ArticlePDF AvailableLiterature Review

The avian subpallium: New insights into structural and functional subdivisions occupying the lateral subpallial wall and their embryological origins

Authors:
  • University of Lleida & IRB Lleida Biomedical Research Institute of Lleida

Abstract and Figures

The subpallial region of the avian telencephalon contains neural systems whose functions are critical to the survival of individual vertebrates and their species. The subpallial neural structures can be grouped into five major functional systems, namely the dorsal somatomotor basal ganglia; ventral viscerolimbic basal ganglia; subpallial extended amygdala including the central and medial extended amygdala and bed nuclei of the stria terminalis; basal telencephalic cholinergic and non-cholinergic corticopetal systems; and septum. The paper provides an overview of the major developmental, neuroanatomical and functional characteristics of the first four of these neural systems, all of which belong to the lateral telencephalic wall. The review particularly focuses on new findings that have emerged since the identity, extent and terminology for the regions were considered by the Avian Brain Nomenclature Forum. New terminology is introduced as appropriate based on the new findings. The paper also addresses regional similarities and differences between birds and mammals, and notes areas where gaps in knowledge occur for birds.
Schematic diagram of the three major, developmental subpallial proliferative zones and their derivatives, comparing mouse and chicken. The three major subpallial divisions are: lateral ganglionic eminence (LGE) in mouse, and the corresponding striatal division (St) in chicken; medial ganglionic eminence (MGE) in mouse, and the corresponding pallidal division (Pa) in chicken; and the preoptic division (PO) in mouse and commissural preoptic division (POC) in chicken. Based on differential gene expression patterns (for example, differential expression of the transcription factors Pax6, Islet1 or Nkx2.1), each major division is subdivided into several subdomains, although some differences are found in the number of subdivisions between mouse and chicken. For example, four LGE subdivisions are found in mouse, but only three striatal subdivisions appear to be present in chicken: 1) dorsal striatal (dorsal st.), 2) ventrointermediate striatal (ventroint. st.) and 3) ventrobasal striatal (ventrobas. st.) subdivisions which may be related to some of the differences found in the mature striatum of these two species. Similarly, although five MGE subdivisions are present in mouse, only three have been described in chicken: 1) dorsal pallidal (dorsal pa.), 2) ventral pallidal (ventral pa.) and 3) caudal pallidal (formerly termed the anterior peduncular area (AEP)) subdivisions. Regarding the preoptic area, two comparable subdivisions are found in mouse and chicken, although here we only refer to one of them (the commissural preoptic subdivision or POC) due to its contribution of cells to the lateral telencephalic wall. See text for more details and list of abbreviations for other abbreviations shown in figure.
… 
Content may be subject to copyright.
Review
The avian subpallium: New insights into structural and
functional subdivisions occupying the lateral subpallial wall
and their embryological origins
Wayne J. Kuenzel
a,
, Loreta Medina
b
, Andras Csillag
c
, David J. Perkel
d
, Anton Reiner
e
a
Department of Poultry Science, Poultry Science Center, University of Arkansas, Fayetteville, Arkansas 72701, USA
b
Department of Experimental Medicine, Universitat de Lleida, Institute of Biomedical Research of Lleida (IRBLLEIDA), 25008 Lleida, Spain
c
Department of Anatomy, Histology and Embryology, Semmelweis University, Faculty of Medicine, H-1094, Budapest, Hungary
BRAIN RESEARCH 1424 (2011) 67101
Corresponding author. Fax: +1 479 575 7139.
E-mail address: wkuenzel@uark.edu (W.J. Kuenzel).
Abbreviations: 6-OHDA, 6-hydroxydopamine; A, Arcopallium; A6, Locus coeruleus, noradrenergic cell group; A8, Dopaminergic cell group;
Ac, AcS, AcC, Nucleus accumbens (shell and core); AChE, Acetylcholine esterase; AEP,Anterior peduncular area; AFP, Anterior forebrain path-
way; AHi, Amygdalo-hippocampal area; AMPA, Glutamate agonist specific for AMPA-gated glutamate receptor subtype; APH, Parahippocam-
pal area; ARCO, Arcopallium; Av, Arcopallium ventrale; AVT, Arginine vasotocin; BDA, biotinylated dextran amine; BG, dBG, vBG, Basal
ganglia, dorsal, ventral BG; BO, Bulbus olfactorius, olfactory bulb; BST, BSTL, BSTLdl,vm, Bed nucleus of the stria terminalis, lateral BST(BSTL),
BSTL pars dorsolateralis (BSTLdl), BSTL pars ventromedialis (BSTLvm); BSTM1, BSTM2, Medial BST sub-nucleus 1(dorsolateralis, dl), sub-nucleus
2(ventromedialis, vm); CA, Commissura anterior, anterior commissure; CCS, Caudocentral septal area; CeA, Central amygdalar nucleus; CGRP,
Calcitonin gene-related peptide; ChAT, Choline acetyltransferase; CHCS (FiHp), Tractus cortico-habenularis and tractus cortico-septalis (fimbria
of hippocampus); CLSt, Caudolateral part of striatum; CoS, Nucleus commissuralis septi, commissural septal nucleus; CPa (HpC), Pallial commis-
sure (hippocampal commissure); CPu, Caudate putamen; CRH, Corticotropin releasing hormone; CSFcn, Cerebrospinal fluid contacting neuron/
s;cSP,SP,ChickensubstanceP;CVO,Circumventricularorgan/s;DA,Dopamine; DARPP, Dopamine and cAMP-regulated phosphoprotein; DIEN,
Diencephalon; DL CPu, Dorsolateral part of caudate putamen; DLM, Medial dorsolateral nucleus of the anterior thalamus; DMA, Anterior dor-
somedial thalamic nucleus; DVR, Dorsal ventricular ridge; DYN, Dynorphin; E14, E16, E18, Embryonic day 14, 16 and 18 in developing chick;
EA, Extended amygdala; EAce, cel, cem, Central EA, lateral part of EAce (EAcel), medial part of EAce (EAcem); EAmes, Medial EA, subpallial part;
EAp, Pallial part of the medial EA; ENK, Enkephalin; GABA, Gamma-aminobutyric acid; GAD, Glutamic acid decarboxylase; GnRH-1, Gonadotro-
pin releasing hormone-1; GP, GPe, GPi, Globus pallidus, GP externus (GP e), GP internus (GPi); H, Hyperpallium (dorsal pallium); HVC, HVC, used as
a letter based name (associative center of the caudal nidopallium); IHA, Interstitial nucleus of the apical hyperpallium; INP, Intrapeduncular nu-
cleus; IP, Interpeduncular nucleus; ITC, Intercalated amygdalar cells; LAMP, Limbic-associated membrane protein; LFB, Lateral forebrain bundle;
LGE, Lateral ganglionic eminence; LMAN, Lateral magnocellular nucleus of the anterior nidopallium; LoC, Locus coeruleus (A6), noradrenergic
cell group; LPS, Pallial-subpallial lamina; LS, Lateral septum; LSO, LSOm, LSOl, Lateral septal organ, pars medialis (LSOm), pars lateralis (LSOl);
LSt, Lateral striatum; M, Mesopallium (lateral pallium); MeA, Medial amygdala; MeAs, Subpallial medial amygd ala; MFB, Medial forebrain bundle;
MGE, MGEcv, Medial ganglionic eminence, caudalventral area; MPTP, 1-methy-4-phenyl-1,2,3,6-tetrahydropyridine; MSIB, Medial septum, inter-
nal band; MSt, Medial striatum (MSt); MStm, Medial striatum, magnocellular part; N, Nidopallium; NBM, Nucleus basalis magnocellularis (of
Meynert), basal magnocellular n; NCPa (NHpC), Bed nucleus of pallial commissure, replaced by NHpC; NDB, NDBh,v, Nucleus of the diagonal
band, horizontal limb, ventricular limb; N-III, Third cranial nerve (oculomotor nerve); NHpC, Nucleus of hippocampal commissure; NMDA, Glu-
tamate agonist specific for NMDA-gated glutamatereceptorsubtype;NOS,Nitricoxidesynthase;NPY,NeuropeptideY;NTS,Nucleustractus
solitarius; OB, Olfactory bulb; OVLT, Organum vasculosum of the lamina terminalis; PO, Preoptic area; PoA, PoAc, Posterior nucleus of the pallial
amygdala,compactdivision;POC,Commissuralpreopticarea;PPN,Pedunculopontinenucleus;Pt,Pretectum;RSv,Ventralreticularsuperior
nucleus; S, Se, Septum; SL, Nucleus septalis lateralis, Lateral septal nucleus; SM, Nucleus septalis medialis, Medial septal nucleus; SNc, Substan-
tia nigra pars compacta; SNr,Substantianigra pars reticulata; SpA, Subpallial amygdaloid area; SpL, Lateral spiriform nucleus; SPV, Supraopto-
paraventricular domain; SS, Somatostatin; SSO, Subseptal organ; ST, Stria terminalis; StC, Striatal capsule; Std, Dorsal striatal subdivision;
STN, Subthalamic nucleus (formerly, anterior nucleus of the ansa lenticularis); Stvb, Stvi, Ventral striatal subdivision, basal domain, interme-
diate domain; Supv, Subparaventricular nucleus (part of suprachiasmatic domain of embryo); TEO, Optic tectum; Th, Thalamus; TH, Tyrosine
hydroxylase; TnA, Nucleus taeniae of the amygdala, currently proposed as the subpallial medial amygdala (MeAS); TPO, Area temporo-
parieto-occipitalis; TRH, Thyroid releasing hormone; TSM, Tractus septopallio-mesencephalicus; TuO, Olfactory tubercle; Tup, Tus, Pallidal
olfactory tubercle, striatal olfactory tubercle; vaf, Ventral amygdalofugal tract (formerly occipitomesencephalic tract); VIA, Ventrointermedi-
ate thalamic area; VL, Lateral ventricle; VP, VPa, Ventral pallidum; VTA, Ventral tegmental area
0006-8993/$ see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.brainres.2011.09.037
Available online at www.sciencedirect.com
www.elsevier.com/locate/brainres
d
Departments of Biology and Otolaryngology, University of Washington, Seattle, Washington 98195-6515, USA
e
Departments of Anatomy and Neurobiology; and, Ophthalmology, University of Tennessee Health Science Center, Memphis,
Tennessee 38163, USA
ARTICLE INFO ABSTRACT
Article history:
Accepted 17 September 2011
Available online 24 September 2011
The subpallial region of the avian telencephalon contains neural systems whose
functions are critical to the survival of individual vertebrates and their species. The
subpallial neural structures can be grouped into five major functional systems, namely
the dorsal somatomotor basal ganglia; ventral viscerolimbic basal ganglia; subpallial
extended amygdala including the central and medial extended amygdala and bed nuclei
of the stria terminalis; basal telencephalic cholinergic and non-cholinergic corticopetal
systems; and septum. The paper provides an overview of the major developmental, neu-
roanatomical and functional characteristics of the first four of these neural systems, all of
which belong to the lateral telencephalic wall. The review particularly focuses on new
findings that have emerged since the identity, extent and terminology for the regions
were considered by the Avian Brain Nomenclature Forum. New terminology is introduced
as appropriate based on the new findings. The paper also addresses regional similarities
and differences between birds and mammals, and notes areas where gaps in knowledge
occur for birds.
