eisch, A.J. (2002) Adult neurogenesis: implications for psychi-
ment binding protein (CREB) in hippocampus produces an atry. Prog. Brain Res., 138: 315–342.
antidepressant effect. Biol. Psychiatry, 49: 753–762. Encinas, J.M., Vaahtokari, A. and Enikolopov, G. (2006) Flu-
Chen, G., Rajkowska, G., Du, F., Seraji-Bozorgzad, N. and oxetine targets early progenitor cells in the adult brain. Proc.
Manji, H.K. (2000) Enhancement of hippocampal neurogen- Natl. Acad. Sci. U.S.A., 103: 8233–8238.
esis by lithium. J. Neurochem., 75: 1729–1734. Eriksson, P.S., Perfilieva, E., Bjork-Eriksson, T., Alborn, A.M.,
Chourbaji, S., Zacher, C., Sanchis-Segura, C., Dormann, C., Nordborg, C., Peterson, D.A. and Gage, F.H. (1998) Ne-
Vollmayr, B. and Gass, P. (2005) Learned helplessness: va- urogenesis in the adult human hippocampus. Nat. Med., 4:
lidity and reliability of depressive-like states in mice. Brain 1313–1317.
Res. Brain Res. Protoc., 16: 70–78. Esposito, M.S., Piatti, V.C., Laplagne, D.A., Morgenstern,
Cryan, J.F., Markou, A. and Lucki, I. (2002) Assessing anti- N.A., Ferrari, C.C., Pitossi, F.J. and Schinder, A.F. (2005)
depressant activity in rodents: recent developments and fu- Neuronal differentiation in the adult hippocampus recapit-
ture needs. Trends Pharmacol. Sci., 23: 238–245. ulates embryonic development. J. Neurosci., 25:
Czeh, B., Michaelis, T., Watanabe, T., Frahm, J., de Biurrun, 10074–10086.
G., van Kampen, M., Bartolomucci, A. and Fuchs, E. (2001) Faber, K.M. and Haring, J.H. (1999) Synaptogenesis in the
Stress-induced changes in cerebral metabolites, hippocampal postnatal rat fascia dentata is influenced by 5-HT1a receptor
volume, and cell proliferation are prevented by antidepres- activation. Brain Res. Dev. Brain Res., 114: 245–252.
sant treatment with tianeptine. Proc. Natl. Acad. Sci. U.S.A., Feldman, S. and Weidenfeld, J. (1993) The dorsal hippocampus
98: 12796–12801. modifies the negative feedback effect of glucocorticoids on
Czeh, B., Welt, T., Fischer, A.K., Erhardt, A., Schmitt, W., the adrenocortical and median eminence CRF-41 responses
Muller, M.B., Toschi, N., Fuchs, E. and Keck, M.E. (2002) to photic stimulation. Brain Res., 614: 227–232.
Chronic psychosocial stress and concomitant repetitive trans- Feldman, S. and Weidenfeld, J. (1999) Glucocorticoid receptor
cranial magnetic stimulation: effects on stress hormone levels antagonists in the hippocampus modify the negative feedback
and adult hippocampal neurogenesis. Biol. Psychiatry, 52: following neural stimuli. Brain Res., 821: 33–37.
1057–1065. Felitti, V.J., Anda, R.F., Nordenberg, D., Williamson, D.F.,
Drevets, W.C., Bogers, W. and Raichle, M.E. (2002) Func- Spitz, A.M., Edwards, V., Koss, M.P. and Marks, J.S. (1998)
tional anatomical correlates of antidepressant drug treatment Relationship of childhood abuse and household dysfunction
assessed using PET measures of regional glucose metabolism. to many of the leading causes of death in adults. The Adverse
Eur. Neuropsychopharmacol., 12: 527–544. Childhood Experiences (ACE) Study. Am. J. Prev. Med., 14:
Drew, M.R. and Hen, R. (in press) Adult hippocampal neuro- 245–258.
genesis as a target for the treatment of depression. CNS Ne- Filippov, V., Kronenberg, G., Pivneva, T., Reuter, K., Steiner,
urol. Disord. Drug Targets. B., Wang, L.P., Yamaguchi, M., Kettenmann, H. and Kem-
Dulawa, S.C., Holick, K.A., Gundersen, B. and Hen, R. (2004) permann, G. (2003) Subpopulation of nestin-expressing pro-
Effects of chronic fluoxetine in animal models of anxiety and genitor cells in the adult murine hippocampus shows
depression. Neuropsychopharmacology, 29: 1321–1330. electrophysiological and morphological characteristics of as-
Duman, R.S. (2002) Structural alterations in depression: cellu- trocytes. Mol. Cell Neurosci., 23: 373–382.
lar mechanisms underlying pathology and treatment of mood van der Flier, W.M., van Buchem, M.A., Weverling-Rijnsburg-
disorders. CNS Spectrosc., 7: 140–142 144–147. er, A.W., Mutsaers, E.R., Bollen, E.L., Admiraal-Behloul,
Duman, R.S. (2004) Depression: a case of neuronal life and F., Westendorp, R.G. and Middelkoop, H.A. (2004) Mem-
death? Biol. Psychiatry, 56: 140–145. ory complaints in patients with normal cognition are asso-
Duman, R.S., Heninger, G.R. and Nestler, E.J. (1997) A mo- ciated with smaller hippocampal volumes. J. Neurol., 251:
lecular and cellular theory of depression. Arch. Gen. Psychi- 671–675.
atry, 54: 597–606. Fossati, P., Coyette, F., Ergis, A.M. and Allilaire, J.F.
Duman, R.S., Malberg, J., Nakagawa, S. and D’sa, C. (2000) (2002) Influence of age and executive functioning on verbal
Neuronal plasticity and survival in mood disorders. Biol. memory of inpatients with depression. J. Affect. Disord., 68:
Psychiatry, 48: 732–739. 261–271.
718

Freund, T.F., Gulyas, A.I., Acsady, L., Gorcs, T. and Toth, K. Herman, J.P., Cullinan, W.E., Young, E.A., Akil, H. and Wat-
(1990) Serotonergic control of the hippocampus via local in- son, S.J. (1992) Selective forebrain fiber tract lesions impli-
hibitory interneurons. Proc. Natl. Acad. Sci. U.S.A., 87: cate ventral hippocampal structures in tonic regulation of
8501–8505. paraventricular nucleus corticotropin-releasing hormone
Garcia, A., Steiner, B., Kronenberg, G., Bick-Sander, A. and (CRH) and arginine vasopressin (AVP) mRNA expression.
Kempermann, G. (2004a) Age-dependent expression of Brain Res., 592: 228–238.
glucocorticoid- and mineralocorticoid receptors on neural Holick, K.A., Lee, D.C., Hen, R. and Dulawa, S.C. (2007)
precursor cell populations in the adult murine hippocampus. Behavioral effects of chronic fluoxetine in BALB/cJ mice do
Aging Cell, 3: 363–371. not require adult hippocampal neurogenesis or the serotonin
Garcia, A.D., Doan, N.B., Imura, T., Bush, T.G. and So- 1A receptor. Neuropsychopharmacology.
froniew, M.V. (2004b) GFAP-expressing progenitors are the Holsboer, F. and Barden, N. (1996) Antidepressants and hypo-
principal source of constitutive neurogenesis in adult mouse thalamic-pituitary-adrenocortical regulation. Endocr. Rev.,
forebrain. Nat. Neurosci., 7: 1233–1241. 17: 187–205.
Garriock, H.A., Delgado, P., Kling, M.A., Carpenter, L.L., Holsboer, F., Liebl, R. and Hofschuster, E. (1982) Repeated
Burke, M., Burke, W.J., Schwartz, T., Marangell, L.B., dexamethasone suppression test during depressive illness.
Husain, M., Erickson, R.P. and Moreno, F.A. (2006) Normalisation of test result compared with clinical improve-
Number of risk genotypes is a risk factor for major depres- ment. J. Affect. Disord., 4: 93–101.
sive disorder: a case control study. Behav. Brain Funct., 2: Huot, R.L., Plotsky, P.M., Lenox, R.H. and Mcnamara, R.K.
24. (2002) Neonatal maternal separation reduces hippocampal
Gilbertson, M.W., Shenton, M.E., Ciszewski, A., Kasai, K., mossy fiber density in adult Long Evans rats. Brain Res., 950:
Lasko, N.B., Orr, S.P. and Pitman, R.K. (2002) Smaller 52–63.
hippocampal volume predicts pathologic vulnerability to Huot, R.L., Thrivikraman, K.V., Meaney, M.J. and Plotsky,
psychological trauma. Nat. Neurosci., 5: 1242–1247. P.M. (2001) Development of adult ethanol preference and
Gold, P.W. and Chrousos, G.P. (2002) Organization of the anxiety as a consequence of neonatal maternal separation in
stress system and its dysregulation in melancholic and atyp- Long Evans rats and reversal with antidepressant treatment.
ical depression: high vs. low CRH/NE states. Mol. Psychi- Psychopharmacology (Berl.), 158: 366–373.
atry, 7: 254–275. Islam, A., Ayer-Lelievre, C., Heigenskold, C., Bogdanovic, N.,
Gould, E., Beylin, A., Tanapat, P., Reeves, A. and Shors, T.J. Winblad, B. and Adem, A. (1998) Changes in IGF-1 recep-
(1999a) Learning enhances adult neurogenesis in the hippo- tors in the hippocampus of adult rats after long-term adre-
campal formation. Nat. Neurosci., 2: 260–265. nalectomy: receptor autoradiography and in situ
Gould, E., Reeves, A.J., Fallah, M., Tanapat, P., Gross, C.G. hybridization histochemistry. Brain Res., 797: 342–346.
and Fuchs, E. (1999b) Hippocampal neurogenesis in adult Jacobs, B.L., Praag, H. and Gage, F.H. (2000) Adult brain
Old World primates. Proc. Natl. Acad. Sci. U.S.A., 96: neurogenesis and psychiatry: a novel theory of depression.
5263–5267. Mol. Psychiatry, 5: 262–269.
Goursaud, A.P., Mendoza, S.P. and Capitanio, J.P. (2006) Do Jayatissa, M.N., Bisgaard, C., Tingstrom, A., Papp, M. and
neonatal bilateral ibotenic acid lesions of the hippocampal Wiborg, O. (2006) Hippocampal cytogenesis correlates to
formation or of the amygdala impair HPA axis responsive- escitalopram-mediated recovery in a chronic mild stress rat
ness and regulation in infant rhesus macaques (Macaca mu- model of depression. Neuropsychopharmacology, 31:
latta)? Brain Res., 1071: 97–104. 2395–2404.
von Gunten, A., Fox, N.C., Cipolotti, L. and Ron, M.A. (2000) Jiang, W., Zhang, Y., Xiao, L., van Cleemput, J., Ji, S.P., Bai,
A volumetric study of hippocampus and amygdala in de- G. and Zhang, X. (2005) Cannabinoids promote embryonic
pressed patients with subjective memory problems. J. Ne- and adult hippocampus neurogenesis and produce anxiolytic-
uropsychiatry Clin. Neurosci., 12: 493–498. and antidepressant-like effects. J. Clin. Invest., 115:
von Gunten, A. and Ron, M.A. (2004) Hippocampal volume 3104–3116.
and subjective memory impairment in depressed patients. Joels, M. and van Riel, E. (2004) Mineralocorticoid and
Eur. Psychiatry, 19: 438–440. glucocorticoid receptor-mediated effects on serotonergic
Haydon, P.G. and Carmignoto, G. (2006) Astrocyte control of transmission in health and disease. Ann. N.Y. Acad. Sci.,
synaptic transmission and neurovascular coupling. Physiol. 1032: 301–303.
Rev., 86: 1009–1031. Kempermann, G., Kuhn, H.G. and Gage, F.H. (1997) More
Hellsten, J., Wennstrom, M., Mohapel, P., Ekdahl, C.T., Ben- hippocampal neurons in adult mice living in an enriched en-
gzon, J. and Tingstrom, A. (2002) Electroconvulsive seizures vironment. Nature, 386: 493–495.
increase hippocampal neurogenesis after chronic corticoster- Kempermann, G., Kuhn, H.G. and Gage, F.H. (1998) Expe-
one treatment. Eur. J. Neurosci., 16: 283–290. rience-induced neurogenesis in the senescent dentate gyrus. J.
Herman, J.P., Cullinan, W.E., Morano, M.I., Akil, H. and Neurosci., 18: 3206–3212.
Watson, S.J. (1995) Contribution of the ventral subiculum to Kendler, K.S., Kuhn, J.W., Vittum, J., Prescott, C.A. and
inhibitory regulation of the hypothalamo-pituitary-adreno- Riley, B. (2005) The interaction of stressful life events and a
cortical axis. J. Neuroendocrinol., 7: 475–482. serotonin transporter polymorphism in the prediction of
719

episodes of major depression: a replication. Arch. Gen. Psy- Liu, S., Wang, J., Zhu, D., Fu, Y., Lukowiak, K. and Lu, Y.M.
chiatry, 62: 529–535. (2003) Generation of functional inhibitory neurons in the
Kendler, K.S., Thornton, L.M. and Gardner, C.O. (2001) Ge- adult rat hippocampus. J. Neurosci., 23: 732–736.
netic risk, number of previous depressive episodes, and Loo, C.K. and Mitchell, P.B. (2005) A review of the efficacy of
stressful life events in predicting onset of major depression. transcranial magnetic stimulation (TMS) treatment for de-
Am. J. Psychiatry, 158: 582–586. pression, and current and future strategies to optimize effi-
Kennedy, S.H., Evans, K.R., Kruger, S., Mayberg, H.S., cacy. J. Affect. Disord., 88: 255–267.
Meyer, J.H., Mccann, S., Arifuzzman, A.I., Houle, S. and Lovic, V., Gonzalez, A. and Fleming, A.S. (2001) Maternally
Vaccarino, F.J. (2001) Changes in regional brain glucose separated rats show deficits in maternal care in adulthood.
metabolism measured with positron emission tomography Dev. Psychobiol., 39: 19–33.
after paroxetine treatment of major depression. Am. J. Psy- Lucassen, P.J., Muller, M.B., Holsboer, F., Bauer, J., Holtrop,
chiatry, 158: 899–905. A., Wouda, J., Hoogendijk, W.J., De Kloet, E.R. and Swaab,
Kimbrell, T.A., Ketter, T.A., George, M.S., Little, J.T., Ben- D.F. (2001) Hippocampal apoptosis in major depression is a
son, B.E., Willis, M.W., Herscovitch, P. and Post, R.M. minor event and absent from subareas at risk for glucocorti-
(2002) Regional cerebral glucose utilization in patients with a coid overexposure. Am. J. Pathol., 158: 453–468.
range of severities of unipolar depression. Biol. Psychiatry, Ma, D.K., Ming, G.L. and Song, H. (2005) Glial influences on
51: 237–252. neural stem cell development: cellular niches for adult ne-
Knigge, K.M. (1961) Adrenocortical response to stress in rats urogenesis. Curr. Opin. Neurobiol., 15: 514–520.
with lesions in hippocampus and amygdala. Proc. Soc. Exp. Macqueen, G.M., Campbell, S., Mcewen, B.S., Macdonald, K.,
Biol. Med., 108: 18–21. Amano, S., Joffe, R.T., Nahmias, C. and Young, L.T. (2003)
Kohler, S.J., Jennings, V., Boklewski, J.L., Degrush, B.J., Course of illness, hippocampal function, and hippocampal
Williams, N.I., Rockcastle, N.J., Cameron, J.L. and Green- volume in major depression. Proc. Natl. Acad. Sci. U.S.A.,
ough, W.T. (2006) Maturation of new granule cells in the 100: 1387–1392.
dentate gyrus of adult macaque monkeys. Soc. Neurosci. Madsen, T.M., Treschow, A., Bengzon, J., Bolwig, T.G.,
Abstr. Lindvall, O. and Tingstrom, A. (2000) Increased neurogen-
Kondoh, M., Shiga, T. and Okado, N. (2004) Regulation of esis in a model of electroconvulsive therapy. Biol. Psychiatry,
dendrite formation of Purkinje cells by serotonin through 47: 1043–1049.
serotonin1A and serotonin2A receptors in culture. Neurosci. Malberg, J.E. and Duman, R.S. (2003) Cell proliferation in
Res., 48: 101–109. adult hippocampus is decreased by inescapable stress: re-
Kornack, D.R. and Rakic, P. (1999) Continuation of neuro- versal by fluoxetine treatment. Neuropsychopharmacology,
genesis in the hippocampus of the adult macaque monkey. 28: 1562–1571.
Proc. Natl. Acad. Sci. U.S.A., 96: 5768–5773. Malberg, J.E., Eisch, A.J., Nestler, E.J. and Duman, R.S.
Kuhn, H.G., Dickinson-Anson, H. and Gage, F.H. (1996) Ne- (2000) Chronic antidepressant treatment increases neuro-
urogenesis in the dentate gyrus of the adult rat: age-related genesis in adult rat hippocampus. J. Neurosci., 20:
decrease of neuronal progenitor proliferation. J. Neurosci., 9104–9110.
16: 2027–2033. Malberg, J.E. and Schechter, L.E. (2005) Increasing hippocam-
Kuhn, H.G., Winkler, J., Kempermann, G., Thal, L.J. and pal neurogenesis: a novel mechanism for antidepressant
Gage, F.H. (1997) Epidermal growth factor and fibroblast drugs. Curr. Pharm. Des., 11: 145–155.
growth factor-2 have different effects on neural progenitors Manev, H., Uz, T., Smalheiser, N.R. and Manev, R. (2001)
in the adult rat brain. J. Neurosci., 17: 5820–5829. Antidepressants alter cell proliferation in the adult brain in
Kvachnina, E., Liu, G., Dityatev, A., Renner, U., Dumuis, A., vivo and in neural cultures in vitro. Eur. J. Pharmacol., 411:
Richter, D.W., Dityateva, G., Schachner, M., Voyno- 67–70.
Yasenetskaya, T.A. and Ponimaskin, E.G. (2005) 5-HT7 re- Matus-Amat, P., Higgins, E.A., Barrientos, R.M. and Rudy,
ceptor is coupled to G alpha subunits of heterotrimeric J.W. (2004) The role of the dorsal hippocampus in the ac-
G12-protein to regulate gene transcription and neuronal quisition and retrieval of context memory representations. J.
morphology. J. Neurosci., 25: 7821–7830. Neurosci., 24: 2431–2439.
Laplagne, D.A., Esposito, M.S., Piatti, V.C., Morgenstern, Mayberg, H.S., Brannan, S.K., Tekell, J.L., Silva, J.A., Mah-
N.A., Zhao, C., van Praag, H., Gage, F.H. and Schinder, urin, R.K., Mcginnis, S. and Jerabek, P.A. (2000) Regional
A.F. (2006) Functional convergence of neurons generated metabolic effects of fluoxetine in major depression: serial
in the developing and adult hippocampus. PLoS Biol., 4: changes and relationship to clinical response. Biol. Psychia-
e409. try, 48: 830–843.
Lauder, J.M. and Krebs, H. (1978) Serotonin as a differenti- Mcewen, B.S. (2005) Glucocorticoids, depression, and mood
ation signal in early neurogenesis. Dev. Neurosci., 1: 15–30. disorders: structural remodeling in the brain. Metabolism, 54:
Leonardo, E.D. and Hen, R. (2006) Genetics of affective and 20–23.
anxiety disorders. Annu. Rev. Psychol., 57: 117–137. Meltzer, L.A., Roy, M., Parente, V. and Deisseroth, K. (2006)
Levinson, D.F. (2006) The genetics of depression: a review. Hippocampal neurogenesis is required for lasting antidepres-
Biol. Psychiatry, 60: 84–92. sant effects of fluoxetine. Soc. Neurosci. Abstr.
720

Meltzer, L.A., Yabaluri, R. and Deisseroth, K. (2005) A role Oleskevich, S., Descarries, L., Watkins, K.C., Seguela, P. and
for circuit homeostasis in adult neurogenesis. Trends Neuro- Daszuta, A. (1991) Ultrastructural features of the serotonin
sci., 28: 653–660. innervation in adult rat hippocampus: an immunocytochem-
Meshi, D., Drew, M.R., Saxe, M., Ansorge, M.S., David, D., ical description in single and serial thin sections. Neurosci-
Santarelli, L., Malapani, C., Moore, H. and Hen, R. (2006) ence, 42: 777–791.
Hippocampal neurogenesis is not required for behavioral Overmier, J.B. and Seligman, M.E. (1967) Effects of inescap-
effects of environmental enrichment. Nat. Neurosci., 9: able shock upon subsequent escape and avoidance respond-
729–731. ing. J. Comp. Physiol. Psychol., 63: 28–33.
Mineur, Y.S., Belzung, C. and Crusio, W.E. (2006) Effects of Overstreet-Wadiche, L.S. and Westbrook, G.L. (2006) Func-
unpredictable chronic mild stress on anxiety and depression- tional maturation of adult-generated granule cells. Hippo-
like behavior in mice. Behav. Brain Res., 175: 43–50. campus, 16: 208–215.
Mineur, Y.S., Prasol, D.J., Belzung, C. and Crusio, W.E. (2003) Patel, T.D. and Zhou, F.C. (2005) Ontogeny of 5-HT1A re-
Agonistic behavior and unpredictable chronic mild stress in ceptor expression in the developing hippocampus. Brain Res.
mice. Behav. Genet., 33: 513–519. Dev Brain Res., 157: 42–57.
Mirescu, C., Peters, J.D. and Gould, E. (2004) Early life ex- Phillips, R.G. and LeDoux, J.E. (1992) Differential contribu-
perience alters response of adult neurogenesis to stress. Nat. tion of amygdala and hippocampus to cued and contextual
Neurosci., 7: 841–846. fear conditioning. Behav. Neurosci., 106: 274–285.
Monje, M.L., Mizumatsu, S., Fike, J.R. and Palmer, T.D. van Praag, H., Kempermann, G. and Gage, F.H. (1999) Run-
(2002) Irradiation induces neural precursor-cell dysfunction. ning increases cell proliferation and neurogenesis in the adult
Nat. Med., 8: 955–962. mouse dentate gyrus. Nat. Neurosci., 2: 266–270.
Monje, M.L., Toda, H. and Palmer, T.D. (2003) Inflammatory Radley, J.J. and Jacobs, B.L. (2002) 5-HT1A receptor antag-
blockade restores adult hippocampal neurogenesis. Science, onist administration decreases cell proliferation in the dent-
302: 1760–1765. ate gyrus. Brain Res., 955: 264–267.
Moser, M.B. and Moser, E.I. (1998) Functional differentiation Ramirez-Amaya, V., Marrone, D.F., Gage, F.H., Worley, P.F.
in the hippocampus. Hippocampus, 8: 608–619. and Barnes, C.A. (2006) Integration of new neurons into
Motl, R.W., Birnbaum, A.S., Kubik, M.Y. and Dishman, R.K. functional neural networks. J. Neurosci., 26: 12237–12241.
(2004) Naturally occurring changes in physical activity are Rao, M.S. and Shetty, A.K. (2004) Efficacy of doublecortin as a
inversely related to depressive symptoms during early ado- marker to analyse the absolute number and dendritic growth
lescence. Psychosom. Med., 66: 336–342. of newly generated neurons in the adult dentate gyrus. Eur. J.
Muller, M.B., Lucassen, P.J., Yassouridis, A., Hoogendijk, Neurosci., 19: 234–246.
W.J., Holsboer, F. and Swaab, D.F. (2001) Neither major Reif, A., Fritzen, S., Finger, M., Strobel, A., Lauer, M., Sch-
depression nor glucocorticoid treatment affects the cellular mitt, A. and Lesch, K.P. (2006) Neural stem cell proliferation
integrity of the human hippocampus. Eur. J. Neurosci., 14: is decreased in schizophrenia, but not in depression. Mol.
1603–1612. Psychiatry, 11: 514–522.
Murray, C.J. and Lopez, A.D. (1996) Evidence-based health Rhodes, J.S., van Praag, H., Jeffrey, S., Girard, I., Mitchell,
policy — lessons from the Global Burden of Disease Study. G.S., Garland Jr., T. and Gage, F.H. (2003) Exercise in-
Science, 274: 740–743. creases hippocampal neurogenesis to high levels but does not
Nakagawa, S., Kim, J.E., Lee, R., Chen, J., Fujioka, T., Ma- improve spatial learning in mice bred for increased voluntary
lberg, J., Tsuji, S. and Duman, R.S. (2002a) Localization of wheel running. Behav. Neurosci., 117: 1006–1016.
phosphorylated cAMP response element-binding protein in Sairanen, M., Lucas, G., Ernfors, P., Castren, M. and Castren,
immature neurons of adult hippocampus. J. Neurosci., 22: E. (2005) Brain-derived neurotrophic factor and antidepres-
9868–9876. sant drugs have different but coordinated effects on neuronal
Nakagawa, S., Kim, J.E., Lee, R., Malberg, J.E., Chen, J., turnover, proliferation, and survival in the adult dentate
Steffen, C., Zhang, Y.J., Nestler, E.J. and Duman, R.S. gyrus. J. Neurosci., 25: 1089–1094.
(2002b) Regulation of neurogenesis in adult mouse hippo- Santarelli, L., Saxe, M., Gross, C., Surget, A., Battaglia, F.,
campus by cAMP and the cAMP response element-binding Dulawa, S., Weisstaub, N., Lee, J., Duman, R., Arancio, O.,
protein. J. Neurosci., 22: 3673–3682. Belzung, C. and Hen, R. (2003) Requirement of hippocampal
Nestler, E.J., Barrot, M., Dileone, R.J., Eisch, A.J., Gold, S.J. neurogenesis for the behavioral effects of antidepressants.
and Monteggia, L.M. (2002) Neurobiology of depression. Science, 301: 805–809.
Neuron, 34: 13–25. Sapolsky, R.M. (1994) The physiological relevance of gluco-
Neumeister, A., Wood, S., Bonne, O., Nugent, A.C., Luckenb- corticoid endangerment of the hippocampus. Ann. N.Y.
augh, D.A., Young, T., Bain, E.E., Charney, D.S. and Dre- Acad. Sci., 746: 294–304 discussion 304–307.
vets, W.C. (2005) Reduced hippocampal volume in Sapolsky, R.M. (2000) Glucocorticoids and hippocampal atro-
unmedicated, remitted patients with major depression versus phy in neuropsychiatric disorders. Arch. Gen. Psychiatry, 57:
control subjects. Biol. Psychiatry, 57: 935–937. 925–935.
Nottebohm, F. (2002) Neuronal replacement in adult brain. Sapolsky, R.M., Armanini, M.P., Sutton, S.W. and Plotsky,
Brain Res. Bull., 57: 737–749. P.M. (1989) Elevation of hypophysial portal concentrations
721

of adrenocorticotropin secretagogues after fornix transec- Silva, R., Lu, J., Wu, Y., Martins, L., Almeida, O.F. and Sousa,
tion. Endocrinology, 125: 2881–2887. N. (2006) Mapping cellular gains and losses in the postnatal
Saxe, M.D., Malleret, G., Vronskaya, S., Mendez, I., Garcia, dentate gyrus: implications for psychiatric disorders. Exp.
A.D., Sofroniew, M.V., Kandel, E.R. and Hen, R. (2007) Neurol., 200: 321–331.
Paradoxical influence of hippocampal neurogenesis on work- Singh, N.A., Clements, K.M. and Singh, M.A. (2001) The effi-
ing memory. Proc. Natl. Acad. Sci. U.S.A., 104: 4642–4646. cacy of exercise as a long-term antidepressant in elderly sub-
Saxe, M.D., Battaglia, F., Wang, J.W., Malleret, G., David, jects: a randomized, controlled trial. J. Gerontol. A Biol. Sci.
D.J., Monckton, J.E., Garcia, A.D., Sofroniew, M.V., Kan- Med. Sci., 56: M497–M504.
del, E.R., Santarelli, L., Hen, R. and Drew, M.R. (2006) Snyder, J.S., Hong, N.S., Mcdonald, R.J. and Wojtowicz, J.M.
Ablation of hippocampal neurogenesis impairs contextual (2005) A role for adult neurogenesis in spatial long-term
fear conditioning and synaptic plasticity in the dentate gyrus. memory. Neuroscience, 130: 843–852.
Proc. Natl. Acad. Sci. U.S.A., 103: 17501–17506. Snyder, J.S., Kee, N. and Wojtowicz, J.M. (2001) Effects of
Saxena, S., Brody, A.L., Ho, M.L., Alborzian, S., Ho, M.K., adult neurogenesis on synaptic plasticity in the rat dentate
Maidment, K.M., Huang, S.C., Wu, H.M., Au, S.C. and gyrus. J. Neurophysiol., 85: 2423–2431.
Baxter Jr., L.R. (2001) Cerebral metabolism in major de- Song, H., Kempermann, G., Overstreet Wadiche, L., Zhao, C.,
pression and obsessive-compulsive disorder occurring sepa- Schinder, A.F. and Bischofberger, J. (2005) New neurons in
rately and concurrently. Biol. Psychiatry, 50: 159–170. the adult mammalian brain: synaptogenesis and functional
Schaaf, M.J., de Kloet, E.R. and Vreugdenhil, E. (2000) integration. J. Neurosci., 25: 10366–10368.
Corticosterone effects on BDNF expression in the hippo- Stockmeier, C.A., Mahajan, G.J., Konick, L.C., Overholser,
campus. Implications for memory formation. Stress, 3: J.C., Jurjus, G.J., Meltzer, H.Y., Uylings, H.B., Friedman, L.
201–208. and Rajkowska, G. (2004) Cellular changes in the postmor-
Scharfman, H.E., Sollas, A.L., Smith, K.L., Jackson, M.B. and tem hippocampus in major depression. Biol. Psychiatry, 56:
Goodman, J.H. (2002) Structural and functional asymmetry 640–650.
in the normal and epileptic rat dentate gyrus. J. Comp. Strange, B. and Dolan, R. (1999) Functional segregation within
Neurol., 454: 424–439. the human hippocampus. Mol. Psychiatry, 4: 508–511.
Schildkraut, J.J. (1965) The catecholamine hypothesis of affec- Strange, B.A., Fletcher, P.C., Henson, R.N., Friston, K.J. and
tive disorders: a review of supporting evidence. Am. J. Psy- Dolan, R.J. (1999) Segregating the functions of human hip-
chiatry, 122: 509–522. pocampus. Proc. Natl. Acad. Sci. U.S.A., 96: 4034–4039.
Schmidt-Hieber, C., Jonas, P. and Bischofberger, J. (2004) Strekalova, T., Spanagel, R., Bartsch, D., Henn, F.A. and
Enhanced synaptic plasticity in newly generated granule cells Gass, P. (2004) Stress-induced anhedonia in mice is associ-
of the adult hippocampus. Nature, 429: 184–187. ated with deficits in forced swimming and exploration. Ne-
Seki, T. and Arai, Y. (1995) Age-related production of new uropsychopharmacology, 29: 2007–2017.
granule cells in the adult dentate gyrus. Neuroreport, 6: Sullivan, P.F., Neale, M.C. and Kendler, K.S. (2000) Genetic
2479–2482. epidemiology of major depression: review and meta-analysis.
Seligman, M.E. and Beagley, G. (1975) Learned helplessness in Am. J. Psychiatry, 157: 1552–1562.
the rat. J. Comp. Physiol. Psychol., 88: 534–541. Tanapat, P., Hastings, N.B., Reeves, A.J. and Gould, E. (1999)
Seminowicz, D.A., Mayberg, H.S., Mcintosh, A.R., Goldapple, Estrogen stimulates a transient increase in the number of new
K., Kennedy, S., Segal, Z. and Rafi-Tari, S. (2004) Limbic- neurons in the dentate gyrus of the adult female rat. J. Ne-
frontal circuitry in major depression: a path modeling met- urosci., 19: 5792–5801.
analysis. Neuroimage, 22: 409–418. Taylor, M.J., Freemantle, N., Geddes, J.R. and Bhagwagar, Z.
Sheline, Y.I. (1996) Hippocampal atrophy in major depression: (2006) Early onset of selective serotonin reuptake inhibitor
a result of depression-induced neurotoxicity? Mol. Psychia- antidepressant action: systematic review and meta-analysis.
try, 1: 298–299. Arch. Gen. Psychiatry, 63: 1217–1223.
Sheline, Y.I., Gado, M.H. and Kraemer, H.C. (2003) Untreated Tozuka, Y., Fukuda, S., Namba, T., Seki, T. and Hisatsune, T.
depression and hippocampal volume loss. Am. J. Psychiatry, (2005) GABAergic excitation promotes neuronal differenti-
160: 1516–1518. ation in adult hippocampal progenitor cells. Neuron, 47:
Sheline, Y.I., Sanghavi, M., Mintun, M.A. and Gado, M.H. 803–815.
(1999) Depression duration but not age predicts hippocampal Vakili, K., Pillay, S.S., Lafer, B., Fava, M., Renshaw, P.F.,
volume loss in medically healthy women with recurrent major Bonello-Cintron, C.M. and Yurgelun-Todd, D.A. (2000)
depression. J. Neurosci., 19: 5034–5043. Hippocampal volume in primary unipolar major depression:
Shirayama, Y., Chen, A.C., Nakagawa, S., Russell, D.S. and a magnetic resonance imaging study. Biol. Psychiatry, 47:
Duman, R.S. (2002) Brain-derived neurotrophic factor pro- 1087–1090.
duces antidepressant effects in behavioral models of depres- Videbech, P. and Ravnkilde, B. (2004) Hippocampal volume
sion. J. Neurosci., 22: 3251–3261. and depression: a meta-analysis of MRI studies. Am. J. Psy-
Shors, T.J., Miesegaes, G., Beylin, A., Zhao, M., Rydel, T. and chiatry, 161: 1957–1966.
Gould, E. (2001) Neurogenesis in the adult is involved in the Videbech, P., Ravnkilde, B., Pedersen, A.R., Egander, A.,
formation of trace memories. Nature, 410: 372–376. Landbo, B., Rasmussen, N.A., Andersen, F., Stodkilde-
722

Jorgensen, H., Gjedde, A. and Rosenberg, R. (2001) The Winocur, G., Wojtowicz, J.M., Sekeres, M., Snyder, J.S. and
Danish PET/depression project: PET findings in patients Wang, S. (2006) Inhibition of neurogenesis interferes with
with major depression. Psychol. Med., 31: 1147–1158. hippocampus-dependent memory function. Hippocampus,
Videbech, P., Ravnkilde, B., Pedersen, T.H., Hartvig, H., 16: 296–304.
Egander, A., Clemmensen, K., Rasmussen, N.A., Andersen, Wiskott, L., Rasch, M.J. and Kempermann, G. (2006) A func-
F., Gjedde, A. and Rosenberg, R. (2002) The Danish PET/ tional hypothesis for adult hippocampal neurogenesis: avoid-
depression project: clinical symptoms and cerebral blood ance of catastrophic interference in the dentate gyrus.
flow. A regions-of-interest analysis. Acta Psychiatr. Scand., Hippocampus, 16: 329–343.
106: 35–44. Wojtowicz, J.M. (2006) Irradiation as an experimental tool in
Vollmayr, B., Simonis, C., Weber, S., Gass, P. and Henn, F. studies of adult neurogenesis. Hippocampus, 16: 261–266.
(2003) Reduced cell proliferation in the dentate gyrus is not Wong, E.Y. and Herbert, J. (2004) The corticoid environment:
correlated with the development of learned helplessness. Biol. a determining factor for neural progenitors’ survival in the
Psychiatry, 54: 1035–1040. adult hippocampus. Eur. J. Neurosci., 20: 2491–2498.
Wang, S., Scott, B.W. and Wojtowicz, J.M. (2000) Heteroge- Wong, E.Y. and Herbert, J. (2006) Raised circulating corticos-
nous properties of dentate granule neurons in the adult rat. J. terone inhibits neuronal differentiation of progenitor cells in
Neurobiol., 42: 248–257. the adult hippocampus. Neuroscience, 137: 83–92.
West, M.J. (1993) Regionally specific loss of neurons in the Yalcin, I., Aksu, F. and Belzung, C. (2005) Effects of desipra-
aging human hippocampus. Neurobiol. Aging, 14: 287–293. mine and tramadol in a chronic mild stress model in mice are
Willner, P. (2005) Chronic mild stress (CMS) revisited: consist- altered by yohimbine but not by pindolol. Eur. J. Pharma-
ency and behavioural-neurobiological concordance in the col., 514: 165–174.
effects of CMS. Neuropsychobiology, 52: 90–110. Yan, W., Wilson, C.C. and Haring, J.H. (1997) Effects of
Willner, P. and Mitchell, P.J. (2002) The validity of animal neonatal serotonin depletion on the development of rat
models of predisposition to depression. Behav. Pharmacol., dentate granule cells. Brain Res. Dev. Brain Res., 98:
13: 169–188. 177–184.
Willner, P., Muscat, R. and Papp, M. (1992) Chronic mild Yuen, E.Y., Jiang, Q., Chen, P., Gu, Z., Feng, J. and Yan, Z.
stress-induced anhedonia: a realistic animal model of depres- (2005a) Serotonin 5-HT1A receptors regulate NMDA recep-
sion. Neurosci. Biobehav. Rev., 16: 525–534. tor channels through a microtubule-dependent mechanism.
Willner, P., Towell, A., Sampson, D., Sophokleous, S. and J. Neurosci., 25: 5488–5501.
Muscat, R. (1987) Reduction of sucrose preference by chronic Yuen, E.Y., Jiang, Q., Feng, J. and Yan, Z. (2005b) Microtu-
unpredictable mild stress, and its restoration by a tricyclic bule regulation of N-methyl-D-aspartate receptor channels in
antidepressant. Psychopharmacology (Berl.), 93: 358–364. neurons. J. Biol. Chem., 280: 29420–29427.
Plate 38.1. A schematic of the dentate gyrus granule cell layer (GCL) illustrating the different ways by which neurogenesis can
influence its structure and function. Boxed panel reveals a cross section of the dentate GCL with the different populations that reside
within it: mature granule cells born during development (light blue), adult-generated mature granule cells (dark blue), adult-born
immature neurons (red) and interneurons (green). Over the lifespan, the GCL may increase in size due to a net addition of new neurons
(A) or may remain unchanged due to a net replacement of developmentally generated granule cells (B). Changes in neurogenesis can
result in increased representation of interneurons (C), a larger pool of adult generated immature neurons (D) or the generation of
mature neurons with distinct physiological and biochemical properties (E). Conceivably, neurogenesis may be altered in any one of
these ways in MDD. Conversely, AD drugs may influence DG function in more than one way to exert their behavioral effects. (For
B/W version, see page 700 in the volume.)
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 39

The dentate gyrus in Alzheimer’s disease

Thomas G. Ohm

Institute of Integrative Neuroanatomy, Department of Clinical Cell and Neurobiology, Charité CCM, 10098 Berlin,
Germany

Abstract: As part of the hippocampus, the dentate gyrus is considered to play a crucial role in associative
memory. The reviewed data suggest that the dentate gyrus withstands the formation of plaques, tangles and
neuronal death until late stages of Alzheimer’s disease (AD). However, changes related to a disconnecting
process, and more subtle intrinsic alterations, may contribute to disturbances in memory and learning
observed in early stages of AD.

Keywords: hippocampus; granule cells; interneuron; Alzheimer’s disease; dentate gyrus; review

Basic anatomy originating from 650,000 to 1,100,000 (Gomez Isla


et al., 1996; West and Slomianka, 1998; von
The human dentate gyrus belongs to the heteroge- Gunten et al., 2006) excitatory entorhinal projec-
neously composed allocortex, which is located tion neurons located in the pre-a layer or — in a
mainly in the antero-medial temporal lobe. As part different nomenclature — layer II. Pre-a belongs to
of the hippocampus, the dentate gyrus is considered the lamina principalis externa, the superimposed
to play a crucial role in associative memory, concept for the superficial layers, which is sepa-
especially with respect to events (‘what happens’) rated by the stratum dissecans from the deeper
(Morris, 2006). To perform this task, the dentate layers, together called lamina principalis interna.
gyrus is highly organized and numerous synaptic Perforant path axons end preferentially within the
interactions take place involving at least 11 types of outer two-thirds of the superficial molecular layer
interneurons, which form an intrinsic network mainly on the (apical) dendrites of the granule
(Freund and Buzsaki, 1996). In cognitively normal cells, but also on dendrites of interneurons such as
humans, 9–22 million (mean: 14.6) (Seress, 1988; the basket cells. Origins of other important affer-
West and Gundersen, 1990; West et al., 1994; ents are cholinergic neurons of the basal forebrain
Bobinski et al., 1996a, 1997; Simic et al., 1997; mainly located in the medial septal and diagonal
Harding et al., 1998; Korbo et al., 2004) densely band nucleus and association fibres from the mossy
packed principal (projection) neurons, the granule cells of sector CA4. These afferents are also at-
cells, form the narrow granule cell layer. tracted to the inner molecular layer where they
The major input of the dentate gyrus comes terminate together with axons stemming from the
through the perforant path representing axons reunion nucleus of the thalamus. In humans, the
relative proportion of the respective input is not
Corresponding author. Tel.: +49 (0)30 450 528202; exactly known. Furthermore, the wide-spreading
Fax: +49 (0)30 450 528913; E-mail: thomas_georg.ohm@ projection from noradrenergic and serotoninergic
charite.de nuclei of the brain stem (nucleus coeruleus and

DOI: 10.1016/S0079-6123(07)63039-8 723


724

oral raphe nuclei), along with some hypothalamic the mossy cell synapses with the dendrite of
(supra- and/or tuberomamillary) projections form a the basket cell, a pyramid-shaped GABAergic,
profuse input to the molecular layer of the dentate parvalbumin-containing interneuron. Their short
gyrus. Neurons from the pre-b layer (also termed axons profusely give rise of collaterals in the gran-
layer III) uses the perforant path for a direct pro- ule cell layer where these branches form inhibitory
jection to the sector CA1 of the Ammon’s horn axosomatic synapses with the granule cells.
(probably encoding information used for spatial Common to both projection neurons and inter-
memory (‘where is it happening’) (Morris, 2006). neurons is their richness in calcium-binding
Here the NMDA receptor-mediated synaptic plas- proteins. Granule cells contain calbindin and
ticity integrates the spatial- and event-related asso- calmyrin, whereas calbindin, parvalbumin or
ciative memory components. Several findings suggest calretinin is found in interneurons.
that the CA3 pyramidal cells, which receive the gran-
ule cell output, aid at ‘pattern completion’ if recall of
spatial information is challenged under situations of Alzheimer’s disease-related histopathology
incomplete data about the environment (Nakazawa
et al., 2002). A hippocampal commissural system The formation and deposition of certain aggregated
(psalterium) is virtually absent (Amaral et al., 1984; proteins is most striking among the several histo-
Demeter et al., 1985) and only a scanty callosal input logical changes tied up to Alzheimer’s disease (AD).
arrives from the contralateral hippocampus. This is in Normally, aggregation of these proteins does not
contrast to a rodent brain and the classical descrip- occur, and even more surprisingly, this material is
tions made by, e.g., Kölliker or de Nó. only rarely found in animals. Two proteins seem to
The granule cells convey the processed informa- be central: tau and Ab.
tion via the mossy fibers, i.e., the granule cells’ Tau is a microtubule associated, primarily
axons, to the cornu ammonis (CA). Here they ter- neuronal-axonal protein involved in microtubule
minate with unusually large and zinc-rich synaptic assembly and stabilisation, processes controlled
boutons on proximal dendrites of pyramidal cells by the degree of tau’s phosphorylation (normal
of the sector CA3 of the hippocampus, thereby 2–3 mol phosphate per mol tau) and O-
forming the stratum lucidum. Mossy fiber collat- GlcNAcetylation. In humans, alternative splicing
erals are abound, terminating on inhibitory local of a single gene generates six isoforms, all of which
circuit neurons of the dentate gyrus’s polymorphic are normally among the most soluble proteins
layer and on the glutamatergic mossy cells known so far, and all of which are involved in AD
( ¼ modified hilar/CA4 pyramidal cells with osten- (Spillantini and Goedert, 1998). After hyper-
tatious dendritic excrescences). In monkey, mossy phosphorylation of some of the 30 potential
fiber collaterals terminate on basal dendrites of phosphorylation sites tau tends to self-polymerise.
granule cells, which may be the morphological ba- Moderate hyperphosphorylation (4–6 mol phos-
sis for a recurrent excitatory feedback loop (Seress phate per mol tau) results in sequestration of nor-
and Frotscher, 1990). Others, however, when per- mal tau whereas higher phosphorylation (10 mol
forming biocytin-tracing on human hippocampal phosphate per mol tau) is associated with filament
tissue obtained from surgery, did not find evidence formation. In AD tau is hyperphosphorylated and
for this (Lim et al., 1997). forms highly characteristic paired helical filaments
The local interneuron’s axon terminates with (PHF; interval between two twists 80 nm, closest
inhibitory synapses on the dendrites of granule width 10 nm and broadest width 20 nm) and — to
cells or on other interneurons. The granule cells a lesser amount — also straight filaments. These
mostly direct their dendrites into the molecular filaments represent the protein backbone of in-
layer (apical dendrites) but in one-third (Lim et al., traneuronal histopathological features, i.e., tangles
1997), — much more abundant than in rodents — and neuropil threads (AD-related ‘tau-pathology’).
also into the polymorphic or plexiform layer (basal These structures can be stained selectively with
dendrites) (Seress and Mrzljak, 1987). The axon of highly sensitive silver stains such as the Gallyas
725

stain that was proven to be as sensitive and selec- the death of more than 90% of the neurons
tive as immunostains (Schönheit et al., 2004). (Gomez Isla et al., 1996). Ghost tangles are
Immunostains, however, allow the determination numerous and may even be cleared in considera-
of the hierarchically occurring phosphorylation ble numbers from the neuropil by astrocytes. This
and thereby the staging of tangles (Uchihara et al., may feign a non-tangle related neuron loss (Kril
2001; Augustinack et al., 2002) unveiling that et al., 2002). In terms of time, the development of
changes in tau phosphorylation precede tangle these six stages was estimated to take almost five
formation (Bancher et al., 1989; Braak et al., 1994). decades (Ohm et al., 1995). Although many neu-
Tangles occur in somata of neurons, which may rons of the cortex may have developed a tangle in
live for years (Bobinski et al., 1998) or even decades these late stages, it is to note that only few of the
with a tangle (Morsch et al., 1999). When the many nerve cell types are vulnerable (Braak and
tangle-bearing neuron has deceased, inert remnants Del Tredici, 2004).
remain for quite some time, termed ‘ghost tangles’. Ab is physiologically formed by proteolytic
Eventually, they become cleared by astrocytes. processing of a receptor-like transmembrane pro-
Ghost tangles are usually less-densely twisted, less tein (APP, amyloid precursor protein) (Dyrks
argyrophilic, affine for acid dyes and Congo red, et al., 1988) through a sequential action of so-
ubiquitinated and even immunoreactive against called beta- and gamma-secretases (Haass, 2004).
anti-Ab-antibodies. Neuropil threads are likewise APP is member of a larger superfamily including
silver-stainable filaments but primarily located several splice-variants of APP and APP-like pro-
within the dendrites of neurons, often before teins (APPLP1 and 2) and which may share com-
the respective soma has formed a tangle (Braak mon functions (Bush et al., 1994). Apparently the
et al., 1986). APP, APPLP1 and APPLP2 showed a similar
The Alzheimer-related tau pathology shows a pattern of expression in the hippocampus with
regular, highly area-specific, lamina-specific and both mRNA and protein in granule cells and
cell type-specific spread through the cortex allow- pyramidal cells of all sectors of the Ammon’s horn
ing their classification into six stages. Neither neu- (McNamara et al., 1998). Ab is extracellularly
ritic plaques nor any other plaque subtype shows a deposited in a variety of states and shapes, collec-
similar degree of reliability in the distribution and tively termed ‘plaques’. When Ab adopts a b-sheet
spread. The Braak-staging [for more details see: state, it can be stained as an amyloid-like material
(Braak and Braak, 1991)], developed on Gallyas- (‘b-amyloid-plaque’) with classical amyloid stains
stained tissue, distinguishes two ‘entorhinal’ (stages such as Congo red or thioflavine. In many cases, a
I and II), two ‘limbic’ (stages III and IV) and two core of b-amyloid is surrounded by a surface of
‘isocortical stages’ (stages V and VI). The stages dystrophic neurites and swollen glial processes,
I–II do not show clinical or insipient signs of AD and forms by this the so-called ‘neuritic plaque’.
(Bancher et al., 1993; Braak et al., 1993). The stage b-pleated Ab, however, represents only a limited
III displays a severe involvement of layers pre-a of amount of all extracellularly deposited Ab-peptide.
the transentorhinal and entorhinal cortices includ- This other Ab forms the ‘diffuse’ or ‘pre-amyloid’
ing first occurrence of ghost tangles as sign of be- plaques and stains either with immunocyto-
ginning neuronal loss and deafferentiation of the chemical tools or a few specific silver stains. The
dentate gyrus. Stage IV is marked off by an abun- often-used term ‘senile plaque’ refers frequently to
dance of ghost tangles in layer pre-a, severer tangle an undefined melange of Ab-states and shapes. The
formation in layer pri-a and now commencing tan- distribution pattern of diffuse and amyloid plaques
gle formation in layer pre-b. Stages V and VI meet is not congruent. Ab deposits in phases that are
the classical neuropathological criteria of AD characterized by a hierarchical neuroanatomical
(Khatchaturian, 1985; Jellinger, 1997; Jellinger pattern (Thal et al., 2002). The relationship
and Bancher., 1997) and generally are associated between plaques and tangles is a matter of debate
with dementia. At stage VI the entorhinal cortex since long. For many years the so-called amyloid
may have an almost denuded layer pre-a, due to cascade hypothesis has dominated. The hypothesis
726

postulates that Ab-plaques are formed before tan- pyramidal cells of the Ammon’s horn) built-up by
gles and may even cause their formation. Only re- PHFs occur only in late stages of AD, i.e., stage VI
cently, work using the well-defined circuitry of the of Braaks’ classification (Braak and Braak, 1991).
hippocampal formation provided strong evidence The percentage of tangle-bearing neurons among
that tangles develop either before or independent all neurons of the granule cell layer is 1.7–4.2 in
from plaques (Schönheit et al., 2004) and follows severe AD (Bobinski et al., 1997). Spherical tau-
an anterograde pattern (Thal et al., 2002; Schön- aggregates consisting of straight filaments
heit et al., 2004). (18–25 nm in diameter), however, are seen in many
Apart from the tau-pathology and the Ab- granule cells when probing severe Alzheimer cases
pathology, there is an inconspicuous pathology, with TG3 (Wakabayashi et al., 1997), a mono-
which, not surprisingly, correlates best with the de- clonal antibody binding a phosphate-dependent
gree of cognitive impairment. This is the loss of epitope in PHFs. Interneurons seem to be very
synapses reported by Davies and colleagues (Davies resistant, perhaps because of their high content of
et al., 1987), later confirmed in many subsequent calcium-binding proteins such as parvalbumin or
electron-microscopical studies and (indirectly) by calretinin (Braak et al., 1991; Nitsch and Ohm,
semi-quantitative immunolabelling of synaptic mar- 1995; Freund and Buzsaki, 1996). However, an
ker proteins (Scheff and Price, 2003). Synapse loss obligatory absence cannot be presumed because in
takes place in terminal zones of neurons prone to frontal cortex 4–6% parvalbumin-containing neu-
develop tangles. Thus, it is not surprising that rons also reacted with an antibody raised against
number of tangles also correlate strongly with the tau from tangles (Iwamoto and Emson, 1991).
cognitive decline. Tangle-bearing neurons express Others have reported only a 0–0.7% subpopula-
baser levels of the message for the synaptic marker tion of parvalbumin-containing neurons of the
protein synaptophysin (Callahan and Coleman, superior frontal gyrus which may have tangles
1995), suggesting that synapse dysfunction and loss (Sampson et al., 1997). In the hippocampus, how-
is not only the self-evident consequence of neuronal ever, only two cells out of 1950 examined parval-
loss but may occur in still living neurons. This is bumin-containing interneurons showed — an even
further indicated by a reduction in the synapse to questionable — tangle formation (Ohm et al.,
neuron ratio in dentate gyrus (48%) (Bertoni- 2002). Also for NOS-containing neurons, it is
Freddari et al., 1996). Interestingly, dendrites which shown that they resist tangle formation. The rel-
lay within plaques show similar numbers of spines ative scarceness of NOS neurons in the dentate
compared to adjacent dendrites or dendritic seg- gyrus does not allow forming a definite opinion
ments which lay in the inter-plaque region of the about this region. However, double-labelling of
neuropil (Einstein et al., 1994). Apart from numeric other hippocampal neurons has revealed that more
loss of synapses there are morphological changes, than 3300 tangle-bearing neurons are negative for
which may differ depending on the nature of the NOS. Starting with Braak-stage IV, some large
participating fibers and cells. From mossy fiber ter- multipolar neurons of the plexiform layer and ad-
minals decreases in number and area of spines and jacent rim of the sector CA4 develop tangles ex-
post-synaptic thickenings are reported, whereas non- tending into proximal dendrites. In stage V, a
mossy fiber terminals seem to behave in an opposite small number of closely arranged tangles are seen
way, i.e., displaying more synapses, spines and post- in modified pyramidal cells, thereby easily distin-
synaptic thickenings (Kiktenko et al., 1997). guishable from the ones in the multipolar neurons
described above. In stage VI, no new type accrues
AD related histopathology in the dentate gyrus but the number of afflicted neurons increases con-
siderably. Neuropil threads are mainly located in
Tau-pathology dendrites of neurons that undergo hyper-
phosphorylation and, subsequently, formation of
Densely packed tangles with a globose shape (in argyrophilic aggregation of tau (Braak and Braak,
contrast to the torch-like shaped tangles, e.g., in 1991). This may indicate that a neuron’s main
727

receptive structure, the dendrite, has lost normal Bertoni-Freddari and colleagues reported highly
structure and function. Beginning with stage III of significant reductions in synaptic length, surface
Braaks’ classification, hyperphosphorylated tau density and numeric density determined in the
diffusely stains the outer molecular layer of the supragranular band (innermost part of the inner
dentate gyrus. At stages V and particularly at stage molecular layer). Later ultrastructural examina-
VI, argyrophilic neuropil threads are seen (Thal et tions have confirmed this (Scheff and Price, 1998),
al., 2000). This corresponds to the beginning of and also assigned it to the outer molecular layer
tangle formation (coarse granular material) in the (Dekosky et al., 1996; Scheff et al., 1996). The
granule cells at stage V, and marked numbers of reduction has been calculated to a 21% decrease in
tangles exhibiting a globose shape in stage VI numeric synaptic densities and a 26% decrease in
(Braak and Braak, 1991). the thickness of the outer molecular layer, and a
27% decrease in numeric synaptic densities and
26% reduction in thickness of the inner molecular
Plaques
layer vs. control, respectively. None of the studies,
however, has analysed Braak-staged material,
Ab-deposits in CA1 occur in phase 2 of Ab depo-
which would have allowed making estimates
sition, the dentate gyrus is involved later, namely in
about the dynamic of the process. Given that
phase 3 (Thal et al., 2002). A correlative analysis
synaptophysin represents synapses, a significant
suggests that phases 1 and 2 are associated with
reduction was visualized correlating to the cogni-
CDR 0, i.e., no signs of cognitive deterioration
tive decline. Early, mild and severe Alzheimer cases
(Thal et al., 2002). Neuritic plaques are seen in a
are accompanied by a 23, 49 and 67% reduction in
zone between the outer two and the inner third of
the outer third of the molecular layer, and a 30, 39
the molecular layer, i.e., within the terminal zone of
and 62% loss in the middle third, respectively. No
the perforant path (Crain and Burger, 1988). They
significant changes are seen in the inner third of the
form later than the tangles in the layer pre-a of the
molecular layer (Masliah et al., 1994). At the time
entorhinal cortex, namely not before Braak-stage
being, we have no information about the relative
IV (Braak and Braak, 1991). The caveat that spot
changes in synapse subtypes, e.g., excitatory vs.
check-like examinations might overlook changes
inhibitory forms. Many transmitters coexist within
is ruled out in studies using close-meshed serial
the same terminal (Torrealba and Carrasco, 2004).
sections throughout the whole hippocampal
It is, therefore, not insentient to consider the
formation (Schönheit et al., 2004). The limbic
possibility of selective defects associated with
stage III with hyperphosphorylated tau highlight-
only one of the transmitters leaving the other
ing the outer molecular layer and neuritic plaques
transmitter and the ‘synapse’ intact.
formed at stage IV probably represents neuro-
Because synapses often terminate on dendritic
pathological signs of insipient AD (Bancher et al.,
spines, changes in spine number, spine density and
1993; Braak et al., 1993). At stage V, the numbers
spine shape have been determined as well. Spine
increase and a row-like appearance becomes clear.
numbers of the dentate gyrus are reduced in AD
A dense ribbon of neuritic plaques (Braak and
cases (Einstein et al., 1994) by 35–44%, depending
Braak, 1991) characterizes the stage VI. Interest-
on the topographical relationship between plaques
ingly, the terminal zone of the mossy fibers is
and dendrites. Dendrites crossing plaques when
almost devoid of neuritic plaques even in late and
drawing through the molecular layer show 9%
severe stages of AD.
fewer spines than those traversing in plaque-free
regions. This difference, however, is statistically
Synaptic changes not significant. As there are no data concerning the
numbers of axons within plaque-free and plaque
In 1988, the first report on AD-related synaptic regions, the putative role of plaques (be it Ab-
loss quantitatively assessed by electron microscopy peptide or Ab-amyloid) on spine maintenance or
was published (Bertoni-Freddari et al., 1988). plasticity in AD remains penumbral. Others have
728

found a reduction in granule cell spine density by 2000). 3[H] kainate, which labels binding sites on
60% (Williams and Matthysse, 1986), or 40% (De mossy fibre terminals, is found reduced in autora-
Ruiter and Uylings, 1987) in Golgi-stained mate- diography of the stratum lucidum (34%)
rial and of 23% in DiO-labelled neurons (Ji et al., (Represa et al., 1988). The dendritic tree (basal as
2003). One should, in this context, keep in mind well as apical) of the CA3 pyramids, however, is
that determinations of spine densities not only de- unchanged (Flood et al., 1987a, b) suggesting no
pend on post- and peri-mortem factors more than overall loss of synaptic contacts. In line herewith,
other morphometrical measures (Uylings et al., non-mossy fibre terminals (of unknown origin)
1986), but also on shrinkage of the dendrite or the form more synapses and are associated with
selection of an appropriate or representative den- a larger area of spines and form post-synaptic
dritic intercept. In addition, Golgi stains do show thickenings (Kiktenko et al., 1997).
only a small proportion of neurons and they may
not be representative. Determinations of the
post-synaptic density protein SAP97 has unveiled Neuronal loss
no significant decrease in AD dentate gyrus
(Wakabayashi et al., 1999). In cognitively normal individuals, 9–22 million
Detailed data on intrinsic connectivity in AD is granule cells are counted with stereological tech-
apparently lacking. It is, however, conceivable, that niques (Seress, 1988; West and Gundersen, 1990;
certain subtypes of dentate interneurons react with West et al., 1994; Bobinski et al., 1996a, 1997;
adaptive changes analogous to those seen under Simic et al., 1997; Harding et al., 1998; Korbo
other pathological circumstances [e.g., epilepsy et al., 2004). The broad range (>factor 2) is prob-
(Magloczky et al., 2000)]. It is to note in this con- ably not only due to a large inter-individual var-
text, however, that epilepsy may be a symptom of iation but also to differences in the sampling
advanced AD (Lozsadi and Larner, 2006). scheme or anatomical delineation in regions where
With respect to output synapses, mossy fibre the dentate gyrus does not display a C-shaped
terminals are the sole ones. They are marked by band. Thus, relative changes seem to be more in-
zinc and dynorphin. Densitometrical analysis of formative. Unfortunately, the picture remains
Timm-stained zinc has found an increase in AD somewhat blurred. The reported neuronal cell loss
(Goldsmith and Joyce, 1995). This may reflect in the granule cell layer ranges from 0% (Korbo et
either an increase in zinc levels per terminal or an al., 2004), 8% (n.s.) (Bobinski et al., 1997), around
increase in terminal with unchanged levels of zinc, 11% (n.s.) (West et al., 1994, 2004), 25% (n.s)
or higher shrinkage in Alzheimer tissue with more (Simic et al., 1997) to 43% (po0.05) (Bobinski
zinc-rich terminal per tissue volume. The copper– et al., 1996b). Curiously, researchers who found
zinc superoxide dismutase, expressed in granule also the 43% loss described the 8% loss. However,
cells, might be unaltered on protein and message the fact that five out of six studies have failed to
level (Ceballos et al., 1991), if the low numbers of show a statistically significant change suggests that
investigated cases were representative. However, a putative dysfunction of the dentate gyrus is not
35% reduction is found in the stratum lucidum of caused by simple loss of granule cells. Interestingly
CA3 for hMTH1, another neuronal protein sug- to note is a transient tendency for more granule
gested to protect from zinc-mediated oxidative cells in pre-clinical Alzheimer (+13%) of Braak
damage (Furuta et al., 2001). Decreases in number stages II–IV (West et al., 2004). This may be the
and area of spines and post-synaptic thickenings result of an increase in neurogenesis, as it appears
are reported with respect to mossy fiber terminals to occur in some Alzheimer brains (Jin et al., 2004).
on CA3 dendrites (Kiktenko et al., 1997). However, granule cell dispersion, occurring in
Chromogranin B, a large dense core vesicle pro- epilepsy, and suggestive for substantial reactive
tein and within the hippocampus particularly rich neurogenesis, does apparently not take place in
in mossy fibres, is found reduced (46%) in AD. Because most granule cells normally express
immunocytostaining intensity (Marksteiner et al., calbindin, calbindin expression and protein levels
729

are of interest. A 25% reduction (not statistically the dentate gyrus and sector CA4 together (Brady
significant) is found in a radioactive in situ hybrid- and Mufson, 1997), whereas others have found a
isation analysis (Maguire-Zeiss et al., 1995) and a slight increase in the dentate gyrus (Satoh et al.,
85% reduction mRNA level is seen in hippocampal 1991). The latter study is flawed by the fact that
homogenate determinations (Iacopino and the total number of cases and immunoreactive cells
Christakos, 1990). The calbindin protein level in is low. In other brain areas, a reduction is found,
whole hippocampal homogenates is reported to be ranging between 20 and 72%, depending on brain
67% (Iacopino and Christakos, 1990). The reduc- region and layer (Arai et al., 1987; Solodkin et al.,
tions may reflect a loss of calbindin-mRNA and 1996; Mikkonen et al., 1999). Again, others differ
subsequently also calbindin occurring before gran- when reporting no loss (Hof et al., 1991; Fonseca
ule cells fade away. Others, pondered that loss of et al., 1993; Sampson et al., 1997; Leuba et al.,
calbindin in many of the granule cells might relate 1998). In general, however, the loss of parvalbu-
to the cognitive decline in AD despite of the fact min-containing neurons is seen only in late
that the total number of granule cells might be un- Alzheimer and especially in those regions, which
altered (Greene et al., 2001; Palop et al., 2003). undergo massive differentiation or loss of principal
Calbindin loss, however, is a well-known feature of neurons. Thus, it is likely that at least those
dentate granule cells in epilepsy (Magloczky et al., GABAergic interneurons, which contain parval-
1997) where it not necessarily associates with bumin, are primarily resistant against the tau
cognitive decline or even dementia. In epilepsy, a pathology and untimely cell death. Another inter-
group with severe hippocampal cell loss (mainly in neuron-associated calcium-binding protein is
CA1), displaying the so-called Ammon’s horn calretinin, which is found in differentially shaped
sclerosis, shows a marked reduction in granule cell neurons of the human dentate gyrus (Nitsch and
calbindin levels. A second group with medial tem- Ohm, 1995). With respect to the AD hippocampus
poral lobe epilepsy (but not with Ammon’s horn no indication for a loss in the numbers of immuno-
sclerosis) displays normal calbindin levels in their stained neurons is reported (Brion and Resibois,
granule cells. Thus, alternatively to a putative 1994). In line herewith, also other brain areas
causal role in cognitive decline, depletion of cal- display an unchanged overall neuronal density of
bindin may increase the resistance of granule cells calretinin-positive neurons (Hof et al., 1993;
(Nagerl et al., 2000). Interestingly, immunocyto- Fonseca and Soriano, 1995; Leuba et al., 1998;
chemical demonstration of calmyrin, another Mikkonen et al., 1999). The observed layer-specific
granule cell-associated calcium-binding protein, increase of small neurons that is concomitantly
and which is lacking in interneurons, showed no found with a decrease in large calretinin-contain-
differences between controls and Alzheimer cases ing neurons (Sampson et al., 1997; Leuba et al.,
in stages IV–V (Bernstein et al., 2005). Also 1998; Mikkonen et al., 1999) supports the notion
neuron-specific enolase is found unaltered that more subtle plastic–adaptive changes occur.
(Wakabayashi et al., 1999). Together this further In this context it is also to note that the intensity of
supports the notion that granule cells withstand the calretinin immunoreaction with a staining
neuronal death in AD. gradient (apical dendrites stronger, basal weaker)
Apart from loss of granule cells, local interneu- is inverted in the entorhinal cortex of AD patients
rons have to be considered, too. Many of these are (Mikkonen et al., 1999). This suggests a plastic and
characterized by the presence of certain calcium- adaptive process. Coincided with calretinin, the
binding protein: parvalbumin, calretinin and cal- innermost part of the molecular layer is normally
bindin. It seems that at least some of the inter- strongly immunolabelled by secretoneurin, a mem-
neuron types (Freund and Buzsaki, 1996) have lost ber of the chromogranin family. In AD, this band
the parvalbumin-phenotype or are even numeri- is significantly reduced (52%), whereas the in-
cally reduced. Data reported, however, are not tensity of the calretinin immunostaining is only
unanimous. A 60% reduction in neuronal density slightly diminished (16%) (Kaufmann et al.,
of parvalbumin-containing neurons is found for 1998). The rest of the inner molecular layer is
730

unchanged, but a 40% decrease in secretoneurin is devoid of CCK-8-positive profiles and only rarely
found in the outer molecular layer. The third cal- they were seen in the angular bundle. The only
cium binding protein of various interneurons is exception seems to be a projection via the perfo-
calbindin, which, however, is also typical for gran- rant path probably originating from pre-a neu-
ule cells (Sloviter et al., 1991; Seress et al., 1993). rons. Changes in CCK are not reported for AD
This implies that determination of protein or mes- hippocampus. Some larger enkephalin-containing
sage levels represent the contribution of both neu- neurons are found in the molecular layer of the
ronal classes. The immunocytochemical analysis of dentate gyrus (Kulmala, 1985). No changes are
the dentate gyrus reveals only few calbindin-con- seen in AD (Kulmala, 1985) when examining the
taining granule cells and interneurons in Al- dentate gyrus or the whole hippocampus (Yew
zheimer brains (Iritani et al., 2001). No changes et al., 1999). However, a Leu- and Met-encephalin-
in the number and distribution of GABAergic receptor binding loss is found in hippocampal ho-
neurons are seen in the hippocampus between mogenates (Rinne et al., 1993). Others have found
controls and AD (Yew et al., 1999). Some inter- a statistically significant 48% loss of m- and a 36%
neurons of the dentate gyrus, mostly located just loss of k-opiod-binding sites whereas no changes
underneath the granule cell layer, contain neuro- have been seen for d-binding sites in quantitative
peptide Y (NPY). These peptidergic large poly- autoradiography (Mathieu-Kia et al., 2001).
morphic neurons are reduced in numbers, to only Prominent terminal-like Substance P staining is
a mild degree in the rostral and middle part of the seen in the dentate gyrus and in CA1–4 fields, and
hippocampus, but evident caudally. Although multipolar immunolabelled neurons are located in
there are neurons with intense immunostaining the polymorphic layer of the dentate gyrus as well
and a preserved dendritic tree, most others, how- as in CA1–4 (Kowall et al., 1993). These neurons
ever, show a loss or regression of their dendritic are NADPH-negative, in contrast to most of the
tree (Chan-Palay et al., 1986). A further interneu- Substance P-containing neurons of the isocortex.
ron subgroup is identified by its content of som- Substance P-receptor immunocytochemistry
atostatin. These neurons are evenly spread shows a light marking of CA2 pyramidal cells
throughout the polymorphic layer, lacking in the and occasional labelling of basket cells in CA1–4.
granule cell layer, and are only occasionally found In AD, staining intensity is reduced in the dentate
in the molecular layer. In AD, only vestiges remain gyrus but sparing the hilar neurons (Kowall et al.,
with greatly diminished dendritic processes 1993). Nitric oxide is generated in neurons with
(Chan-Palay, 1987). This is in line with biochemi- NADPH-diaphorase activity (NOS-neurons). The
cal findings, which have determined a 70% loss of distribution and morphological types does not
somatostatin in hippocampal tissue (Davies and differ significantly between controls and Alzheimer’
Terry, 1981). Also in other cortex areas, disease patients in all sectors of the Ammon’s horn
somatostatin-containing interneurons show a and subiculum (Hyman et al., 1992). Also the total
decreased numeric density (>70%) (Kumar, number of NOS neurons is unaltered. However,
2005) and/or cytoskeletal changes reminiscent of whereas in CA4 and CA3 a considerable decline
tau-pathology (van de Nes et al., 2002). Together (though not statistically significant) in the numbers
this suggests that this population is more vulner- is found, the subiculum shows a light increase.
able than other interneurons. Multipolar CCK-8- Only few NOS-neurons are occasionally found in
containing neurons are occasionally found in the the dentate gyrus, i.e., in 2 of 20 examined hippo-
molecular layer and numerous polymorphic ones campi. Double labelling has revealed that more
in the anterior part of the dentate gyrus, mainly in than 3300 tangle-bearing neurons are negative for
the subgranular zone of the polymorphic layer NOS. The planimetrically determined size of the
(Lotstra and Vanderhaeghen, 1987). In contrast to neurons does not show changes (exception CA1)
animals, CCK-8-positive fibers seem to emanate but regressive changes are found in the form
almost exclusively from intrahippocampal neu- of foreshortened dendrites and distorted and
rons, because fimbria and alveus were found beaded axons. NADPH-diaphorase activity is
731