© 2011 Elsevier B.V. All rights reserved.
Keywords:
Basal ganglia
Striatum
Pallidum
Basal forebrain
Extended amygdala
Corticopetal system
Contents
1. Introduction .......................................................... 69
2. Organization of the subpallium................................................ 70
2.1. Definition and components of the subpallium ................................... 70
2.2. Developmental origin ................................................. 70
3. Four neural systems occupying the lateral subpallial wall ................................. 75
3.1. Dorsal somatomotor basal ganglia .......................................... 75
3.1.1. Medial striatum ................................................ 75
3.1.2. Lateral striatum ................................................ 78
3.1.3. Intrapeduncular nucleus ........................................... 78
3.1.4. Globus pallidus................................................. 79
3.1.5. Functional considerations for dorsal basal ganglia . . . .......................... 79
3.1.6. Area X: A specialized songbird striatal structure. .............................. 81
3.2. Ventral viscerolimbic basal ganglia.......................................... 82
3.2.1. Olfactory tubercle ............................................... 82
3.2.2. Nucleus accumbens core and shell ...................................... 83
3.2.3. Ventral pallidum................................................ 85
3.2.4. Functional considerations for ventral basal ganglia . . . .......................... 85
3.3. Subpallial amygdaloid nuclei: The extended amygdala central and medial amygdala, and bed nuclei
of the stria terminalis ................................................. 86
3.3.1. Central extended amygdala and BSTL .................................... 86
3.3.2. Functional considerations for the central extended amygdala complex .................. 88
3.3.3. Medial extended amygdala and BSTM .................................... 89
3.3.4. Functional considerations for subpallial medial amygdala and BSTM ................... 90
3.4. Basal telencephalic cholinergic and non-cholinergic corticopetal system ..................... 90
3.4.1. Nucleus basalis magnocellularis ....................................... 91
3.4.2. Nucleus of the diagonal band ......................................... 91
3.4.3. Nucleus commissuralis septi (Commissural septal nucleus) ........................ 92
3.4.4. Functional considerations for corticopetal system . . . .......................... 92
4. Conclusions .......................................................... 92
Acknowledgments ......................................................... 92
References.............................................................. 92
68 BRAIN RESEARCH 1424 (2011) 67101
1. Introduction
A revised terminology of the avian forebrain was proposed in
2004 (Reiner et al., 2004b) as a result of an international Nomen-
clature Forum held at Duke University in July of 2002. The revi-
sion became necessary due to the longstanding existence of an
inappropriate terminology for many forebrain structures based
upon persistent but outdated and erroneous assumptions of
homology to mammalian brain structures (Jarvis et al., 2005;
Reiner, 2005; Reineret al., 2004b). Historically, a single structure,
the basal ganglia, was thought to comprise most of the fore-
brain of birds (Edinger, 1908; Elliot-Smith, 1901; Herrick, 1956).
Forty years ago, a seminal hypothesis was put forth by Harvey
Extended amygdala and BST
Central extended amygdala (SpA,BSTL)
Medial extended amygdala (MeAs,BSTM1,BSTM2)
Ventral viscerolimbic BG(TuO,AcS,AcC,VP)
Dorsal somatomotor BG(MSt,LSt,INP,GP)
NEURALSYSTEMSINTHESUBPALLIUM
Basal telencephalic corticopetal system
(NBM,NDB,CoS)
Septum and septal neuroendocrine system
(SM,SL,NHpC,LSO,OVLT,SSO)
LPS
ARCO
DIEN
Fig. 1 Schematic diagram of five neural systems comprising the avian subpallium.
Red Dorsal somatomotor basal ganglia: structures include the lateral striatum (LSt), medial striatum (MSt), globus pallidus (GP),
intrapeduncular nucleus (INP) and pallialsubpallial lamina (LPS; the dorsal border of LSt and MSt).
Tan Ventral viscerolimbic basal ganglia: structures include the olfactory tubercle (TuO), nucleus accumbens (shell and core, AcS, AcC) and
ventral pallidum (VP).
Blue Extended amygdala and bed nuclei of the stria terminalis: structures include the central extended amygdalasubpallial amygdaloid area (SpA),
striatal capsule (StC) and lateral bed nucleus of the stria terminalis (BSTL); and, the medial extended amygdalasubpallial medial
amygdala (MeAs) and medial bed nucleus of the stria terminalis (BSTM1, BSTM2).
Green Basal telencephalic cholinergic and non-cholinergic corticopetal system: structures include the basal magnocellular nucleus (NBM), diagonal
band nucleus (NDB) and commissural septum (CoS).
Yellow Septum and septal neuroendocrine systems: structures include the medial Septum (SM), lateral septum (SL), nucleus of the hippocampal
commissure (NHpC) and circumventricular organs lateral septal organ (LSO), organum vasculosum of the lamina terminalis (OVLT)
and subseptal organ (SSO).
Other abbreviations: ARCO arcopallium, DIEN diencephalon.
69BRAIN RESEARCH 1424 (2011) 67101
Karten that the basal ganglia occupied a much more restricted
basal region of the avian telencephalic hemispheres (Karten,
1969; Nauta and Karten, 1970). Numerous subsequent studies
confirmed this hypothesis, and more broadly showed that the
entire basal telencephalon (the subpallium) in birds contained
homologues of many cell groups found in the subpallial telen-
cephalon in mammals. One of the major goals of the Nomencla-
ture Forum was to rename regions of the telencephalon in birds
with terms that accurately reflect the new understanding
of their homology to mammalian telencephalic regions.
This resulted in the re-naming of nearly all structures in the
two major regions (i.e. the pallium and subpallium) of the
avian telencephalon. Since the time of the Forum, new ana-
tomical, electrophysiological, embryological, behavioral and
gene expression data have accrued on the organization and
development of the somatic basal ganglia in birds, as well
as the visceral basal ganglia, subpallial amygdala, basal fore-
brain cholinergic system, and septum. The new findings ex-
tend the understanding of these regions, especially with
respect to their function and mammalian homologues. The
present review seeks to provide an updated synthesis of the
avian subpallium, focusing on the lateral subpallial wall.
2. Organization of the subpallium
2.1. Definition and components of the subpallium
The subpallium in birds consists of telencephalic structures
ventral to the pallialsubpallial lamina (LPS), medial to the
arcopallium/pallial amygdala, and rostro-dorsal to the dien-
cephalon (Fig. 1). Subpallial structures can be organized into
five major groups (Table 1), based on developmental, topo-
graphic, neurochemical, hodological and functional criteria.
Four groups are located lateral to the ventral horns of the lat-
eral ventricles, while the fifth group resides medial to the ven-
tral horns of the lateral ventricles thereby occupying the
medial wall of subpallium (Fig. 1). The five groups and their
constituent major structures are:
1) Dorsal Somatomotor Basal Ganglia, consisting of the later-
al striatum, dorsal part of medial striatum, the globus pal-
lidus, and the intrapeduncular nucleus.
2) Ventral Viscerolimbic Basal Ganglia, consisting of the ol-
factory tubercle, ventral part of medial striatum, nucleus
accumbens, and ventral pallidum.
3) Extended Amygdala, consisting of two components: (1) central
extended amygdala which includes the so-called subpallial
amygdala, an apparent central amygdala homologue, and
the lateral bed nucleus of the stria terminalis, and, (2) the me-
dial extended amygdala comprising the subpallial medial
amygdala homologue (i.e. the subpallial nucleus taeniae) and
the medial bed nucleus of the stria terminalis.
4) Basal Telencephalic Cholinergic and Non-cholinergic Cor-
ticopetal System, consisting of: the nucleus basalis magno-
cellularis, the nucleus of the diagonal band, and nucleus
commissuralis septi.
5) Septum and Septal Neuroendocrine Systems, consisting of:
the medial septum, the lateral septum, the nucleus of the
hippocampal commissure (previously known as the bed
nucleus of the pallial commissure), the lateral septal
organ, the organum vasculosum of the lamina terminalis,
and the subseptal organ.
Of these five systems, the first four residing withinthe later-
al wall of subpallium will be reviewed here because of the close
associations with intratelencephalic circuitry, motor function,
and reward-motivated learning. The Septum and Septal Neuro-
endocrine System occurring medially between the ventral
horns of the lateral ventricles and more closely associated
with downstream diencephalon neuroendocrine functions
and related social behaviors, will be reviewed in a later paper.
2.2. Developmental origin
The subpallium in both birds and mammals forms in the basal
telencephalic anlage during development, and gives rise to the
structures of the mature basal telencephalon, which are thus re-
ferred to as subpallial structures (Cobos et al., 2001b; Holmgren,
1925; Källén, 1951, 1953, 1962; Marín and Rubenstein, 2001;
Table 1 Structures comprising the subpallium.
Structures within the lateral wall of subpallium
1. Somatic basal ganglia
a. Somatic striatum (striatal region)
i. Dorsal and ventral parts of medial striatum
ii. Lateral striatum
iii. Nucleus intrapeduncularis
iv. Area X (mixed striatal and pallidal region)
b. Globus pallidus (GP; pallidal region)
2. Limbic basal ganglia
a. Limbic striatum (striatal region)
i. Ventral part of medial striatum
ii. Nucleus accumbens shell and core
iii. Olfactory tubercle rostral and dorsolateral
b. Limbic pallidum (pallidal region)
i. Ventral pallidum
ii. Olfactory tubercle posterior and caudomedial
3. Extended amygdala complex
a. Extended central amygdala complex (mixed striatal and
pallidal region)
i. Subpallial amygdaloid area
ii. Lateral bed nucleus of stria terminalis
b. Extended medial amygdala complex (pallidal region)
i. Nucleus taeniae amygdala (subpallial medial amygdala)
ii. Medial bed nucleus of stria terminalis
4. Cholinergic corticopetal system (commissural preoptic areal
derivatives)
a. Nucleus basalis magnocellularis
b. Nucleus of the diagonal band (mixed commissural preoptic
area and pallidal region)
c. Nucleus of the septal commissure
Structures within the medial wall of subpallium
1. Septum and septal neuroendocrine system
a. Medial septum
b. Lateral septum
c. Nucleus of hippocampal commissure (formerly termed the
bed nucleus of pallial commissure)
d. Circumventricular organs: (lateral septal organ, organum
vasculosum of the lamina terminalis and subseptal organ)
70 BRAIN RESEARCH 1424 (2011) 67101
Pombero and Martínez, 2009; Puelles et al., 2000; Striedter,
1997). Current gene expression and fate mapping data indicate
that the subpallium possesses three major radially-oriented his-
togenetic zones (Fig. 2) during development: 1) a dorsal, striatal
zone termed the lateral ganglionic eminence (LGE) in mammals,
which produces striatal subdivisions of the dorsal and ventral
basal ganglia and the subpallial amygdala; 2) a ventral, pallidal
subdivision called the medial ganglionic eminence (MGE) in
mammals, which produces pallidal subdivisions of the dorsal
and ventral basal ganglia and the subpallial amygdala; and
3) the preoptic subdivision (POC) in the telencephalic stalk,
which contributes to the subpallial amygdala, and gives rise
to most of the cholinergic cells of the basal forebrain cortico-
petal system (Abellán and Medina, 2008, 2009; Cobos et al.,
2001a,b; Flames et al., 2007; García-López et al., 2008; Marín
and Rubenstein, 2001; Puelles et al., 2000, 2007; Redies et al.,
2001). Recent data in the chicken and mouse indicate that
each of these three zones is divided into further subzones,
each giving rise to different parts of the striatum, pallidum
and subpallial amygdala, as detailed below.
The entire subpallium in birds and mammals is defined by
expression of the transcription factors Dlx2/5 and the neuro-
genetic gene Mash1 (or its ortholog in other vertebrates).
These are involved in regulating the production of GABAergic
neurons, which are the predominant and defining neuron
type of the subpallium (Abellán and Medina, 2009; Garda
et al., 2002; Puelles et al., 2000; Wullimann and Mueller,
2004). By contrast, the predominant and defining neuron
type of the pallium is glutamatergic (Abellán et al., 2009).
The striatal subdivision of the developing subpallium is dis-
tinct from the pallidal and preoptic zones in that it expresses
Gsh2,Pax6 and the LIM-only gene Lmo4, in addition to Dlx2/5
and Mash1. This is particularly evident using coronal sections
of embryonic chick(c) hybridized for cLmo4 (Figs. 3CF) and c
substance P (cSP) (Figs. 3AB) which mark major striatal compo-
nents of the dorsal somatomotor and ventral viscerolimbic
basal ganglia. By contrast, the pallidal and the preoptic
zones additionally express the transcription factors Nkx2.1,
Lhx6 and Lhx7/8, with the preoptic subdivision also showing
strong ventricular expression of Sonic hedgehog (Abellán and
Fig. 2 Schematic diagram of the three major, developmental subpallial proliferative zones and their derivatives, comparing
mouse and chicken. The three major subpallial divisions are: lateral ganglionic eminence (LGE) in mouse, and the corresponding
striatal division (St) in chicken; medial ganglionic eminence (MGE) in mouse, and the corresponding pallidal division (Pa) in
chicken; and the preoptic division (PO) in mouse and commissural preoptic division (POC) in chicken. Based on differential gene
expression patterns (for example, differential expression of the transcription factors Pax6, Islet1 or Nkx2.1), each major division is
subdivided into several subdomains, although some differences are found in the number of subdivisions between mouse and
chicken. For example, four LGE subdivisions are found in mouse, but only three striatal subdivisions appear to be present in
chicken: 1) dorsal striatal (dorsal st.), 2) ventrointermediate striatal (ventroint. st.) and 3) ventrobasal striatal (ventrobas. st.)
subdivisions which may be related to some of the differences found in the mature striatum of these two species. Similarly,
although five MGE subdivisions are present in mouse, only three have been described in chicken: 1) dorsal pallidal (dorsal pa.),
2) ventral pallidal (ventral pa.) and 3) caudal pallidal (formerly termed the anterior peduncular area (AEP)) subdivisions. Regarding
the preoptic area, two comparable subdivisions are found in mouse and chicken, although here we only refer to one of them
(the commissural preoptic subdivision or POC) due to its contribution of cells to the lateral telencephalic wall. See text for more
details and list of abbreviations for other abbreviations shown in figure.
71BRAIN RESEARCH 1424 (2011) 67101
Medina 2009; Flames et al., 2007; García-López et al., 2008;
Garda et al., 2002; Puelles et al., 2000). Thus, each of the
three major subpallial histogenetic zones (striatal, pallidal
and preoptic) expresses a unique and defining combination
of genes that are thought to control development of those re-
gions. Note that the so-called caudal ganglionic eminence of
mammals, previously proposed to represent a separate sub-
pallial subdivision (Nery et al., 2002), is now thought to repre-
sent the caudal parts of both the lateral and medial ganglionic
eminences (Flames et al., 2007; García-López et al., 2008).
Studies in mice and other neurogenetic models have begun
to clarify the role of these genes and their hierarchical interac-
tions during development. Gsh1/2 and Nkx2.1 are among the
earliest transcription factors expressed in the mouse subpal-
lium, and they play key roles in patterning and specification
of the striatal and pallidal subdivisions, respectively (Sussel
et al., 1999; Yun et al., 2001, 2003). The role of Gsh1/2 in striatal
formation is evidenced by the absence or malformation of the
striatum in Gsh2-null mice (Yun et al., 2001)orGsh1/2-double
null mice (Yun et al., 2003). Similarly, the role of Nkx2.1 in pal-
lidal development is evidenced by the severe pallidal malfor-
mation in Nkx2.1-null mice (Sussel et al., 1999). Gsh1/2 and
Nkx2.1 are transcription factors that regulate the expression
of downstream transcription factors, such as Mash1,Dlx1/2,
Lhx6, and Lhx7/8 (Sussel et al., 1999; Toresson et al., 2000;
Yun et al., 2001, 2003). Mash1 and Dlx1/2 have been shown by
studies in mice to be involved in the neurogenesis and differ-
entiation of subpallial GABAergic neurons (Cobos et al., 2005;
Long et al., 2009a,b; Stühmer et al., 2002a,b), while Lhx7/8
plays a key role in the differentiation of cholinergic neurons
(Zhao et al., 2003). Most of these transcription factors are
expressed in chicken subpallium in patterns identical to
those in mouse (e.g., Dlx2/5,Nkx2.1,Lhx6,Lhx7/8). While it
seems likely their function is similar, this has not been direct-
ly demonstrated. Nonetheless, the correlated expression of
Dlx2/5 and GAD67 (a synthetic enzyme for GABA) in chicken
and Xenopus suggests that Dlx transcription factors play a
role in differentiation of GABAergic neurons in the forebrain
of nonmammalian vertebrates as well (Abellán and Medina,
2009; Brox et al., 2003). In addition, the correlated expression
of Lhx7/8 and ChAT (the synthetic enzyme for acetylcholine)
in chicken suggests that Lhx7/8 plays a role in the differentia-
tion of telencephalic cholinergic neurons in birds, much like it
does in mammals (Abellán and Medina, 2009). Moreover,
Nkx2.1-knockdown in Xenopus suggests that this transcription
factor also plays an evolutionarily conserved role in pallidal
specification and in the regulation of Lhx7/8 expression (van
der Akker et al., 2008).
The striatal and pallidal zones possess subdomains that are
characterized by expression of distinct combinations of devel-
opmental regulatory genes, with each giving rise to specificsub-
populations of subpallial neurons (Abellán and Medina, 2009;
Flames et al., 2007). In chicken, the developing striatal progeni-
tor zone includes three subdivisions. Early in development
there are only two, a dorsal and ventral, with the ventral then
later subdividing into separate ventrointermediate and ventro-
basal zones to make three distinct striatal progenitor regions.
These three striatal subdomains have been identified in
mouse as well, and mice additionally possess a fourth zone,
not evident in birds, interposed between the initial dorsal and
ventral subdomains. In both chickens and mice, the two ventral
subdivisions are the source of the bulk of the dorsal and ventral
striatum. The dorsal subdivision is characterized by expression
of the transcription factor Pax6 in both mitotic and postmitotic
cells, and the ventral subdivision expresses the transcription
factor Islet1 in the subventricular zone and mantle (Abellán
and Medina, 2009; Puelles et al., 2000; Stenman et al., 2003;
Yun et al., 2001). Based on the expression of Pax6 and correlation
with data on mammals, the dorsal striatal subdivision also
appears to be the source of GABAergic and dopaminergic in-
terneurons that migrate into the olfactory bulb (Abellán and
Medina, 2009). The dorsal striatal subdivisions may thus
also be the source of catecholaminergicdopaminergic neurons
in the olfactory tubercle of some birds (Abellán and Medina,
2009; Roberts et al., 2002), and in the striatum of mice and pri-
mates (Huot and Parent, 2007; Marín et al., 2005). The dorsal
striatal subdomain also produces part of the lateral striatum
in chicken (Abellán and Medina, 2009), and similarly in mouse
Fig. 3 AF present low-magnification digital images of frontal
sections of the telencephalon of chicken at prehatching stages
(E18) or at hatching (P0), hybridized for cSP (A, B), cLmo4 (CF).
The images show the major subdivisions of the developing
somatic and limbic basal ganglia, with striatal subdivisions
enriched in cSP and cLmo4. Scale bar=1 mm in A (applies to A,
B), and Scalebar=1 mm in C (applies to CF). AcC = accumbens
core; AcS = accumbens shell; StC = striatal capsule; Tup =
pallidal olfactory tubercle; Tus = striatal olfactory tubercle.
From Abellán and Medina (2009). See list of abbreviations for
other abbreviations shown in this figure.
72 BRAIN RESEARCH 1424 (2011) 67101
some neurons of the dorsolateral striatum (Toresson and
Campbell, 2001; Yun et al., 2003). At caudal hemispheric levels
in both chicken and mouse, this dorsal striatal subdivision
also appears to produce Pax6-expressing striatal neurons for
the central amygdala in both chicken and mouse (Abellán and
Medina, 2009; Puelles et al., 2000; Tole et al., 2005). Finally, a
thin group of neurons directly below the palliosubpallial lamina
also appears to be part of the dorsal striatal subdivision. This re-
gion distinctly expresses Lmo4 (Abellán and Medina, 2009)and
the Enc1 gene (ectodermal and neural crest cortex), a homolog
of Kelch aDrosophila melanogaster gene essential for oogenesis
(García-Calero and Puelles, 2009; Puelles et al., 2007), and it
overlaps the subpallial area expressing Pax6. Since intercalated
cell masses of the mammalian amygdala derive from the dorsal
LGE based on gene expression and fatemap data (Kaoru et al.,
2010; Tole et al., 2005; Waclaw et al., 2010), the Lmo4-expressing
striatal capsular region in chicken has been proposed to be ho-
mologous to the amygdaloid intercalated cell masses of mam-
mals (Abellán and Medina, 2009).