histochemically visualized and displays a striking no changes are seen in the dentate gyrus in quan-
loss in the terminal zone of the perforant path in titative autoradiography (Dewar et al., 1991) or a
AD (Rebeck et al., 1993). In controls, this zone reduction in the outer molecular layer is found,
stands out clear due to the high level of reaction together with an increase in the polymorphic layer
product and gives the molecular layer a sublam- (Geddes et al., 1992). However, Kd and Bmax may
inated aspect. As with NOS antibody staining, have changes, making an interpretation difficult.
only few neurons are stained with NADPH- Again, more recent studies give a more defined
diaphorase activity in the granule cell layer. CA4 picture. Glu-R1 is found unaltered in an in situ
appeared to be unaltered. hybridisation study (Pellegrini-Giampietro et al.,
1994) and in immunocytochemical investigations
(Aronica et al., 1998). Others, however, find a
The excitatory–inhibitory neurochemistry subtly modified picture. Glu-R1 immunoreactive
neurons are apparently unaltered in number
Because the dentate gyrus receives a glutamatergic (Ikonomovic et al., 1995; Aronica et al., 1998;
excitatory input from the layer pre-a and transmits Wakabayashi et al., 1999) but show an increase in
processed information to CA3, the glutamatergic staining intensity in the molecular layer (Hyman
system has attracted researchers and numerous ne- et al., 1994; Ikonomovic et al., 1995; Wakabayashi
urochemical studies have been performed. Early et al., 1999) especially in the inner (Hyman et al.,
studies using microdissected specimens of the 1994) and supragranular part (Ikonomovic et al.,
termination zone of the perforant path, and subse- 1995), and in the plexiform layer (Ikonomovic
quent determinations of the glutamate content et al., 1995). This increase is on one hand attributed
therein, have demonstrated an 83% reduction in to an increase in the number of immunoreactive
AD (Hyman et al., 1987). More recent studies have fibers but on the other hand due to an increase of
determined possible changes at a higher definition immunoreactivity. The latter may also explain why
in terms of receptor subtypes and (sub)regional the proportion of Alzheimer cases displaying a sig-
distribution. In contrast to significant decreases in nal (specific immunoreactivity over background) is
other parts of the hippocampus (or the hippocam- about twice as high as of controls (Hyman et al.,
pus as a whole), the dentate gyrus shows no sig- 1994). With respect to CA3 (and CA4), i.e., the
nificant changes for NMDA-R1, 2A and 2B terminal zone of the granule cells, an increase is
message or protein (despite of trends toward de- found (Aronica et al., 1998), which seems to follow
creased levels) as determined by Western blotting of the progression of the Braak-stages (though not
micropunches (Wakabayashi et al., 1999; Mishizen- statistically significant) (Carter et al., 2004). Curi-
Eberz et al., 2004), radioactive in situ hybridisation ously, Western blotting shows a significant 70%
(Ulas and Cotman, 1997) and immunocytochemis- reduction in the dentate gyrus of AD patients,
try (Aronica et al., 1998). An immunocytochemical which is accompanied by virtually identical levels of
study with Braak-staged tissue has found no neuron-specific enolase (Wakabayashi et al., 1999).
changes in early stages but an increase in staining The latter suggests that no major granule cell loss
intensity at stages IV and later in all CA fields has taken place and is in line with most of the cell
concomittant with a decrease in staining of the counting studies. The apparent contradiction be-
outer molecular layer (Ikonomovic et al., 1999). tween the Western blot and immunocytochemical
NMDA-receptor message and protein is located on data obtained from identical cases (Wakabayashi et
neurons, i.e., granule cells and pyramidal cells, al., 1999) may be due to unmasking of epitopes or
showing expression differences between the hippo- increasing of affinity, which is not observed under
campal subfields. NMDA-R1, -2A and -2B message denaturation conditions in Western blotting.
is highest in granule cells, followed by CA2 and 3, Glu-R2 and Glu-R2/3 immunoreactivity is found
protein highest in CA1 for NMDA-R1 and -2A and to be unaltered (Hyman et al., 1994) or increased in
granule cells for 2B (Mishizen-Eberz et al., 2004). the (inner) molecular layer and plexiform layer of
Likewise AMPA receptors have been examined and the dentate gyrus (Ikonomovic et al., 1995; Aronica
732

et al., 1998). The number of Glu-R2 immunoreac- found unaltered or even displaying a slight in-
tive neurons in the dentate gyrus is reduced crease. Micropunches of hippocampi subjected to
(Ikonomovic et al., 1995). Microdissected punches Western blotting of b1, b2, a1 and a5 subunits of
of hippocampi, grouped according to the Braak- GABAA receptors show no differences between
classification into stages 0–II, stages III–IV and controls and Alzheimer individuals. A significant
stages V–VI, and subjected to Western blotting, has decrease is seen only for a5 in CA3 (20%) (Ri-
contained comparable amounts of immunoreactive ssman et al., 2003), whereas a5-ligand binding
protein (Carter et al., 2004). Interestingly, however, show only in CA1 a significant 27% decrease (Ho-
double-labelling of Glu-R2 and tangles in the well et al., 2000). Immunocytochemical analyses
entorhinal cortex and subiculum/CA1 by the give similar results for GABAA-R-b2/3 (Mizukami
MC1 antibody has resulted in a complementary et al., 1997) and GABAA-R-a1 (Singer et al., 2006)
pattern: number of Glu-R2 immunoreactive cells when comparing cases with Braak-stages I–VI. In
diminished when the number of MC1 positive (i.e., severe stages, the GABAA-R-a1 is found decreased
tangle-bearing cells) increased (Ikonomovic et al., locally on soma and dendritic surface of mossy
1997). Others, however, report similar immunore- cells. Labelling of interneurons with GABAA-R-a1
activity in both tangle-bearing and tangle-free neu- or GABAA-R-b2/3 remains unaltered though they
rons for any of the studied Glu-R subtypes (Hyman display shrunken processes. GABAB-R1, however,
et al., 1994). Glu-R4, though relatively sparse in the show a transient but significant increase in stages II
hippocampus, is found unaltered in terms of stain- and IV in sectors CA4 and CA3, whereas a 64%
ing pattern and intensity. In sharp contrast to loss of immunoreactive neurons of sector CA1
staining with Glu-R1 and Glu-R2 antibodies, the takes place between stages I–II and II–IV which
immunoreaction is not localised to dendrites or ax- further proceeds to 70% in stages V–VI (Iwakiri et
ons but highlighted neuronal somata including al., 2005). The finding that between stages I–II and
CA3 pyramidal cells and granule cells (Hyman III—IV, only 12% of Nissl-stained neurons are lost
et al., 1994). Quantitative 3[H] kainate autoradiog- (and 39% in stages V–VI) suggests that receptor
raphy reveals a statistically significant 54% reduc- loss precedes neuronal death. In situ hybridisation
tion in the supragranular layer (Represa et al., for GABAA-R-b2 in cases with Braak stages I–VI
1988). Kainate receptors can be visualised with an- shows — in line with the protein findings — no
tibodies raised against Glu-R5/6/7. Immunoreac- changes. In contrast, GABAA-R-b3 message is
tive neurons are granule cells, and mossy cells as reduced in all hippocampal sectors but not in the
well as pyramidal cells of the CA sectors. The som- CA4 (Mizukami et al., 1998). In a parenthetical
atodendritic compartment is labelled. Other neu- note GABAA-R-a1 message localisation and signal
rons marked are non-pyramidal cells of the stratum intensity is reported to be apparently unaltered
oriens and some glial cells in the alveus and fimbria. in Alzheimer hippocampus (Pellegrini-Giampietro
No significant changes are seen in Alzheimer brains et al., 1994).
(Aronica et al., 1998).
With respect to the inhibitory GABAergic sys-
tem autoradiographic studies show a similar Kd for Disconnected dentate gyrus
3
[H] GABA hippocampal binding sites between
controls and Alzheimer cases (Chu et al., 1987). From the above it becomes conceivable that the
GABAA receptors are found unaltered, whereas classical histopathological signs of AD, which cor-
GABAB binding sites are significantly reduced in relate strongly with the degree of dementia (tan-
the molecular layer of the dentate gyrus, the stra- gles), are late events. This also holds true for
tum pyramidale and lacunosum-moleculare of the neuronal cell loss. In contrast, clinical signs such as
sector CA1 (Chu et al., 1987). A similar picture is dementia may develop much earlier. In terms of the
seen with 3[H] flunitrazepame except for the reduc- Braak classification tangles and neuropil threads
tion in the molecular layer of the gyrus dentatus evolve in stage V or later, whereas loss of cognitive
(Jansen et al., 1990; Penney et al., 1990), which is and memory function as assessed by the mini
733

mental state examination (MMSE) test may occur marked rarefaction of white matter along the
already in stage III (Bancher et al., 1993; Braak course of the perforant tract together with a loss
et al., 1993). The dentate gyrus is considered to of oligodendrocytes thereabout (Morys et al.,
play a crucial role in associative memory, especially 1994). In vivo imaging by MRI scans in psycho-
with respect to events (‘what happens’) (Morris, metrically assessed individuals supports this. A
2006). This raises the possibility that changes which decrease in white matter volume in the parahippo-
lead to interruption or perturbation of the granule campal gyrus that includes the perforant path is
cells’ input isolates the dentate gyrus and causes found already in amnesic mild cognitive impaired
impairment of learning and memory (Hyman et al., individuals (the MMSE scores suggest that this
1984). To discharge one single granule cell, 400 might represent very mild AD) (Petersen et al.,
axons of the perforant path have to convey the exci- 2006; Stoub et al., 2006).
tatory input from the pre-a neurons (McNaughton Post-synaptically, the loss of perforant path in-
et al., 1991). This suggests that loss of entorhinal put seems to induce changes in granule cells. Gran-
pre-a neurons may be sufficient to cause learning ule cells remodel their dendritic tree in AD as
and memory defects. Support for this concept determined in quantitative morphometrical analy-
comes from several lines of evidence. Correlative ses on Golgi stained (Flood et al., 1985, 1987a, b)
clinico-neuropathological studies are, unfortu- and with Lucifer yellow intracellularly filled (Ein-
nately, not unambiguous. In general, the numbers stein et al., 1994) neurons. In the Golgi studies, the
of tangles in entorhinal neurons correlate with both granule cells’ apical dendrites of severe Alzheimer
Braak staging and deterioration of cognition cases are compared to controls. The controls con-
(Garcia-Sierra et al., 2000, 2001). An early study sist of three differentially aged groups of non-
in 1996 indicated that a 57% loss of layer pre-a demented individuals. The middle-aged group
neurons (650,000 in CDR 0 control cases vs. (mean of age 52.2972.60 SEM), the old-aged
285,000 in CDR 0.5 individuals) is associated with group (comparable to the Alzheimer group)
a Clinical Dementia Rating (CDR) score of 0.5, (73.4072.36 years) and the very old-aged group
which represents beginning dementia (Gomez Isla (90.2071.85 years), however, have various degrees
et al., 1996). At that time, the total entorhinal of AD-related histopathology. Flood and col-
neuron loss is only 30%. This is confirmed by a leagues have found a significantly reduced total
35 and 50% pre-a layer neuron loss in CDR 0.5 dendritic length (37%) and a significantly reduced
cases (Kordower et al., 2001; Price et al., 2001), average segment length (dendritic length/number of
respectively, but not in a newer study (only 1–2% segments; 34%) but no significant changes in
loss) (Hof et al., 2003). Although there is no direct total segment numbers or width of the fascia den-
translating between CDR and Braak scores, it is tata (granule cell layer + molecular layer) when
suggested that a CDR 0.5 may represent Braak- comparing the old-aged group with the Alzheimer
stage III and a MMSE range of 26–29 (Hof et al., group. Interestingly, the Alzheimer group has only
2003). Severe AD (CDR 3–5, MMSE 0 or Braak a 16% shorter total dendritic length than the mid-
stage VI) is found associated with a 90% loss of dle-aged group, which has the lowest degree of
pre-a neurons (Gomez Isla et al., 1996) and a AD-related neuropathology (no changes or only
laminar-specific spongiosis in the perforant path ‘very slight or slight changes’). This difference as
termination zone (Duyckaerts et al., 1998) as well well that between the very old group and the
as ALZ50 immunoreactivity highlights the perfo- Alzheimer group is not statistically significant in
rant path terminal zone (Hyman et al., 1988). The post-hoc tests. It is tempting to speculate that the
immunocytochemical detection of the intrinsic pre- difference in the dendritic tree measures between
synaptic vesicle protein synaptophysin, suggests a the middle-aged group (with no or only very few
functional or terminal loss in the termination zone AD-related neuropathological changes) and the old
of the perforant path (Cabalka et al., 1992; group (with some more changes but no clinical
Masliah et al., 1994). Demonstration of the my- signs of AD) might represent an early stage in the
elinated fibres by the Weil-stain has indicated a evolution of AD where regenerative plasticity is
734

dominant (Flood et al., 1985). Later, when differ- comparative morphometrical analysis of apical
entiation proceeds the balance between regenerative and basal dendrites of parvalbumin-containing
(progressive) and degenerative (regressive) plasticity GABAergic interneurons located within or directly
should shift towards the formation of a reduced at the border of the granule cell layer has been
dendritic arborisation. Similar to the Golgi studies, made in Braak-staged cases (Ohm et al., 2002). A
the analysis of Lucifer yellow-filled apical granule non-significant and transient increase in dendritic
cell dendrites unveils a statistically significant re- length, branch order and number of segments of
duction in dendritic length (by 50%) and also a the apical dendrites is reported for an early Braak-
shorter average segment length (19%) (Einstein et stage (stage II) followed by a statistically significant
al., 1994). Proliferative signs, however, such as in- decline of the respective measures in stages IV and
creased number of dendritic twigs or unusual ex- V. Basal dendrites, in contrast, remained stable.
crescences are not observed. Because apical dendrites receive entorhinal input
Spine numbers are reduced in AD cases by from pre-a neurons whereas basal dendrites do not
35–60% (Williams and Matthysse, 1986; De Ruiter (Zipp et al., 1989), the dendrite-specific change
and Uylings, 1987; Einstein et al., 1994; Ji et al., suggests an input-specific plastic change. A
2003). However, it is likely that spine plasticity will 35–50% pre-a layer neuron loss is reported for
also occur. In early stages, first loss of afferents may mild cognitive decline (Kordower et al., 2001; Price
induce signals, which allow the local sprouting in et al., 2001), which may represent Braak-stage III
order to attract ‘new’ synaptic input from (Hof et al., 2003). It is therefore likely that a sub-
neighboured afferents. Given that 400 axons stantial loss of input (>50%) is required to induce
(synapses) may be necessary to discharge a granule the regressive changes which may eventually lead
cell (McNaughton et al., 1991), this may be an at- to the interneuron’s own death. It may also play an
tempt to keep the neuronal net properly working. important role how fast this input decreases. In-
Additionally, the existing terminals are statistically terestingly, the transient increase in dendritic
significantly enlarged in terms of apposition length length, segment numbers and branches (Ohm et
(+18% in outer molecular layer and +12% in the al., 2002) parallels a transient tendency for more
inner molecular layer) and — as a mild trend — as granule cells in pre-clinical Alzheimer (+13%) of
well as total synaptic contact area [in outer and in- Braak stages II–IV (West et al., 2004). The latter
ner molecular layer +4% (n.s.)] (Scheff et al., 1996; may be the result of an increase in neurogenesis, as
Scheff and Price, 1998) probably to serve the same it appears to occur in some Alzheimer brain (Jin et
purpose. Later, when differentiation proceeds, the al., 2004), the former as an attempt to sprout.
balance between regenerative (progressive) and de-
generative (regressive) plasticity shifts towards the Closing remark
formation of synaptically stripped dendrites (Scheff
and Price, 2003). AP180, a protein considered to be The available data suggest that the dentate gyrus
crucially involved in neuronal clathrin-mediated withstands the formation of plaques, tangles and
synaptic recycling, is expressed in granule cells and neuronal death until late stages of AD. However,
stains the molecular layer of cognitive normal changes related to a disconnecting process, mainly
humans homogeneously. In AD, a loss of granule from loss of entorhinal input, and more subtle in-
cell body staining but an increase in the staining trinsic alterations may contribute to disturbances
intensity and width of the inner molecular layer is in memory and learning observed already in early
reported (Yao et al., 1999). This suggests that gran- stages of AD.
ule cells show plastic adaptation probably respond-
ing to challenges on their dendritic input.
References
Apart from disconnection of principal cells, in-
terneurons may also become disconnected from the Amaral, D.G., Insausti, R. and Cowan, W.M. (1984) The com-
entorhinal input. This may alter the balance be- missural connections of the monkey hippocampal formation.
tween excitatory and inhibitory effects. A J. Comp. Neurol., 224: 307–336.
735

Arai, H., Emson, P.C., Mountjoy, C.Q., Carassco, L.H. and Braak, E., Braak, H. and Mandelkow, E.M. (1994) A sequence
Heizmann, C.W. (1987) Loss of parvalbumin-immunoreac- of cytoskeleton changes related to the formation of neuro-
tive neurones from cortex in Alzheimer-type dementia. Brain fibrillary tangles and neuropil threads. Acta Neuropathol.
Res., 418: 164–169. Berl., 87: 554–567.
Aronica, E., Dickson, D.W., Kress, Y., Morrison, J.H. and Braak, E., Strotkamp, B. and Braak, H. (1991) Parvalbumin-
Zukin, R.S. (1998) Non-plaque dystrophic dendrites in immunoreactive structures in the hippocampus of the human
Alzheimer hippocampus: a new pathological structure adult. Cell Tissue Res., 264: 33–48.
revealed by glutamate receptor immunocytochemistry. Braak, H. and Braak, E. (1991) Neuropathological stageing of
Neuroscience, 82: 979–991. Alzheimer-related changes. Acta Neuropathol. (Berl.), 82:
Augustinack, J.C., Schneider, A., Mandelkow, E.M. and 239–259.
Hyman, B.T. (2002) Specific tau phosphorylation sites cor- Braak, H., Braak, E., Grundke-Iqbal, I. and Iqbal, K. (1986)
relate with severity of neuronal cytopathology in Alzheimer’s Occurrence of neuropil threads in the senile human brain and
disease. Acta Neuropathol. (Berl.), 103: 26–35. in Alzheimer’s disease: a third location of paired helical
Bancher, C., Braak, H., Fischer, P. and Jellinger, K.A. (1993) filaments outside of neurofibrillary tangles and neuritic
Neuropathological staging of Alzheimer lesions and intellec- plaques. Neurosci. Lett., 65: 351–355.
tual status in Alzheimer’s and Parkinson’s disease patients. Braak, H. and Del Tredici, K. (2004) Poor and protracted
Neurosci. Lett., 162: 179–182. myelination as a contributory factor to neurodegenerative
Bancher, C., Brunner, C., Lassmann, H., Budka, H., disorders. Neurobiol. Aging, 25: 19–23.
Jellinger, K., Wiche, G., Seitelberger, F., Grundke Iqbal, I., Braak, H., Duyckaerts, C., Braak, E. and Piette, F. (1993)
Iqbal, K. and Wisniewski, H.M. (1989) Accumulation of Neuropathological staging of Alzheimer-related changes
abnormally phosphorylated tau precedes the formation of correlates with psychometrically assessed intellectual status.
neurofibrillary tangles in Alzheimer’s disease. Brain Res., In: Corain B., Iqbal K., Nicolini M., Winblad B., Wisniewski
477: 90–99. H. and Zatta P. (Eds.), Alzheimer’s Disease: Advances in
Bernstein, H.G., Blazejczyk, M., Rudka, T., Gundelfinger, Clinical and Basic Research. Wiley, Chicester, pp. 131–137.
E.D., Dobrowolny, H., Bogerts, B., Kreutz, M.R., Kuznicki, Brady, D.R. and Mufson, E.J. (1997) Parvalbumin-immunore-
J. and Wojda, U. (2005) The Alzheimer disease-related cal- active neurons in the hippocampal formation of Alzheimer’s
cium-binding protein Calmyrin is present in human forebrain diseased brain. Neuroscience, 80: 1113–1125.
with an altered distribution in Alzheimer’s as compared to Brion, J.P. and Resibois, A. (1994) A subset of calretinin-
normal ageing brains. Neuropathol. Appl. Neurobiol., 31: positive neurons are abnormal in Alzheimer’s disease. Acta
314–324. Neuropathol. (Berl.), 88: 33–43.
Bertoni-Freddari, C., Fattoretti, P., Casoli, T., Caselli, U. and Bush, A.I., Pettingell, W.-H.J., de Paradis, M., Tanzi, R.E. and
Meier-Ruge, W. (1996) Deterioration threshold of synaptic Wasco, W. (1994) The amyloid beta-protein precursor and its
morphology in aging and senile dementia of Alzheimer’s mammalian homologues. Evidence for a zinc-modulated
type. Anal. Quant. Cytol. Histol., 18: 209–213. heparin-binding superfamily. J. Biol. Chem., 269: 26618–26621.
Bertoni-Freddari, C., Meier-Ruge, W. and Ulrich, J. (1988) Cabalka, L.M., Hyman, B.T., Goodlett, C.R., Ritchie, T.C.
Quantitative morphology of synaptic plasticity in the aging and Vanhoesen, G.W. (1992) Alteration in the pattern of
brain. Scanning Microsc., 2: 1027–1034. nerve terminal protein immunoreactivity in the perforant
Bobinski, M., Wegiel, J., Tarnawski, M., de Leon, M.J., pathway in Alzheimer’s disease and in rats after entorhinal
Reisberg, B., Miller, D.C. and Wisniewski, H.M. (1998) Du- lesions. Neurobiol. Aging, 13: 283–291.
ration of neurofibrillary changes in the hippocampal pyrami- Callahan, L.M. and Coleman, P.D. (1995) Neurons bearing
dal neurons. Brain Res., 799: 156–158. neurofibrillary tangles are responsible for selected synaptic
Bobinski, M., Wegiel, J., Tarnawski, M., Reisberg, B., de Leon, deficits in Alzheimer’s disease. Neurobiol. Aging, 16:
M.J., Miller, D.C. and Wisniewski, H.M. (1997) Relation- 311–314.
ships between regional neuronal loss and neurofibrillary Carter, T.L., Rissman, R.A., Mishizen-Eberz, A.J., Wolfe,
changes in the hippocampal formation and duration and B.B., Hamilton, R.L., Gandy, S. and Armstrong, D.M.
severity of Alzheimer disease. J. Neuropathol. Exp. Neurol., (2004) Differential preservation of AMPA receptor subunits
56: 414–420. in the hippocampi of Alzheimer’s disease patients according
Bobinski, M., Wegiel, J., Wisniewski, H.M., Tarnawski, M., to Braak stage. Exp. Neurol., 187: 299–309.
Reisberg, B., de Leon, M.J. and Miller, D.C. (1996a) Ceballos, I., Javoy-Agid, F., Delacourte, A., Defossez, A.,
Neurofibrillary pathology — correlation with hippocampal Lafon, M., Hirsch, E., Nicole, A., Sinet, P.M. and Agid, Y.
formation atrophy in Alzheimer disease. Neurobiol. Aging, (1991) Neuronal localization of copper-zinc superoxide di-
17: 909–919. smutase protein and mRNA within the human hippocampus
Bobinski, M., Wegiel, J., Wisniewski, H.M., Tarnawski, M., from control and Alzheimer’s disease brains. Free Radic.
Reisberg, B., de Leon, M.J. and Miller, D.C. (1996b) Res. Commun., 12–13(Pt 2): 571–580.
Neurofibrillary pathology — correlation with hippocampal Chan-Palay, V. (1987) Somatostatin immunoreactive neurons
formation atrophy in Alzheimer disease. Neurobiol. Aging, in the human hippocampus and cortex shown by immuno-
17: 909–919. gold/silver intensification on vibratome sections: coexistence
736

with neuropeptide Y neurons, and effects in Alzheimer-type Flood, D.G., Guarnaccia, M. and Coleman, P.D. (1987b) Den-
dementia. J. Comp. Neurol., 260: 201–223. dritic extent in human CA2-3 hippocampal pyramidal neu-
Chan-Palay, V., Lang, W., Haesler, U., Köhler, C. and rons in normal aging and senile dementia. Brain Res., 409:
Yasargil, G. (1986) Distribution of altered hippocampal 88–96.
neurons and axons immunoreactive with antisera against Fonseca, M. and Soriano, E. (1995) Calretinin-immunoreactive
neuropeptide Y in Alzheimer’s-type dementia. J. Comp. neurons in the normal human temporal cortex and in Al-
Neurol., 248: 376–394. zheimer’s disease. Brain Res., 691: 83–91.
Chu, D.C.M., Penney Jr., J.B. and Young, A.B. (1987) Quan- Fonseca, M., Soriano, E., Ferrer, I., Martinez, A. and Tunon,
titative autoradiography of hippocampal GABAB and T. (1993) Chandelier cell axons identified by parvalbumin-
GABAA receptor changes in Alzheimer’s disease. Neurosci. immunoreactivity in the normal human temporal cortex and
Lett., 82: 246–252. in Alzheimer’s disease. Neuroscience, 55: 1107–1116.
Crain, B.J. and Burger, P.C. (1988) The laminar distribution of Freund, T.F. and Buzsaki, G. (1996) Interneurons of the
neuritic plaques in the fascia dentata of patients with hippocampus. Hippocampus, 6: 347–470.
Alzheimer’s disease. Acta Neuropathol. (Berl.), 76: 87–93. Furuta, A., Iida, T., Nakabeppu, Y. and Iwaki, T. (2001) Ex-
Davies, C.A., Mann, D.M.A., Sumpter, P.Q. and Yates, P.O. pression of hMTH1 in the hippocampi of control and
(1987) A quantitative morphometric analysis of the neuronal Alzheimer’s disease. Neuroreport, 12: 2895–2899.
and synaptic content of the frontal and temporal cortex in Garcia-Sierra, F., Hauw, J.J., Duyckaerts, C., Wischik, C.M.,
patients with Alzheimer’s disease. J. Neurol. Sci., 78: Munoz, J. and Mena, R. (2000) The extent of neurofibrillary
151–164. pathology in perforant pathway neurons is the key determi-
Davies, P. and Terry, R.D. (1981) Cortical somatostatin-like nant of dementia in the very old. Acta Neuropathol. (Berl.),
immunoreactivity in cases of Alzheimer’s disease and senile 100: 29–35.
dementia of the Alzheimer type. Neurobiol. Aging, 2: 9–14. Garcia-Sierra, F., Wischik, C.M., Harrington, C.R., Luna-Mu-
De Ruiter, J.P. and Uylings, H.B.M. (1987) Morphometric and noz, J. and Mena, R. (2001) Accumulation of C-terminally
dendritic analysis of fascia dentata granule cells in human truncated tau protein associated with vulnerability of the
aging and senile dementia. Brain Res., 402: 217–229. perforant pathway in early stages of neurofibrillary pathol-
Dekosky, S.T., Scheff, S.W. and Styren, S.D. (1996) Structural ogy in Alzheimer’s disease. J. Chem. Neuroanat., 22: 65–77.
correlates of cognition in dementia: quantification and as- Geddes, J.W., Ulas, J., Brunner, L.C., Choe, W. and Cotman,
sessment of synapse change. Neurodegeneration, 5: 417–421. C.W. (1992) Hippocampal excitatory amino acid receptors in
Demeter, S., Rosene, D.L. and van Hoesen, G.W. (1985) In- elderly, normal individuals and those with Alzheimer’s dis-
terhemispheric pathways of the hippocampal formation, pre- ease: non-N-methyl-D-aspartate receptors. Neuroscience, 50:
subiculum, and entorhinal and posterior parahippocampal 23–34.
cortex in the rhesus monkey: the structure and organization Goldsmith, S.K. and Joyce, J.N. (1995) Alterations in hippo-
of the hippocampal commissures. J. Comp. Neurol., 233: campal mossy fiber pathway in schizophrenia and Al-
30–47. zheimer’s disease. Biol. Psychiatry, 37: 122–126.
Dewar, D., Chalmers, D.T., Graham, D.I. and McCulloch, J. Gomez Isla, T., Price, J.L., McKeel Jr., D.W., Morris, J.C.,
(1991) Glutamate metabotropic and AMPA binding sites are Growdon, J.H. and Hyman, B.T. (1996) Profound loss of
reduced in Alzheimer’s disease — an autoradiographic study layer II entorhinal cortex neurons occurs in very mild Al-
of the hippocampus. Brain Res., 553: 58–64. zheimer’s disease. J. Neurosci., 16: 4491–4500.
Duyckaerts, C., Colle, M.A., Seilhean, D. and Hauw, J.J. Greene, J.R., Radenahmad, N., Wilcock, G.K., Neal, J.W. and
(1998) Laminar spongiosis of the dentate gyrus: a sign of Pearson, R.C. (2001) Accumulation of calbindin in cortical
disconnection, present in cases of severe Alzheimer’s disease. pyramidal cells with ageing; a putative protective mechanism
Acta Neuropathol. (Berl.), 95: 413–420. which fails in Alzheimer’s disease. Neuropathol. Appl.
Dyrks, T., Weidemann, A., Multhaup, G., Salbaum, J.M., Neurobiol., 27: 339–342.
Lemaire, H.G., Kang, J., Müller-Hill, B., Masters, C.L. and von Gunten, A., Kovari, E., Bussiere, T., Rivara, C.B., Gold,
Beyreuther, K. (1988) Identification, transmembrane orien- G., Bouras, C., Hof, P.R. and Giannakopoulos, P. (2006)
tation and biogenesis of the amyloid A4 precursor of Cognitive impact of neuronal pathology in the entorhinal
Alzheimer’s disease. EMBO J., 7: 949–957. cortex and CA1 field in Alzheimer’s disease. Neurobiol.
Einstein, G., Buranosky, R. and Crain, B.J. (1994) Dendritic Aging, 27: 270–277.
pathology of granule cells in Alzheimer’s disease is unrelated Haass, C. (2004) Take five — BACE and the gamma-secretase
to neuritic plaques. J. Neurosci., 14: 5077–5088. quartet conduct Alzheimer’s amyloid beta-peptide genera-
Flood, D.G., Buell, S.J., Defiore, C.H., Horwitz, G.J. and tion. EMBO J., 23: 483–488.
Coleman, P.D. (1985) Age-related dendritic growth in Harding, A.J., Halliday, G.M. and Kril, J.J. (1998) Variation in
dentate gyrus of human brain is followed by regression in hippocampal neuron number with age and brain volume.
the ‘oldest old’. Brain Res., 345: 366–368. Cereb. Cortex, 8: 710–718.
Flood, D.G., Buell, S.J., Horwitz, G.J. and Coleman, P.D. Hof, P.R., Bussiere, T., Gold, G., Kovari, E., Giannakopoulos,
(1987a) Dendritic extent in human dentate gyrus granule cells P., Bouras, C., Perl, D.P. and Morrison, J.H. (2003) Stereo-
in normal aging and senile dementia. Brain Res., 402: 205–216. logic evidence for persistence of viable neurons in layer II of
737

the entorhinal cortex and the CA1 field in Alzheimer disease. expression in Alzheimer’s patients: association with Braak
J. Neuropathol. Exp. Neurol., 62: 55–67. staging. Acta Neuropathol. (Berl.), 109: 467–474.
Hof, P.R., Cox, K., Young, W.G., Celio, M.R., Rogers, J. and Iwamoto, N. and Emson, P.C. (1991) Demonstration of ne-
Morrison, J.H. (1991) Parvalbumin-immunoreactive neurons urofibrillary tangles in parvalbumin-immunoreactive inter-
in the neocortex are resistant to degeneration in Alzheimer’s neurones in the cerebral cortex of Alzheimer-type dementia
disease. J. Neuropathol. Exp. Neurol., 50: 451–462. brain. Neurosci. Lett., 128: 81–84.
Hof, P.R., Nimchinsky, E.A., Celio, M.R., Bouras, C. and Jansen, K.L.R., Faull, R.L.M., Dragunow, M. and Synek, B.L.
Morrison, J.H. (1993) Calretinin-immunoreactive neocortical (1990) Alzheimer’s disease — changes in hippocampal N-
interneurons are unaffected in Alzheimer’s disease. Neurosci. methyl-D-aspartate, quisqualate, neurotensin, adenosine,
Lett., 152: 145–148. benzodiazepine, serotonin and opioid receptors — an auto-
Howell, O., Atack, J.R., Dewar, D., Mckernan, R.M. and Sur, radiographic study. Neuroscience, 39: 613–627.
C. (2000) Density and pharmacology of alpha5 subunit-con- Jellinger, K.A. (1997) Neuropathological staging of Alzheimer-
taining GABA(A) receptors are preserved in hippocampus of related lesions: the challenge of establishing relations to age.
Alzheimer’s disease patients. Neuroscience, 98: 669–675. Neurobiol. Aging, 18: 369–375.
Hyman, B.T., van Hoesen, G.W. and Damasio, A.R. (1987) Jellinger, K.A. and Bancher, C. (1997) Proposals for re-eval-
Alzheimer’s disease: glutamate depletion in the hippocampal uation of current autopsy criteria for the diagnosis of Al-
perforanth pathway zone. Ann. Neurol., 22: 37–40. zheimer’s disease. Neurobiol. Aging, 18: S55–S65.
Hyman, B.T., Van Horsen, G.W., Damasio, A.R. and Barnes, Ji, Y., Gong, Y., Gan, W., Beach, T., Holtzman, D.W. and
C.L. (1984) Alzheimer’s disease: cell-specific pathology iso- Wisniewski, T. (2003) Apolipoprotein E isoform-specific reg-
lates the hippocampal formation. Science, 225: 1168–1170. ulation of dendritic spine morphology in apolipoprotein E
Hyman, B.T., Kromer, L.J. and van Hoesen, G.W. (1988) A transgenic mice and Alzheimer’s disease patients. Neurosci-
direct demonstration of the perforath pathway terminal zone ence, 122: 305–315.
in Alzheimer’s disease using the monoclonal antibody Alz-50. Jin, K., Peel, A.L., Mao, X.O., Xie, L., Cottrell, B.A., Henshall,
Brain Res., 450: 392–397. D.C. and Greenberg, D.A. (2004) Increased hippocampal
Hyman, B.T., Marzloff, K., Wenniger, J.J., Dawson, T.M., neurogenesis in Alzheimer’s disease. Proc. Natl. Acad. Sci.
Bredt, D.S. and Snyder, S. (1992) Relative sparing of nitric U.S.A., 101: 343–347.
oxide synthase-containing neurons in the hippocampal for- Kaufmann, W.A., Barnas, U., Humpel, C., Nowakowski, K.,
mation in Alzheimer’s disease. Ann. Neurol., 32: 818–820. DeCol, C., Gurka, P., Ransmayr, G., Hinterhuber, H., Win-
Hyman, B.T., Penney, J.-B.J., Blackstone, C.D. and Young, kler, H. and Marksteiner, J. (1998) Synaptic loss reflected by
A.B. (1994) Localization of non-N-methyl-D-aspartate gluta- secretoneurin-like immunoreactivity in the human hippo-
mate receptors in normal and Alzheimer hippocampal for- campus in Alzheimer’s disease. Eur. J. Neurosci., 10:
mation. Ann. Neurol., 35: 31–37. 1084–1094.
Iacopino, A.M. and Christakos, S. (1990) Specific reduction of Khatchaturian, Z.S. (1985) Diagnosis of Alzheimer’s disease.
calcium-binding protein (28-kilodalton calbindin-D) gene ex- Arch. Neurol., 31: 545–548.
pression in aging and neurodegenerative diseases. Proc. Natl. Kiktenko, A.I., Uranova, N.A. and Denisov, D.V. (1997)
Acad. Sci. U.S.A., 87: 4078–4082. Quantitative characteristics of changes in synaptic contacts in
Ikonomovic, M.D., Mizukami, K., Davies, P., Hamilton, R., the hippocampus in Alzheimer’s disease. Neurosci. Behav.
Sheffield, R. and Armstrong, D.M. (1997) The loss of Physiol., 27: 681–682.
GluR2(3) immunoreactivity precedes neurofibrillary tangle Korbo, L., Amrein, I., Lipp, H.P., Wolfer, D., Regeur, L., Os-
formation in the entorhinal cortex and hippocampus of Al- ter, S. and Pakkenberg, B. (2004) No evidence for loss of
zheimer brains. J. Neuropathol. Exp. Neurol., 56: 1018–1027. hippocampal neurons in non-Alzheimer dementia patients.
Ikonomovic, M.D., Mizukami, K., Warde, D., Sheffield, R., Acta Neurol. Scand., 109: 132–139.
Hamilton, R., Wenthold, R.J. and Armstrong, D.M. (1999) Kordower, J.H., Chu, Y., Stebbins, G.T., Dekosky, S.T., Co-
Distribution of glutamate receptor subunit NMDAR1 in the chran, E.J., Bennett, D. and Mufson, E.J. (2001) Loss and
hippocampus of normal elderly and patients with Al- atrophy of layer II entorhinal cortex neurons in elderly peo-
zheimer’s disease. Exp. Neurol., 160: 194–204. ple with mild cognitive impairment. Ann. Neurol., 49:
Ikonomovic, M.D., Sheffield, R. and Armstrong, D.M. (1995) 202–213.
AMPA-selective glutamate receptor subtype immunoreactiv- Kowall, N.W., Quigley, B.-J.J., Krause, J.E., Lu, F., Kosofsky,
ity in the hippocampal formation of patients with Al- B.E. and Ferrante, R.J. (1993) Substance P and substance P
zheimer’s disease. Hippocampus, 5: 469–486. receptor histochemistry in human neurodegenerative dis-
Iritani, S., Niizato, K. and Emson, P.C. (2001) Relationship of eases. Regul. Pept., 46: 174–185.
calbindin D28K-immunoreactive cells and neuropathological Kril, J.J., Patel, S., Harding, A.J. and Halliday, G.M. (2002)
changes in the hippocampal formation of Alzheimer’s dis- Neuron loss from the hippocampus of Alzheimer’s disease
ease. Neuropathology, 21: 162–167. exceeds extracellular neurofibrillary tangle formation. Acta
Iwakiri, M., Mizukami, K., Ikonomovic, M.D., Ishikawa, M., Neuropathol. (Berl.), 103: 370–376.
Hidaka, S., Abrahamson, E.E., Dekosky, S.T. and Asada, T. Kulmala, H.K. (1985) Immunocytochemical localization of
(2005) Changes in hippocampal GABABR1 subunit enkephalin-like immunoreactivity in neurons of human
738