Based on genetic fate mappingof Islet1 progenitors in mouse,
most neurons of the striatal part of the basal ganglia (including
the majority of projection neurons of dorsal and ventral stria-
tum) derive from the two parts of the ventral striatal subdivi-
sion or ventral LGE (Stenman et al., 2003; Waclaw et al., 2010;
reviewed in Medina and Abellán, 2009), and this appears to be
true in chicken as well (Abellán and Medina, 2009). In contrast,
most striatal interneurons in mouse and chicken derive from
the pallidal or preoptic subpallial progenitor zones (Abellán
and Medina, 2009; Cobos et al., 2001a; Marín et al., 2000;see
below). The striatal part of the developing basal ganglia in
birds, however, differs from that in mammals in the number
of regions making up the striatal progenitor zone. As noted
above, this region of the developing subpallium in birds comes
to consist of separate dorsal, ventrointermediate and ventroba-
sal subdomains, while inmammals an additional zone isinter-
posed between the dorsal and two ventral striatal zones.
This additional zone may giverise to the striosomal dorsal stria-
tal compartment in mammals, which may explain why this
compartment seems absent in birds (Abellán and Medina,
2009; Flames et al., 2007). The ventrointermediate striatal-
equivalent in mammals (LGE3 in mouse) then gives rise to the
projection neurons of the matrix compartment of dorsal stria-
tum, and in birds to the dorsal medial striatum and lateral stri-
atum. The ventrobasal striatal zone (corresponding to LGE4 in
mouse), in turn, gives rise to nucleus accumbens in birds and
mammals (Abellán and Medina, 2009). The ventrobasal zone
also appears to produce the INP in birds (Abellán and Medina,
2009).
The pallidal progenitor zone of mouse and chicken also
has molecularly distinct subdivisions (Abellán and Medina,
2009; Flames et al., 2007; García-López et al., 2008). Again, the
development of the pallidal subpallial regions is more com-
plex in mouse than in chicken, with mouse having five dis-
tinct progenitor subdomains and birds only three evident
ones (Abellán and Medina, 2009; Flames et al., 2007), although
one of the avian pallidal progenitor subdomains appears to
correspond to two MGE zones. Of those in mammals, the dor-
salmost seems absent in birds, and the second and third MGE
zones give rise to most neurons of the globus pallidus, and
correspond to the dorsal pallidal zone of chick, which also
gives rise to globus pallidus. As will be discussed below,
these various pallidal zones also contribute to the bed nucleus
of the stria terminalis and extended amygdala. In addition, re-
cent data indicate that the adult mouse globus pallidus in-
cludes a neuron subpopulation that originates in the striatal
progenitor zone (Nóbrega-Pereira et al., 2010). The neurons of
this subpopulation of the globus pallidus do not contain par-
valbumin, although they are GABAergic (Nóbrega-Pereira et
al., 2010), and likely represents a striatal-type subpopulation
of projection neurons that contain calbindin and enkephalin,
and project back to the striatum (Hoover and Marshall, 2002;
Medina and Abellán, 2009). A similar neuron subpopulation
also appears to be present in the chicken globus pallidus
(Abellán and Medina, 2009; Molnaret al., 1994). The mammalian
ventral pallidum possesses a rostral, precommissural part that
arises from MGE4, and a commissural ventral pallidum that
arises from MGE2/3 (Abellán and Medina, 2009; García-López
et al., 2008). The avian ventral pallidum, however, appears to
arise only from the MGE4-equivalent (i.e. the ventral pallidal
zone), and is thus homologous only to the precommissural ven-
tral pallidum (Abellán and Medina, 2009; García-López et al.,
2008). Both mouse and chicken possess a caudoventral pallidal
subdivision (MGEcv), which is included as a caudal part of the
fifth progenitor subdomain of the medial ganglionic eminence
(pMGE5; Flames et al., 2007), and is sometimes called the anteri-
or peduncular area (AEP) (Abellán and Medina, 2009; García-
López et al., 2008), that appears to giverise to part of themedial
extended amygdala, as discussed in more detail in a later
section.
Based on Sonic hedgehog (Shh) expression, the commissural
preoptic area (POC) of both mouse and chicken also appears to
contribute cells to the subpallial medial amygdala and other
parts of the medial extended amygdala, and this has been
shown experimentally in the mouse (Bupesh et al., 2011a).
Pallidal (MGEcv) and POC cells intermingle in the subpallial
medial amygdala of chicken, resembling in this respect the
anterior part of the medial amygdala of mammals (Abellán
and Medina, 2009; García-López et al., 2008). Recent data indi-
cate that the Lhx7/8 transcription factor is particularly highly
expressed in the POC of mouse and chicken (including the
ventricular zone where most progenitor cells are located)
and that this subdivision may be the source of many cholin-
ergic cells of the basal forebrain, including the corticopetal
neurons of the basal magnocellular nucleus (NBM) and
those overlapping the pallidum, as well as most of the cholin-
ergic interneurons of the striatum (Abellán and Medina, 2008;
2009; García-López et al., 2008). Data in mouse indicate that
Lhx8 (also called Lhx7/8)isrequiredfordevelopmentofmost
cholinergic neurons of the telencephalon, since they are
missing in Lhx7/8 knockout mouse (Manabe et al., 2007;
Zhao et al., 2003). The influence of Lhx7/8 is partially mediat-
ed by the transcription factor Islet1, which is expressed in sub-
ventricular and postmitotic cells of POC. This interpretation is
consistent since POC-derived cholinergic neurons of basal fore-
brain, pallidum and striatum are absent in conditional
forebrain-specific Islet1-null mouse (Elshatory and Gan, 2008).
Because most cholinergic neuronsof the medial septum/diago-
nal band nuclei are preserved in conditional Islet1-null mice
(Elshatory and Gan, 2008), these may derive from the rostroven-
tral MGE, a subdomain that expresses Lhx7/8 in its ventricular
73BRAIN RESEARCH 1424 (2011) 67101
and subventricular zones, but not Islet1, in mouse and chicken
(Abellán and Medina, 2009; Flames et al., 2007; reviewed in Me-
dina and Abellán, 2009).
In summary, developmental and anatomical data suggest
the following categorization of lateral subpallial structures
(Table 1). Striatal regions derived from the striatal progenitor
zone are the lateral and medial striatum, accumbens core and
shell, intrapeduncular nucleus, part of the central extended
amygdala, and part of the olfactory tubercle. The pallidal regions
derived from the pallidal progenitor zone are the globus palli-
dus, ventral pallidum, part of the central and medial extended
amygdala, and some neuron subpopulations of the diagonal
band nuclei. The POC derivatives of the lateral subpallial wall
are the basal magnocellular nucleus, the cholinergic neurons
Fig. 4 AD. A series of schematic line drawings of midtelencephalic transverse brain sections of pigeon and rat. A. The
outdated interpretation of the organization of the telencephalon in birds and the outdated nomenclature that view engendered
for the telencephalon of birds. B. The longstanding interpretation of mammalian telencephalic organization and the
established nomenclature consistent with that view. C. The current interpretation of the organization of avian telencephalon
and the outdated avian telencephalic nomenclature, which highlights the inappropriateness of this nomenclature. D. The
current interpretation of the organization of avian telencephalon and new avian telencephalic nomenclature adopted by the
Avian BrainNomenclature Forum that is consistent with current findings on telencephalicorganization in birds. In each schematic
interpretation of telencephalic organization, the speckled region represents pallium, the striped region represents striatum, and
the checked region represents globus pallidus. The new avian terminology (D) avoids the erroneous misimpressions about
correspondences between avian and mammalian telencephala perpetuated by the old nomenclature. Hp = hippocampus.
From Reiner et al. (2004a, 2004b).
74 BRAIN RESEARCH 1424 (2011) 67101
invading the pallidum and striatum, and part of the medial ex-
tended amygdala (Abellán and Medina, 2008, 2009; García-López
et al., 2008; Marín and Rubenstein, 2001). Although cells of each
subpallial nucleus/area primarily derive and primarily migrate
radially from a specific striatal, pallidal or preoptic subdivision,
there are important migrations that cross subdivision bound-
aries within the subpallium, or cell immigration from outside
the subpallium, leading to a mixed cellular composition in
most subpallial areas (Abellán and Medina, 2009; Cobos et al.,
2001a,b). This is the case for some components of the extended
amygdala, as will be discussed in more detail below. A strik-
ing case of tangential cell movement inside the avian subpal-
lium is the massive migration of cells from the pallidal
proliferative zone into the medial striatum demonstrated in
chickens (Abellán and Medina, 2009). This will be discussed
in more detail in the sections on the striatal part of the
somatic basal ganglia and in the section on area X, a striatal
nucleus of songbirds that contains both striatal and pallidal
neurons (Farries and Perkel, 2002; Person et al., 2008; Reiner
et al., 2004a). Finally, in both chicken and mouse, the subpal-
lium is the source of tangential migrations to the pallium,
providing it with various neurochemically distinct subpopu-
lations of GABAergic interneurons (Abellán and Medina,
2009; Cobos et al., 2001a, b; Marín and Rubenstein, 2001,
2003). The subpallium also appears to produce some cholin-
ergic interneurons for the pallium (Abellán and Medina,
2009).
Hence, the developing vertebrate subpallium, at least in
tetrapods, appears to commonly possess the same three
major progenitor zones or their correspondents: striatal, palli-
dal and POC (Fig. 2). Markers useful in identifying progenitor
zones have likewise been shown to give rise to homologous
subpallial cell groups of the lateral telencephalic wall in
birds and mammals, as well as reptiles and amphibians
(Bachy et al., 2002; Brox et al., 2003; Fernández et al., 1998;
González et al., 2002a,b; Moreno et al., 2004, 2008a,b,c, 2009).
In the following sections, we review recent progress on the
dorsal basal ganglia, the ventral basal ganglia, the subpallial
amygdala/extended amygdala, and the basal forebrain corti-
copetal system of birds.
3. Four neural systems occupying the lateral
subpallial wall
3.1. Dorsal somatomotor basal ganglia
Discovery in the 1960s of the abundance of dopamine and ace-
tylcholinesterase in the avian subpallium began to reshape
understanding of the location and extent of the avian basal
ganglia (Juorio and Vogt, 1967; Karten, 1969; Nauta and Karten,
1970; Spooner and Winters, 1966). The outdated and current
understanding of components of the avian telencephalon in
comparison to that of mammals is shown in Fig. 4. Immunos-
taining with an antibody to tyrosine hydroxylase (TH) clearly
demonstrated that the avian basal ganglia did not occupy
the majority of the telencephalon as suggested by Nissl stain-
ing (Figs. 4 and 5C) rather it occupied the ventromedial region
similar to mammals (Figs. 4B, D). The avian dorsal somatomo-
tor basal ganglia and its major structures in relation to the
pallium are shown in Figs. 4D and 5. Details of each of its com-
ponents (medial striatum, lateral striatum, globus pallidus,
and intrapeduncular nucleus) which have been more exten-
sively studied than other parts of the avian subpallium are
reviewed below. In this section, we also discuss area X of the
songbird medial striatum.