hippocampal formation: effects of ageing and Alzheimer’s the entorhinal cortex in Alzheimer’s disease. Neuroscience, 92:
disease. Neuropathol. Appl. Neurobiol., 11: 105–115. 515–532.
Kumar, U. (2005) Expression of somatostatin receptor subtypes Mishizen-Eberz, A.J., Rissman, R.A., Carter, T.L.,
(SSTR1-5) in Alzheimer’s disease brain: an immunohisto- Ikonomovic, M.D., Wolfe, B.B. and Armstrong, D.M.
chemical analysis. Neuroscience, 134: 525–538. (2004) Biochemical and molecular studies of NMDA recep-
Leuba, G., Kraftsik, R. and Saini, K. (1998) Quantitative dis- tor subunits NR1/2A/2B in hippocampal subregions
tribution of parvalbumin, calretinin, and calbindin D-28k throughout progression of Alzheimer’s disease pathology.
immunoreactive neurons in the visual cortex of normal and Neurobiol. Dis., 15: 80–92.
Alzheimer cases. Exp. Neurol., 152: 278–291. Mizukami, K., Grayson, D.R., Ikonomovic, M.D., Sheffield,
Lim, C., Blume, H.W., Madsen, J.R. and Saper, C.B. (1997) R. and Armstrong, D.M. (1998) GABAA receptor beta 2 and
Connections of the hippocampal formation in humans: I. The beta 3 subunits mRNA in the hippocampal formation of
mossy fiber pathway. J. Comp. Neurol., 385: 325–351. aged human brain with Alzheimer-related neuropathology.
Lotstra, F. and Vanderhaeghen, J.J. (1987) Distribution of Brain Res. Mol. Brain Res., 56: 268–272.
immunoreactive cholecystokinin in the human hippocampus. Mizukami, K., Ikonomovic, M.D., Grayson, D.R., Rubin,
Peptides, 8: 911–920. R.T., Warde, D., Sheffield, R., Hamilton, R.L., Davies, P.
Lozsadi, D.A. and Larner, A.J. (2006) Prevalence and causes of and Armstrong, D.M. (1997) Immunohistochemical study of
seizures at the time of diagnosis of probable Alzheimer’s GABA(A) receptor beta2/3 subunits in the hippocampal for-
disease. Dement. Geriatr. Cogn. Disord., 22: 121–124. mation of aged brains with Alzheimer-related neuropatho-
Magloczky, Z., Halasz, P., Vajda, J., Czirjak, S. and Freund, logic changes. Exp. Neurol., 147: 333–345.
T.F. (1997) Loss of Calbindin-D28K immunoreactivity from Morris, R.G.M. (2006) Elements of a neurobiological theory of
dentate granule cells in human temporal lobe epilepsy. hippocampal function: the role of synaptic plasticity, synap-
Neuroscience, 76: 377–385. tic tagging and schemas. Eur. J. Neurosci., 23: 2829–2846.
Magloczky, Z., Wittner, L., Borhegyi, Z., Halasz, P., Vajda, J., Morsch, R., Simon, W. and Coleman, P.D. (1999) Neurons
Czirjak, S. and Freund, T.F. (2000) Changes in the distribu- may live for decades with neurofibrillary tangles. J.
tion and connectivity of interneurons in the epileptic human Neuropathol. Exp. Neurol., 58: 188–197.
dentate gyrus. Neuroscience, 96: 7–25. Morys, J., Sadowski, M., Barcikowska, M., Maciejewska, B.
Maguire-Zeiss, K.A., Li, Z.W., Shimoda, L.M. and Hamill, and Narkiewicz, O. (1994) The second layer neurones of the
R.W. (1995) Calbindin D28k mRNA in hippocampus, supe- entorhinal cortex and the perforant path in physiological
rior temporal gyrus and cerebellum: comparison between ageing and Alzheimer’s disease. Acta Neurobiol. Exp.
control and Alzheimer disease subjects. Brain Res. Mol. (Wars.), 54: 47–53.
Brain Res., 30: 362–366. Nagerl, U.V., Mody, I., Jeub, M., Lie, A.A., Elger, C.E. and
Marksteiner, J., Lechner, T., Kaufmann, W.A., Gurka, P., Beck, H. (2000) Surviving granule cells of the sclerotic human
Humpel, C., Nowakowski, C., Maier, H. and Jellinger, K.A. hippocampus have reduced Ca(2+) influx because of a loss
(2000) Distribution of chromogranin B-like immunoreactiv- of calbindin-D(28k) in temporal lobe epilepsy. J. Neurosci.,
ity in the human hippocampus and its changes in Alzheimer’s 20: 1831–1836.
disease. Acta Neuropathol. (Berl.), 100: 205–212. Nakazawa, K., Quirk, M.C., Chitwood, R.A., Watanabe, M.,
Masliah, E., Mallory, M., Hansen, L., Deteresa, R., Alford, M. Yeckel, M.F., Sun, L.D., Kato, A., Carr, C.A., Johnston, D.,
and Terry, R. (1994) Synaptic and neuritic alterations during Wilson, M.A. and Tonegawa, S. (2002) Requirement for
the progression of Alzheimer’s disease. Neurosci. Lett., 174: hippocampal CA3 NMDA receptors in associative memory
67–72. recall. Science, 297: 211–218.
Mathieu-Kia, A.M., Fan, L.Q., Kreek, M.J., Simon, E.J. and van de Nes, J.A., Sandmann-Keil, D. and Braak, H. (2002)
Hiller, J.M. (2001) Mu-, delta- and kappa-opioid receptor Interstitial cells subjacent to the entorhinal region expressing
populations are differentially altered in distinct areas of somatostatin-28 immunoreactivity are susceptible to devel-
postmortem brains of Alzheimer’s disease patients. Brain opment of Alzheimer’s disease-related cytoskeletal changes.
Res., 893: 121–134. Acta Neuropathol. (Berl.), 104: 351–356.
McNamara, M.J., Ruff, C.T., Wasco, W., Tanzi, R.E., Nitsch, R. and Ohm, T.G. (1995) Calretinin immunoreactive
Thinakaran, G. and Hyman, B.T. (1998) Immunohistochem- structures in the human hippocampal formation. J. Comp.
ical and in situ analysis of amyloid precursor-like protein-1 Neurol., 360: 475–487.
and amyloid precursor-like protein-2 expression in Alzheimer Ohm, T.G., Müller, H., Braak, H. and Bohl, J. (1995) Close-
disease and aged control brains. Brain Res., 804: 45–51. meshed prevalence rates of different stages as a tool to un-
McNaughton, B.L., Barnes, C.A., Mizumori, S.J.Y., Green, cover the rate of Alzheimer’s disease-related neurofibrillary
E.J. and Sharp, P.E. (1991) In: Morell F. (Ed.), Kindling and changes. Neuroscience, 64: 209–217.
Synaptic Plasticity. Birkhäuser, Basel, pp. 110–123. Ohm, T.G., Münch, S., Schönheit, B., Zarski, R. and Nitsch, R.
Mikkonen, M., Alafuzoff, I., Tapiola, T., Soininen, H. and Mi- (2002) Transneuronally altered dendritic processing of tan-
ettinen, R. (1999) Subfield- and layer-specific changes in parv- gle-free neurons in Alzheimer’s disease. Acta Neuropathol.
albumin, calretinin and calbindin-D28K immunoreactivity in (Berl.), 103: 437–443.
739

Palop, J.J., Jones, B., Kekonius, L., Chin, J., Yu, G.Q., Raber, Scheff, S.W., Sparks, D.L. and Price, D.A. (1996) Quantitative
J., Masliah, E. and Mucke, L. (2003) Neuronal depletion of assessment of synaptic density in the outer molecular layer of
calcium-dependent proteins in the dentate gyrus is tightly the hippocampal dentate gyrus in Alzheimer’s disease.
linked to Alzheimer’s disease-related cognitive deficits. Proc. Dementia, 7: 226–232.
Natl. Acad. Sci. U.S.A., 100: 9572–9577. Schönheit, B., Zarski, R. and Ohm, T.G. (2004) Spatial and
Pellegrini-Giampietro, D.E., Bennett, M.V. and Zukin, R.S. temporal relationships between plaques and tangles in Al-
(1994) AMPA/kainate receptor gene expression in normal zheimer-pathology. Neurobiol. Aging, 25: 697–711.
and Alzheimer’s disease hippocampus. Neuroscience, 61: Seress, L. (1988) Interspecies comparison of the hippocampal
41–49. formation shows increased on the regio superior in the Am-
Penney, J.B., Maragos, W.F., Greenamyre, J., Debowey, D.L., mon’s horn of the human brain. J. Hirnforsch., 29: 35–340.
Hollingsworth, Z. and Young, A.B. (1990) Excitatory amino Seress, L. and Frotscher, M. (1990) Morphological variability is
acid binding sites in the hippocampal region of Alzheimer’s a characteristic feature of granule cells in the primate fascia
disease and other dementias. J. Neurol. Neurosurg. Psychi- dentata: a combined Golgi/electron microscope study. J.
atry, 53: 314–320. Comp. Neurol., 293: 253.
Petersen, R.C., Parisi, J.E., Dickson, D.W., Johnson, K.A., Seress, L., Gulyas, A.I., Ferrer, I., Tunon, T., Soriano, E. and
Knopman, D.S., Boeve, B.F., Jicha, G.A., Ivnik, R.J., Smith, Freund, T. (1993) Distribution, morphological features, and
G.E., Tangalso, E.G., Braak, H. and Kokmen, E. (2006) synaptic connections of parvalbumin- and calbindin D28k-
Neuropathologic features of amnestic mild cognitive impair- immunoreactive neurons in the human hippocampal forma-
ment. Arch. Neurol., 63: 645–646. tion. J. Comp. Neurol., 337: 208–230.
Price, J.L., Ko, A.I., Wade, M.J., Tsou, S.K., Mckeel, D.W. Seress, L. and Mrzljak, L. (1987) Basal dendrites of granule
and Morris, J.C. (2001) Neuron number in the entorhinal cells are normal features of the fetal and adult dentate gyrus
cortex and CA1 in preclinical Alzheimer disease. Arch. Ne- of both monkey and human hippocampal formations. Brain
urol., 58: 1395–1402. Res., 405: 169–174.
Rebeck, G.W., Marzloff, K. and Hyman, B.T. (1993) The pat- Simic, G., Kostovic, I., Winblad, B. and Bogdanovic, N. (1997)
tern of NADPH-diaphorase staining, a marker of nitric oxide Volume and number of neurons of the human hippocampal
synthase activity, is altered in the perforant pathway terminal formation in normal aging and Alzheimer’s disease. J. Comp.
zone in Alzheimer’s disease. Neurosci. Lett., 152: 165–168. Neurol., 379: 482–494.
Represa, A., Duyckaerts, C., Tremblay, E., Hauw, J.J. and Ben Singer, D., Lehmann, J., Hanisch, K., Hartig, W. and
Ari, Y. (1988) Is senile dementia of the Alzheimer type as- Hoffmann, R. (2006) Neighbored phosphorylation sites as
sociated with hippocampal plasticity? Brain Res., 457: PHF-tau specific markers in Alzheimer’s disease. Biochem.
355–359. Biophys. Res. Commun., 346: 819–828.
Rinne, J.O., Lonnberg, P., Marjamaki, P., Molsa, P., Sako, E. Sloviter, R.S., Sollas, A.L., Barbaro, N.M. and Laxer, K.D.
and Paljarvi, L. (1993) Brain methionine- and leucine- (1991) Calcium-binding protein (Calbindin-D28K) and parv-
enkephalin receptors in patients with dementia. Neurosci. albumin immunocytochemistry in the normal and epileptic
Lett., 161: 77–80. human hippocampus. J. Comp. Neurol., 308: 381–396.
Rissman, R.A., Mishizen-Eberz, A.J., Carter, .L., Wolfe, B.B., Solodkin, A., Veldhuizen, S.D. and van Hoesen, G.W. (1996)
De Blas, A.L., Miralles, C.P., Ikonomovic, M.D. and Arm- Contingent vulnerability of entorhinal parvalbumin-containing
strong, D.M. (2003) Biochemical analysis of GABA(A) re- neurons in Alzheimer’s disease. J. Neurosci., 16: 3311–3321.
ceptor subunits alpha 1, alpha 5, beta 1, beta 2 in the Spillantini, M.G. and Goedert, M. (1998) Tau protein pathology
hippocampus of patients with Alzheimer’s disease neuropa- in neurodegenerative diseases. Trends Neurosci., 21: 428–433.
thology. Neuroscience, 120: 695–704. Stoub, T.R., Detoledo-Morrell, L., Stebbins, G.T., Leurgans,
Sampson, V.L., Morrison, J.H. and Vickers, J.C. (1997) The S., Bennett, D.A. and Shah, R.C. (2006) Hippocampal dis-
cellular basis for the relative resistance of parvalbumin and connection contributes to memory dysfunction in individuals
calretinin immunoreactive neocortical neurons to the pathol- at risk for Alzheimer’s disease. Proc. Natl. Acad. Sci. U.S.A.,
ogy of Alzheimer’s disease. Exp. Neurol., 145: 295–302. 103: 10041–10045.
Satoh, J., Tabira, T., Sano, M., Nakayama, H. and Tateishi, J. Thal, D.R., Holzer, M., Rub, U., Waldmann, G., Gunzel, S.,
(1991) Parvalbumin-immunoreactive neurons in the human Zedlick, D. and Schober, R. (2000) Alzheimer-related tau-
central nervous system are decreased in Alzheimer’s disease. pathology in the perforant path target zone and in the hip-
Acta Neuropathol. (Berl.), 81: 388–395. pocampal stratum oriens and radiatum correlates with onset
Scheff, S.W. and Price, D.A. (1998) Synaptic density in the and degree of dementia. Exp. Neurol., 163: 98–110.
inner molecular layer of the hippocampal dentate gyrus in Thal, D.R., Rub, U., Orantes, M. and Braak, H. (2002) Phases
Alzheimer disease. J. Neuropathol. Exp. Neurol., 57: of A beta-deposition in the human brain and its relevance for
1146–1153. the development of AD. Neurology, 58: 1791–1800.
Scheff, S.W. and Price, D.A. (2003) Synaptic pathology in Al- Torrealba, F. and Carrasco, M.A. (2004) A review on electron
zheimer’s disease: a review of ultrastructural studies. Ne- microscopy and neurotransmitter systems. Brain Res. Brain
urobiol. Aging, 24: 1029–1046. Res. Rev., 47: 5–17.
740

Uchihara, T., Nakamura, A., Yamazaki, M. and Mori, O. West, M.J. and Gundersen, H.J.G. (1990) Unbiased stereolog-
(2001) Evolution from pretangle neurons to neurofibrillary ical estimation of the number of neurons in the human
tangles monitored by thiazin red combined with Gallyas hippocampus. J. Comp. Neurol., 296: 1–22.
method and double immunofluorescence. Acta Neuropathol. West, M.J., Kawas, C.H., Stewart, W.F., Rudow, G.L. and
(Berl.), 101: 535–539. Troncoso, J.C. (2004) Hippocampal neurons in pre-clinical
Ulas, J. and Cotman, C.W. (1997) Decreased expression of Alzheimer’s disease. Neurobiol. Aging, 25: 1205–1212.
N-methyl-D-aspartate receptor 1 messenger RNA in select West, M.J. and Slomianka, L. (1998) Total number of neurons
regions of Alzheimer brain. Neuroscience, 79: 973–982. in the layers of the human entorhinal cortex. Hippocampus,
Uylings, H.B.M., Ruiz-Marcos, A. and van Pelt, J. (1986) The 8: 69–82.
metric analysis of three-dimensional dendritic tree patterns: a Williams, R.S. and Matthysse, S. (1986) Age-related changes
methodological review. J. Neurosci. Methods, 18: 127–151. in Down syndrome brain and the cellular pathology of
Wakabayashi, K., Hansen, L.A., Vincent, I., Mallory, M. and Alzheimer disease. Prog. Brain Res., 70: 49–67.
Masliah, E. (1997) Neurofibrillary tangles in the dentate Yao, P.J., Morsch, R., Callahan, L.M. and Coleman, P.D.
granule cells of patients with Alzheimer’s disease, Lewy body (1999) Changes in synaptic expression of clathrin assembly
disease and progressive supranuclear palsy. Acta protein AP180 in Alzheimer’s disease analysed by
Neuropathol. (Berl.), 93: 7–12. immunohistochemistry. Neuroscience, 94: 389–394.
Wakabayashi, K., Narisawa-Saito, M., Iwakura, Y., Arai, T., Yew, D.T., Li, W.P., Webb, S.E., Lai, H.W. and Zhang, L.
Ikeda, K., Takahashi, H. and Nawa, H. (1999) Phenotypic (1999) Neurotransmitters, peptides, and neural cell adhesion
down-regulation of glutamate receptor subunit GluR1 in molecules in the cortices of normal elderly humans and Al-
Alzheimer’s disease. Neurobiol. Aging, 20: 287–295. zheimer patients: a comparison. Exp. Gerontol., 34: 117–133.
West, M.J., Coleman, P.D., Flood, D.G. and Troncoso, J.C. Zipp, F., Nitsch, R., Soriano, E. and Frotscher, M. (1989)
(1994) Differences in the pattern of hippocampal neuronal Entorhinal fibers form synaptic contacts on parvalbumin-
loss in normal ageing and Alzheimer’s disease. Lancet, 344: immunoreactive neurons in the rat fascia dentata. Brain Res.,
769–772. 495: 161–166.
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 40

Hippocampal atrophy and disconnection in incipient


and mild Alzheimer’s disease

Leyla deToledo-Morrell, Travis R. Stoub and Changsheng Wang

Department of Neurological Sciences, Rush University Medical Center, Chicago, IL 60612, USA

Abstract: Quantitative imaging techniques allow the in vivo investigation of age and disease related
changes in the brain and their relation to cognitive function. In this chapter we review imaging evidence
indicating that the entorhinal cortex and hippocampus show atrophy very early in Alzheimer’s disease
(AD) and in individuals who are at risk of developing AD compared to age appropriate controls. Fur-
thermore, the extent and rate of atrophy of the entorhinal cortex, a brain region pathologically involved
very early in the disease process, can predict who among the elderly will develop AD. Techniques that
assess the integrity of white matter further demonstrate that alterations in the parahippocampal white
matter in the region that includes the perforant path could partially disconnect the dentate gyrus and other
hippocampal subfields from incoming sensory information. Such partial disconnection and degradation in
transmission of sensory information in people at risk of AD and in patients with very mild AD could
contribute to the memory dysfunction associated with the early stages of the disease.

Keywords: entorhinal cortex; perforant path; imaging; aging; memory

Introduction The hippocampal dentate gyrus has been impli-


cated as the sub-region most sensitive to the effects
High-resolution structural magnetic resonance im- of advancing age (Small et al., 2004). Although
aging (MRI) techniques provide a unique tool for synaptic loss in the dentate molecular layer has
examining alterations in brain anatomy in vivo been demonstrated in patients with Alzheimer’s
during healthy aging and in age-related diseases. disease (AD) (Scheff and Price, 2003; Scheff et al.,
In addition, such techniques allow us to (a) exam- 2006), cell loss associated with the disease seems
ine the relation between alterations in given to be more prominent in the CA1 region of the
brain regions and the sequential development of hippocampus (West et al., 1994, 2000). In addi-
behavioral symptoms in degenerative diseases, and tion, AD-related pathological changes such as
(b) delineate the specific role of certain mesial neurofibrillary tangles are most notable in the
temporal lobe structures, such as the entorhinal subiculum and the CA1 region (Van Hoesen and
cortex and hippocampus in human memory Hyman, 1990). Unfortunately, until recently, in
function, because of the age or disease-related vivo structural imaging techniques did not have
degeneration in those structures. the resolution to differentiate sub-regions of the
hippocampal formation (hippocampus proper and
Corresponding author. Tel.: +1 (312) 942 5399; dentate). However, a recently published structural
Fax: +1 (312) 563 3570; E-mail: ldetoled@rush.edu MRI protocol acquired with a high-field (4 Tesla)

DOI: 10.1016/S0079-6123(07)63040-4 741


742

magnet holds promise for differentiating hippo- brain. At the level of the foramen magnum, a
campal sub-fields in vivo in future investigations straight line was drawn from the inner surface of
(Mueller et al., 2007). the clivus to the occipital bone.
In this chapter, we will review the evidence Entorhinal cortex volume was quantified with
indicating that there is atrophy of both the the use of a protocol developed and validated
entorhinal cortex and hippocampus in the very in our laboratory, technical details of which
early stages of AD, as well as in individuals with are presented in Goncharova et al. (2001). The
amnestic mild cognitive impairment (MCI) who advantage of this protocol is that entorhinal vol-
are at high risk of developing AD (Petersen et al., ume is measured from the same oblique coronal
1999; Petersen, 2000). These elderly individuals sections most commonly used for hippocampal
have impairment in memory function, but do not volumetry, so that one of these two adjacent
meet criteria for dementia. Furthermore, we will structures is not over estimated at the expense of
demonstrate that white matter changes in the the other.
region of the parahippocampal gyrus that includes Briefly, both entorhinal and hippocampal
the perforant path in people with amnestic MCI volumes were computed separately for the right
and mild AD could exacerbate the memory dys- and left hemispheres from coronal slices reformat-
function characteristic of the initial stages of AD, ted to be perpendicular to the long axis of the
by partially disconnecting the hippocampus from hippocampus. For the entorhinal cortex, tracing
incoming sensory information. began with the first section in which the gyrus
Since the MRI techniques in the experiments ambiens, amygdala and the white matter of the
from our laboratory to be described were similar, parahippocampal gyrus first appeared visible.
they will be detailed first. The dorsomedial border in rostral sections was
the sulcus semiannularis and in caudal sections the
subiculum. The shoulder of the collateral sulcus
Acquisition and quantitation of MRI data was used as the lateral border, a somewhat
conservative criterion that allowed consistency in
All MR images were acquired on a 1.5 Tesla Gen- tracings and avoided the use of different lateral
eral Electric Signa scanner with the manufacturer’s borders, depending on individual differences in the
three-dimensional (3D) Fourier transform spoiled depth of the collateral sulcus (Insausti et al., 1998).
gradient recalled (SPGR) pulse sequence. The ac- The last section traced was three 1.6 mm slices
quisition parameters for the T1 weighted sequence rostral to the image in which the lateral geniculate
were as follows: 124 contiguous images acquired in first appeared visible.
the coronal plane, 1.6 mm-thick sections, matrix ¼ The protocol and validation procedures used for
256  192, field of view ¼ 22 cm, TR/TE ¼ 34/7, quantifying hippocampal volume were published
flip angle ¼ 351, signals averaged ¼ 1. previously (Wilson et al., 1996; deToledo-Morrell
The analyze software package (Mayo Clinic et al., 1997). For hippocampal volumetry, tracings
Foundation, Rochester, MN, USA) was used for started with the first section where the hippocam-
determining the volume of regions of interest and pus could be clearly differentiated from the
for co-registering sequential scans. Both the amygdala by the alveus and included the fimbria,
entorhinal cortex and hippocampal formation the dentate gyrus, the hippocampus proper and the
were manually segmented as described below. subiculum. Tracings continued on all consecutive
To correct for individual differences in brain size, slices until the slice before the full appearance of
entorhinal and hippocampal volumes were divided the fornix. Investigators involved in MRI analyses
by total intracranial volume (an accepted measure were blinded to clinical information until all such
of pre-morbid brain size) derived from sagittaly analyses were completed.
reformatted 5 mm slices. To compute intracranial Figure 1 shows sample tracings of the entorhinal
volume, the inner table of the cranium was traced cortex and hippocampal formation. Both regions
in consecutive sagittal sections spanning the whole of interest were normalized using the following
743

Fig. 1. A sample coronal image illustrating the segmentation of the entorhinal cortex (outlined, right side) and hippocampus (outlined,
left side).

formula: (absolute volume in mm3/intracranial entorhinal and hippocampal volumes in healthy


volume in mm3)  1000. elderly controls with no cognitive impairment, in
people who presented at the clinic with cognitive
Hippocampal and entorhinal cortex atrophy in complaints, but who did not meet criteria for
incipient and mild AD dementia and in patients with very mild AD.
Interestingly, the two patient groups differed sig-
The entorhinal cortex and hippocampus are part nificantly from controls in entorhinal volume, but
of the mesial temporal lobe memory system not from each other; in contrast, they differed
(Squire and Zola-Morgan, 1991; Young et al., from each other as well as from controls in hip-
1997). The entorhinal cortex receives sensory input pocampal volume, with the very mild AD cases
from the neocortex and provides the hippocampus showing the greatest atrophy.
with multi-modal sensory information via the per- Recently, interest has increased in directly
forant path. These brain regions have received comparing the accuracy with which the volumes
special attention in both post mortem and in vivo of these two mesial temporal lobe structures
investigations on the pathophysiology of AD, can differentiate those who convert from a non-
since memory dysfunction is one of the earliest demented status to AD (i.e., those who are at risk
hallmarks of the disease. of developing AD). The study described below was
Post mortem pathological studies have impli- undertaken to examine this question using an
cated the entorhinal cortex and the transentorhinal initial baseline MRI in a group of participants
region as sites that are involved in the early stages diagnosed with amnestic MCI (deToledo-Morrell
of AD and in individuals with MCI (Braak and et al., 2004).
Braak, 1991, 1995; Gomez-Isla et al., 1996; Braak The participants consisted of 27 elderly individ-
et al., 1998; Kordower et al., 2001). In vivo MRI uals (Z65 years of age) who received a clinical
investigations have also demonstrated atrophy of diagnosis of amnestic MCI during their baseline
both the entorhinal cortex and hippocampus not evaluation. They had a significant deficit in the
only in patients diagnosed with mild AD, but also memory domain, but did not meet criteria for de-
in individuals with amnestic MCI and those with mentia. After the baseline diagnosis of MCI, all
subjective cognitive complaints (Convit et al., participants were followed with yearly clinical
1997; Jack et al., 1997, 1999; de Leon et al., evaluations. During a 36-month follow-up period,
1996; deToledo-Morrell et al., 1997, 2000a, b; 10 of the 27 participants originally diagnosed with
Killiany et al., 2000, 2002; Xu et al., 2000; amnestic MCI converted to AD.
Dickerson et al., 2001; Du et al., 2001; Jessen All clinical evaluations were carried out at
et al., 2006; Saykin et al., 2006). the Rush Alzheimer’s Disease Center (RADC,
In an earlier study from our laboratory (Di- Chicago, IL) as previously described (Wilson et al.,
ckerson et al., 2001) we compared MRI-derived 1996; deToledo-Morrell et al., 1997). Briefly, the
744

evaluation incorporated the Consortium to Estab- predict conversion to AD. In these analyses, total
lish a Registry for Alzheimer’s Disease (CERAD, normalized entorhinal and hippocampal volumes
Morris et al., 1989) procedures and included a were used as predictors to determine the extent to
medical history, neurological examination, neuro- which each region of interest contributed sepa-
psychological testing, informant interview and rately to describing group differences. The results
blood tests. demonstrated that although both total entorhinal
Exclusion criteria for entry into the study con- and total hippocampal volumes were independent
sisted of evidence of other neurologic, psychiatric predictors of conversion [w2(1) ¼ 7.3, p ¼ 0.007,
or systemic conditions that could cause cognitive odds ratio per 0.1 unit volume change ¼ 2.024 for
impairment (e.g., stroke, alcoholism, major de- the entorhinal cortex and w2(1) ¼ 6.2, p ¼ 0.013,
pression, a history of temporal lobe seizures). The odds ratio ¼ 1.318 for the hippocampus], ent-
clinical diagnosis of probable AD in the longitu- orhinal volume was a better predictor with a con-
dinal clinical evaluations followed NINCDS/AD- cordance rate of 90.6%. Thus, for every 0.1 unit
RDA guidelines (McKhann et al., 1984); it decrease in entorhinal cortex volume, the odds of
required a history of cognitive decline and neuro- conversion doubled, while the same unit decrease
psychological test evidence of impairment in at in hippocampal volume increased the chances of
least two cognitive domains, one of which had to conversion by only 30%.
be memory. Similar analyses were carried out on entorhinal
Demographic data and Mini Mental State and hippocampal volumes separated by hemi-
Examination (MMSE, Folstein et al., 1975) scores sphere. In this case, the right hemisphere ent-
for the MCI-converters and non-converters are orhinal volume best predicted conversion with a
presented in Table 1. The maximum score on the concordance rate of 93.5%. Unlike these findings,
MMSE is 30, with scores Z27 being considered hemispheric differences in hippocampal volume
within the normal range. The individuals who did not lead to differential predictions regarding
developed AD (converters) did not differ from the conversion.
non-converters in age, but there was a significant The in vivo anatomical results presented here
difference between them in MMSE scores and in are in agreement with post mortem pathological
years of education. findings and underscore the early involvement of
Total (right+left) normalized entorhinal and the entorhinal cortex in AD. Taken together,
hippocampal volumes derived from the baseline the results just described indicate that entorhinal
MRI for MCI converters and non-converters volume is a good predictor of conversion from
are shown in Fig. 2. Converters differed signifi- MCI to AD. In agreement with these results,
cantly from non-converters in total entorhinal others have also reported that baseline entorhinal
[t(25) ¼ 3.090, p ¼ 0.005] and hippocampal volume differentiated those at risk for developing
[t(25) ¼ 2.442, p ¼ 0.022] volume. Multivariate lo- AD better than hippocampal volume (Killiany
gistic regression analyses adjusted for education et al., 2002).
were carried out to examine how well baseline Our results also demonstrated that the right
entorhinal and hippocampal volumes could hemisphere entorhinal volume was the best

Table 1. Demographic characteristics of participants

MCI converters (N ¼ 10) MCI non-converters (N ¼ 17)

Age (years) 82.7 (74.5) 81.1 (78.1)


(75.6–89.2) (66.1–98.0)
Education (years) 18.4 (72.1) 15.2 (73.1)
Female/male 6/4 9/8
MMSE score 26.1 (71.4) 28.0 (71.8)
 Significantly differently from each other (po0.01).
745

Fig. 2. Box plot comparing MRI-derived normalized entorhinal (left hand side) and hippocampal (right hand side) volumes in
participants with MCI who converted to a diagnosis of Alzheimer’s disease, in contrast to those who did not. The central box shows
the data between the upper and lower quartiles, with the median represented by the line. The height of the line is the interquartile range
(IQR); the ‘‘whiskers’’ extend from the upper and lower quartiles to a distance of 1.5 IQR away or to the most extreme data point
within that range, whichever is closer (adapted with permission from deToledo-Morrell et al., 2004).

predictor of risk of AD. Very similar findings have development of AD. Similarly, Rodrigue and Raz
been recently reported by another laboratory (2004) found that the rate of entorhinal atrophy in
(Tapiola et al., 2006). The reason for this hemi- healthy elderly people was associated with greater
spheric difference in entorhinal volume in those memory decline.
who are at risk of developing AD is not clear since,
in general, post mortem pathological investigations
do not carry out hemispheric comparisons. MRI Predictors of risk of incident AD: a
It should be pointed out that changes in the longitudinal study
entorhinal cortex in elderly individuals seem to be
predictive of conversion not only from MCI to As discussed above, a large number of cross-sec-
AD, but also of the transition from normal cog- tional MRI investigations have demonstrated at-
nition to MCI or AD. In an interesting report, de rophy of both the hippocampus and entorhinal
Leon and his colleagues (de Leon et al., 2001) used cortex in patients with mild to moderate AD and
MRI guided resting state [2-18F]-2-deoxy-D-glu- in people with cognitive impairment who are at
cose/positron-emission tomography to assess met- risk of developing AD. Longitudinal studies that
abolic changes in given brain regions of interest. measured the volumes of these two mesial tempo-
They found that a reduction in entorhinal cortex ral lobe structures have shown the annual rate of
metabolism was predictive of cognitive decline atrophy of both structures to be greater in patients
among cognitively normal elderly subjects. Al- with AD and in those with MCI compared to eld-
though obtained with a different imaging modal- erly control subjects (Jack et al., 1998, 2000, 2004;
ity, these results are very much in line with our Cardenas et al., 2003; Du et al., 2003, 2004).
findings and provide further evidence for the very However, most of these longitudinal investigations
early involvement of the entorhinal cortex in the used scans from two-time points separated by 1–5
746

years for determining rate of atrophy, rather than significant predictor of event time, all models were
multiple yearly scans. education adjusted.
The experiment from our laboratory described The results of the proportional odds models
above examined MRI markers of AD using a demonstrated that initial diagnosis of MCI was an
baseline scan and longitudinal clinical evaluations. important predictor of incident AD [w2(1) ¼
In the next study, we used proportional odds 16.569, po0.0001; Fig. 3A]. Furthermore, both
models to assess the relationship between rate of baseline entorhinal volume and its rate of atrophy
atrophy of mesial temporal lobe structures were independent predictors of incident AD; their
and risk of incident AD among non-demented addition to initial diagnosis improved the model
elderly individuals (Stoub et al., 2005). Rate of [w2(2) ¼ 10.904, po0.0043]. Lower baseline ent-
atrophy of regions of interest was derived from orhinal volume was associated with greater risk of
yearly scans. AD (see Fig. 3B); for every 0.25 unit of decrease in
Data reported here were obtained from 58 non- baseline entorhinal volume, the odds of incident
demented elderly participants (age 65 or older) AD increased by a factor of 4.98. Similarly, in-
who were followed with annual clinical evaluations creased rate of decline in entorhinal volume was
as well as high-resolution MRI scans for up to 5 associated with greater risk of AD (see Fig. 3C);
years (baseline and 5 years of follow-up). The for every 0.02 unit of increase in rate of entorhinal
clinical evaluations and MRI acquisition and anal- atrophy, the likelihood of developing AD in-
ysis protocols were as described above. Twenty- creased by a factor of 2.18.
three of the 58 participants received a diagnosis of Surprisingly, hippocampal volume was not an
amnestic MCI during the baseline evaluation independent predictor of risk of AD; the addition
and 35 were healthy controls with no cognitive of baseline hippocampal volume and its slope of
impairment. decline to initial diagnosis did not improve the
The number of years until a diagnosis of AD model based on initial diagnosis alone
was modeled using proportional odds models [w2(2) ¼ 2.701, p ¼ 0.26].
(Kalbfleish and Prentice, 1980). For this purpose, APOE e4 allele status, a risk factor for AD, was
AD was treated as an all absorbing state, with available for 53 of 58 participants (see Table 2).
onset at the first evaluation at which NINCDS/ When the analyses described above were carried
ADRDA criteria (McKhann et al., 1984) were out on data from these 53 participants, results
met. For participants who did not develop AD, the
number of years to last clinical diagnosis was the Table 2. Baseline demographic characteristics of participants
event time, which was coded as censored. Covari-
Non-converters Converters to AD
ates considered were demographic variables (age,
sex, years of education), baseline diagnosis (MCI N 44 14
or control), and volumetric measures (normalized Sex
total hippocampal and entorhinal volumes). Male 16 4
Female 28 10
Volumes of regions of interest derived from scans Age (years) 80 (76) 81 (76)
following a diagnosis of AD were not included in MMSE 28.5 (71.5) 26.9 (71.8)
the analyses (for further analytic details, see Stoub Education (years) 15 (73) 18 (73)
et al., 2005). APOE status
Fourteen of the 58 non-demented participants 4/4 0 0
3/4 8 5
developed AD during the follow-up period; three 3/3 25 6
of these 14 participants entered the study as con- 2/4 0 2
trol subjects with no cognitive impairment and the 2/3 7 0
rest with a diagnosis of amnestic MCI. Demo- 2/2 0 0
graphic characteristics for participants who did Not available 4 1
and did not develop AD are presented in Table 2. Significantly different p ¼ 0.002.
 Significantly different p ¼ 0.019.
Since education, but not age, was found to be a
747

remained similar to those already detailed. How- models based on initial diagnosis and entorhinal
ever, APOE e4 allele status (i.e., the presence or and hippocampal volume [w2(2) ¼ 1.972, p ¼ 0.16],
absence of any e4 allele) was not found to be an although the presence or absence of any APOE e4
important predictor of incident AD when added to allele by itself was a predictor of event time
[w2(1) ¼ 5.00, p ¼ 0.025]. These results are in line
with those from a prior report (Bennett et al.,
2003) which demonstrated that the effect of the e4
allele, which was strongly associated with the like-
lihood of clinical AD, was no longer significant
after controlling for the effects of AD pathology,
suggesting that such pathology mediates the effect
of allele status on clinical disease. In our experi-
ments, MRI-derived hippocampal and entorhinal
atrophy could be viewed as surrogate markers of
the underlying pathology.
The major finding of this study is that among
non-demented individuals, the initial size of the
entorhinal cortex and its rate of atrophy are
significant risk factors for AD, with initial small
size and steep declines in volume being associated
with incident AD. In contrast, baseline size and
rate of hippocampal atrophy were not found to be
independent risk factors. Other investigators have
reported that hippocampal atrophy at baseline in
patients with MCI is associated with risk of AD
(Jack et al., 1999). Although our results may,
at first sight, seem contrary to this report, it is
important to point out that in this paper the in-
vestigators assessed only MRI-based hippocampal
atrophy and did not directly compare hippocam-
pal and entorhinal volumes in their models.