3.1.1. Medial striatum
The medial striatum (MSt) together with the lateral striatum
(LSt) has been known for many years to share numerous neu-
rochemical, developmental and hodological traits with the
mammalian dorsal striatum, and on this basis (including the
Fig. 5 Images of transverse sections of pigeon brain
immunolabeled for A. Substance P (SP) and B. Choline
acetyltransferase (ChAT). Transverse section of chicken brain
immunolabeled for C. Tyrosine hydroxylase (TH). Note the
enrichment of the ventral pallidum (VP) in SP fibers and ChAT
neurons and the paucity of SP fibers and relative paucity of
ChAT neurons in the lateral part of the bed nucleus ofthe stria
teriminalis (BSTL). In B, the field of cholinergic neurons
spanning the VP and lateral forebrain bundle (LFB) represents
the basal magnocellular cholinergic cell group (NBM). Image C
shows the enrichment of striatal parts of the basal ganglia in
fibers containing TH, and identifies the parts of the dorsal
somatic basal ganglia. Abbreviations: GP = globus pallidus;
LSt lateral striatum; MSt = medial striatum; SL = lateral
septal nucleus; TSM = tractus septopallio-mesencephalicus.
Scale bar= 1 mm in A and C (scale bar in A applies to A, B). See
list of abbreviations for other abbreviations shown in this
figure.
75BRAIN RESEARCH 1424 (2011) 67101
presence of similar structures in reptiles) considered to be ho-
mologous to the mammalian caudate-putamen (Reiner et al.,
2004a). The MSt and LSt differ in their connectivity, and are
not considered strictly homologous to caudate and putamen,
respectively, in a one-to-one manner. Thus, the MSt and LSt
need to be discussed separately. In addition to embryological
origin, shared traits that show MSt and LSt are striatal which
include a neuropil rich in acetylcholinesterase (AChE), choline
acetyltransferase-containing terminals (ChAT; Medina and
Reiner, 1994), dopamine (DA) containing terminals (revealed
by TH immunostaining), substance P (SP) containing peri-
karya and processes, and enkephalin (ENK) containing peri-
karya and processes (Fig. 5; reviewed in Reiner et al., 1998a).
Concordant with its rich dopaminergic innervation, MSt and
LSt have also been shown to be rich in D1
A
,D1
B
, and D2 dopa-
mine receptors (Casto and Ball, 1994; Dietl and Palacios,
1988; Kubikova et al., 2010; Richfield et al., 1987; Schnabel
and Braun, 1996; Schnabel et al., 1997; Stewart et al., 1996; Sun
and Reiner, 2000), and the dopamine receptor signaling pro-
tein DARPP32 (Bálint et al., 2004; Reiner et al., 1998b). Associ-
ated with its rich cholinergic innervation, MSt and LSt are
rich in muscarinic receptors (Dietl et al., 1988; Kohler et al.,
1995; Wächtler and Ebinger, 1989). In these regards, MSt re-
sembles mammalian striatum, as well as in the expression
of the developmentally regulated genes noted above.
One underlying reason for the neurochemical similarity be-
tween avian MSt and mammalian striatum is that they consist
of the same major neuron types. This includes GABAergic pro-
jection neurons, which make up the bulk of MSt neurons.
These projection neurons possess spiny dendrites, and about
half also contain enkephalin while the other half show co-
localization of substance P (SP) and dynorphin (Anderson and
Reiner, 1990a, 1991; Brauth et al., 1983; Reiner and Anderson,
1990; Reiner et al., 1983, 1984; Sun et al., 2005; Veenman and
Reiner, 1994). These neurons account for the SP+ and enkepha-
linergic projections of MSt to the dopaminergic neurons of the
substantia nigra pars compacta and the GABAergic neurons of
the substantia nigra pars reticulata (Anderson et al., 1991; Medina
et al., 1995; Veenman and Reiner, 1994), of which the SP+ projec-
tion is more prominent. The MSt neurons projecting to the sub-
stantianigraandtheventraltegmentalareasendtheiraxons
through the ventral pallidum, and some have terminations
there (Kitt and Brauth, 1981; Medina and Reiner, 1997; Person
et al., 2008). Interneurons make up most of the remainder of
MSt neurons, although the percent projection neuron versus
interneuron composition of MSt is not certain. The interneurons
include: (1) large, aspiny cholinergic neurons that contain choline
acetyltransferase (Medina and Reiner, 1994); (2) medium-sized
aspiny GABAergic interneurons co-localized with somatostatin
and neuropeptide Y (NPY) (Anderson and Reiner, 1990b); and
(3) medium-sized aspiny neurons containing GABA, the
calcium-binding protein parvalbumin, and the neurotensin-
related hexapeptide LANT6 (Reiner and Anderson, 1993; Reiner
and Carraway, 1987). These neurons appear to have similar elec-
trophysiological properties to their mammalian counterparts
(Farries and Perkel, 2000; Farries et al., 2005b). In mammals, a
fourth type of interneuron contains GABA and the calcium bind-
ing protein calretinin (Bennett and Bolam, 1993; Figueredo-
Cárdenas et al., 1996),andthiscelltypeisrareorabsentinbird
MSt (Laverghetta et al., 2005).
One cell type that is present in MSt but absent in mamma-
lian striatum is a neuron type that appears pallidal in its de-
velopmental derivation, neurochemistry, connectivity and
physiology. As noted in the section above, during develop-
ment a massive migration of cells occurs from the pallidal
subdivision into the medial striatum in chickens. These cells
can be recognized as pallidal since they express Nkx2.1,Lhx6
and/or Lhx7/8, and they are distributed throughout MSt and
continuous with the far more concentrated pallidal neurons of
the globus pallidus, which have a similar gene expression sig-
nature (Abellán and Medina, 2009). Some of these neurons de-
rived from the pallidal proliferative zone are likely to constitute
striatal interneurons, such as those containing parvalbumin/
LANT6 or somatostatin/neuropeptide Y (Carrillo and Doupe,
2004; Cobos et al., 2001a). The neurons expressing Nkx2.1,
Lhx6 and/or Lhx7/8 are, however, more numerous than the par-
valbumin/LANT6 or somatostatin/neuropeptide Y interneuron
populations (Abellán and Medina, 2009; Anderson and Reiner,
1990b; Reiner and Carraway, 1987), and at least some of them
may be those shown to project to intralaminar thalamus in
chickens and zebra finches, and possess pallidal electrophysiol-
ogy (Farries and Perkel, 2002; Farries et al., 2005b; Person et al.,
2008; Reiner et al., 2004a). Given the robustness of this trait in
these two avian species, it seems likely that it is a general
trait of MSt in birds. Consistent with this, the MSt in pigeons
is rich in GABAergic woolly fibers that also contain either SP
or enkephalin, especially along its medial wall (Reiner et al.,
1983, 1984, 1998b). The woolly fiber pattern of SP+ and enke-
phalinergic terminal distribution is characteristic of pallidal
neurons but not striatal interneurons. The putative intra-MSt
pallidal neuron dendrite targets of this woolly fiber input are
not rich in parvalbumin or LANT6 (Reiner and Carraway,
1987), and thus the pallidal neurons receiving this input in
MSt have not been reported previously. The developmental
data together with the neuropeptide immunolabeling data sug-
gest that pallidal neurons may be numerous and widespread in
MSt, but low in the conventional pallidal neuron markers par-
valbumin or LANT6, at least in adults. While it is known that
some of these pallidal MSt neurons project to intralaminar
thalamus, it is unknown if they have other projection targets
as well.
The MSt also evidences a medial-lateral difference in
neurochemical traits and connectivity. For example, medial
MSt receives its dopaminergic input from the A10 dopami-
nergic neurons of the ventral tegmental area (VTA, Figs. 6A,
B) (Bailhache and Balthazart, 1993; Kitt and Brauth, 1986b;
Puelles and Medina, 1994; Reiner et al., 1994), while more lat-
eral MSt receives its dopaminergic input from the A9 dopami-
nergic neurons of the substantia nigra pars compacta (SNc,
Figs. 6A, C; Bailhache and Balthazart, 1993; Durstewitz et al.,
1998, 1999; Karle et al., 1996; Kitt and Brauth, 1986b; Medina
and Reiner, 1995; Metzger et al., 1996; Puelles and Medina,
1994; Reiner et al., 1994, 1998a,b; Smeets and Reiner, 1994;
Székely et al., 1994). Similar parcellation has been observed
in the preferential origin of the projections to the ventral
tegmental area or substantia nigra from the medial versus
lateral MSt, respectively (Mezey and Csillag, 2002). The MSt
also receives major excitatory afferents from pallial regions.
As in mammals, the corticostriatalprojection utilizes gluta-
mate, an excitatory amino acid neurotransmitter (Csillag et
76 BRAIN RESEARCH 1424 (2011) 67101
al., 1997; Ding and Perkel, 2004; Ding et al., 2003; Farries et al.,
2005a; Reiner et al., 2001; Veenman and Reiner, 1996). More
lateral MSt receives pallial input from somatic regions, such as
those involved in somatosensory, visual, auditory and motor
function (Brauth et al., 1978; Karten and Dubbeldam, 1973; Notte-
bohm et al., 1976; Veenman et al., 1995; Wild, 1987; Wild et al.,
1993). By contrast, medial MSt appears more viscerolimbic,
since its pallial input arises from such regions as hippocampus
and olfactory bulb. Similarly, its excitatory thalamic input also
arises from midline, more viscerolimbic intralaminar nuclei
than does that to lateral MSt (Veenman et al., 1997). Thus, the
two parts of MSt may differ in function, with medial MSt more
viscerolimbic than lateral MSt. One possibility is that medial
MSt is comparable to mammalian striosomes, since both are
poor in calbindin (Bálint and Csillag, 2007; Roberts et al., 2002).
Abellán and Medina (2009), however, have suggested that birds
may lack the part of the dorsal striatal proliferative zone that
gives rise to striosomes. Thus, the medial MSt viscerolimbic
traits may be comparable to those of medial mammalian cau-
date. Nonetheless, MSt seemingly differs from mammalian cau-
date in that its striatal neurons mainly project to the nigra and
very few to globus pallidus. Additionally, a ventral and caudal
part of what has been termed MSt may be part of the viscerolim-
bic striatum comparable to part of the shell of nucleus
Fig. 6 Sources of afferent inputs to the basal ganglia in chick brain. A. Brain atlas plate A3.4 (Kuenzel and Masson, 1988).
Boxed areas show locations of images 6B and 6C. B. Ventral tegmental area (VTA) or A10 dopaminergic cell group. Scale
bar= 200 μm. C. Substantia nigra pars compacta (SNc) or A9 dopaminergic cell group. Scale bar =200 μm. D. Chick brain atlas
plate A2.4; boxed area shows location of photomicrograph 6E. E. A8 dopaminergic cell group. Scale bar=200 μm. F. Chick brain
atlas plate A1.4. Boxed area shows location of photomicrograph 6G. G. Locus coeruleus or A6 noradrenergic cell group. Scale
bar= 100 μm. Neurons in figure immunolabeled with antibody to tyrosine hydroxylase.