Fig. 3. Estimates of the probability of not yet having been di-


agnosed with Alzheimer’s disease as a function of years of fol-
low-up for two different populations. The dotted line represents
the reference population and the solid line depicts the hypo-
thetical population. The reference population in all panels con-
sists of people who have no cognitive impairment (NCI) at
baseline, have an average education level of 16 years, whose
normalized entorhinal cortex volume is 1.00 (chosen to be close
to the overall mean of 0.97027 at baseline), and for whom the
slope of the rate of entorhinal atrophy is 0.03 (which is close
to the mean slope of 0.03074). For the model shown in A, the
hypothetical population is different from the reference only in
that the initial diagnosis is MCI, not NCI. In B, the hypothet-
ical population is different from the reference in terms of base-
line entorhinal cortex volume, which was 25% smaller than 1 or
0.75. The hypothetical population shown in C is different from
the reference population only in that entorhinal cortex volume
declines faster with a slope of 0.05 rather than 0.03 (adapted
with permission from Stoub et al., 2005).
748

As discussed earlier in this chapter, post mortem in investigations on the pathophysiology of incip-
studies have suggested that AD-related pathology ient or clinically diagnosed AD and the anatomical
may start in the transentorhinal and entorhinal underpinnings of the memory dysfunction associ-
regions and then spread to the hippocampus ated with the disease.
(Braak and Braak, 1991, 1995; Braak et al., Up to this point we have concentrated on a
1998). The fact that shrinkage of the entorhinal description of structural alterations detected in
cortex, but not of the hippocampus was a predic- vivo in the entorhinal cortex and hippocampus. In
tor of AD in our study is in support of these post the next experiment (Stoub et al., 2006) we turned
mortem findings and suggests that the extent of our attention to assessing the contribution of al-
entorhinal involvement in the disease process may terations in the parahippocampal white matter, in
precede hippocampal involvement. addition to entorhinal and hippocampal atrophy,
to the memory dysfunction in people with amnes-
tic MCI who are at risk of developing AD.
Hippocampal disconnection due to parahippocampal The participants in the study consisted of 40
white matter changes in incipient AD: relation to older individuals (mean age 77.977.5 years) who
memory dysfunction met criteria for amnestic MCI and 50 healthy con-
trols (mean age 78.176.0 years) with no cognitive
As discussed up to this point, previous histopath- impairment. Clinical evaluations and acquision of
ological findings (Hyman et al., 1984; Braak and MRI scans were as described before. White matter
Braak, 1991, 1995; Gomez-Isla et al., 1996; Braak volume changes throughout the brains of individ-
et al., 1998; Mufson et al., 1999; Kordower et al., uals with amnestic MCI, compared to age-
2001), as well as in vivo MRI results indicate that matched healthy controls were assessed with vox-
the entorhinal cortex and hippocampus are path- el-based morphometry (VBM, Good et al., 2001).
ologically involved very early in patients with AD Automated techniques such as VBM that provide
and in those with MCI who are at high risk for quick, unbiased means of comparing changes in
developing AD. More specifically, two studies white matter at the voxel level are especially useful
(Gomez-Isla et al., 1996; Kordower et al., 2001) when there is no a priori knowledge of where in the
found a loss of entorhinal cortex layer II neurons brain such white matter changes may be taking
in patients with AD and in those with MCI, com- place. In addition to white matter changes, we as-
pared to elderly controls. These neurons receive sessed differences between the groups in entorhinal
multimodal sensory input from primary sensory cortex and hippocampal volume using manual
and association cortices and project this informa- segmentation and volumetric measurement of the
tion to the hippocampus via the perforant path, two structures since, in our hands, these measures
a white matter tract located in the anterior medial are better at detecting subtle changes in small gray
portion of the parahippocampal gyrus (Van matter regions.
Hoesen and Pandya, 1975; Van Hoesen et al., Group differences in whole-brain white matter
1975; Amaral et al., 1987). volumes were assessed with the two-sample t sta-
The loss of layer II neurons in the entorhinal tistic within the Statistical Parametric Mapping
cortex and their axons is important since it would software (SPM 99); the significance threshold was
result in alterations in information flow to the set at 0.001. Regions of statistically significant
hippocampus. In addition, damage to the white difference were identified according to the Mon-
matter of the parahippocampal gyrus could dis- treal Neurological Institute template and then
rupt afferent connections to the entorhinal cortex converted to Talariach coordinates (Talairach
and ultimately disconnect multi-modal sensory in- and Turnoux, 1988) with the use of MNITOTAL
put to the hippocampus, information vital to the software. The Talairach coordinates were trans-
formation of new memories. These changes in formed to actual brain areas by using the Talai-
white matter may be detectable in vivo with im- rach Daemon system (Lancaster et al., 2000).
aging techniques, but have received less attention These coordinates were used to construct regions
749

of interest that were then applied to individual In addition to the parahippocampal white matter
white matter density maps to extract individual change, the group of MCI participants also
volume values. showed significant entorhinal and hippocampal
In order to relate structural changes to memory atrophy (po0.001 for both) similar to what our
function, individual memory Z scores were calcu- laboratory and others have reported. The atrophy
lated based on combined performance on declar- in both structures was bilateral.
ative memory tasks using seven declarative For predicting memory function from MRI
memory scores. These consisted of immediate measures, each of the three MRI indices (i.e., total
and delayed recall of the East Boston Story (Al- parahippocampal white matter volume, entorhinal
bert et al., 1991) and of Story A from the Logical cortex volume and hippocampal volume) was en-
Memory of the Wechsler memory scale — Revised tered singly into a regression analysis that used as
(Wechsler, 1987). An additional test involved the the dependent variable individual memory Z
learning and retention of a 10-word list from the scores. Each of the three anatomical measures
CERAD battery (Morris et al., 1989). The three was found to be a significant predictor of memory
scores for this test included Word List Memory function [F(1,88) ¼ 24.08, po0.001 for hippocam-
(the total number of words immediately recalled pal volume; F(1,88) ¼ 10.35, po0.002 for
after each of three consecutive presentations of the entorhinal cortex volume; and F(1,88) ¼ 11.11,
list), Word List Recall (the number of words re- po0.001 for parahippocampal white matter
called after a delay) and Word List Recognition volume]. However, when all three MRI measures
(the number of words correctly recognized in a were entered simultaneously into a multiple
four-alternative, forced-choice format, adminis- regression model, only hippocampal and white
tered after Word List Recall). matter volume measures were found to be signifi-
The demographic characteristics of all partici- cant predictors of memory function [t(86) ¼ 3.01,
pants are listed in Table 3. The amnestic MCI p ¼ 0.003 and t(86) ¼ 2.22, po0.029 respectively].
group did not differ from healthy controls in age The major contribution of these data is that,
or education, but had significantly lower [t(88) ¼ in addition to a reduction in the size of the hip-
6.479, po0.001] MMSE scores and memory Z pocampus, a decrease in white matter volume in
scores [t(88) ¼ 9.921, po0.001], which is not the region of the parahippocampal gyrus that
surprising. includes the perforant path significantly contrib-
As shown in Fig. 4, there was a significant de- utes to the memory dysfunction in people with
crease (po0.001) in white matter volume in par- MCI. The perforant path that supplies the hip-
ticipants with MCI compared to controls in the pocampus with multi-modal information is path-
anterior-medial aspect of both parahippocampal ologically involved very early in AD (Hyman et
gyri, in the region that includes the perforant path. al., 1984). The investigation by Hyman and his
collaborators was carried out on post mortem
tissue and demonstrated that in patients with
Table 3. Demographic characteristics and memory Z scores of very mild AD, there is disconnection of the hip-
participants
pocampus from sensory cortical inputs as a re-
Healthy MCI sult of loss of layer II cells in the entorhinal
controls cortex. The authors hypothesized that such
disconnection could contribute to memory dys-
Number of participants 50 40
Age (years) 78.1 (76.0) 77.9 (77.5)
function during the very early stages of AD;
Education (years) 15.0 (72.8) 16.2 (73.1) however, the report did not include cognitive
Male/female 15/35 16/24 data proximal to death. Our results provide an
MMSE score 29.0 (70.9) 27.2 (71.6) in vivo demonstration of hippocampal discon-
Memory Z score 0.5337 0.5948 nection resulting from white matter changes in
(70.4681) (70.6111)
the region of the perforant path that impacts
 po0.001.
memory function.
750

Fig. 4. Color map showing significant (p ¼ 0.001) voxels of decreased white matter density in participants with amnestic MCI
compared with controls, superimposed on coronal and sagittal slices of a template based on data for all subjects in the study. The
image is masked to include white matter regions. The colors correspond to the t values shown on the color bar. Note the bilaterality of
significant differences in the white matter of the parahippocampal gyrus (adapted with permission from Stoub et al., 2006). (See Color
Plate 40.4 in color plate section.)

White matter volume change may reflect not normal, as indicated by lower FA values com-
only loss of afferent and efferent fibers in the re- pared to age appropriate controls.
gion of the parahippocampal gyrus, but also par-
tial demyelination in remaining fibers. Recently
developed diffusion tensor imaging (DTI) that de- Conclusions
tects microstructural alterations in normal appear-
ing white matter could aid in determining the In this chapter, we reviewed evidence from in vivo
microstructural changes in remaining fibers that MRI investigations that demonstrate significant
may further degrade impulse transmission. DTI entorhinal cortex and hippocampal atrophy not
detects microstructural alterations in white matter only in patients with very mild AD, but also in
by measuring the directionality of molecular diffu- individuals who are at risk of developing AD.
sion (fractional anisotropy, FA). Highly organized Atrophy in these mesial temporal lobe structures
and well myelinated white matter tracts have high important for memory function can predict who
FA because diffusion is constrained by the tract’s will develop AD a number of years prior to a
cellular organization. As white matter is damaged, clinical diagnosis. Recently, there has been in-
FA decreases due to decreased anisotropic diffu- creased interest in investigating alterations in white
sion. In fact, in the most recent study from our matter that could lead to degradation in informa-
laboratory (Wang et al., 2006), using a high res- tion flow due to partial disconnection of different
olution DTI protocol, we demonstrated that in brain regions. We have shown that white matter
very mild AD cases there is not only loss of white volume loss in the region of the parahippocampal
matter volume in the region of the parahippocam- gyrus contributes significantly to memory dysfunc-
pal gyrus, but that remaining fibers are not tion in individuals with incipient AD. The exact
751

underlying mechanism of the white matter volume brain change in cognitive impairment and dementia. Ne-
loss in the parahippocampal region cannot be de- urobiol. Aging, 24(4): 537–544.
Convit, A., de Leon, M.J., Tarshish, C., DeSanti, S., Tsui, W.,
termined in vivo with the tools currently available
Rusinek, H. and George, A. (1997) Specific hippocampal
to us. One may speculate that a decrease in white volume reductions in individuals at risk for Alzheimer’s
matter fibers in this region, reflected in white mat- disease. Neurobiol. Aging, 18(2): 131–138.
ter volume change, may cause a disruption of in- Dickerson, B.C., Goncharova, I., Sullivan, M.P., Forchetti, C.,
put to the dentate gyrus and hippocampus proper Wilson, R.S., Bennett, D.A., Beckett, L.A. and deToledo-
from the entorhinal cortex. Because these incom- Morrell, L. (2001) MRI-derived entorhinal and hippocampal
atrophy in incipient and very early Alzheimer’s disease.
ing fibers arise from cells in the entorhinal cortex, Neurobiol. Aging, 22(5): 747–754.
entorhinal atrophy may, in part, be the origin of Du, A.T., Schuff, N., Amend, D., Laakso, M.P., Hsu, Y.Y.,
the decrease in white matter volume. Future in Jagust, W.J., Yaffe, K., Kramer, J.H., Reed, B., Norman, D.,
vivo imaging studies using high field scanners are Chui, H.C. and Weiner, M.W. (2001) MRI of entorhinal
cortex and hippocampus in mild cognitive impairment and
necessary to fully elaborate the changes post-
Alzheimer’s disease. J. Neurol. Neurosurg. Psychiatry, 71(4):
synaptic to the entorhinal input; such changes in 441–447.
response to entorhinal atrophy are likely to in- Du, A.T., Schuff, N., Kramer, J.H., Ganzer, S., Zhu, X.P.,
clude the dentate gyrus. Jagust, W.J., Miller, B.L., Reed, B.R., Mungas, D., Yaffe,
K., Chui, H.C. and Weiner, M.W. (2004) Higher atrophy rate
of entorhinal cortex than hippocampus in AD. Neurology,
62(3): 422–427.
Acknowledgments Du, A.T., Schuff, N., Zhu, X.P., Jagust, W.J., Miller, B.L.,
Reed, B.R., Kramer, J.H., Mungas, D., Yaffe, K.,
This research was supported by grants P01 Chui, H.C. and Weiner, M.W. (2003) Atrophy rates of ent-
AG09466, P30 AG10161 and R01 AG17917 from orhinal cortex in AD and normal aging. Neurology, 60(3):
481–486.
the National Institute on Aging, National Insti- Folstein, M.F., Folstein, S.E. and McHugh, P.R. (1975) ‘‘Mini-
tutes of Health. mental state.’’ A practical method for grading the mental
status of patients for the clinician. J. Psychiatr. Res., 12(3):
189–192.
References Gomez-Isla, T., Price, J.L., McKeel, D.W., Morris, J.C., Grow-
don, J.H. and Hyman, B.T. (1996) Profound loss of layer. II.
Albert, M., Smith, L., Scherr, P., Taylor, J., Evans, D. and Entorhinal cortex neurons occurs in very mild Alzheimer’s
Funkenstein, H. (1991) Use of brief cognitive tests to identify disease. J. Neurosci., 16(14): 4491–4500.
individuals in the community with clinically diagnosed Good, C.D., Johnstrude, I.S., Ashburner, J., Henson, R.N.,
Alzheimer’s disease. Int. J. Neurosci., 57(3–4): 67–178. Friston, K.J. and Frakoviak, R.S. (2001) A voxel-based
Amaral, D.G., Insausti, R. and Cowan, W.M. (1987) The ent- morphometric study of ageing in 465 normal adult human
orhinal cortex of the monkey. I. Cytoarchitectonic organiza- brains. Neuroimage, 14(1): 21–36.
tion. J. Comp. Neurol., 264(3): 326–355. Goncharova, I.I., Dickerson, B.C., Stoub, T.R. and deToledo-
Bennett, D.A., Wilson, R.S., Schneider, J.A., Evans, D.A., Morrell, L. (2001) MRI of entorhinal cortex: a reliable pro-
Aggarwal, N.T., Arnold, S.E., Cochran, E.J., Berry-Kravis, tocol for volumetric measurement. Neurobiol. Aging, 22(5):
E. and Bienias, J.L. (2003) Apolipoprotein E epsilon4 allele, 737–745.
AD pathology, and the clinical expression of Alzheimer’s Hyman, B.T., Van Hoesen, G.W., Damasio, A.R. and Barnes,
disease. Neurology, 60(2): 246–252. C.L. (1984) Alzheimer’s disease: cell specific pathology iso-
Braak, H. and Braak, E. (1991) Neuropathological staging of lates the hippocampal formation. Science, 225(4667):
Alzheimer-related changes. Acta Neuropathol., 82(4): 1168–1170.
239–259. Insausti, R., Juottonen, K., Soininen, H., Insausti, A.M., Part-
Braak, H. and Braak, E. (1995) Staging of Alzheimer’s disease- anen, K., Vainio, P., Laakso, M.P. and Pitkanen, A. (1998)
related neurofibrillary changes. Neurobiol. Aging, 16(3): MR volumetric analysis of the human entorhinal, perirhinal
271–288. and temporopolar cortices. Am. J. Neuroradiol., 19(4):
Braak, H., Braak, E., Bohl, J. and Bratzke, H. (1998) Evolution 659–671.
of Alzheimer’s disease related cortical lesions. J. Neural Jack Jr., C.R., Petersen, R.C., Xu, Y., O’Brien, P.C., Smith,
Trans. Suppl., 54: 97–106. G.E., Ivnik, R.J., Boeve, B.F., Tangalos, E.G. and Kokmen,
Cardenas, V.A., Du, A.T., Hardin, D., Ezekiel, F., Weber, P., E. (2000) Rates of hippocampal atrophy correlate with
Jagust, W.J., Chui, H.C., Schuff, N. and Weiner, M.W. change in clinical status in aging and AD. Neurology, 55(4):
(2003) Comparison of methods for measuring longitudinal 484–489.
752

Jack Jr., C.R., Petersen, R.C., Xu, Y., O’Brien, P.C., Smith, Alzheimer’s disease: report of the NINCDS-ADRDA work
G.E., Ivnik, R.J., Tangalos, E.G. and Kokmen, E. (1998) group under the auspices of department of health and human
Rate of medial temporal lobe atrophy in typical aging and services task force on Alzheimer’s disease. Neurology, 34(7):
Alzheimer’s disease. Neurology, 51(4): 993–999. 939–944.
Jack Jr., C.R., Petersen, R.C., Xu, Y.C., O’Brien, P.C., Smith, Morris, J.C., Heyman, A., Mohs, R.C., van Belle, G., Fill-
G.E., Ivnik, R.J., Boeve, B.F., Waring, S.C., Tangalos, E.G. enbaum, G., Mellits, E.D. and Clark, C. (1989) The Con-
and Kokmen, E. (1999) Prediction of AD with MRI-based sortium to Establish a Registry for Alzheimer’s Disease
hippocampal volume in mild cognitive impairment. Neurol- (CERAD). Part I. Clinical and neuropsychological assess-
ogy, 52(7): 1397–1403. ment of Alzheimer’s disease. Neurology, 39(9): 1159–1165.
Jack Jr., C.R., Petersen, R.C., Xu, Y.C., Waring, S.C., O’Brien, Mueller, S.G., Stables, L., Du, A.T., Schuff, N., Truran, D.,
P.C., Tangalos, E.G., Smith, G.E., Ivnik, R.J. and Kokmen, Cashdollar, N. and Weiner, M.W. (2007) Measurement of
E. (1997) Medial temporal atrophy on MRI in normal aging hippocampal subfields and age-related changes with high
and very mild Alzheimer’s disease. Neurology, 49(3): 786–794. resolution MRI at 4T. Neurobiol. Aging, 28(5): 719–726.
Jack Jr., C.R., Schlung, M.M., Gunter, J.L., O’Brien, P.C., Mufson, E.J., Chin, E.Y., Cochran, E.J., Beckett, L.A., Ben-
Weigand, S.D., Knopman, D.S., Boeve, B.F., Ivnik, R.J., nett, D.A. and Kordower, J.H. (1999) Entorhinal cortex
Smith, G.E., Cha, R.H., Tangalos, E.G. and Petersen, R.C. beta-amyloid load in individual with mild cognitive impair-
(2004) Comparison of different MRI brain atrophy measures ment. Exp. Neurol., 158(2): 469–490.
with clinical disease progression in AD. Neurology, 62(4): Petersen, R.C. (2000) Aging, mild cognitive impairment and
591–600. Alzheimer’s disease. Dementia, 18: 789–805.
Jessen, F., Feyen, L., Freymann, K., Tepest, R., Maier, W., Petersen, R.C., Smith, G.E., Waring, S.C., Ivnik, R.J., Tanga-
Heun, R., Schild, H.H. and Scheef, L. (2006) Volume reduc- los, E.G. and Kokmen, E. (1999) Mild cognitive impairment:
tion of the entorhinal cortex in subjective memory impair- clinical characterization and outcome. Arch. Neurol., 56(3):
ment. Neurobiol. Aging, 27(12): 1751–1756. 303–308.
Kalbfleish, J.D. and Prentice, R.L. (1980) The Statistical Anal- Rodrigue, K.M. and Raz, N. (2004) Shrinkage of the entorhinal
ysis of Failure Time Data. Wiley, New York. cortex over five years predicts memory performance in
Killiany, R.J., Gomez-Isla, T., Moss, M., Kikinis, R., Sandor, healthy adults. J. Neurosci., 24(4): 956–963.
T., Jolesz, F., Tanzi, R., Jones, K., Hyman, B.T. and Albert, Saykin, A.J., Wishart, H.A., Rabin, L.A., Flashman, L.A.,
M.S. (2000) Use of structural magnetic resonance imaging to West, J.D., McHugh, T.L. and Mamourian, A.C. (2006)
predict who will get Alzheimer’s disease. Ann. Neurol., 47(4): Older adults with cognitive complaints show brain atrophy
430–439. similar to that of amnestic MCI. Neurology, 67(5): 834–842.
Killiany, R.J., Hyman, B.T., Gomez-Isla, T., Moss, M.B., Scheff, S.W. and Price, D.A. (2003) Synaptic pathology in
Kikinis, R., Jolesz, F., Tanzi, R., Jones, K. and Albert, M.S. Alzheimer’s disease: a review of ultrastructural studies.
(2002) MRI measures of entorhinal cortex vs. hippocampus Neurobiol. Aging, 24(8): 1029–1046.
in preclinical AD. Neurology, 58(8): 1188–1196. Scheff, S.W., Price, D.A., Schmitt, F.A. and Mufson, E.J.
Kordower, J.H., Chu, Y., Stebbins, G.T., DeKosky, S.T., Co- (2006) Hippocampal synaptic loss in early Alzheimer’s dis-
chran, E.J., Bennett, D. and Mufson, E.J. (2001) Loss and ease and mild cognitive impairment. Neurobiol. Aging,
atrophy of layer. II. Entorhinal cortex neurons in elderly 27(10): 1372–1384.
people with mild cognitive impairment. Ann. Neurol., 49(2): Small, S.A., Chawla, M.K., Buonocore, M., Rapp, P.R. and
202–213. Barnes, C.A. (2004) Imaging correlates of brain function in
Lancaster, J.L., Woldorff, M.G., Parsons, L.M., Liotti, M., monkeys and rats isolates a hippocampal subregion differ-
Freitas, C.S., Rainey, L., Kochunov, P.V., Nickerson, D., entially vulnerable to aging. Proc. Natl. Acad. Sci. U.S.A.,
Mikiten, S.A. and Fox, P.T. (2000) Automated Talairach 101(18): 7181–7186.
atlas labels for functional brain mapping. Hum. Brain Stoub, T.R., Bulgakova, M., Leurgans, S., Bennett, D.A.,
Mapp., 10(3): 120–131. Fleischman, D., Turner, D.A. and deToledo-Morrell, L.
de Leon, M.J., Convit, A., George, A.E., Golomb, J., DeSanti, (2005) MRI predictors of risk of Alzheimer’s disease: a lon-
S., Tarshish, C., Rusinek, H., Bobinski, M., Ince, C., Miller, gitudinal study. Neurology, 64(9): 1520–1524.
D. and Wisniewski, H. (1996) In vivo structural studies of the Stoub, T.R., deToledo-Morrell, L., Stebbins, G.T., Leurgans,
hippocampus in normal aging and in incipient Alzheimer’s S., Bennett, D.A. and Shah, R.C. (2006) Hippocampal dis-
disease. Ann. N.Y. Acad. Sci., 777: 1–13. connection contributes to memory dysfunction in individuals
de Leon, M.J., Convit, A., Wolf, O.T., Tarshish, C.Y., DeSanti, at risk for Alzheimer’s disease. Proc. Natl. Acad. Sci. U.S.A.,
S., Rusinek, H., Tsui, W., Kandil, E., Sherer, A.J., Roche, 103(26): 10041–10045.
A., et al. (2001) Prediction of cognitive decline in normal Squire, L.R. and Zola-Morgan, S. (1991) The medial temporal
elderly subjects with 2-[(18)F]fluoro-2-deoxy-D-glucose/posi- lobe memory system. Science, 253(5026): 1380–1386.
tron emission tomography (FDG/PET). Proc. Natl. Acad. Talairach, J. and Turnoux, P. (1988) Co-Planar Stereotaxic
Sci. U.S.A., 98(19): 10966–10971. Atlas of the Human Brain. Thieme, New York.
McKhann, G., Drachman, D., Folstein, M., Katzman, R., Tapiola, T., Pennanen, C., Tapiola, M., Tervo, S., Kivipelto,
Price, D. and Stadlan, E.M. (1984) Clinical diagnosis of M., Hänninen, T. and Pihlajamäki, M., et al. (2006) MRI of
753

hippocampus and entorhinal cortex in mild cognitive im- (area 35) cortices of the rhesus monkey. II. Frontal lobe
pairment: a follow-up study. Neurobiol. Aging, (in press). afferents. Brain Res., 95(1): 25–38.
deToledo-Morrell, L., Dickerson, B., Sullivan, M.P., Spanovic, Wang, C., Medina, D., Stebbins, G., Shah, R., Bammer, R.,
C., Wilson, R.S. and Bennett, D. (2000a) Hemispheric differ- Moseley, M. and deToledo-Morrell, L. (2006) Changes in the
ences in hippocampal volume predict verbal and spatial microstructural integrity of parahippocampal white matter
memory performance in patients with Alzheimer’s disease. tracts in Alzheimer’s disease. Soc. Neurosci. Abstr., on line.
Hippocampus, 10(2): 136–142. Wechsler, D. (1987) Wechsler Memory Scale: Revised Manual.
deToledo-Morrell, L., Goncharova, I., Dickerson, B., Wilson, Psychological Corporation, San Antonio, TX.
R.S. and Bennett, D.A. (2000b) From healthy aging to West, M.J., Coleman, P.D., Flood, D.G. and Troncoso, J.C.
Alzheimer’s disease: in vivo detection of entorhinal cortex (1994) Differences in the pattern of hippocampal neuronal
atrophy. Ann. N.Y. Acad. Sci., 911: 240–253. loss in normal aging and Alzheimer’s disease. Lancet,
deToledo-Morrell, L., Sullivan, M.P., Morrell, F., Wilson, 344(8925): 769–772.
R.S., Bennett, D.A. and Spencer, S. (1997) Alzheimer’s dis- West, M.J., Kawas, C.H., Martin, L.J. and Troncoso, J.C.
ease: in vivo detection of differential vulnerability of brain (2000) The CA1 region of the human hippocampus is a hot
regions. Neurobiol. Aging, 18(5): 463–468. spot in Alzheimer’s disease. Ann. N.Y. Acad. Sci., 908:
deToledo-Morrell, L., Stoub, T.R., Bulgakova, M., Wilson, 255–259.
R.S., Bennett, D.A., Leurgans, S., Wuu, J. and Turner, D.A. Wilson, R.S., Sullivan, M.P., deToledo-Morrell, L., Stebbins,
(2004) MRI-derived entorhinal volume is a good predictor of G.T., Bennett, D.A. and Morrell, F. (1996) Association of
conversion from MCI to AD. Neurobiol. Aging, 25(9): memory and cognition in Alzheimer’s disease with volumetric
1197–1203. estimates of temporal lobe structures. Neuropsychology,
Van Hoesen, G.W. and Hyman, B.T. (1990) Hippocampal for- 10(4): 459–463.
mation: anatomy and the patterns of pathology in Al- Xu, Y., Jack, C.R., O’Brien, P.C., Kokmen, E., Smith,
zheimer’s disease. Prog. Brain Res., 83: 445–457. G.E., Ivnik, R.J., Boeve, B.F., Tangalos, R.G. and Peter-
Van Hoesen, G.W. and Pandya, D.N. (1975) Some connections sen, R.C. (2000) Usefulness of MRI measures of entorhinal
of the entorhinal (area 28) and perirhinal (area 35) cortices of cortex versus hippocampus in AD. Neurology, 54(9):
the rhesus monkey. III. Efferent connections. Brain Res., 1760–1767.
95(1): 39–59. Young, B., Otto, T., Fox, G.D. and Eichenbaum, H. (1997)
Van Hoesen, G.W., Pandya, D.N. and Butters, N. (1975) Memory representation within the parahippocampal region.
Some connections of the entorhinal (area 28) and perirhinal J. Neurosci., 17(13): 5183–5195.
Plate 40.4. Color map showing significant (p ¼ 0.001) voxels of decreased white matter density in participants with amnestic MCI
compared with controls, superimposed on coronal and sagittal slices of a template based on data for all subjects in the study. The
image is masked to include white matter regions. The colors correspond to the t values shown on the color bar. Note the bilaterality of
significant differences in the white matter of the parahippocampal gyrus (adapted with permission from Stoub et al., 2006). (For B/W
version, see page 750 in the volume.)
H.E. Scharfman (Ed.)
Progress in Brain Research, Vol. 163
ISSN 0079-6123
Copyright r 2007 Elsevier B.V. All rights reserved

CHAPTER 41

Epileptogenesis in the dentate gyrus:


a critical perspective

F. Edward Dudek1, and Thomas P. Sutula2

1
Department of Physiology, University of Utah School of Medicine, 420 Chipeta Way, Suite 1700, Salt Lake City,
UT 84108, USA
2
Department of Neurology, University of Wisconsin, Madison, WI, USA

Abstract: The dentate gyrus has long been a focal point for studies on the molecular, cellular, and
network mechanisms responsible for epileptogenesis in temporal lobe epilepsy (TLE). Although several
hypothetical mechanisms are considered in this chapter, two that have garnered particular interest and
experimental support are: (1) the selective loss of vulnerable interneurons in the region of the hilus and
(2) the formation of new recurrent excitatory circuits after mossy fiber sprouting. Histopathological data
show that specific GABAergic interneurons in the hilus are lost in animal models of TLE, and several
lines of electrophysiological evidence, including intracellular analyses of postsynaptic currents, support
this hypothesis. In particular, whole-cell recordings have demonstrated a reduction in the frequency of
miniature inhibitory postsynaptic currents in the dentate gyrus and other areas (e.g., CA1 pyramidal
cells), which provides relatively specific evidence for a reduction in GABAergic input to granule cells.
These studies support the viewpoint that modest alterations in GABAergic inhibition can have significant
functional impact in the dentate gyrus, and suggest that dynamic activity-dependent mechanisms of
GABAergic regulation add complexity to this local synaptic circuitry and to analyses of epileptogenesis.
In regard to mossy fiber sprouting, a wide variety of experiments involving intracellular or whole-cell
recordings during electrical stimulation of the hilus, glutamate microstimulation, and dual recordings
from granule cells support the hypothesis that mossy fiber sprouting forms new recurrent excitatory
circuits in the dentate gyrus in animal models of TLE. Similar to previous studies on recurrent excitation
in the CA3 area, GABA-mediated inhibition and the intrinsic high threshold of granule cells in the
dentate gyrus tends to mask the presence of the new recurrent excitatory circuits and reduce the likelihood
that reorganized circuits will generate seizure-like activity. How cellular alterations such as neuron loss in
the hilus and mossy fiber sprouting influence functional properties is potentially important for under-
standing fundamental aspects of epileptogenesis, such as the consequences of primary initial injuries,
mechanisms underlying network synchronization, and progression of intractability. The continuous
nature of the axonal sprouting and formation of recurrent excitation could account for aspects of the
latent period and the progressive nature of the epileptogenesis. Future studies will need to identify

Corresponding author. Tel.: +1 (801)587-5880;


Fax: +1 (801)581-8075; E-mail: ed.dudek@hsc.utah.edu

DOI: 10.1016/S0079-6123(07)63041-6 755


756

precisely how these hypothetical mechanisms and others contribute to the process whereby epileptic
seizures are initiated or propagated through an area such as the dentate gyrus. Finally, in addition to
its unique features and potential importance in epileptogenesis, the dentate gyrus may also serve as a
model for other cortical structures in acquired epilepsy.

Keywords: epilepsy; seizure; granule cell; hilus; interneuron; axon sprouting

Introduction A definition of epileptogenesis

Overview Epileptogenesis is broadly considered to be the


collective molecular-, cellular-, and systems-level
For decades, the dentate gyrus has attracted the mechanisms whereby spontaneous recurrent sei-
attention of epilepsy researchers, and this interest zures develop after a brain insult in a manner that
has led to many publications that focus on this is time-dependent. For purposes of this review, we
particular brain structure. The high level of will consider epileptogenesis in a more specific and
interest stems from the development of several restrictive sense (Stables et al., 2003), rather than
concepts that have been derived from traditional the loose and more general meaning often associ-
brain science and from both basic and clinical ated with the term. The more restrictive definition
epilepsy research. The goal of this chapter is to of epileptogenesis does not include: (1) acute sei-
consider some of the key hypotheses, concepts zures associated with brain injury, (2) pharmaco-
and controversies in the field of epilepsy that are logical treatments in animals or isolated tissue that
centered on the role of the dentate gyrus in induce seizure-like activity, or (3) genetic abnor-
‘‘epileptogenesis’’ and on how different possible malities that are associated with increased seizure
mechanisms in the dentate gyrus may be altered susceptibility. While those processes are of interest
in a manner that might lead to temporal lobe ep- for understanding the acute mechanisms of net-
ilepsy (TLE). Epileptogenesis is broadly defined as work synchronization and are sometimes included
the biological process responsible for the devel- in the broader less restrictive definition of epile-
opment of spontaneous recurrent seizures after an ptogenesis, we believe for purposes of this critical
acquired brain insult or a primary abnormality analysis that focusing on the more restrictive defi-
such as a mutation or developmental defect. nition has advantages for understanding how
Even focusing on epileptogensis alone, the field chronic epilepsy develops and progresses. We be-
is too extensive and complicated to review in this lieve the term epileptogenesis also applies to the
brief chapter. Thus, our aim is to target a few progressive worsening of the epilepsy — in terms
basic concepts in epilepsy research, and provide of the increased frequency and/or severity of sei-
commentary in those areas where we believe zures — after the first of the spontaneous recurrent
there are particular problems in the field and/or seizures. A critical debate in the field is whether
where a different or modified perspective would acquired epilepsy (versus genetic epilepsy) requires
be useful. We will provide only minimal discus- actual brain damage and neuronal loss; some
sion of neurogenesis and GABA-A receptors as researchers believe, particularly in regard to
they relate to epileptogenesis in the dentate acquired pediatric epilepsy (e.g., after hypoxia,
gyrus because these topics are amply covered by prolonged febrile seizures, or status epilepticus
Parent (this volume) and Coulter and Carlson (SE) in children and immature animals), that frank
(this volume), respectively. Similarly, a detailed injury with neuronal death does not need to be
discussion of the electrophysiological studies of present for the development of chronic epilepsy
the human dentate gyrus in relation to epilepto- (i.e., epileptogenesis) to occur. This issue has been
genesis is provided by Williamson and Patrylo controversial, particularly in research on the dent-
(this volume). ate gyrus. Many studies have regarded increased
757

seizure susceptibility as a surrogate marker (e.g., Sloviter, 1991; Sloviter et al., 2003) that the
for epileptogenesis, and so another critical issue loss of hilar neurons (or ‘‘endfolium sclerosis’’ —
involves the question: how good are the data that considered to be a subset of the more general
an animal shows epileptogenesis (i.e., becomes histopathological condition of mesial temporal
‘‘epileptic’’ with spontaneous recurrent seizures)? sclerosis; Margerison and Corsellis, 1966) is the
central feature or minimum essential substrate of
TLE. Several studies on animal models of trau-
Concepts of epileptogenesis pertinent to the dentate matic brain injury (e.g., Lowenstein et al., 1992;
gyrus Toth et al., 1997) and ischemia (e.g., Crain et al.,
1988; Hsu and Buzsaki, 1993; Williams et al.,
Several questions about epileptogenesis in the 2004), which may lead to epilepsy, have found that
dentate gyrus have captured the attention of the hilus is particularly prone to neuronal injury;
the epilepsy research community during the last therefore, interest has focused on the vulnerability
decade; these concepts will serve as targets for of the different types of hilar neurons, and how
discussion in terms of interpretation of the pres- this might alter the electrophysiological properties
ently available literature and suggestions for future of the dentate network. At least two separate
research. The hippocampus has long been seen as concepts or hypotheses have emerged concerning
a critical structure in TLE, particularly as an the consequences of hilar neuron loss and the
epileptic focus (i.e., presence of interictal spikes), possible development of increased network excit-
as an epileptogenic zone (site of seizure initiation), ability: (1) a reduction in the number of interneu-
and thus, as a primary target for resections aimed rons (e.g., Obenaus et al., 1993; Houser and
at surgical treatment of intractable TLE (see Esclapez, 1996) and/or (2) a loss of excitatory
Engel, 1989 for review of the earlier literature). input to interneurons (i.e., the ‘‘dormant basket
This emphasis on the hippocampus in TLE, and by cell hypothesis’’; e.g., Sloviter, 1987; Sloviter et al.,
extrapolation of the dentate gyrus, however, has 2003).
been challenged from many perspectives (e.g., see In a separate but related line of research, at least
Gloor, 1992; Bertram, 1997; Bertram et al., 1998). four groups (de Lanerolle et al., 1989; Sutula et al.,
Because the dentate gyrus has traditionally been 1989; Houser et al., 1990; Babb et al., 1991) found
viewed as the first stage or ‘‘gate’’ into the classical that the presence of mesial temporal sclerosis in
trisynaptic circuit of the hippocampus, a hippo- specimens from surgical resections for intractable
campo-centric view has led to a dentate-centric TLE was correlated with ‘‘mossy fiber sprouting,’’
focus in the minds of many researchers. This is not where the Timm stain method densely stains the
to dismiss the concept that the hippocampus and inner molecular layer, which is not present in
dentate gyrus are likely important in TLE; rather, normal tissue (e.g., Tauck and Nadler, 1985; see
it is to acknowledge that other areas may be as Sutula and Dudek, this volume). Although the
important (or even more important), and that the dentate gyrus is strategically located, and both
mechanisms that occur during epileptogenesis in hilar neuron loss and mossy fiber sprouting are
the dentate gyrus may also be present in other common features of TLE, the dentate gyrus and its
brain regions. The observation of ‘‘maximal dent- granule cells are clearly not required for epilepto-
ate activation’’ strengthened the dentate ‘‘gate’’ genesis and both hilar neuron loss and mossy fiber
concept in epilepsy research, and provided a sprouting in the dentate gyrus do not appear to be
hypothetical electrophysiological basis for how necessary or sufficient for spontaneous recurrent
the dentate could serve a ‘‘closed gate’’ versus an seizures to occur. For reasons that we will discuss
‘‘open gate’’ function (Heinemann et al., 1992; and that are also developed in another chapter
Lothman et al., 1992; Stringer and Lothman, (Sutula and Dudek, this volume), these observa-
1992). tions do not diminish the potential importance of
Another basis for interest in the dentate gyrus understanding how the dentate gyrus contributes
by epilepsy researchers has been the hypothesis to hippocampal epileptogenesis and the syndromes
758

of TLE. In the sections that follow, we will con- their time course, these events are almost certainly
sider how some of the features of the organization synchronous action potentials (i.e., small popula-
and electrophysiology of the dentate gyrus are tion spikes) that are likely to be much larger in
specifically pertinent to the role of the dentate the dentate gyrus and hippocampus than other
gyrus in epileptogenesis. seizure-prone areas because of their laminar struc-
ture (Bragin et al., 2002). As a result of the laminar
organization, it is and will continue to be difficult
Epileptogenesis in the dentate gyrus: the laminar to assess where electrographic seizures actually
organization of the hippocampus leads to large field start, and whether an area in the hippocampus,
potentials and thus an increased ability to detect or in another structure, is the actual initiation
epileptiform events site. Nonetheless, the ability to analyze different
components of the field potentials in the dentate
While the dentate gyrus and hippocampus are gyrus (and hippocampus proper) in response to
commonly involved in TLE, which is the most electrical stimulation of synaptic inputs and during
common type of intractable epilepsy, the high spontaneous seizures provides important potential
interest and potential importance of this region of avenues for research.
limbic circuitry derives from the frequent obser-
vation of interictal discharges and apparent onset
of seizures during intrahippocampal depth record- Epileptogenesis in the dentate gyrus: the normal
ings both clinically and in animal models of TLE, dentate gyrus has a high threshold for seizure
in addition to the characteristic histopathological generation, so it may not be particularly prone to
(and clinical imaging) abnormalities (see Engel, chronic epileptogenesis
1989 for classical overview). An important point is
that because of the highly organized and laminar It is widely believed that the most important
structure of the hippocampus, including the dent- epileptogenic structures are particularly suscepti-
ate gyrus, synchronous activity in any particular ble to seizure generation in the normal brain, and
part of the hippocampus generates especially large these structures are even more prone to ‘‘hyperex-
field potentials in the extracellular space. There- citability’’ in the epileptic brain, where the term
fore, compared to other structures with less hyperexcitability describes a general increase in
packed and layered cell bodies, such as the am- response to a particular stimulus or enhanced
ygdala and entorhinal cortex, epileptiform events tendency to generate repetitive synchronous neu-
(including interictal spikes and seizures) would be ronal discharges manifesting as a burst of popu-
expected to be much more readily observed in the lation spikes. If this is the case, one could argue
hippocampus and dentate gyrus than other parts that the dentate gyrus — as studied in the normal
of the cerebral cortex. Thus, a seizure may start in brain — should be relatively low on the list of
a nearby structure (e.g., amygdala or entorhinal candidate structures for epileptogenesis. Although
cortex) with less lamination that generates a the dentate gyrus is a unique structure with a
smaller and more-difficult-to-detect field potential, high degree of neuronal homogeneity among
and then spread to the dentate gyrus and other the granule cells and heterogeneity in the hilus, it
hippocampal areas, where the relatively large- has long been known that the dentate granule
amplitude field potentials of the seizure might be cells are highly resistant to epileptiform activity
first detected. For example, considerable interest after treatment with convulsive agents, such as
has focused on the occurrence of ‘‘fast ripples,’’ GABA-A receptor antagonists (e.g., Fricke and
which is vernacular that refers to high-frequency Prince, 1984). In the normal brain, the dentate
electrographic oscillations (i.e., approximately gyrus appears to be the least susceptible part of the
200–500 Hz), as markers of epileptic seizure onset hippocampus to generate either brief interictal-like
in the dentate gyrus and elsewhere in the hippo- bursts or prolonged seizure-like activity, at least
campus and parahippocampal areas. Based on when compared to the CA1 and CA3 areas. This
759

resistance to epileptiform activity is widely fails, the dentate may allow propagation of activity
believed to be due to strong GABAergic inhibi- into other structures that are projection targets
tion, but the relatively negative resting membrane from the hippocampus (Heinemann et al., 1992;
potential, high threshold for action potential Lothman et al., 1992; Stringer and Lothman,
generation, and high degree of spike accommoda- 1992). There is evidence that the ‘‘gate’’ function
tion to maintained depolarization (e.g., see Staley of the dentate gyrus may operate in an ‘‘all or
et al., 1992) are probably equally as important none’’ fashion, as implied by the observation of
reasons why the dentate granule cells are relatively ‘‘maximal dentate activation’’ associated with
seizure-resistant under normal conditions. Many propagation of epileptiform activity. One basis
studies in epilepsy research have not considered for the ‘‘gate’’ concept is the relatively high thresh-
these fundamental properties of the normal dent- old for excitation of the dentate, and according to
ate gyrus; and on these grounds, the dentate gyrus this perspective, reduction in the ‘‘gate’’ function
may be an unlikely area for seizure onset during of the dentate gyrus could be an epileptogenic
epileptogenesis (i.e., be an epileptogenic zone), mechanism that would promote increased excita-
because seizure threshold of the dentate gyrus bility (i.e., hyperexcitability) to perforant path
often appears high compared to other structures stimuli, transforming synchronous excitatory
even when tested in chronically epileptic animals. postsynaptic potentials (EPSPs) with superim-
The transformation of the synaptic interactions in posed action potentials into a large burst of ac-
the dentate during epileptogenesis could, however, tion potentials. This mechanism could apply to
be more complex and important than is apparent synaptic inputs that include interictal spikes and
from the previous discussion. Indeed, if properties actual seizures; thus, it has been considered that
such as a more negative resting membrane poten- the ‘‘gate’’ to the hippocampus normally restricts
tial, higher action potential threshold, lower or blocks epileptiform activity. While this ‘‘gating’’
spontaneous firing rate, and resistance to repeti- property may restrict epileptiform activity, the
tive firing endow the dentate gyrus with resistance parallel pathway from the entorhinal cortex pro-
to synchronous network discharge and epilepto- jecting directly to the CA3 and CA1 areas may still
genesis, then processes such as hilar neuronal transmit discharges into the hippocampus regard-
loss and formation of recurrent excitatory circuits less of the properties of the dentate gyrus. Other
by mossy fiber sprouting could be especially studies suggest that the dentate gyrus can serve as
important in the transformation of the dentate a filter whereby activity is blocked from entering
gyrus into a more epileptogenic structure. the dentate under some conditions but not others
(e.g., see Iijima et al., 1996; Behr et al., 1998; ver-
sus Ang et al., 2006). Although the dentate gyrus
Models of epileptogenesis in the dentate gyrus: may not be necessary for transmission of signals
evidence for filtering and gating properties in the into the hippocampus, it is not fully understood
dentate gyrus how the dentate gyrus processes electrical infor-
mation in relation to normal hippocampal inte-
The perforant path projection from the entorhinal gration (see Kesner, this volume) or how the
cortex to the dentate gyrus is the first synapse of synaptically reorganized dentate gyrus affects the
the classic tri-synaptic hippocampal circuit, transmission of seizures into the hippocampus in
and this synapse to the granule cells is generally animal models of epilepsy or patients with TLE
believed to be relatively resistant to transmission (see Sutula and Dudek, this volume). Thus, the
of activity into CA3 (i.e., the dentate can be dentate gyrus presumably plays an important role
considered to be a gate that is usually closed; see in epileptogenesis, even though the entorhinal cor-
Hsu, this volume). The ‘‘gate’’ property of the tex appears to project ‘‘downstream’’ to the CA3
dentate gyrus is often thought to impede the prop- and CA1 areas of the hippocampus. The concept
agation of normal electrical activity and seizures that the dentate simply functions as a ‘‘gate’’ that
into hippocampus, and when this gating property normally blocks seizure activity in the entorhinal
760

cortex from propagating into the hippocampus potentially important functions regarding sponta-
may, however, be an oversimplification and neous interictal spikes and seizures versus electri-
require a more sophisticated assessment in the cally stimulated events.
complex network of serial and parallel pathways
projecting into the hippocampus and to parahip-
pocampal structures, such as the subiculum, which Animal models of epileptogenesis during TLE:
has also been implicated in TLE (Cohen et al., alterations of the dentate gyrus
1998).
Models and hypothetical mechanisms of
epileptogenesis
Epileptogenesis in the dentate gyrus: repeated
activation of the dentate gyrus can promote Animal models simulating histopathological
propagation of seizures into the hippocampus features of human TLE in the dentate gyrus have
been used for the study of chronic epileptogenesis.
The high seizure threshold of the normal dentate The two most widely used models of epileptogen-
gyrus becomes dramatically reduced after it has esis in adult animals, SE and kindling, have been
experienced a few electrically induced afterdis- used to examine processes of epileptogeneis fol-
charges (i.e., maximal dentate activation), and so lowing an initial injury (e.g., SE) and repeated
the dentate appears to be highly sensitive to pre- brief seizures (i.e., kindling). Both of these induc-
vious electrical activity (Heinemann et al., 1992; tion scenarios mimic common but distinctive clin-
Lothman et al., 1992; Stringer and Lothman, ical features of human TLE, namely syndromes
1992). Even a single afterdischarge in the dentate that follow an initial precipitating injury and
gyrus increases synaptic transmission for periods cryptogenic cases that gradually progress to in-
of as long as 3 months, and induction of long-term tractability. These models have provided experi-
potentiation increases susceptibility to evoked net- mental opportunities to assess two general types of
work activity and afterdischarges (Sutula and hypothetical mechanisms derived from histopath-
Steward, 1986, 1987; Sayin et al., 1999). These ological alterations in the dentate gyrus commonly
forms of network plasticity occur over a relatively associated with human TLE that would change the
short time frame and are therefore unlikely to ac- balance of excitation and inhibition, specifically
count for chronic epileptogenesis, but relatively reducing GABAergic inhibition and/or increasing
long-lasting alterations could promote network glutamatergic excitation. The hypothetical reduc-
synchronization in the dentate gyrus and hippo- tion in inhibition has been proposed to be due
campal pathways. The potential role of the dentate to a decrease in inhibitory interneurons and/or a
in modulating, reducing, or filtering some forms of reduction of excitatory synaptic input to inhibitory
electrical activity and possibly single seizures interneurons (i.e., the ‘‘dormant basket cell’’
would be degraded by activity-dependent enhance- hypothesis). Both of these effects could occur as
ment of synaptic transmission in granule cells, by a result of the histopathological observation of a
allowing more propagation of clusters of seizures loss of hilar neurons. The hypothesis of enhanced
into the hippocampus proper. Therefore, although excitation in the dentate gyrus has focused on new
some forms of entorhinal cortical activity can nor- recurrent excitatory circuits associated with mossy
mally bypass the dentate gyrus during propagation fiber sprouting, as detected by increased Timm
into the hippocampus, it remains possible that the stain product in the inner molecular layer of the
altered ability of repetitive seizures or seizure clus- dentate gyrus. Although there are many differ-
ters to spread into the hippocampus after maximal ences between the kindling and SE models, both
dentate activation or kindling may be a critical show some of the histopathological alterations in
feature of epileptogenesis during TLE. The present the dentate gyrus associated with TLE, and have
approaches with hippocampal slices, or even intact at least some utility to address these hypothetical
anaesthetized preparations, might fail to identify mechanisms of epileptogenesis.
761

Models of epileptogenesis based on kindling not recapitulate the key histopathological features
of TLE (Sloviter, 1991). Based on an interpreta-
Kindling is a progressive decrease in the threshold tion of Margerison and Corsellis (1966) (see also
for induction of afterdischarges to daily electrical Meldrum and Bruton, 1992), some workers have
stimulations of the amygdala or other limbic struc- argued that the SE models have too little damage
tures. This animal model progressively develops a in areas that are completely damaged in TLE
chronic irreversible state where low-intensity stim- (e.g., CA1 and hilus), but too much damage in
uli trigger prolonged afterdischarges and seizures. extrahippocampal areas. Although the hilus
Kindling initially induces low-level apoptosis in was seen to be the most common area for severe
the hilus (Bengzon et al., 1997) and cumulatively neuronal loss (about 65% of patients) in a histo-
results in progressive loss of neurons not only pathological study on autopsy tissue from patients
in the hilus, but in CA3 and CA1 in a pattern institutionalized due to epilepsy and other poten-
resembling human hippocampal sclerosis (Cavazos tial disorders, Margerison and Corsellis (1966)
et al., 1994). Kindling also induces mossy fiber also emphasized that extrahippocampal damage is
sprouting, and the Timm stain in the molecular commonly observed in many patients with classic
layer progressively increases over many months features of mesial temporal scerosis. For example,
with prolonged kindling (Cavazos et al., 1991), Margerison and Corsellis (1966) state that their
although the density of staining in the inner study of the whole brain ‘‘has emphasized the
molecular layer is generally less than is typically need to see beyond the temporal lobes and to take
seen in SE models where epileptogenesis occurs into account the possible importance of damage
over several weeks and months. in other parts of the brain as well.’’ More recent
imaging studies have emphasized the more wide-
spread injury to areas that communicate with the
Models of epileptogenesis based on SE
hippocampus, such as the amygdala and the ent-
orhinal and perirhinal cortex (see Cascino, 2005),
The two main approaches that have been invoked
although other brain areas are damaged in some
to induce SE and subsequent epileptogenesis
patients (Margerison and Corsellis, 1966). Thus,
involve injection of chemotoxins or electrical
the hippocampus in general and the dentate gyrus
stimulation. Kainic acid (i.e., kainate; Ben-Ari,
in particular are brain regions of considerable
1985; Nadler, 1991) and pilocarpine (Turski et al.,
interest in relation to human TLE, but other areas
1983; with or without lithium pretreatment)
are also involved. Furthermore, it is not clear that
are the most commonly used chemotoxins, but
the SE models are actually such a poor reflection
repetitive electrical stimulation of certain limbic
of TLE, as some would suggest.
structures can also induce SE (e.g., Lothman et al.,
1990). The repetitive seizures during SE cause
neuronal loss in the hilus of the dentate gyrus and
Histopathological alterations in the dentate gyrus of
in the CA3 and CA1 areas of the hippocampus,
animal models of epileptogenesis
plus other brain areas. The neuronal death in the
hilus is thought to contribute to mossy fiber
The models of TLE based on SE often have
sprouting (see Sutula and Dudek, this volume).
more damage to other brain areas than those TLE
patients with rather focal hippocampal and para-
Specific neuronal injury in the hilus and hippocampal lesions, but the SE models show many
hippocampus versus more extensive brain damage histopathological similarities to human TLE in the
associated with human TLE dentate gyrus. The SE models typically show exten-
sive histopathological damage to the hilus of the
The extrahippocampal damage associated with the dentate gyrus (Fig. 1) and to the CA3 and/or CA1
models based on SE is often severe and extensive, areas of the hippocampus that is roughly similar to
and thus it has been argued that the SE models do human TLE. These models have been employed
762

Fig. 1. Nissl staining to show hilar neurons in a control rat versus a rat with kainate-induced epilepsy. (A and B) Sections in the septal
hippocampus show some loss of hilar neurons in a kainate-treated rat (B) compared to control rat (A). (C and D) Near the temporal
end of the hippocampus, the loss of hilar neurons in the kainate-treated rat was severe (D). m: molecular layer; g: granule layer; h:
hilus; CA3: CA3 pyramidal cell layer; bar ¼ 500 mm. Adapted with permission from Buckmaster and Dudek (1997) Fig. 1, p. 388.

extensively, because they are relatively easy to use clinical features of some TLE syndromes, including
(although mortality can be a problem), and they onset with a delay after initial injury and progres-
show histopathological alterations in the hilus of the sion to intractability. It should be emphasized that
dentate gyrus that broadly reflect some of the fea- not all cases of human TLE develop after an ob-
tures of human TLE. In the SE models, spontane- vious initial injury and not all cases progress to in-
ous seizures develop with an apparent latent period tractability, so neither the SE models nor kindling
and progressively increase in seizure severity and are universally applicable to the broad range of
frequency. The SE models provide a contrast to syndromes of human TLE. It should be further em-
kindling, where neuronal injury is more gradually phasized that most histopathological studies on hu-
induced in the hilus and other regions, mossy fiber mans with TLE are based on patients who have
sprouting gradually progresses as a function of the suffered from intractable epilepsy for many years or
number of evoked seizures, and spontaneous sei- even decades, while research on animals (e.g., the SE
zures develop more slowly, typically after 90–100 models) does not usually involve periods greater
seizures evoked by daily stimulation (Sayin et al., than a few weeks or months.
2003). When spontaneous seizures are present in
more extensively kindled animals, mossy fiber Damage to the hilus is a common characteristic of
sprouting is more prominent and interneurons are TLE
lost in the hilus (Sayin et al., 2003). Thus, the kin-
dling and SE models recapitulate certain features of The dormant basket cell hypothesis
the histopathological changes and seizure suscepti-
bility characteristic of TLE, but with rates of in- Many studies over the last two decades have
duction and progression that distinctively mimic aimed to address the issue of which types of
763

neurons (see Houser, this volume) are lost in the chronic epilepsy (i.e., SE models), Bernard et al.
hilus in human tissue samples and in animal mod- (1998) (see also Esclapez et al., 1997) have pro-
els of TLE (particularly the SE models). Although vided several lines of evidence that basket cells are
certain vulnerable GABAergic interneurons are not ‘‘dormant’’ in TLE. In addition, studies on
lost, many other GABAergic neurons remain in- miniature and spontaneous inhibitory postsynap-
tact after a repetitive stimulation protocol to the tic currents (mIPSCs and sIPSCs) of dentate gran-
perforant path that causes ‘‘hyperexcitability’’ ule cells from the kainate model provide evidence
of dentate field potentials (Sloviter, 1987, 1991). that interneurons are spontaneously active even
Because many inhibitory basket cells were still when iontopic gluatamatergic receptors are
present within the dentate gyrus in human TLE blocked. When in vitro electrophysiological exper-
and in animal models, Sloviter (1987, 1991) pro- iments were conducted on hippocampal slices after
posed the ‘‘dormant basket cell’’ hypothesis, which the pharmacological blockade of ionotropic glut-
essentially states that TLE is associated with loss amatergic receptors to isolate mIPSCs and sIPSCs
of excitatory mossy cells that normally project to from glutamatergic synaptic circuits (Shao and
GABAergic basket cell interneurons; therefore; the Dudek, 2005), the frequency of sIPSCs in dentate
basket cells in the hilus of TLE patients and an- granule cells at o1 week and >3 months (i.e.,
imal models of TLE have lost excitatory synaptic from chronically epileptic rats) after kainate-
input (i.e., become ‘‘dormant’’), thus leading to the induced SE was not significantly different from
hyperexcitability. This hypothesis, however, is controls, and was much higher than when the
based on the relatively acute effects of repetitive slices were bathed in tetrodotoxin to block so-
extracellular stimulation (i.e., after hours and a dium-mediated action potentials and isolate mI-
few days), when chronic epileptic seizures either do PSCs. The observation that blocking action
not occur or are quite rare. Furthermore, nearly potentials with tetrodotoxin greatly reduced the
all of the electrophysiological data presented in frequency of IPSCs (i.e., the frequency of sIPSCs
support of this hypothesis involve in vivo paired- was much higher than mIPSCs) in all groups
pulse experiments with extracellular stimulation indicates that the interneurons that project to
and recording techniques using a range of inter- granule cells in kainate-treated rats (and controls)
pulse intervals and repetitive stimulation frequen- are clearly not ‘‘dormant’’ (i.e., they are sponta-
cies (e.g., Sloviter, 1987; Sloviter et al., 2003) that neously active) either o1 week or >3 months (i.e.,
are difficult to interpret and are generated by from chronically epileptic rats) after kainate-in-
complicated physiological processes at many duced SE, even when recorded in isolated slices
other cellular sites and levels (e.g., the perforant where GABAergic inhibitory circuits are intact and
path-to-granule cell synapse; GABA-A receptors, bathed in pharmacological agents that block glut-
chloride ion homeostasis, etc). A reduction of amatergic EPSCs. Although this approach relies
afferent input to GABAergic neurons has been on mIPSCs and sIPSCs in granule cells as a meas-
reported in the CA1 area (Bekenstein and Loth- ure of the upstream GABAergic inhibitory net-
man, 1993), but this study did not directly assess work, it is not affected by potential differences in
excitation to interneurons and alterations of a va- the responses to extracellular stimulation, nor is it
riety of physiological processes are likely to con- influenced by simultaneous alterations in glut-
tribute to a net defect in inhibitory mechanisms. amatergic synapses or circuits. This analysis is
For example, Doherty and Dingledine (2001) likely also somewhat of a simplification, since it
found a reduced excitatory drive onto interneu- has focused on somatic inhibition (i.e., the primary
rons in the dentate gyrus after SE, and this effect target of basket cells), but other studies with com-
was due to a use-dependent mechanism involving bined dendritic and somatic recording in the CA1
group II metabotropic glutamate receptors. Sev- area of slices from pilocarpine-treated rats have
eral other studies have addressed the dormant provided evidence for differential effects of epile-
basket cell hypothesis. Based on a series of exper- ptogenesis on dendritic and somatic inhibition
iments using hippocampal slices from animals with (Cossart et al., 2001). It remains possible, however,
764

that basket cells and some other remaining inter- forebrain ischemia have shown that selectively
neurons may experience a significant reduction of vulnerable neurons, including cells that are
excitatory synaptic input due to the loss of input likely GABAergic interneurons, are damaged.
from glutamatergic neurons in the hilus and else- Similar studies, focusing on immunohistochemical
where, and this loss of excitatory input could techniques, identified several different types of
importantly affect the ‘‘balance’’ of synaptic GABAergic interneurons that were damaged after
excitation and inhibition. Experiments with traumatic brain injury, and many of the injured
paired-pulse stimulation, particularly in intact interneuron types corresponded to those lost in
animals, lack the mechanistic resolution to pro- TLE. These histopathological alterations in inter-
vide evidence supporting the dormant basket cell neuron populations may contribute to the hyper-
hypothesis. Given the numerous studies at the cel- excitability often seen shortly after experimental
lular and synaptic levels indicating that surviving SE and other brain insults (e.g., Lowenstein et al.,
interneurons have significant spontaneous and 1992). Several other mechanisms, however, can
evoked activity in SE models, and the interpretive hypothetically be responsible for or contribute to
ambiguity and limitations of paired-pulse meas- the hyperactive responses to extracellular stimula-
urements using a range of interpulse intervals and tion within days after experimental SE, including
repetitive stimulation, overall the evidence for dor- alterations arising from direct SE-induced changes
mancy of basket cells as a mechanism of epilepto- in chloride distribution and other GABA-A
genesis is weak and indirect. Quantitative receptor-mediated mechanisms (e.g., Kapur et al.,
electrophysiological studies with whole-cell 1999).
recordings of EPSCs to assess changes in the ex- Several laboratories have used high-resolution
citatory synaptic inputs to the different types of whole-cell recordings of mIPSCs to test more di-
GABAergic interneurons during epileptogenesis, rectly the hypothesis of a loss of interneuron in-
preferably using several different animal models of put to granule cells (Kobayashi and Buckmaster,
TLE (i.e., including but not limited to kindling and 2003; Shao and Dudek, 2005) and CA1 pyramidal
SE models), could be useful to further evaluate cells (Wierenga and Wadman, 1999), and they
this hypothesis. have found that the frequency of mIPSCs is re-
duced in these models, which is consistent with a
loss of inhibitory GABAergic terminals (i.e.,
Loss of interneurons death of GABAergic interneurons) in TLE. Sim-
ilar data showing a reduction in the frequency of
Whereas the dormant basket cell hypothesis has mIPSCs have been reported after lateral fluid
not garnered support from recent electrophysio- percussion injury (Toth et al., 1997). An impor-
logical experiments using modern recording tech- tant advantage of this approach is that it is
niques, data with a wide range of morphological independent of the complexities and uncertainties
and physiological techniques from many labora- of extracellular stimulation within a multisynaptic
tories in human tissue and animal models directly circuit. Because all of the experimental record-
support the hypothesis that GABAergic inhibition ings were conducted in tetrodotoxin to block
undergoes alterations in association with modest action potential-mediated activity (and in some
loss of specific interneurons in TLE. Early studies cases, ionotropic glutamatergic antagonists to
in a variety of hippocampal and cortical areas block fast EPSPs), the number of other potential
using various immunohistochemical and in situ confounding mechanisms and difficult-to-control
hybridization techniques have shown that some variables was reduced. In contrast to the lack of
GABAergic interneurons (i.e., a relatively small evidence for dormancy of basket cells, the evi-
number, and only specific types) are lost in TLE dence from direct experiments on several animal
(Fig. 2; e.g., Sloviter, 1987; Obenaus et al., 1993; models from several laboratories that epilepto-
Buckmaster and Dudek, 1997; Gorter et al., 2001). genesis is associated with a modest but measur-
Immunohistochemical data and silver stains after able reduction in inhibition to dentate granule
765

Fig. 2. GAD mRNA-containing neurons in the dentate gyrus of a control (A) and pilocarpine-treated (B) rat. (A) In the normal
dentate gyrus, numerous GAD mRNA-labeled neurons are present in the hilus (H). Many labeled neurons are also located along the
inner border of the granule cell layer (G) and in the molecular layer (M). (B) In a pilocarpine-treated animal, the number of labeled
neurons throughout the hilus (H) is substantially reduced. Some GAD mRNA-containing neurons persist along the base of the granule
cell layer (G) and in the molecular layer (M); scale bar ¼ 200 mm. Adapted with permission from Obenaus et al. (1993) Fig. 4, p. 4476.

cells associated with a loss of specific types of Mossy fiber sprouting is also a common feature of
vulnerable interneurons in the hilus of the dentate TLE
gyrus (and also in CA1) is quite strong. Because
this reduction in the number of specific interneu- Axon sprouting and increased recurrent excitation
rons ultimately would be expected to enhance the
effects of excitatory synaptic input to the granule Mossy fiber sprouting has attracted considerable
cell population, and because assessments of GAB- attention because the Timm stain shows a dra-
Aergic inhibition using extracellular stimulation matic and reproducible change in the distribution
depend heavily on the intensity of the stimulation, of the mossy fiber axons in the inner molecular
the types of analyses that employ electrical stim- layer of the dentate gyrus associated with TLE
ulation of an afferent input such as the perforant (Fig. 3). The ease and relative simplicity of this
path are problematic. Furthermore, the presence staining technique and the apparent clarity of the
of fewer interneurons will lead to complex effects results have drawn attention to the dentate gyrus
during the repetitive activation expected to occur and this particular form of synaptic reorganization
as multiple interictal spikes and seizures begin of a local-circuit pathway. The fact that numerous
to occur in the entorhinal cortex and dentate laboratories have independently observed Timm
gyrus. stain in the inner molecular layer of human tissue
766

Fig. 3. The Timm stain showed dark labeling in the granule cell layer and inner molecular layer in rats with kainate-induced epilepsy
versus control rats. (A) Hippocampal sections from a control rat had relatively little Timm stain (with light cresyl violet counter-
staining) in the granule cell and inner molecular layers. (B–E) Sections of the middle region along the septotemporal axis of the
hippocampus from rats with kainate-induced epilepsy displayed different degrees of abnormal Timm staining in the granule cell layer
(B shows the least and E the most). Arrows indicate regions of Timm staining in the inner molecular layer. M: molecular layer; g:
granule layer; h: hilus; CA3: CA3 pyramidal cell layer; bar ¼ 500 mm. Adapted with permission from Buckmaster and Dudek (1997)
Fig. 7, p. 395.