77BRAIN RESEARCH 1424 (2011) 67101
accumbens (Abellán and Medina, 2009). This region, however,
may not be strictly equivalent to mammalian accumbens shell
since unlike shell it is very rich in cholinergic neurons, dopami-
nergic terminals and SP+ and ENK+ woolly fibers. Accumbens
core and shell are discussed in more detail in the viscerolimbic
section.
3.1.2. Lateral striatum
As noted above, the LSt of birds possesses the neurochemistry
and many of the neuron types characteristic of mammalian
dorsal striatum, including neuropil enrichment in AChE, cho-
line acetyltransferase-containing terminals, SP+ neurons, enke-
phalinergic neurons, D1, D1
A
,D1
B
, and D2 dopamine receptors,
DARPP32 and muscarinic receptors. As true of MSt, the dopa-
mine receptors and DARPP32 enrichment in LSt reflect the
prominent dopaminergic input, in this case from the substantia
nigra pars compacta (Figs. 6A, C) and A8 (Figs. 6D, E) tegmental
cell groups. The LSt, except for its lateral edge, which may be-
long to the olfactory tubercle, receives its excitatory input main-
ly from somatic pallial and thalamic regions (Veenman et al.,
1995, 1997). This feature of its input is consistent with its output
circuitry to the globuspallidus, which gives rise to an output to
motor pallium in the Wulst (thought to be an M1 homologue)
via the avian motor thalamus (the ventrointermediate area, or
VIA). The LSt neurons giving rise to this are likely to be SP+,
and one of the so-called direct pathwayoutputs of the stria-
tum for facilitating movement (Jiao et al., 2000). The enkephali-
nergic neurons of the LSt, by contrast, have been shown to give
rise to the indirect pathwayout of the avian striatum, which
projects to pallidal neurons innervating thesubthalamic nucle-
us (Jiao et al., 2000; Person et al., 2008). This circuit appears to be
involved in suppressing unwanted movements, as also the case
in mammals (Jiao et al., 2000). With respect to its motor role as
well as its presumed output to pallidal neurons projecting to
intralaminar thalamus, the LSt may be comparable to the dor-
solateral mammalian striatum. Moreover, the striatal neurons
of primate dorsolateral striatum more heavily project to the
globus pallidus than to nigra (vesque and Parent, 2005; Parent
et al., 1995), as is also true of avian LSt. While at one time it
seemed the avian LSt had no projections to the substantia
nigra, recent evidence shows that some LSt neurons project to
substantia nigra pars compacta and pars reticulata (Mezey and
Csillag, 2002; Person et al., 2008).
Although the LSt contains most of the same striatal neuron
types as the MSt, there are some notable differences. For exam-
ple, while the LSt contains abundant SP+ spiny projection neu-
rons, ENK+ spiny projection neurons and parvalbuminergic
interneurons, immunolabeling suggests that LSt contains far
fewer cholinergic and SS/NPY interneurons than MSt (Anderson
and Reiner, 1990b; Medina and Reiner, 1995; see Fig. 6 and 13 in
Abellán and Medina, 2009) or they may even be absent (Person
et al., 2008). Nonetheless, the LSt is rich in cholinergic terminals
and muscarinic receptors (Dietl et al., 1988; Kohler et al., 1995;
Medina andReiner, 1995; Wächtler and Ebinger, 1989). Addition-
ally, in situ hybridization for ChAT mRNA in day-old hatching
chicks reveals some LSt cells positive for ChAT (Abellán and
Medina, 2009), although the labeling is weaker than in MSt. It
may be that immunostaining does not effectively label the
somata of these cells or perhaps there is a developmental
change in ChAT expression. Regardless, it is possible that the
sparse and weaklylabeled somata of LSt give rise to a densecho-
linergic innervation. Alternatively, cholinergic neurons of MSt,
globus pallidus or the intrapeduncular nucleus (see below)
could provide the cholinergic innervation to LSt. Cholinergic in-
nervation is important in regulating striatal projection neuron
function and synaptic plasticity (Kreitzer and Malenka, 2008).
The paucity of unambiguous resident cholinergic neurons in
LSt (i.e. motor striatum) in birds indicates that the role of cholin-
ergic neurons in striatal plasticity may differ from that in mam-
mals. The functional implication of the apparent absence of
somatostatinergic/NPY+ interneurons in LSt is uncertain.
3.1.3. Intrapeduncular nucleus
The intrapeduncular nucleus (INP), located below the inferior
margin of the avian globus pallidus, is named for its location
within the lateral forebrain bundle. Karten and Dubbeldam
(1973) originally thought that its position resembled that of the
mammalian internal pallidal segment, but subsequent immuno-
labeling studies showed that it lacked the pallidal-type neurons
and the SP/DYN-containing striatal input characteristic of the in-
ternal pallidal segment (Anderson and Reiner, 1990a; Reiner and
Carraway, 1987; Reiner et al., 1983, 1999; Veenman and Reiner,
1994). More recent studies have shown that the INP contains
densely packed GABAergic spiny neurons that express
DARPP32 (Reiner et al., 1998b; Schnabel et al., 1997; Sun et al.,
2005), and a very similar glutamate receptor profile to the stria-
tum (Wada et al., 2004). Moreover, the INP has recently been
found to develop from the ventralmost striatal progenitor zone,
to show continuity with the medial striatum, and to contain neu-
rons expressing Lmo4 and cell-surface proteins (Cadherin-8)
characteristic of striatal neurons (Abellán and Medina, 2009).
The striatal neurons of INP include both SP-containing and enke-
phalinergic neurons. It also contains neurons derived from the
pallidal and preoptic proliferative zones (identifiable as expres-
sing Lhx6 and/or Lhx7/8), as typical of the medial and lateral stri-
atum proper (Abellán and Medina, 2009). The INP neurons
expressing Lhx6 and/or Lhx7/8 may include the many cholinergic
neurons and the few parvalbuminergic/LANT6 neurons it
contains (Medina and Reiner, 1994; Reiner and Carraway,
1987). Like LSt, INP is poor in somatostatin/NPY interneurons
(Anderson and Reiner, 1990b). Although the data thus suggest a
largely striatal nature for the INP, it also possesses some traits
that differ from those of MSt and LSt. For example, INP is
much richer than either MSt or LSt in LAMP, a limbic system
marker (Yamamoto and Reiner, 2005), which is consistent
with its derivation from the ventralmost partof the striatalsec-
tor of the developing subpallium, from which much of limbic
striatum derives. The INP also contains many more immigrant
cholinergic neurons than either MSt or LSt (Abellán and Medina,
2009). These cholinergic neurons appear to belong to the corti-
copetal system (Medina and Reiner, 1994). Finally, although
seemingly a striatal territory, the INP is poor in dopaminergic
terminals and dopamine receptors, making its enrichment in
DARPP32 somewhat puzzling.
Given then that INP is a striatal derivative containing neu-
rons with striatal traits, on what basis should it be assigned to
dorsal striatum rather than ventral striatum, especially with
its derivation from the ventral part of the striatal proliferative
zone and its enrichment in LAMP? The relevant feature of INP
that warrants this classification is that the INP is part of the
78 BRAIN RESEARCH 1424 (2011) 67101
circuitry controlling motor function via descending motor
projections to the midbrain tectum. Karten and Dubbeldam
(1973) had suggested that both globus pallidus and INP might
project to the lateral spiriform nucleus of the pretectum (SpL).
In later studies, SpL was found to project to deep layers of the
optic tectum (Reiner et al., 1982a,b), and its neurons were
shown to have pallidal morphology and neurochemistry (Reiner
and Anderson, 1993; Reiner and Carraway, 1987; Veenman and
Reiner, 1994). At that time, retrograde labeling with horseradish
peroxidase only confirmed the GP projection to SpL but not the
INP projection. More recent studies using biotinylated dextran
amine (BDA) as the tracer have shown that neurons in INP
give rise to a descending projection through the diencephalon
(Jiao et al., 2000). Recent studies indicate that neurons in INP
as well as some in LSt project to the SpL (Reiner and Medina,
unpub. obs). The LSt flanking the INP receives a major projec-
tion from the internal division of the entopallium, a thalamore-
cipient zone, within the telencephalon, of the tectofugal visual
system (Krützfeldt and Wild, 2005). The neurons in LSt at least
contain SP, and are thus direct pathway neurons. SpL also re-
ceives input from globus pallidus neurons (Medina and Reiner,
1997; Reiner et al., 1982a), which themselves receive input
from enkephalinergic LSt neurons (Reiner and Medina, unpub.
obs.). These results indicate that the INP striatal neurons, to-
gether with SP+ neurons of LSt give rise to a direct pathway
type output to SpL (Reiner et al., 1998a). A facilitatory role in tec-
tally mediated head and eye movements seems likely for this
circuit, based on its connectivity.
3.1.4. Globus pallidus
The globus pallidus (GP) in both mammals and birds is neuro-
chemically distinct fromstriatum. For example, the globus palli-
dus has a low densityof dopaminergicand cholinergic terminals
and little acetylcholinesterase (Karten and Dubbeldam, 1973;
Reiner et al., 1994). Correspondingly, it is poor in dopaminergic
and muscarinic receptors (Casto and Ball, 1994; Dietl and
Palacios, 1988; Dietl et al., 1988; Kohler et al., 1995; Kubikova
et al., 2010; Richfield et al., 1987; Schnabel and Braun, 1996;
Schnabel et al., 1997; Stewart et al., 1996; Sun and Reiner,
2000; Wächtler and Ebinger, 1989). The globus pallidus in
birds and mammals is distinguished by the presence of
large, aspiny GABAergic neurons, many of which also contain
parvalbumin and LANT6 (Reiner and Anderson, 1993; Reiner
and Carraway, 1987; Veenman and Reiner, 1994), and a dense
mat of woolly fiber terminals containing GABA and SP, or
GABA and enkephalin, whichend on the aspiny GABAergic neu-
rons (Reiner et al., 1998a). The pallidal neurons are rich in recep-
tor types related to their predominant inputs. For example, the
pallidal neurons of GP are rich in GABA receptors, as a conse-
quence of their GABAergic input from LSt (Veenman et al.,
1994). Similarly, as a consequence of the glutamatergic input
from subthalamic nucleus, pallidal GP neurons are rich in
NMDA-type and AMPA-type glutamate receptors (Jiao et al.,
2000; Laverghetta et al., 2005; Wada et al., 2004).