from patients with intractable epilepsy (de Lane- spike bursts several weeks or months after kain-
rolle et al., 1989; Sutula et al., 1989; Houser et al., ate treatment, particularly when slices from rats
1990; Babb et al., 1991) supports the concept that with chronic epilepsy are bathed in GABA-A
it is one of the changes likely to be associated with receptor antagonists and/or high potassium,
the broadly defined concept of synaptic reorgan- versus similar experiments on slices from control
ization. Animal models, ranging from pilocarpine- animals where hilar stimulation evokes one anti-
to kainate-treated rats to the kindling model, have dromic action potential and sometimes an EPSP
all revealed a time-dependent increase in Timm (Tauck and Nadler, 1985; Cronin et al., 1992;
stain in the inner molecular layer of the dentate Wuarin and Dudek, 1996; Patrylo and Dudek,
gyrus. As reviewed elsewhere in this volume 1998; Hardison et al., 2000, Lynch and Sutula,
(Sutula and Dudek, this volume), several lines of 2000). The advantage of hilar over perforant path
electrophysiological and ultrastructural data sug- stimulation in these earlier studies was that elec-
gest that most if not all of the new synapses in the trical stimulation of the hilus would be expected
‘‘reorganized’’ dentate gyrus are excitatory. For to activate the mossy fiber axons of the granule
example, several laboratories have shown that hi- cells themselves versus the powerful monosynaptic
lar (and perforant path) stimulation can lead to inputs of the perforant path, but in either case,
delayed or long-latency EPSPs and prolonged the key result is that the EPSPs in slices from rats
767

with robust mossy fiber sprouting occurred with a consequent generation and propagation of
longer and more variable latency than would be epileptiform activity in the dentate gyrus was time
expected from a direct monosynaptic input. These dependent (Wuarin and Dudek, 2001), they did
experiments were founded conceptually on the not directly address the issue of monosynaptic
studies of recurrent excitation in the CA3 area connections between individual granule cells. Dual
(Traub and Wong, 1982, 1983; Miles and Wong, intracellular recordings provided direct evidence
1987), where low-intensity stimulation triggered that mossy fiber sprouting in rats with pilocarpine-
all-or-none synaptic bursts with a long–and vari- induced epilepsy is associated with monosynaptic
able latency and paired intracellular recordings recurrent excitatory connections among dentate
showed multisynaptic interactions, only when granule cells (Scharfman et al., 2003). The most
GABA-A mediated inhibition was blocked. The direct and quantitative ultrastructural studies also
advantage of performing these experiments during point to a high preponderance of new excitatory
blockade of GABA-A mediate inhibition is based synapses to dentate granule cells versus interneu-
not only on the previous work showing that rons (Zhang and Houser, 1999; Buckmaster et al.,
GABA-mediated inhibition has a ‘‘masking’’ effect 2002). Although synaptic reorganization may have
on recurrent excitation (Christian and Dudek, a unique importance in the dentate gyrus, a more
1988; Traub and Wong, 1982, 1983; Miles and likely scenario is that axon sprouting and the for-
Wong, 1987), but also because it controls at least mation of new recurrent excitatory circuits is a
partially for epileptogenesis-associated differences widespread phenomenon after injury that occurs
in GABA-mediated inhibition (see above), partic- throughout many areas of the hippocampus and
ularly since the experimental design in these stud- neocortex during epileptogenesis. For example,
ies compared slices from the animal model of numerous laboratories have provided morpholog-
epilepsy with exactly equivalent slices from control ical and electrophysiological evidence that a sim-
animals. Electrical stimulation, however, activates ilar mechanism of axonal sprouting and enhanced
axons of passage and can thus be nonspecific. recurrent excitation occurs in the CA1 area during
More direct experiments using glutamate micro- epileptogenesis (Meier and Dudek, 1996; Perez et
stimulation techniques that do not activate fibers al., 1996; Esclapez et al., 1999; Smith and Dudek,
of passage, including glutamate microdrops 2001, 2002; Shao and Dudek, 2004). As expected
(Wuarin and Dudek, 1996; Lynch and Sutula, from previous experiments performed on recurrent
2000) and focal photoactivation of caged gluta- excitatory circuits in the CA3 area (Traub
mate (Molnar and Nadler, 1999; Wuarin and and Wong, 1982, 1983; Miles and Wong, 1987;
Dudek, 2001), showed that (1) relatively selective Christian and Dudek, 1988), several laboratories
stimulation of dentate granule cells could cause have shown that reduction of GABA-A mediated
EPSCs in other granule cells in normal medium, inhibition or an increase in extracellular potassium
(2) these recurrent excitatory circuits increased in has an unmasking effect on the new local recurrent
density and effectiveness over the ensuing months excitatory circuits in the dentate gyrus. The un-
after experimental SE, and (3) when the slices were derlying concept is that GABA-A mediated inhi-
bathed in bicuculline, the new local excitatory cir- bition, along with the membrane properties of
cuits of the granule cells could generate novel net- the granule cells (e.g., the highly negative resting
work bursting to focal uncaging of glutamate in potentials, high threshold, and resistance to repet-
the granule cell layer (but not elsewhere) several itive firing), will tend to impede the likelihood of
months after SE, but not at earlier times and not in multisynaptic interactions, which is the prime
slices from control preparations. Although these mechanism by which increased recurrent excita-
experiments with glutamate microdrops and focal tion would be expected to lead to seizure activity.
uncaging of glutamate used more specific tech- This concept may provide a hypothetical explana-
niques than extracellular electrical stimulation to tion for how seizures might not occur under some
activate granule cells, and showed that the forma- circumstances, but would emerge under others;
tion of new recurrent excitatory circuits and the that is, seizures would be generated and/or
768

propagated when synaptic inhibition is slightly reorganization in other areas of the hippocampus
and transiently depressed and/or extracellular and temporal lobe may involve new inhibitory cir-
potassium increased in concentration, as this par- cuits from sprouted mossy fibers to interneurons,
ticular set of changes would allow multisynaptic the preponderance of evidence from many labora-
interactions to generate a seizure. tories with a wide range of techniques using several
different animal models point to the conclusion
that mossy fiber sprouting leads predominantly to
Increased recurrent excitation versus increased new recurrent excitatory circuits.
inhibition as a consequence of mossy fiber sprouting

Although most experiments have focused on and Mossy fiber sprouting and increased seizure
supported the hypothesis that mossy fiber sprout- susceptibility
ing during chronic epileptogenesis leads to in-
creased recurrent excitation among granule cells, Many studies have noted the close association
some data suggest that the sprouted mossy fibers between the presence of mossy fiber sprouting and
preferentially synapse on interneurons (e.g., Slovi- decreased seizure threshold (i.e., in the kindling
ter, 1992; Kotti et al., 1997; Harvey and Sloviter, model) or the presence of spontaneous recurrent
2005). The hypothesis is that the ‘‘dormant’’ bas- seizures (i.e., in the models of epileptogenesis based
ket cells, which have presumably lost excitatory on SE). Nonetheless, many tissue specimens from
synaptic input during repetitive seizures, become patients with intractable TLE lack sprouting, and
re-innervated by the sprouted mossy fibers. The seizure frequency does not appear to be related to
histopathological data include light micrographs the degree of mossy fiber sprouting in animal mod-
of basket cells surrounded by Timm stain, but as els (e.g., Buckmaster and Dudek, 1997). This issue
pointed out previously by Ribak and Peterson is complex, however, because the distribution of
(1991), similar micrographs can be obtained from mossy fiber sprouting is not homogeneous (e.g.,
the temporal pole of the hippocampus of normal more robust sprouting generally occurs in the tem-
animals, and the light-microscopic techniques do poral versus septal hippocampus in animal mod-
not show whether mossy fibers near interneurons els). Plus, axonal sprouting and the formation of
actually synapse with them. Ultrastructural obser- new recurrent excitatory circuits likely occur in
vations have reported mossy fiber synapses on in- many other structures in addition to the dentate
terneurons (Cavazos et al., 2003; see Sutula and gyrus, and synaptic reorganization in these other
Dudek, this volume), but the number of synaptic areas is difficult to detect. Gorter et al. (2001) re-
contacts from mossy fibers to interneurons has not ported that rats subjected to amygdala-stimulated
been rigorously evaluated and there is no quanti- SE could be separated into two groups: (1) animals
tative evidence that there is an increase over nor- that showed a progressive increase in seizure fre-
mal levels. Indeed, contacts by sprouted mossy quency over time and that had robust mossy fiber
fibers on interneurons appear to be infrequent if sprouting, and (2) rats with a low seizure frequency
not rare (see Sutula and Dudek, this volume). and little or no mossy fiber sprouting. Longo and
Most of the electrophysiological evidence for Mello (1997) have reported that pretreatment with
the ‘‘hyper-inhibition’’ hypothesis is essentially the protein-synthesis inhibitor cycloheximide
based on variations of the paired-pulse technique blocks mossy fiber sprouting but does not prevent
(Sloviter, 1992; Buckmaster and Dudek, 1997; the development of spontaneous recurrent seizures.
Wilson et al., 1998; Harvey and Sloviter, 2005), Williams et al. (2002), however, found that cyclo-
which is potentially confounded by technical issues heximide pretreatment had no detectable effect on
(e.g., see Waldbaum and Dudek, 2005, 2006). mossy fiber sprouting. In this study, some rats did
Thus, although mossy fiber sprouting in the dent- not experience SE (i.e., had few if any convulsive
ate gyrus- and sprouting-based synaptic seizures during pilocarpine treatment), but still had
769

at least a few subsequent spontaneous seizures with Does mossy fiber sprouting inhibit seizures?
little or no Timm stain in the inner molecular layer,
consistent with the data of Gorter et al. (2001) After spontaneous seizures in pilocarpine-treated
suggesting that mossy fiber sprouting is not neces- rats, c-fos expression has been reported to occur
sary for the development of occasional spontane- preferentially in interneurons versus dentate
ous seizures. It should not be surprising that the granule cells (Harvey and Sloviter, 2005), which
dentate gyrus and mossy fiber are not required for has been interpreted to mean that mossy fiber
limbic seizures (for more extensive discussion on sprouting leads to new synapses onto inhibitory
this point, see Sutula and Dudek, this volume). interneurons rather than to granule cells (i.e.,
More recently, Toyoda and Buckmaster (2005) hyper-inhibition), and that mossy fiber sprouting
also attempted to determine if cycloheximide suppresses rather than promotes seizures. Harvey
blocked mossy fiber sprouting after pilocarpine-in- and Sloviter (2005) evaluated c-fos staining 1 h
duced SE; they infused cycloheximide directly into after a spontaneous seizure, while in similar stud-
the area around the hippocampus for prolonged ies, Peng and Houser (2005) euthanized animals
periods from an osmotic minipump, provided 15 min after a seizure and found Fos expression
evidence that the cycloheximide was present in preferentially in granule cells. Peng and Houser
the area around the dentate gyrus, and also ob- (2005) found that staining was observed in inter-
served robust Timm stain in the inner molecular neurons when animals were euthanized after 1 h,
layer (i.e., cycloheximide had no detectable effect the Fos staining in interneurons persisted longer
on mossy fiber sprouting). It is not clear how one than in granule cells, and the enhanced Fos ex-
pretreatment injection of cycloheximide would be pression in the granule cells was not observed
expected to block the weeks/months-long continu- when a second seizure occurred shortly after a
ous process that is responsible for sprouting of the previous seizure. The results of Harvey and Slovi-
mossy fibers and the formation of new synaptic ter (2005) could have arisen if they studied animals
contacts with dentate granule cells. Presently, that had had more than one seizure in close
therefore, the proposal that cycloheximide blocks temporal proximity. In addition, progressive acti-
mossy fiber sprouting is problematic. The obser- vation of granule cells and interneurons during
vation of spontaneous seizures before little or any spontaneous seizures would be expected to occur
mossy fiber sprouting has occurred (i.e., between 5 over hundreds of milliseconds, and seizures typi-
and 7 days after kainate-induced SE in the study of cally last up to 3 min; therefore, c-fos expression
Hellier et al., 1999), and other data where animals lacks the temporal resolution to differentiate
have shown spontaneous seizures with little or no whether granule cells or interneurons are sequenti-
evidence of mossy fiber sprouting weeks after pi- ally or preferentially activated during spontaneous
locarpine treatment (Williams et al., 2002), suggest seizures. Harvey and Sloviter (2005) also proposed
that other mechanisms (e.g., loss of interneurons) that population spikes are highly prevalent in the
besides mossy fiber sprouting in the dentate gyrus dentate gyrus during spontaneous seizures that
can lead to the development of spontaneous occur a few days after SE, but are rare or nonex-
seizures after experimental SE. Although axonal istent when spontaneous seizures occur weeks
sprouting after neuronal injury in other brain areas later, as might be expected if inhibition increased
could account for the occurrence of spontaneous in the dentate gyrus over time after SE. Alternative
seizures, particularly in the cases of spontaneous mechanisms besides increased inhibition are
seizures within a few weeks after SE, a more likely possible, including depolarization inactivation, as
explanation is that the death of vulnerable inter- occurs during paroxysmal depolarization shifts af-
neurons during the SE (e.g., see Houser and ter blockade of GABA-A receptors. Thus, it is not
Esclapez, 1996) decreased seizure threshold (i.e., clear whether these data apply to the question of
increased seizure susceptibility) and led to occa- whether mossy fiber sprouting leads to hyper-in-
sional seizures. hibition and is potentially compensatory.
770

Reorganization of GABAergic neurons synchronization demand continuing attention to


a diverse range of potential mechanisms and in-
Another hypothesis, based on increased teractions spanning molecular, cellular, and sys-
immunohistochemical staining of GABAergic ter- tems levels, the insights gained from study of
minals in human tissue and in animal models of neuronal loss and mossy fiber sprouting during the
TLE is whether other remaining GABAergic ter- last two decades have been a substantial advance
minals might sprout new synaptic connections to and a major accomplishment for epilepsy research.
partly (or totally) restore GABA-mediated inhibi-
tion (e.g., see Davenport et al., 1990; Bausch,
2005). If this occurred, one would expect an in- References
crease in the frequency of mIPSCs with time after
experimental SE, but this was not seen in rats with Ang, C.W., Carlson, G.C. and Coulter, D.A. (2006) Massive
kainate-induced epilepsy (Shao and Dudek, 2005); and specific dysregulation of direct cortical input to the hip-
however, other techniques may reveal axon pocampus in temporal lobe epilepsy. J. Neurosci., 26:
sprouting, synaptic reorganization and recovery 11850–11856.
Babb, T.L., Kupfer, W.R., Pretorius, J.K., Crandall, P.H. and
of inhibition during epileptogenesis. Levesque, M.F. (1991) Synaptic reorganization by mossy
fibers in human epileptic fascia dentata. Neuroscience, 42:
351–363.
The dentate gyrus as a model for studies of Bausch, S.B. (2005) Axonal sprouting of GABAergic interneu-
rons in temporal lobe epilepsy. Epilepsy Behav., 7: 390–400.
epileptogenesis associated with TLE
Behr, J., Lyson, K.J. and Mody, I. (1998) Enhanced propaga-
tion of epileptiform activity through the kindled dentate
Although the dentate gyrus and the hilus clearly gyrus. J. Neurophysiol., 79: 1726–1732.
have an important and probably have a unique Bekenstein, J.W. and Lothman, E.W. (1993) Dormancy of in-
role in hipocampal integration and epileptogene- hibitory interneurons in a model of temporal lobe epilepsy.
sis, they may also serve as a model for other Science, 259: 97–100.
Ben-Ari, Y. (1985) Limbic seizure and brain damage produced
cortical structures where epileptogenesis involves by kainic acid: mechanisms and relevance to human temporal
neuronal death and synaptic reorganization in ad- lobe epilepsy. Neuroscience, 14: 375–404.
dition to possible alterations in neurotransmitter Bengzon, J.K.Z., Elmér, E., Nanobashvili, A., Kokaia, M. and
receptors and voltage-gated channels. Although Lindvall, O. (1997) Apoptosis and proliferation of dentate
epilepsy is caused by a great variety of age-depend- gyrus neurons after single and intermittent limbic seizures.
Proc. Natl. Acad. Sci., 94: 10432–10437.
ent etiologies and the mechanisms appear to be Bernard, C., Esclapez, M., Hirsch, J.C. and Bernard, C. (1998)
equally heterogeneous, the cellular and network Interneurones are not so dormant in temporal lobe epilepsy:
mechanisms of (1) a loss of vulnerable interneu- a critical reappraisal of the dormant basket cell hypothesis.
rons, and (2) axon sprouting with progressive for- Epilepsy Res., 32: 93–103.
Bertram, E.H. (1997) Functional anatomy of spontaneous sei-
mation of new recurrent excitatory circuits have
zures in a rat model of limbic epilepsy. Epilepsia, 38: 95–105.
the most support from independent experiments Bertram, E.H., Zhang, D.X., Mangan, P., Fountain, N. and
in numerous laboratories and multiple animal Rempe, D. (1998) Functional anatomy of limbic epilepsy: a
models. Although other molecular, cellular and proposal for central synchronization of a diffusely hyperex-
network mechanisms have been implicated in the citable network. Epilepsy Res., 32: 194–205.
dentate gyrus, these two cellular mechanisms that Bragin, A., Mody, I., Wilson, C.L. and Engel, J. (2002) Local
generation of fast ripples in epileptic brain. J. Neurosci., 22:
have been extensively documented in the dentate 2012–2021.
gyrus may be generally applicable in many other Buckmaster, P.S. and Dudek, F.E. (1997) Neuron loss, granule
temporal and neocortical structures. Precisely how cell axon reorganization, and functional changes in the dent-
these mechanisms and others in the dentate ate gyrus of epileptic kainate-treated rats. J. Comp. Neurol.,
385: 385–404.
gyrus and elsewhere lead to the epileptic seizures
Buckmaster, P.S., Zhang, G.F. and Yamawaki, R. (2002) Axon
associated with TLE will require further studies. sprouting in a model of temporal lobe epilepsy creates a pre-
While the heterogeneity of TLE etiologies and the dominantly excitatory feedback circuit. J. Neurosci., 22:
mechanisms underlying neuronal and network 6650–6658.
771

Cascino, G.D. (2005) Temporal lobe epilepsy: more than hip- Gloor, P. (1992) Role of the amygdala in temporal lobe epi-
pocampal pathology. Epilepsy Curr., 5: 187–189. lepsy. In: Aggleton J.P. (Ed.), The Amygdala: Neurobiolog-
Cavazos, J.E., Das, I. and Sutula, T.P. (1994) Neuronal loss ical Aspects of Emotion, Memory, and Mental Dysfunction.
induced in limbic pathways by kindling: evidence for induc- Wiley-Liss, Inc., New York, pp. 505–538.
tion of hippocampal sclerosis by repeated brief seizures. J. Gorter, J.A., van Vliet, E.A., Aronica, E. and Lopes da Silva,
Neurosci., 14: 3106–3121. F.H. (2001) Progression of spontaneous seizures after status
Cavazos, J.E., Golarai, G. and Sutula, T.P. (1991) Mossy fiber epilepticus is associated with mossy fibre sprouting and ex-
synaptic reorganization induced by kindling: time course of tensive bilateral loss of hilar parvalbumin and somatostatin-
development, progression, and permanence. J. Neurosci., 11: immunoreactive neurons. Eur. J. Neurosci., 13: 657–669.
2795–2803. Hardison, J.L., Okazaki, M.M. and Nadler, J.V. (2000) Modest
Cavazos, J.E., Zhang, P., Qazi, R. and Sutula, T.P. (2003) increase in extracellular potassium unmasks effect of recur-
Ultrastructural features of sprouted mossy fiber synapses in rent mossy fiber growth. J. Neurophysiol., 84: 2380–2389.
kindled and kainic acid-treated rats. J. Comp. Neurol., 458: Harvey, B.D. and Sloviter, R.S. (2005) Hippocampal granule
272–292. cell activity and c-Fos expression during spontaneous sei-
Christian, E.P. and Dudek, F.E. (1988) Characteristics of local zures in awake, chronically epileptic, pilocarpine-treated rats:
excitatory circuits studied with glutamate microapplication in Implications for hippocampal epileptogenesis. J. Comp. Ne-
the CA3 area of rat hippocampal slices. J. Neurophysiol., 59: urol., 488: 442–463.
90–109. Heinemann, U., Beck, H., Dreier, J.P., Ficker, E., Stabel, J. and
Cohen, I., Navarrao, V., Clemenceau, S., Baulac, M. and Miles, Zhang, C.L. (1992) The dentate gyrus as a regulated gate
R. (1998) On the origin of interictal activity in human tem- for the propagation of epileptiform activity. Epilepsy Res.
poral lobe epilepsy in vitro. Science, 298: 1418–1421. Suppl., 7: 273–280.
Cossart, R., Dinocourt, C., Hirsch, J.C., Merchan-Perez, A., Hellier, J.L., Patrylo, P.R., Dou, P., Nett, M., Rose, G.M. and
De Felipe, J., Ben-Air, Y., Esclapez, M. and Bernard, C. Dudek, F.E. (1999) Assessment of inhibition and epilepti-
(2001) Dendritic but not somatic GABAergic inhibition form activity in the septal dentate gyrus of freely behaving
is decreased in experimental epilepsy. Nat. Neurosci., 4: rats during the first week after kainate treatment. J. Neuro-
52–62. sci., 19: 10053–10064.
Coulter, D.A. and Carlson, G.C. (this volume) Functional reg- Houser, C.R. (this volume) Interneurons of the dentate gyrus:
ulation of the dentate gyrus by GABA-mediated inhibition. An overview of cell types, terminal fields and neurochemical
In: Scharfman, H. (Ed.), The Dentate Gyrus. Progress in identity. In: Scharfman, H. (Ed.), The Dentate Gyrus.
Brain Research. Elsevier, Amsterdam. Progress in Brain Research. Elsevier, Amsterdam.
Crain, B.J., Westerkam, W.D., Harrison, A.H. and Nadler, J.V. Houser, C.R. and Esclapez, M. (1996) Vulnerability and plas-
(1988) Selective neuronal death after transient forebrain is- ticity of the GABA system in the pilocarpine model of spon-
chemia in the Mongolian gerbil: a silver impregnation study. taneous recurrent seizures. Epilepsy Res., 26: 207–218.
Neuroscience, 27: 387–402. Houser, C.R., Miyashiro, J.E., Swartz, B.E., Walsh, G.O.,
Cronin, J., Obenaus, A., Houser, C.R. and Dudek, F.E. (1992) Rich, J.R. and Delgado-Escueta, A.V. (1990) Altered pat-
Electrophysiology of dentate granule cells after kainate-in- terns of dynorphin immunoreactivity suggest mossy fiber
duced synaptic reorganization of the mossy fibers. Brain reorganization in human hippocampal epilepsy. J. Neurosci.,
Res., 573: 305–310. 10: 267–282.
Davenport, D.J., Brown, W.J. and Babb, T.L. (1990) Sprouting Hsu, M. and Buzsaki, G. (1993) Vulnerability of mossy fiber
of GABAergic and mossy fiber axons in dentate gyrus fol- targets in the rat hippocampus to forebrain ischemia. J. Ne-
lowing intrahippocampal kainate in the rat. Exp. Neurol., urosci., 13: 3964–3979.
109: 180–190. Iijima, T., Witter, M.P., Ichikawa, M., Tominage, T., Kajiwara,
Doherty, J. and Dingledine, R. (2001) Reduced excitatory drive R. and Matsumoto, G. (1996) Entorhinal-hippocampal in-
onto interneurons in the dentate gyrus after status epileptic- teractions revealed by real-time imaging. Science, 272:
us. J. Neurosci., 21: 2048–2057. 1176–1179.
Engel, J. (1989) Seizures and Epilepsy (Contemporary Neurol- Kapur, J., Haas, K.F. and Macdonald, R.L. (1999) Physiolog-
ogy Series). F.A. Davis Company, Philadelphia, PA. ical properties of GABAA receptors from acutely dissociated
Esclapez, M., Hirsch, J.C., Ben Ari, Y. and Bernard, C. (1999) rat dentate granule cells. J. Neurophysiol., 81: 2464–2471.
Newly formed excitatory pathways provide a substrate for Kesner, R.P. (this volume) A behavioral analysis of dentate
hyperexcitability in experimental temporal lobe epilepsy. J. gyrus function. In: Scharfman, H. (Ed.), The Dentate Gyrus.
Comp. Neurol., 408: 449–460. Progress in Brain Research. Elsevier, Amsterdam.
Esclapez, M., Hirsch, J.C., Khazipov, R., Ben Ari, Y. and Kobayashi, M. and Buckmaster, P.S. (2003) Reduced inhibition
Bernard, C. (1997) Operative GABAergic inhibition in hip- of dentate granule cells in a model of temporal lobe epilepsy.
pocampal CA1 pyramidal neurons in experimental epilepsy. J. Neurosci., 23: 2440–2452.
Proc. Natl. Acad. Sci. U.S.A., 94: 12151–12156. Kotti, T., Riekkinen, P.J. and Miettinen, R. (1997) Character-
Fricke, R.A. and Prince, D.A. (1984) Electrophysiology of ization of target cells for aberrant mossy fiber collaterals in
dentate gyrus granule cells. J. Neurophysiol., 51: 195–209. the dentate gyrus of epileptic rat. Exp. Neurol., 146: 323–330.
772

de Lanerolle, N.C., Kim, J.H., Robbins, R.J. and Spencer, Perez, Y., Morin, F., Beaulieu, C. and Lacaille, J.C. (1996)
D.D. (1989) Hippocampal interneuron loss and plasticity in Axonal sprouting of CA1 pyramidal cells in hyperexcitable
human temporal lobe epilepsy. Brain Res., 495: 387–395. hippocampal slices of kainate-treated rats. Eur. J. Neurosci.,
Longo, B.M. and Mello, L.E. (1997) Blockade of pilocarpine- 8: 736–748.
or kainate-induced mossy fiber sprouting by cycloheximide Ribak, C.E. and Peterson, G.M. (1991) Intragranular mossy
does not prevent subsequent epileptogenesis in rats. Neuro- fibers in rats and gerbils form synapses with the somata
sci. Lett., 226: 163–166. and proximal dendrites of basket cells in the dentate gyrus.
Lothman, E.W., Bertram, E.H., Kapur, J. and Stringer, J.L. Hippocampus, 1: 355–364.
(1990) Recurrent spontaneous hippocampal seizures in the Sayin, U., Osting, S., Hagen, J. and Rutecki, P.S.T. (2003)
rat as a chronic sequela to limbic status epilepticus. Epilepsy Spontaneous seizures and loss of axo-axonic and axo-somatic
Res., 6: 110–118. inhibition induced by repeated brief seizures in kindled rats.
Lothman, E.W., Stringer, J.L. and Bertram, E.H. (1992) The J. Neurosci., 23: 2759–2768.
dentate gyrus as a control point for seizures in the hippo- Sayin, U., Rutecki, P. and Sutula, T. (1999) Alterations in
campus and beyond. Epilepsy Res. Suppl., 7: 301–313. nmda dependent currents in granule cells of the dentate gyrus
Lowenstein, D.H., Thomas, M.J., Smith, D.H. and McIntosh, contribute to the induction but not the permanence of
T.K. (1992) Selective vulnerability of dentate hilar neurons kindling. J. Neurophysiol., 81: 564–574.
following traumatic brain injury: a potential mechanistic link Scharfman, H.E., Sollas, A.L., Berger, R.E. and Goodman,
between head trauma and disorders of the hippocampus. J. J.H. (2003) Electrophysiological evidence of monosynaptic
Neurosci., 12: 4846–4853. excitatory transmission between granule cells after seizure-
Lynch, M. and Sutula, T. (2000) Recurrent excitatory connec- induced mossy fiber sprouting. J. Neurophysiol., 90:
tivity in the dentate gyrus of kindled and kainic acid-treated 2536–2547.
rats. J. Neurophysiol., 83: 693–704. Shao, L.R. and Dudek, F.E. (2004) Increased excitatory synap-
Margerison, J.H. and Corsellis, J.A. (1966) Epilepsy and the tic activity and local connectivity of hippocampal CA1
temporal lobes. A clinical, electroencephalographic and ne- pyramidal cells in rats with kainate-induced epilepsy. J.
uropathological study of the brain in epilepsy, with particular Neurophysiol., 92: 1366–1373.
reference to the temporal lobes. Brain, 89: 499–530. Shao, L.-R. and Dudek, F.E. (2005) Changes in mIPSCs and
Meier, C.L. and Dudek, F.E. (1996) Spontaneous and stimu- sIPSCs after kainate treatment: evidence for loss of inhibitory
lation-induced synchronized burst afterdischarges in the iso- input to dentate granule cells and possible compensatory re-
lated CA1 of kainate-treated rats. J. Neurophysiol., 76: sponses. J. Neurophysiol., 94: 952–960.
2231–2239. Sloviter, R.S. (1987) Decreased hippocampal inhibition and
Meldrum, B.S. and Bruton, C.J. (1992) Epilepsy. In: Adams a selective loss of interneurons in experimental epilepsy.
J.H. and Duchen L.W. (Eds.), Greenfield’s Neuropathology. Science, 235: 73–76.
Oxford, New York, pp. 1246–1283. Sloviter, R.S. (1991) Permanently altered hippocampal struc-
Miles, R. and Wong, R.K. (1987) Inhibitory control of local ture, excitability, and inhibition after experimental status
excitatory circuits in the guinea-pig hippocampus. J. Physiol., epilepticus in the rat: the ‘‘dormant basket cell’’ hypothesis
388: 611–629. and its possible relevance to temporal lobe epilepsy. Hippo-
Molnar, P. and Nadler, J.V. (1999) Mossy fiber-granule cell campus, 1: 41–66.
synapses in the normal and epileptic rat dentate gyrus studied Sloviter, R.S. (1992) Possible functional consequences of synap-
with minimal laser photostimulation. J. Neurophysiol., 82: tic reorganization in the dentate gyrus of kainate-treated rats.
1883–1894. Neurosci. Lett., 137: 91–96.
Nadler, J.V. (1991) Kainic acid as a tool for the study of tem- Sloviter, R.S., Zappone, C.A., Harvey, B.D., Bumanglag, A.V.,
poral lobe epilepsy. Life Sci., 29: 2031–2042. Bender, R.A. and Frotscher, M. (2003) ‘‘Dormant basket cell’’
Obenaus, A., Esclapez, M. and Houser, C.R. (1993) Loss of hypothesis revisited: relative vulnerabilities of dentate gyrus
glutamate decarboxylase mRNA-containing neurons in the mossy cells and inhibitory interneurons after hippocampal
rat dentate gyrus following pilocarpine-induced seizures. J. status epilepticus in the rat. J. Comp. Neurol., 459: 44–76.
Neurosci., 13: 4470–4485. Smith, B.N. and Dudek, F.E. (2001) Short- and long-term
Parent, J.M. (this volume) Adult neurogenesis in the intact and changes in CA1 network excitability after kainate treatment
epileptic dentate gyrus. In: Scharfman, H. (Ed.), The Dentate in rats. J. Neurophysiol., 85: 1–9.
Gyrus. Progress in Brain Research. Elsevier, Amsterdam. Smith, B.N. and Dudek, F.E. (2002) Network interactions me-
Patrylo, P.R. and Dudek, F.E. (1998) Physiological unmasking diated by new excitatory connections between CA1 pyram-
of new glutamatergic pathways in the dentate gyrus of hip- idal cells in rats with kainate-induced epilepsy. J.
pocampal slices from kainate-induced epileptic rats. J. Ne- Neurophysiol., 87: 1655–1658.
urophysiol., 79: 418–429. Stables, J.P., Bertram, E., Dudek, F.E., Holmes, G., Mathern,
Peng, Z. and Houser, C.R. (2005) Temporal patterns of Fos G., Pitkanen, A. and White, H.S. (2003) Therapy discovery
expression in the dentate gyrus after spontaneous seizures in for pharmacoresistant epilepsy and for disease-modifying
a mouse model of temporal lobe epilepsy. Epilepsy Curr., 6: therapeutics: summary of the NIH/NINDS/AES Models II
57–58. Workshop. Epilepsia, 44: 1472–1478.
773

Staley, K.J., Otis, T.S. and Mody, I. (1992) Membrane prop- In: H. Scharfman (Ed.), The Dentate Gyrus. Progress in
erties of dentate gyrus granule cells: comparison of sharp Brain Research. Elsevier, Amsterdam.
microelectrode and while-cell recordings. J. Neurophysiol., Waldbaum, S. and Dudek, F. (2005) Parametric assessment of
67: 1346–1358. the consistency of the paired-pulse protocol for studying
Stringer, J.L. and Lothman, E.W. (1992) Bilateral maxi- hippocampal inhibition and hyperexcitability. Soc. Neurosci.
mal dentate activation is critical for the appearance of an Abstr., 31: 276.12.
afterdischarge in the dentate gyrus. Neuroscience, 46: Waldbaum, S. and Dudek, F.E. (2006) Increased paired-pulse
309–314. suppression after low-dose application of the selective
Sutula, T., Cascino, G., Cavazos, J., Parada, I. and Ramirez, L. GABAA receptor antagonist SR-95531, GABAzine: The
(1989) Mossy fiber synaptic reorganization in the epileptic paired-pulse technique can provide misleading evidence for
human temporal lobe. Ann. Neurol., 26: 321–330. ‘‘hyperinhibition’’ when GABAA-mediated inhibition is de-
Sutula, T.P. and Dudek, F.E. (this volume) Unmasking recur- creased. Soc. Neurosci. Abstr., 32: 794.18.
rent excitation generated by mossy fiber sprouting in the ep- Wierenga, C.J. and Wadman, W.J. (1999) Miniature inhibitory
ileptic dentate gyrus: an emergent property of a complex postsynaptic currents in CA1 pyramidal neurons after kin-
system. In: Scharfman, H. (Ed.), The Dentate Gyrus. dling epileptogenesis. J. Neurophysiol., 82: 1352–1362.
Progress in Brain Research. Elsevier, Amsterdam. Williams, P.A., Pou, P. and Dudek, F.E. (2004) Epilepsy and
Sutula, T.P. and Steward, O. (1986) Quantitative analysis of synaptic reorganization in a model of perinatal hypoxia-is-
synaptic potentiation during kindling of the perforant path. chemia. Epilepsia, 45: 1210–1218.
J. Neurophysiol., 56: 732–746. Williams, P.A., Wuarin, J.-P., Dou, P., Ferraro, D.J. and
Sutula, T.P. and Steward, O. (1987) Facilitation of kindling by Dudek, F.E. (2002) Reassessment of the effects of cyclohexi-
prior induction of long-term potentiation in the perforant mide on mossy fiber sprouting and epileptogenesis in the pi-
pathway. Brain Res., 420: 109–117. locarpine model of temporal lobe epilepsy. J. Neurophysiol.,
Tauck, D.L. and Nadler, J.V. (1985) Evidence of functional 88: 2075–2087.
mossy fiber sprouting in hippocampal formation of kainic Williamson, A. and Patrylo, P.R. (this volume) Physiological
acid-treated rats. J. Neurosci., 5: 1016–1022. studies of human dentate granule cells. In: Scharfman, H.
Toth, Z., Hollrigel, G.S., Gorcs, T. and Solesz, I. (1997) In- (Ed.), The Dentate Gyrus. Progress in Brain Research. Else-
stantaneous perturbation of dentate interneuronal networks vier, Amsterdam.
by a pressure wave-transient delivered to the neocortex. J. Wilson, D.L., Kahn, S.U., Engel, J., Isokawa, M., Babb, T.L.
Neurosci., 17: 8106–8117. and Behnke, E.J. (1998) Paired pulse suppression and
Toyoda, I. and Buckmaster, P.S. (2005) Prolonged infu- facilitation in human epileptogenic hippocampal formation.
sion of cycloheximide does not block mossy fiber sprouting Epilepsy Res., 31: 211–230.
in a model of temporal lobe epilepsy. Epilepsia, 46: Wuarin, J.P. and Dudek, F.E. (1996) Electrographic seizures
1017–1020. and new recurrent excitatory circuits in the dentate gyrus of
Traub, R.D. and Wong, R.K. (1982) Cellular mechanism of hippocampal slices from kainate-treated epileptic rats. J. Ne-
neuronal synchronization in epilepsy. Science, 216: 745–747. urosci., 16: 4438–4448.
Traub, R.D. and Wong, R.K. (1983) Synchronized burst dis- Wuarin, J.P. and Dudek, F.E. (2001) Excitatory synaptic input
charge in disinhibited hippocampal slice. II. Model of cellular to granule cells increases with time after kainate treatment. J.
mechanism. J. Neurophysiol., 49: 459–471. Neurophysiol., 85: 1067–1077.
Turski, W.A., Cavalheiro, E.A., Schwarz, M., Czuczwar, S.J., Zhang, N. and Houser, C.R. (1999) Ultrastructural localization
Kleinrok, Z. and Turski, L. (1983) Limbic seizures produced of dynorphin in the dentate gyrus in human temporal lobe
by pilocarpine in rats: behavioural, electroencephalographic epilepsy: a study of reorganized mossy fiber synapses. J.
and neuropathological study. Behav. Brain Res., 9: 315–335 Comp. Neurol., 405: 472–490.
Subject Index

a-Adrenoceptors 302 Seizure susceptibility 689–690


a-Amino-3-hydroxy-5- methylisoxazole-4- Spatial learning and memory 689
propionic acid (AMPA) 401 Allocentric coordinate 620
Aberrant neurogenesis hypothesis 536 Allopregnanolone 405
Acetylcholine 63–78, 461 Alzheimer’s disease (AD) 463–464
Acetylcholinesterase (AChE) Dentate gyrus 723–734
Reorganization after entorhinal denervation Disconnected DG 732–734
509–510 Excitatory–inhibitory systems 731–732
Activity-regulated cytoskeleton (Arc) 671 Histopathology 724–726
Arc-associated protein (Arc) as synapse Entorhinal cortex atrophy 743–745
consolidation mediator 453, 455–457 Hippocampal atrophy 741–751
Adrenalectomy (ADX) 355, 358–361 MRI 741–743
Adrenoceptors 300–302 Neuropeptide Y (NPY) and its receptors in
a-Adrenoceptors 302 292–293
b-Adrenoceptors 301 Somatostatin (SST) levels in 268–269
Adult neurogenesis in the intact and epileptic DG b-Amyloid degradation by 268
529–537 White matter changes 748–750
Hippocampal neurogenesis Amygdalo-entorhinal cortex (AE) 46
Modulation 532 Amyloid b deposition 725–726
Seizure-induced 533–534 Androgen receptors (AR) 399, 407
Integration and function of adult born DGCs Animal models of epilepsy 139, 183–195
530–531 Antidepressants (ADs)
Neonatal and adult dentate granule cell Neurogenic effects, effects on neurogenesis
neurogenesis 530 707–708
Regulation 531–533 Serotonin-dependence 708–710
Afferent and efferent connections 17, 32–34 Aplysia 309
Afferent fiber lamination formation 136–139 Apolipoprotein E receptor 2 (ApoER2) 135
Cholinergic and GABAergic afferents from Apolipoprotein E receptor 3 (ApoER3) 148
medial septum 138–139 Apoptosis 355, 357, 362
Commissural/associational (C/A) fibers Chronic stress and 364
137–138 Arachidonoylethanolamide (AEA) 321,
Entorhinal afferents 136–137 331
Non-hippocampal afferent connections 33 Archicortex 23
Afterhyperpolarization (AHP) 186 Astrocytes 513
Aging effects Astrocytic b-adrenoceptors 301
GABAergic circuitry 682–687 Astroglial cells 404, 513–514
Glutamatergic circuitry 687–689 Autoassociative memory function 579
on dentate circuitry and function 679–690 Autoradiograms 671
on dentate filter function 679–681 Axo-axonic cells 644–645

775
776

Axon sprouting 542 Anatomical connectivity from DG to 10,


Axonal degeneration techniques 44 109–111
Axonal reorganization, after entorhinal CA3 backprojection to DG 627–636 (see also
denervation 505–510 separate entry)
Lesion-induced mossy fiber (MF) sprouting
Basal forebrain inputs 18–19 99–100
Basket cells 155, 157–159, 644, 762–763, 762–764 Lesions 572–574
Basolateral amygdala 622 MF trajectory and termination in 86–97
Behavioral analysis of dentate gyrus 567–575 Pyramidal cells 577–588
Conjunctive encoding 567–568 Role in spatial pattern separation 568–572
Encoding vs. retrieval of spatial information CA3 backprojection to DG 627–636
572–574 Associative networks and reverberatory circuits
Spatial pattern separation 568–572 635
Biotinylated dextran amine (BDA) 69 Backprojection 629–633
Blood-derived cells 513 Anatomical evidence 629–630
Braak-staging 725, 733–734 Physiological evidence 630–633
Brain-derived neurotrophic factor (BDNF) 115, Functional implications 633–635
373–378, 400 Hippocampal information processing 633–634
Arc 456 Pathological conditions 634–635
Consolidation 454–456, 460–461 Ca2+ channels
Effect on neurogenesis 378 Ca2+/calmodulin-dependent protein kinase
Effects on learning and memory 378 (CaMKII) 478
Effects on synaptic transmission 377–378 Ca2+/phospholipid-dependent protein kinase
Gene regulation 375–376 (PKC) 477
in Huntington’s disease 385 in human DGCs 186
in neurodegenerative diseases 385 in mossy fiber (MF) transmission 112
in neuropsychiatric disease 385–386 Interleukin 1-b (IL-1-b) and 348–349
in pain 384–385 Cajal-Retzius cells 137, 140, 144
in Rett syndrome 385 Calbindin 728–729
Localization, transport and release 374–375 Calcitonin generelated peptide (CGRP)-
Physiologic regulation of 377 immunoreactive neurons 66, 323
Role in GABAergic synapses 377 Calretinin 226–227, 508
Roles during development 376–377 Labeling comparison in mouse and rat 227
Seizure 376 cAMP response element binding protein (CREB)
Translation control in dendrites 457–460 267, 427
Brain slice methods 111 cAMP-dependent protein kinase (PKA) 478–479
Brain-specific Na+-dependent inorganic cAMP-dependent signaling pathways 267
phosphate cotransporter (BNPI) 75–76 Carbachol (CCh) 329
BNPI/VGLUT1 75–76 CART-immunoreactive mossy cells 31–32
DNPI/VGLUT2 76–77 Catecholaminergic brainstem-dentate connections
VGLUT3 75, 77–78 71–74
Brainstem inputs 19 Serotonergic afferents 71–73
Brodman’s areas 28a and 28b 46 CB1 receptors 319–321
Endogenous ligands
CA1 (regio superior) 4–5, 418–419 (arachidonoylethanolamide) 321
CA1 interneurons 207 Expression and distribution 322, 324–326
CA2 4–5, 418–419 Coexpression with other receptor subtypes
CA3 (regio inferior) 30, 123, 207, 418–419, 422 326
777

Immunohistochemical studies 323 Anatomy and architecture 3–22, 418–421


Role in GABAergic neurons 322–324 as a filter or gate 601–611
CB2 receptors 319–321 Dentate gate vs. filter 605–607
C-fos protein 604, 769 Dentate filter function 607–608
cGMP-dependent protein kinase (PKG) 479 Behavioral analysis
Chandelier cells 155, 163 CA3 backprojection to (see CA3)
Chemokines 515–516 Cell and fiber layers development in 133–140
Cholecystokinin (CCK) 30, 101, 115, 201, 225 Cortical signals transformation in
Choline acetyltransferase (ChAT) 65–66, Extrinsic afferent systems to 63–78 (see also
138–139 Extrinsic afferent systems to DG)
Cholinergic neurons 18–19, 63–78, 133–140 Interneurons of 217–229
Chronic stress effects in dentate gyrus 362–367 Modeling 639–655
Code conversion 582–585 Neurotrophins in 371–387
Commissural/association (C/A) fibers 63–78 Norepinephrine and 299–312
Conjunctive encoding 567–568 Opioid systems in 245–258
Cornu ammonis 594, 724 Plastic processes in 417–442 (see also Plasticity)
Cortical hem Pro-inflammatory cytokines and their effects in
in DG and hippocampus development 144 339–350 (see also Pro-inflammatory
Corticosteroids, in DG 355–367 cytokines)
Dentate function after chronic stress 362–367 Depolarization-induced suppression of inhibition
Dentate function in the absence of 357–361 (DSI) 319, 327–329
Cell turnover and morphology 357–359 Depression 464–465
Physiology 359–361 NPY and its receptors in 293–294
Functional relevance of corticosteroid effects in Detonator synapses 111, 422, 433
DG 365–367 Development of dentate gyrus 133–140 (see also
Dose dependence 365–366 Reelin)
in health and disease 366–367 Afferent fiber 136–139
Physiological variations in corticosteroid levels Functional considerations 139–140
361–362 Granule cell generation 134–136
Cell turnover 361 Diacylglycerol lipase (DAGL) 321
Physiology 361–362 Diffusion tensor imaging (DTI) 750
Prolonged exposure, damage due to 355 DiI-labeled granule neurons 170
Systems activated by stress 355–357 Dopamine 63–78, 461
Corticotrophin 311 Dormant basket cell hypothesis of epilepsy 204
Cortistatin (CST) 265 Dorsal raphe nucleus (DRN) 703
Crossed entorhinodentate fibers 507, 510 Dorsolateral (DLE) entorhinal cortex 46
Cycloheximide 769 Doublecortin (DCX) 157
Cytokines 339 (see also Pro-inflammatory Dual-labeling immunohistochemical techniques
cytokines) 323
Dynorphins 113, 245–258
Dehydroepiandrosteronesulfate (DHEAS) 401 Anatomical distribution of 249–250
Dendritic changes, following entorhinal Immunoreactivity 256
denervation 514–516 (see also Location in hippocampal formation 248
Transneuronal changes) Prodynorphin-derived opioids 246
Dense-core vesicles 251 a-Neo-endorphin 246
Densitometry 671 b-Neo-endorphin 246
Dentate gyrus (DG) (see also Hippocampal DG) Dynorphin A (1–8) 246
Aging-related changes in 683–685 Dynorphin A (1–17) 246
778

Dynorphin B (1–13) 246 Adult neurogenesis in 529–537


[Leu5]-enkephalin 246 Epilepsy models, neurotrophins roles in 379–381
Leumorphin 246 NPY and its receptors in 292–293
Epileptic DG
Electroconvulsive therapy (ECT) 293 Human dentate characteristics of 185
Embryonic marginal zone 593 Mossy fiber sprouting in 541–558 (see also
Emx2 gene 146 Mossy fiber sprouting)
Endocannabinoids (ECs) in DG 319–332 Epileptogenesis 650–652, 755–770
CB1 receptors 320–321 (see also separate entry) Definition 756–757
CB2 receptor 320–321 in DG 757–760
Markers of EC system 322–327 Filtering and gating properties 759–760
Modulating glutamatergic transmission in DG Propagation of seizures 760
329–330 TLE, animal models 760–762 (see also
Non-CB receptor targets of 330–331 Temporal lobe epilepsy)
Physiological role in dentate 327–331 Histopathological alterations 761–762
Role in neurogenesis 330 Kindling 761
Enkephalins 113, 245–258 Status epilepticus 761
Anatomical distribution of 247–249 Epileptic animals
Location in hippocampal formation 248 Pharmacology and subunit expression
Proenkephalin-derived opioids 246 alterations 240
C-terminally extended forms of [Met5]- Synaptic GABAA receptor expression
enkephalin 246 Changes in 240
[Leu5]-enkephalin 246 Upregulation in 240
[Met5]-enkephalin 246 Tonic GABAA current 241
Entorhinal afferents 136–137 Zinc sensitivity 240
Entorhinal cortex 3, 45–46 Estradiol 402
Entorhinal cortex lesions (ECL) 503–519 17b-Estradiol 401, 405
Projection to DG 17–18 Estradiol treatment 256
Subdivisions 46 Estrogen receptor (ER) 399, 402
Entorhinal denervation, structural reorganization Effect on seizures 405
of DG after 501–519 ERa distribution in DG 406–408
Acetylcholinesterase (AChE) reorganization ERb distribution in DG 408
509–510 Subcellular localization 407
Axonal reorganization 505–510 Estrous cycle 402
Commissural/associational (C/A) mossy cell Eukaryotic elongation factor 2 (eEF2) 458–459
axons 505, 508 Experimental autoimmune encephalomyelitis
Crossed entorhinodentate fibers, reorganization (EAE) 342
510 Extended preparation 44
Dendritic changes 514–516 (see also Extracellular signal-regulated kinase (ERK) 480
Transneuronal changes) Extracellular unit recording 304–305
Fiber systems reorganization 505–510 Extrasynaptic GABAA receptors 235
GABAergic C/A projection 509 Alterations in animal models of epilepsy 240
Glial changes 511–514 (see also Epileptic animals)
Immune response 518 Altered zinc sensitivity 240
Neuroanatomical species differences 503–505 Composition and function of 236
Entorhinal-dentate projection 43–58 (see also Expressed by dentate granule cells 236–237
Perforant path) Upregulation in granule cells of epileptic
Epilepsy 371 (see also Epileptic DG) animals 240
779

Extrinsic afferent systems to DG 17–19, 63–78 GABAergic commissural projection 74–75


Afferent projections 17 GABAergic inhibition 548–551
Basal forebrain inputs 18–19 in human DGCs 190
BNPI/VGLUT1 75–76 GABAergic interneurons 14–16, 63–78, 164
Brainstem inputs 19 GABAergic neurons 322–324, 582, 587, 770
CA1–CA3 subfields 64 Septohippocampal cholinergic systems and
Catecholaminergic brainstem-dentate 68–69
connections 71–74 Transporters, co-localization 118–119
Commissural connections 74–75 GABA-mediated inhibition (see also
DNPI/VGLUT2 76–77 Extrasynaptic GABAA receptors)
Entorhinal cortex projection 17–18 Functional regulation of DG 235–241
Glutamatergic innervation 75 Gatekeeper function 237–240
Noradrenergic and dopaminergic afferents Tonic and phasic inhibition in dentate granule
73–74 cells 237
Septo-hippocampal connections 65–69 Genetic regulation of DG development 143–150
Supramammillary input 19, 69–71 Cortical hem 144
VGLUT3 75, 77–78 Developmental signaling systems in adult DG
150
Fadrozole 405 Emx2 146
Fatty acid amide hydrolase (FAAH) 321 Lhx5 146
Feedforward interneurons 304–305 Mutants with defects in development 148
Field excitatory postsynaptic potential (fEPSP) Neurogenic niche 147–148
359–360 Transcription factors 144–146
Filopodia 167–180 Glia
Filtering function, dentate-hilar 607–608 Glia-guided neuronal migration 134
Fimbria stimulation 630 Role of glia in synaptic remodeling 404
Flinder’s Sensitive Line (FSL) 293 Glial changes, following entorhinal denervation
Fluoxetine 293, 708 511–514
Fornix 618 Astroglia 513–514
Fractional anisotropy (FA) 750 Blood-derived cells 513
Functional cellular connectivity 202–203 Microglia 511–513
Functional significance of laminated organization Oligodendroglia and NG2-positive cells 514
139–140 Glial fibrillary acid protein (GFAP) 35, 136, 404
Fusiform cells (spiny and aspiny) in hilus Global remapping 299, 308, 312, 592
155–164 Glucocorticoid receptor (GR) 355, 357
Glutamate 113
g-EEG 299–312 Glutamatergic innervation of DG 63–78
g-Frequency oscillation 207 Glutamatergic transmission in DG
g-Oscillations 621 Transport, co-localization 118–119
GABA (g-aminobutyric acid) neurons/GABAergic Glutamic acid decarboxylase (GAD) 65, 77,
neurons 217, 603–604 217–229, 641, 765
Axonal arborizations and postsynaptic targets Double labeling 220
of 221–223 GAD expression 116–117
Distribution of 218–221 GAD mRNA-labeled neurons distribution
GABAergic afferents from medial septum 218
138–139 GAD65 219
GABAergic C/A projection after entorhinal Glycogen metabolism 299–312
denervation 509 Gonadal hormones 256, 399
780

Granule cells, dentate (DGCs) 8–11, 643–644, Connectivity of 202–203


664–666 Firing patterns 205
Apical dendrites of 156 Functional identification and activity in vivo
Biophysical properties of 664–665 199–213
Cell layer 5 Slow oscillations of 210–213
Comparative anatomy 24–28 Hilar somatotatin (SST) interneurons 269–270
Co-release of glutamate and GABA 120–121 Excitotoxicity 273–274
Density, regulation 148–150 Hippocampal information processing 633–634
Description 155–157 Hippocampal interneurons 14
Glutamate as transmitter for 11 Hippocampal mossy fibers and terminals
Golgi-impregnated granule cells 27 Supra- and infrapyramidal bands of 87–90
Granule cell association hypothesis 203 Human DG 23–28
Granule cell generation 133–140 Granule cells, physiological studies of
Compact granule cell layer formation 183–195
134–136 Huntington’s disease 385
Sites of 134 Hypothalamic-pituitary-adenocortical (HPA)
in primates 26 system 705–707
Mossy cell 13–14
Mossy fibers 9–10 IL-1 receptor-associated kinase (IRAK) 346
NPY actions 289 Infrapyramidal 134, 87–90, 134
Pyramidal basket cell 11–13 Inhibitory postsynaptic potential (IPSP) 111
Single unit recordings 665–666 Interleukin 18 (IL-18) 349–350
Tonic and phasic inhibition 237 Interleukin 1-b (IL-1-b)
Granule neuron dendrites Action in DG 345–350
Development of 168–173 Ca2+ channels 348–349
Maturation 173–176 IL-1 signalling 346
Morphological development and maturation of Synaptic plasticity 346–348
167–180 Interneurons of DG 14, 110, 157, 217–229
Primary period of 167 Classification 217
Temporal and spatial gradients of 168 GABA as primary neurotransmitter 217–229
(see also GABA)
HCN current 186–187 GABAergic interneurons 14–16
Hebbian plasticity 434–435 Intracellular labeling of 217–229
Hebb-Williams maze 571–574 Neurochemical identity 223–228
Heterosynaptic LTD 437 Intracellular labeling techniques 174, 217–229
Hilar cells with axonal projections to the perforant
path (HIPP) cells 15–16, 222, 583–584, 645 k-Opioid receptors (KORs) 246–247
Hilar commissural-association pathway-related Immunoreactivity 254
(HICAP) cells 56, 164, 223, 645–646 Kainate model 405
Hilar interneurons 303–304 Kainic acid 118, 761, 405
NPY effects on 289–290 Lateral perforant path (LPP) 299–300, 307–308,
Hilar mossy cells 199–213, 630 426, 462–463, 493–494
Anatomy 200–202 Learning and memory, BDNF effect on 378
Basic properties 200 Lef1 mutants 147
b/g-Oscillations 208–210 Lesioning techniques 503
y-Oscillations 207 Lesion-induced sprouting 85–102
g-Frequency oscillation 207 Leumorphin 246
Cellular properties of 203–204 Lhx5 gene 146
781

LIM domain kinase (LIMK) gene 455 Microglia 511–513


Locus coeruleus (LC) 300, 305–306 MicroRNAs (miRNAs) 459–460
Activation 305 Mild cognitive impairment (MCI) 742, 748–749
Glutamatergic activation of 304 Mineralocorticoid receptor (MR) 355–356
Modulation of NE 306–311 Mini mental state examination (MMSE) 732–733,
Stimulation 304–307 744
Long-term depression (LTD) 417, 436–438, 453, Modeling DG 639–655
473–488 Molecular layer perforant path-associated cells
Associative LTD 437, 438 (MOPP) 15, 56, 223
Depotentiation related to 437–438 Monoglyceride lipase (MGL) 321
Heterosynaptic LTD 429, 437, 493–494 Mossy cells 13–14, 159–161, 644 (see also Hilar
Homosynaptic LTD 438 mossy cells)
Induction, cellular mechanisms, comparison Mossy fiber sprouting in epileptic DG 541–558
473–494 (see also Seizure-induced mossy fiber
in sparse coding 438 sprouting)
mGluR-induced LTD 489–491 Mossy fibers (MFs), dentate 9–10, 58–102, 568
Long-term potentiation (LTP) 245–258, 417, Anatomy 110–111
423–436, 578, 590 Development and cognitive function
Long-term memory (LTM) 616 Development 97–98
Low-frequency stimulation (LFS) 111 Genetics and breeding experiments 99
LRP family (LRP6) 146 in CA3 (regio inferior)
Lucifer yellow 733–734 Intra- and supragranular mossy fiber collaterals
94–95
Major depressive disorder (MDD) 697 (see also Mossy fiber LTP 432
Depression) Mossy fiber sprouting 183–195
Genes, environment 703–707 Plasticity of 431–436
5-HT system and hippocampal neurogenesis Projections to CA1 (regio superior) 91–92
703–705 Projections to CA3 10
Neurogenesis, stress effects on 705–707 Projections to the hilus 9–10, 92, 95–96
Hippocampal dysfunction and atrophy 701–702 Sprouting 99–100, 765–769
Mass-associated TLE (MaTLE) 184 Structural plasticity 85–102
Maximal dentate activation (MDA) 603 Supra- and infrapyramidal 87–90
Medial (ME) entorhinal cortex 46, 591–592, 619 Synaptic transmission 109–123
Medial perforant path (MPP) 299–301, 305, Terminals 287–289, 581–582
307–308, 426 Tetanic stimulation 590
Medial septum/diagonal band of Broca (MSDB) Timm staining 93–94
65 Trajectory and termination 86–97
MSDB cholinergic innervation 65–67 Transverse and longitudinal trajectories of
MSDB GABAergic innervation 67–68 90–91
Medial temporal lobe sclerosis (MTS) 183–184, Visualization 86
189 Zinc in 33
Median raphe 63–78
Metabotropic glutamate receptors (mGluRs) a-Neo-endorphin 246
193 b-Neo-endorphin 246
Metaplasticity 438–441 Na+ currents 186
Methylazoxymethanolacetate (MAM) 89, 712 Naloxone 568
mGluR5 receptor 345, 349 NE (see Norepinephrine)
mGluR-dependent LTD 482, 489–491 Neonates 167–180
782

Neprilysin 268 Neuronal cell death, TNF-a and 341


Nerve growth factor (NGF) 115, 371–373 Neuronal migration 134
Neuroanatomical organization, DG 3–22 Neuronal network, connectivity matrix for 642
Cell types and their connectivity 8–19 (see also Neurons of DG 14–17
Granule cells; Mossy cells; Pyramidal Axo-axonic cell 15
basket cells) Long-spined cell 16
Comparative neuroanatomy 6–8 of the molecular layer 15
among rat, monkey, humans 6–8 Neurons of polymorphic cell layer 15–17
Extrinsic inputs 17–19 Neuropeptide Y (NPY) in
Layers of, 5 DG 115, 227–228, 285–294, 686
Granule cell layer 5 (see also granule cells, (see also Y1 receptors; Y2 receptors;
dentate) Y4 receptors; Y5 receptors)
Molecular layer 5 Effects on hilar interneurons 289–290
Polymorphic layer (hilus) 5, 9–10 Effects on LTP 290
Long-spined cell 16 Effects on neurogenesis 291–292
Nissl and Timm’s staining 6 Effects on synaptosomes 290–291
Septotemporal axis 5 Electrophysiological effects of 287–294
Neurochemical identity, of interneurons in the fascia dentata 286–287
223–228 in disease 292–294
Calretinin 226–227 Alzheimer’s disease (AD) 292–293
Cholecystokinin (CCK) 225 Depression 293–294
Neuropeptide Y (NPY) 227–228 Epilepsy 292–293
Parvalbumin (PV)-labeled neurons 224–225 Neurons containing NPY and their synaptic
Somatostatin 225–226 contacts 286
NeuroD mutant 149–150 Neuropil 702
Neurodegenerative diseases, BDNF role in 385 Neuroprotection 399
Neurogenesis 143–144, 355, 464, 662, 697–715 Neuropsychiatric disease, BDNF role in 385–386
and antidepressants 707 Neurotransmitter transport of human dentate
and MDD (see Major depressive disorder) granule cells 191–193
and sex steroids 402–403 Neurotrophins (NT) in DG 371–387
Antidepressant treatments, neurogenic effects of Brain-derived neurotrophic factor (BDNF)
707–708 (see also separate entry)
BDNF effect on 378 Nerve growth factor (NGF) 372–373
Endocannabinoid (EC) role in 330 Neurotrophin-3 (NT-3) 371, 378–379
NPY effects on 291–292 Neurotrophin-4/5 371, 379
Opioids and 256–257 Role in epilepsy models 379–381 (see also
Preclinical studies 710–713 Epilepsy)
Antidepressant (AD) drugs, behavioral effects Seizure regulation of 379
of 712–713 Signal transduction 372
Learning, hippocampal neurogenesis in Structure 371–372
711–712 NF-kB 340, 342
Neurogenesis blockade and depression Nissl staining 762
symptoms 711 Nissl and Timm’s staining 6
Stress based depression paradigms NMDAR/mGluR-dependent LTD 474–481
710–711 Ca2+/calmodulin-dependent protein kinase
Serotonin-dependent ADs and hippocampal (CaMKII) 478
neurogenesis 708–710 Ca2+/phospholipid-dependent protein kinase
Neurokinin B 115 (PKC) 477
783

cAMP-dependent protein kinase (PKA) 478–479 Granule cell neurogenesis 662–663


Cascades involved with 475 Granule cell numbers 662
cGMP-dependent protein kinase (PKG) 479 Granule cell synaptic contacts, changes in
Expression 476 663–664
Extracellular signal-regulated kinase (ERK) 480 Immediate early genes 670–672
Induction 475 In situ hybridization 672
Kinase pathways 477 Perforant path-granule cell synapse 666–670
Phosphatase pathways 476 Novel environment 311
Phospholipase A2 480 Nuclear translocation 134
Presynaptic vesicular release 479–480
NMDAR-dependent LTD 481–482 d-Opioid receptors (DORs) 246–247, 251–252
mGluR-dependent LTD 482 (see also separate Opioid physiology 254–255
entry) Subcellular locations of 249
N-methyl-D-aspartate receptors (NMDAR) 119, m-Opioid receptors (MORs) 246–247, 252–253
175, 401, 474–488, 567 (see also NMDAR- Offline behavioral states 579
dependent LTD; NMDAR/mGluR- Oligodendrocytes 158, 514
dependent LTD) Opioid systems in DG 245–258
Non-hippocampal afferent connections 33 b-Endorphin 246
Non-human primates, 23–38 Anatomical distribution of 247–251
Non-neuronal cells 511 Endomorphin 246
Non-self/self duality 619–620 Gonadal steroids and 256
Norepinephrine (NE) 299–312, 461 Neurogenesis and 256–257
Adrenoceptor distribution 300–302 (see also Opioid peptides 246 (see also Dynorphins;
Adrenoceptors) Enkephalins)
Hypotheses 299 Release 251
Locus coeruleus(LC)-NE modulation and Role in mossy fiber LTP 432–433
environmental events 310–312 Opioid physiology 254–255
NE innervation 300 Opioid receptors 246–247
NE-induced plasticity 306–310 ORL1 247
LC firing and NE release patterns 309–310 Prenatal morphine and 257
Pathway selectivity 307–308 Seizures and 255–256
Physiology 302–306 Subcellular locations of 249
EEG recording 305
Evoked potential recording 305–306 p55 TNFR 340, 342
Extracellular unit recording 304–305 Pain, BDNF role in 384–385
Glycogen metabolism 302–303 Parachlorophenylalanine (PCPA) 703–704
Hilar interneurons 303–304 Paradoxical TLE (PTLE) 184
Intracellular recording 303 Parahippocampal region 43–58
Release modulation by glutamate and vice-versa Parvalbumin (PV)-labeled neurons 224–225,
310 514–515
Normal aging, hippocampal granule cells in Parvalbumin-immunostained basket cells in
661–674 human DG 37
Aged granule cells Pathway selectivity, in NE-induced plasticity
Biophysical properties of 664–665 307–308
Cholinergic responses in vitro in 668 Pediatric epilepsy 756
Single unit recordings in granule cells 665–666 Perforant path 17, 43–58
DG, vulnerability of 672–673 Contralateral entorhinal-dentate projection
Granule cell dendrites 664 43–58
784

Longitudinal organization 43, 54–58 Subcellular localization of 407


Nomenclature 45–46 Pro-inflammatory cytokines, and their effects in
Radial organization in 47–54 DG 339–350 (see also Interleukin 1-b;
Synaptic organization of 56–57 Interleukin 18; Tumour necrosis factor-a)
Transection 503–519 in CNS, distribution 340
in mice 503 Projection cells
Transverse organization in 51, 53 Ultrastructure and synaptic connectivity of
Phaseolus vulgaris-leucoagglutinin (PHAL) 155–164
tracing 508, 510 Pyramidal basket cell, dentate 11–13, 158, 160
Phenylchloromethamphetamine (PCA) 704
Phosphatase pathways 476 Radial glial cells in the adult DG 157
Phospholipase A2 480 Radial glial scaffold 135
Physiological studies of human dentate granule Radial organization in perforant path 47–52,
cells 183–195 52–54
GABAergic inhibition 190 Rate remapping 592
Membrane properties 184–187 Receptor tyrosine kinases (RTKs) 372
Neuromodulation 193–195 (see also Recurrent excitation, in DG 551–558
Neuromodulation) Circuit formation 554
Synaptic properties 187–193 Emergent property 557
Picrotoxin 238 Functional effect 558
Pilocarpine 761 Inhibition state 551–553
Plaques 727 Reelin 133–140, 148, 536
Plasticity 555–556 Effect on granule cell migration 135
at mossy fiber (MF) CA3 synapse 121–122 Effect on granule cell orientation 135
Excitatory and inhibitory circuits 555–556 Effect on radial glial scaffold 135
in somatostain (SST) receptors expression Reinnervation 507
272–273 REM (rapid eye movement) 602
in the DG 417–442 Rett syndrome, BDNF role in 385
Metaplasticity 438–441
of MFs dentate mossy fibers 85–102 Schaffer collateral axons 473–494, 628
Neural grafting, testing MF growth and Seizure disorders 689–690
100–101 Hippocampal-dependent seizures 690
Sex steroids and 402–405 Seizure regulation
Synaptic plasticity 423–436 of BDNF expression 376
Polymorphic layer (hilus) 5, 9–10 of NGF expression 373
Mossy fibers 9–10 of trk receptor expression 374
Neurons of 15–17 trk receptor activation following 381–384
Postmitotic neurons 530 Seizures and hippocampal neurogenesis 533–534
Postsynaptic targets of GABA neurons 221–223 Caudal subventricular zone (SVZ) 535–536
Pregnenolone 405 Functional significance 534–536
Prenatal morphine 257 Mechanisms 536–537
Presynaptic mechanisms in synaptic consolidation Temporal lobe epilepsy (TLE) models 533
462 Seizure-induced mossy fiber sprouting
Presynaptic vesicular release 479–480 541–558
Progenitor cells 529–537 Selective serotonin reuptake inhibitor (SSRI)
Progesterone (PR) 399, 402 708–709
3a,5a-tetrahyroprogesterone 405–406 Septal fibers 18–19
Distribution in DG 409 Septo-hippocampal (SH) connections 65–69
785

GABAergic systems and, interactions between in anesthetized rats 306–307


68–69 In vitro 307
Medial septum/diagonal band (MSDB) Spinophilin 403–404
cholinergic innervation of DG 65–67 Spoiled gradient recalled (SPGR) pulse 742
MSDB GABAergic innervation of DG 67–68 Sprouting 507, 534, 541–558 (see also Mossy fiber
Septotemporal axis 5 sprouting)
Serotonin 63–78, 403, 461, 532 Cholinergic 509
Serotonin-dependent antidepressants 708 Commissural/associational fibers 507–508
Sex steroids Fiber systems 507
and dentate physiology 400–402 Reinnervation process 507
and dentate plasticity 402–405 (see also Mossy fibers (MFs), lesion-induced 99–100
Plasticity) Species difference in 517–518
and neurogenesis 402–403 Status epilepticus (SE) 533
and synaptic remodeling 403–405 Stem cell 529–530
and the DG 399–410 Stimulus/response couplet 616–617
Effect, mediation of 406–410 Stimulus-induced mGluR-dependent LTD 488
ERa distribution in DG 406–408 Stress
ERb distribution in DG 408 Exposure of rat to 356
Local steroid synthesis, role of 409–410 Systems activated by 355–357
PR distribution in DG 409 Subcortical and commissural afferents 63–78
Electrophysiological studies 401 Subiculum 43–58
in adulthood 401 Supragranular mossy fibers 94–95, 156
Neuroprotective effects 405–406 Supramamillo-dentate connections (SUM) 69–71
Subcortical mediation of sex steroid effects Supramammillary area (SUM) 63–78
410 Suprapyramidal bands
Sharp waves-ripple complexes (SPW) 207 of DG 134
Slow oscillations, of hilar mossy cells 210–213 of hippocampal mossy fibers and
Somatostatin (SST) in DG 225–226, 265–278, 686 terminals 87–90
Activity-dependent expression of 270–272 Synapse 505
as a subset of hilar interneurons 269–270 Synaptic consolidation 453–465
Cortistatin (CST) and 265 (see also Brain-derived neurotrophic factor)
Distribution 266 Extrinsic modulation of 461–462
Electrophysiological effects of 274–277 Functions and clinical implications of 463–465
in hippocampus 270 Alzheimer’s disease 463–464
Antiepileptic properties in 271 Depression 464–465
Inhibiting LTP 275 Neurogenesis 464
K+ channels activated by 266 Lateral perforant path (LPP) 462–463
Levels in Alzheimer’s disease 268–269 Mechanisms, functions, and therapeutic
SST receptor expression implications 453–465
in DG 272 Presynaptic mechanisms 462
Plasticity in 272–273 Synaptic consolidation 453
Synaptic potentiation reduction by 265 Translocation 460
Spatial information encoding vs. retrieval Synaptic plasticity in DG 299, 311–312, 423–436
572–574 in hippocampus, forms
Spatial learning 689 LTD 344
Spikes, dentate 586 LTP 344
Potentiation IL-1b and 346–348
and pairing requirements 308 TNF-a and 343–345
786

Synaptic potentiation Tianeptine 708


and spike potentiation in vitro 307 Timm staining 18, 760, 765–766
Synaptic remodeling 399 and species differences, in mossy fibers 93–94
Glia in 404 of DG 25
Sex steroids and 403–405 Nissl and Timm’s staining 6
Synaptic transmission Timm stained hippocampal sections
and TNF-a 343 of cat 88
BDNF effects on 377–378 of dog 88
Synaptic transmission, MFs of European hedgehog 88
CA3-interneuron synapses 123 of guinea pig 88
Communication from DG to CA3 area of man 88
109–123 of rat 88
‘Detonator synapse’ 111 T-maze 622
GABA transporters, co-localization 118–119 TNF receptors (TNFR) 340
Glutamate transporters, co-localization kB (NF-kB) 340
118–119 p55 TNFR 340
Long-term plasticity at 121–122 Topographical organization 43–58
Multiple neurotransmitters/modulators in Tracing techniques 508
113–117 Transcription factors,
Pre- and post-synaptic constituents 114, patterning the cortex and
119–120 hippocampus 144–146
Presynaptic modulation 117–118 Transneuronal changes
Transmission at MF-CA3 pyramidal neuron Cellular mechanisms in 515
synapse, quantal nature of 113 Chemokines role in 515–516
Transmitter release, properties of 112–113 Degeneration 512–516
Voltage clamp methods to the study 111 in granule cells 514–515
Synaptodendrosome (SD) 458 in parvalbumin-positive neurons 514–515
Synaptophysin immunohistochemistry 687 Transplants 85–102
Tri-synaptic circuit/pathway 63, 109, 417
y-EEG 303, 311 Tropomyosinrelated kinase (trk) receptors 372,
y-Oscillations 621 381–384, 460
Tau 724 Tumour necrosis factor (TNF) 339–350 (see also
Temporal lobe epilepsy (TLE) 183, 204, 240–241, TNF receptors)
533–537, 755
Animal models of 183–195 Vasopressin 311
Dormant basket cell hypothesis 762–764 Very low-density lipoprotein receptor (VLDLR)
GABAergic neurons, reorganization of 770 135, 148
Hilar damage 762 Vesicular glutamate transporters
Interneurons, loss of 764–765 (VGLUT), 75–76
Mass-associated TLE (MaTLE) 184 BNPI/VGLUT1 75–76
Mossy fiber sprouting 765–769 DNPI/VGLUT2 76–77
Axon sprouting 765–768 VGLUT1, 75–76, 325
Recurrent excitation vs inhibition 768 VGLUT2, 75
Seizure susceptibility 768–769 VGLUT3, 75, 77–78
Paradoxical TLE (PTLE) 184 Voltage sensitive dye imaging
Pilocarpine model 534–536 techniques 238
Tetrodotoxin (TTX) 454–455
Thorny excrescences 28 Wnt signaling pathway 146–148
787

Y1 receptors 285–294 Z scores 749


Y2 receptors 285–294 Zinc (Zn2+)
Y4 receptors 286 in mossy fibers 33
Y5 receptors 286 Zinc-sensitive synaptic GABAA receptors 240–241

You might also like