The GABAergic neurons of the globus pallidus co-
containing parvalbumin, which represent the majority of pal-
lidal neurons in mammals and birds, derive from the pallidal
progenitor zone (Abellán and Medina, 2009; Nóbrega-Pereira
et al., 2010; Xu et al., 2008). In mammals, globus pallidus neu-
rons remain in situ relatively near the lateral ventricle and
inferior and medial to the striatum. In birds, the pallidal neu-
rons migrate as a stream from their ventral position, sweep-
ing along an arching ventromedial to dorsolateral course to
invade the striatal sector between INP and LSt to form the glo-
bus pallidus (Abellán and Medina, 2009; Puelles et al., 2007).
Some pallidal neurons remain in residence along this course,
and one medial and ventral cluster forms the ventral palli-
dum. Nonetheless, the neurons of the globus pallidus and
ventral pallidum derive from different sectors of the pallidal
anlage, as noted above (Abellán and Medina, 2009). The entire
migratory course of the pallidal neurons remains poor in
striatal markers in adult birds, and largely contains a mix of
pallidal GABAergic neurons and cholinergic neurons. The
migration of pallidal neurons to the globus pallidus is, howev-
er, not precise and many pallidal neurons come to reside in
MSt as well, as noted above. One consequence of pallidal inva-
sion into a lateral striatal territory is that some striatal neu-
rons remain in residence in the formerly striatal territory
that has been taken overby pallidal neurons to create
avian globus pallidus. Hence, SP+ and enkephalinergic spiny
neurons are present in low abundance in avian GP (Abellán
and Medina, 2009; Molnar et al., 1994; Reiner et al., 1983,
1984). In mammals, the globus pallidus also contains a sub-
population of enkephalinergic neurons, which have descend-
ing projections with collaterals to the striatum (Hoover and
Marshall, 1999, 2002; Kita and Kita, 2001). The embryonic origin
of these cells may be the striatal subdivision (Nóbrega-Pereira
et al., 2010). In mice, this pallidal subpopulation is GABAergic,
but does not express the pallidal gene Nkx2.1, and is spared
in Nkx2.1 knockout mouse (Nóbrega-Pereira et al., 2010), in
which the pallidum is missing (Sussel et al., 1999). On the
other hand, the cholinergic neurons that invade the pallidum
are derived from the preoptic subpallial subdivision (POC) in
both mouse and chicken (Abellán and Medina, 2009; García-
López et al., 2008; Nóbrega-Pereira et al., 2010). Since these
neurons belong to the corticopetal system, they will be treated
in a later section.
The globus pallidus in mammals consists of two distinct
segments that differ in the type of striatal neuron from
which they receive input, and that differ in their projection
targets. The external pallidal segment (GPe) receives enkepha-
linergic striatal input (from indirect pathway striatal neurons)
and projects mainly to the subthalamic nucleus (STN), tha-
lamic reticular nucleus and the substantia nigra pars reticu-
lata, while the internal pallidal segment (GPi) receives SP+
striatal input (from direct pathway striatal neurons) and pro-
jects mainly to intralaminar and motor thalamus. In birds,
the globus pallidus is a singular structure that contains both
of these pallidal neuron types, as evidenced by the overlap
of SP+ and enkephalinergic inputs to the avian GP and as evi-
denced by the projection of the avian GP to the targets of both
the mammalian GPe and GPi (Medina and Reiner, 1997). These
projections have been detailed in prior papers (e.g. Farries et
al., 2005a; Medina and Reiner, 1997) and will be discussed in
the following section on functional organization of the avian
basal ganglia.
3.1.5. Functional considerations for dorsal basal ganglia
Models of the functional organization of the somatic basal
ganglia in mammals have recognized two parallel output
79BRAIN RESEARCH 1424 (2011) 67101
circuits that have opposing functions in motor control and
that interact with one another. These two output circuits orig-
inate from differing sets of striatal neurons and are called the
indirect (from enkephalinergic striatal neurons) and direct
pathways (from SP+ striatal neurons) (Albin et al., 1989;
Delong, 1990; Gerfen, 1992). The connections and functions
of these two circuits have been detailed elsewhere and will
not be reviewed here. Notably, the structure comprising the
dorsal basal ganglia in birds is organized functionally into
similar direct and indirect pathways. This too has been
Fig. 7 The songbird anterior forebrain circuit. A. A simplified, neuroanatomical schematic representation of the avian song
system. The direct motor pathway arises from nucleus HVC (proper name), which projects to the robust nucleus of the
arcopallium (RA), which in turn projects to the tracheosyringeal half of the hypoglossal nucleus (nXIIts) and other respiratory
premotor neurons in the brainstem. Another projection from HVC leads to area X of the MSt. Area X projects to the medial
portion of the dorsolateral nucleus of the thalamus (DLM), which projects to the lateral magnocellular nucleus of the anterior
nidopallium (LMAN). LMAN sends a projection to RA, with axon collaterals projecting to area X. As part of the MSt, area
X receives a strong dopaminergic input from the ventral tegmental area (VTA). Axon collaterals of the area X projection to DLM
terminate in the ventral pallidum (VP), an area that sends projections to the region of the dopaminergic midbrain (VTA) that
ultimately projects to area X. B. Combined electrophysiological and cellmorphological classification of neurons of area X. There
are four classes of neurons that correspond to mammalian striatal neurons: spiny neurons; cholinergic; fast spiking (FS); and
low-threshold spiking (LTS). In addition, there is a neuron with pallidal properties, an aspiny fast firing cell (AF).
Images and traces from Farries and Perkel (2002).
80 BRAIN RESEARCH 1424 (2011) 67101
reviewed extensively elsewhere, and the circuit details will
not be repeated here (Reiner et al., 1998a). Of present interest
are two major ways in which the directindirect pathway
plan differs between birds and mammals. First, while mam-
mals have two SP+ direct output pathways (to GPi and to the
SNr), birds possess three, one from MSt SP+ neurons to SNr,
one from LSt SP+ neurons to GPi-type neurons of globus palli-
dus, and one from presumptive SP+ neurons of LSt and INP to
SpL. Secondly, it is not yet certain how the pallidal neurons of
MSt fit into this circuit diagram. It is known that some of these
neurons project to intralaminar thalamus, but it is unknown if
they have additional projection targets, such as STN or SNr.
Since both SP+ and enkephalinergic woolly fibers are present
in MSt, it seems likely that MSt pallidal neurons include both
GPe-type and GPi-type pallidal neurons. Despite these differ-
ences, the similarities between mammals and birds in the func-
tional organization of the dorsal basal ganglia are extensive.
Lesion studies and pharmacological manipulations in birds re-
inforce this view (Cheng and Long, 1974; Goodman and Stitzel,
1977; Goodman et al., 1982, 1983; Koster, 1957; Rieke, 1980,
1981, 1982; Sanberg and Mark, 1983; Schwarcz et al., 1979).
Given these data and that avian species are similar to humans
in being bipedal, warm-blooded, and capable of complex
motor behaviors that can be readily measured, birds can poten-
tially serve as useful models to explore the pathophysiology of
basal ganglia-related disorders such as Parkinson's disease,
Huntington's disease,obsessivecompulsive disorder and Tour-
ette syndrome.
3.1.6. Area X: A specialized songbird striatal structure
The role of the anterior forebrain pathway (AFP) in song learn-
ing in songbirds has become of significant interest to basal
ganglia researchers, due to the involvement of a specialized
part of the songbird basal ganglia in this circuit. The serially
connected components of the AFP include the following: nu-
cleus HVC (proper name) of the pallium; area X of the MSt;
the medial portion of the dorsolateral thalamic nucleus
(DLM), a specialized intralaminar thalamic nucleus; and the
lateral magnocellular nucleus of the anterior nidopallium
(LMAN). These nuclei are serially connected with strictly ipsi-
lateral projections (Fig. 7A). Moreover, area X, like the sur-
rounding striatum, receives a strong dopaminergic input
from the substantia nigra and VTA (Figs. 6AC; Bottjer, 1993;
Gale and Perkel, 2005; Soha et al., 1996). The AFP is required
for song learning and adult song plasticity. Therefore its role
appears comparable to that of corticobasal gangliacortical
circuits in the learning and execution of motor sequences in
mammals (Bottjer et al., 1989; Brainard and Doupe, 2000;
Scharff and Nottebohm, 1991; Sohrabji et al., 1990; Williams
and Mehta, 1999; reviewed in Doupe et al., 2005). Reinforcing
this notion is the observation that the language-related gene
of humans FoxP2 is expressed at very high levels in striatum
and area X (Fisher and Scharff, 2009; Haesler et al., 2004,
2007; Teramitsu et al., 2004; White et al., 2006).
Area X contains a full complement of neurons typical of stri-
atum such as spiny neurons, cholinergic interneurons, parval-
buminergic interneurons and somatostatinergic interneurons
(Farries and Perkel, 2002; Reiner et al., 2004a). As true of stria-
tum, the vast majority of area X neurons are spiny neurons
and essentially identical electrophysiologically to mammalian
striatal spiny projection neurons (Fig. 7B; Farries and Perkel,
2002). In the zebra finch brain, high expression of D1A, D1B
and D2 dopamine receptors has been found in striatum, includ-
ing area X. Within area X, 77% of neurons expressed D1A, 70%
expressed D2, and half appear to express both D1A and D2 re-
ceptors (Kubikova et al., 2010). As noted above, area X receives
acorticostriatalinput from the HVC (Ding and Perkel, 2004;
Ding et al., 2003; Farries et al., 2005b). In contrast to any part of
striatum that has been described in any mammalian species,
however, area X projects directly to the thalamus (Bottjer
et al., 1989; Okuhata and Saito, 1987; Parent and Hazrati, 1995).
In mammals, it is a pallidal structure of the basal ganglia, the
globus pallidus that projects to the thalamus. Area X has been
discovered, however, to adhere to this pallidal output rule in a
surprising way. The neurons of area X that project to the medial
dorsolateral nucleus of the anterior thalamus (DLM) are pallidal
in their traits large, few in number, aspiny and pallidal in
their neurochemistry, electrophysiology (Fig. 7B) and input
from area X spiny neurons (Bottjer et al., 1989; Farries and
Perkel, 2002; Goldberg and Fee, 2010; Goldberg et al., 2010;
Luo and Perkel 1999; Reiner et al., 2004a). The mixing of striatal
and pallidal neurons in area X is not unique to area X in song-
birds, sincethe MSt region surrounding area X also containspal-
lidal neurons projecting to the DLM (Farries et al., 2005b; Reiner
et al., 2004a).It is important to note, however, that not all prop-
erties of area X pallidal neurons are typical of those in mamma-
lian pallidum. For example, area X neurons that project to DLM
express enkephalin (Carrillo and Doupe, 2004). As noted above,
chicken MSt also contains pallidal neurons, as part of the after-
math of the extensive migration of pallidal neurons dorsally to-
ward GP during development. In this light then, the mixing of
spiny striatal and pallidal aspiny neurons is not unique to area
X of songbirds, but is instead a seemingly widespread feature
of avian MSt. Songbird area X also projects to the ventral palli-
dum, via collaterals of axons projecting to the thalamic nucleus
DLM, which then provides a route by which area X can modu-
late dopaminergic input back to itself during song learning
and modification ventral pallidum projects to regions of the
ventral tegmental area and substantia nigra pars compacta
that project back to area X (Gale and Perkel, 2010; Gale et al.,
2008).
The key qualitative difference between area X and medial
striatum is the apparent lack of an extrastriatal projection of
the spiny neurons of area X. In MSt in zebra finches, chickens
and pigeons, spiny neurons project to extrastriatal targets, as
in mammalian striatum. To date, no substantial projection
beyond area X itself has been identified for area X spiny neu-
rons (Person et al., 2008). Instead, they contact and inhibit pal-
lidal projection neurons within area X (Farries et al., 2005a,
2005b; Reiner et al., 2004a). The fraction of pallidal area X neu-
rons that project to DLM versus VP remains uncertain (Farries
et al., 2005a, 2005b; Goldberg and Fee, 2010; Leblois et al., 2009;
Reiner et al., 2004a).
Finally, we have tentatively grouped area X with somatic
striatum because of its location in rostrolateral MSt, a somatic
territory, and its role in a somatic motor function, song learn-
ing and in modulating song variability during learning
(Ölveczky et al., 2005) or in adulthood (Kao et al., 2005; Leblois
et al., 2010). The interconnections of area X with ventral palli-
dum and VTA raise the possibility that it is perhaps allied with
81BRAIN RESEARCH 1424 (2011) 67101
viscerolimbic striatum. More detail on its developmental der-
ivation from the striatal proliferative zone and on its neuro-
chemistry (e.g. LAMP) is needed to judge its classification as
somatic, viscerolimbic, or mixed.
3.2. Ventral viscerolimbic basal ganglia
The ventral (viscerolimbic) basal ganglia in mammals pos-
sesses similar subdivisions, neurochemistry and input
output relations to that of the dorsal somatomotor basal
ganglia. For example, it too consists of striatal subdivisions
(nucleus accumbens and superficial olfactory tubercle) rich
in spiny GABAergic projection neurons containing either sub-
stance P or enkephalin, and a pallidal subdivision (ventral
pallidum and deep olfactory tubercle) to which the striatal
subdivision projects. As true of the dorsal basal ganglia, the
striatal part of the ventral basal ganglia receives pallial and
dopaminergic input, and the pallidal part gives rise to major
outputs of the ventral basal ganglia. Unlike dorsal basal gang-
lia, the inputs and outputs of the ventral basal ganglia are
more related to visceral brain regions and functions. Consis-
tent with this, the ventral basal ganglia appears more in-
volved in reward and motivation underlying appetitive
behavior (Groenewegen and Uylings, 2000; Kelley, 1999). De-
spite the many similarities to dorsal basal ganglia cellular or-
ganization and connectivity, and a basic understanding of
the role of the ventral basal ganglia in reward-based behavior
and functions, neuronal circuit models akin to those for the
somatic basal ganglia have not yet been developed to provide
a neuronal circuit-level explanation for the functions of the
ventral basal ganglia. As reward-motivated processes are a
fundamental attribute of vertebrate behavior, it is not sur-
prising that birds too possess a viscerolimbic basal ganglia
(Reiner et al., 2004b). Although no distinct cytoarchitectonic
boundary distinguishes the dorsal and ventral basal ganglia,
the ventral basal ganglia in both mammals and birds, none-
theless, can be distinguished from the dorsal basal ganglia
by a number of neurochemical distinctions, as detailed
below. The components of the avian viscerolimbic basal
ganglia including the olfactory tubercle, nucleus accumbens
and ventral pallidum were recognized by the Avian Brain No-
menclature Forum, but several gaps in the understanding of
thesestructureswerenotedandsomehavesincebeenfilled.
Notably, the boundaries for structural components are more
clearly e stablished by developmental and neurochemical stud-
ies, including a distinction between the two main subdivisions
of nucleus accumbens. Additionally, it is now evident that, like
in mammals, the avian olfactory tubercle possesses striatal
and pallidal subdivisions.
3.2.1. Olfactory tubercle
The olfactory tubercle (TuO) in mammals forms a prominent
bulge (hence the term tubercle) at the base of the telencepha-
lon, and is distinguished from other structures of the viscero-
limbic basal ganglia by its prominent olfactory bulb input. The
most ventral part of the olfactory tubercle that receives olfac-
tory bulb input in mammals is striatal in derivation, neuro-
chemistry and cytology, while a deeper lying portion is
pallidal. This deeper lying pallidal part of the olfactory tuber-
cle is cytoarchitectonically continuous with the ventral palli-
dum (see below). The striatal part of TuO contains medium-
sized spiny neurons, as typical of striatal subpallium, and
the medium-sized neurons project to the deeper lying pallidal
TuO neurons and to the ventral pallidum (Alheid and Heimer,
1988; Heimer et al., 1976). The small-celled islets of Calleja in-
terposed between the pallidal and striatal parts of the olfacto-
ry tubercle (Puelles et al., 2007) are derived from Lmo4-
expressing neurons of the dorsalmost part of the LGE (Abellán
and Medina, 2009). A ventral subpallial region receiving olfac-
tory bulb input also is present in birds (Reiner and Karten,
1985), although it does not form a distinct ventral telencephal-
ic bulge (Figs. 8A, B). The Nomenclature Forum recognized this
as the olfactory tubercle in birds, noted its hodological and
neurochemical similarities to mammalian olfactory tubercle,
and described it as striatal in nature (Reiner et al., 2004b).
Recent developmental studies suggest, however, that the
TuO in birds has both striatal and pallidal domains (Abellán
and Medina, 2009; Puelles et al., 2000, 2007). Based on the
expression of Pax6 and Lmo4, but the absence of Nkx2.1,Lhx6,
Lhx7/8, and Shh, the rostral and dorsolateral parts of the olfac-
tory tubercle appear to be striatal derivatives, arising from
the avian homologues of mammalian LGE1 and LGE3/LGE4,
respectively. The transcription factor cLmo4 further identifies
cell aggregates in the striatal part of the avian TuO that
resemble the islands of Calleja, which derive from the homo-
logue of mammalian LGE1 (Abellán and Medina, 2009; Puelles
et al., 2007). By contrast, the caudal and ventrolateral TuO in
birds express pallidal markers reflecting an origin from the
pallidal domain. The striatal nature of rostral and dorsolateral
TuO is consistent with its neurochemistry, since it possesses
numerous medium-sized neurons containing the striatal
spiny projection neuron neuropeptide markers enkephalin
and substance P (Abellán and Medina, 2009; den Boer-Visser
and Dubbeldam, 2002; Molnar et al., 1994; Reiner et al., 1983,
Fig. 8 Sections of chick brain showing selected structures of
the viscerolimbic basal ganglia (BG). A. Sagittal section near
midline and B. Transverse section showing the ventral
portion of medial striatum (MSt) and tuberculum olfactorium
(TuO). 8B, C, D. Transverse sections of BG depicting major
components of the ventral viscerolimbic BG including the
tuberculum olfactorium (TuO), medial striatum (MSt), and
ventral pallidum (VP). Immunolabeled with antibody to
substance P. Scale bars for 8A8D =1.0 mm. Refer to list of
abbreviations for names of other structures identified.
82 BRAIN RESEARCH 1424 (2011) 67101
1984). The striatal TuO has also been shown to contain pre-
sumptive interneurons possessing NADPH-diaphorase and
nitric oxide synthase (NOS), as true in mammals as well
(Brüning, 1993; Brüning et al., 1994; Panzica et al., 1994;
Vincent et al., 1983). The striatal TuO additionally is rich in
calcitonin gene-related peptide (CGRP)-containing fibers in bud-
gerigar and quail (Lanuza et al., 2000; Roberts et al., 2002). Mam-
malian TuO also possesses CGRP+ fibers at rostral levels (Kawai
et al., 1985). Additionally, the striatal TuO has been reported to
be rich in thyroid hormone releasing hormone-containing
fibers in ducks (Jozsa et al., 1988). The pallidal nature of caudal
and ventrolateral TuO in birds is consistent with the many
GABAergic, substance-P containing, and enkephalinergic
fibers and terminals it contains (Veenman and Reiner, 1994),
and the presence of many neurons containing the pallidal
neurotensin-related hexapeptide LANT6 (Reiner and Carraway,
1987). A striato-pallidal connectivity of avian TuO is suggested
by the GABAergic, substance-P containing, and enkephalinergic
fibers and terminals in pallidal TuO. In addition, striato-pallidal
connectivity of avian TuO has been indicated in pigeon by a
projection of striatal TuO to the ventral pallidum (Medina and
Reiner, 1997). The TuO in birds has been shown to have wide-
spread extratelencephalic projections to hypothalamus and
midbrain dopaminergic neurons (Medina and Reiner, 1997), as
also true in mammals. The relative contribution of striatal
and pallidal TuO neurons to downstream projection areas is
uncertain.
Consistent with its striatal nature, the rostral and dorsolateral
TuO receive a dopaminergic input from the ventral tegmental
area and substantia nigra of the midbrain (Figs. 6AC; Kitt and
Brauth, 1986b; Moons et al., 1994; Panzica et al., 1994, 1996), and
contain numerous receptors and second messengers associated
with this input, including D1A, D1B and D2 dopamine receptors,
and DARPP-32 (Ball et al., 1995; Dietl and Palacios, 1988; Durste-
witz et al., 1999; Sun and Reiner, 2000). In mammals, the piriform
cortex and hippocampal formation are sources of pallial input to
the TuO (Heimer and Wilson, 1975). Similarly, in pigeons the piri-
form cortex has been shown to project to the TuO (Bingman et al.,
1994). The dorsomedial avian hippocampus has been shown to
project indirectly to the avian TuO, projecting to the medial and
lateral septum, with medial (striatal) septum having reciprocal
connections with the TuO (Atoji and Wild, 2004). Other demon-
strated sources of input to striatal TuO in birds include the lateral
bed nucleus of the stria terminalis, the lateral hypothalamic area,
the parabrachial region, the nucleus of the solitary tract, and the
dorsal motor nucleus of the vagus (Arends et al., 1988; Atoji et al.,
2006; Berk and Hawkin, 1985; Wild et al., 1990), structures associ-
ated with the autonomic nervous system. These various inputs
to striatal TuO resemble those reported in mammals. Finally,
avian TuO also contains numerous immigrant pallial/cortical de-
rived glutamatergic neurons, as also true in mammals (Abellán
and Medina, 2009; Abellán et al., 2009; Puelles et al., 2000; Stried-
ter et al., 1998). Their functional significance is unknown.
3.2.2. Nucleus accumbens core and shell
The ventral striatum of mammals contains a prominent stria-
tal structure called the nucleus accumbens. This structure is a
major target of the mesolimbic dopaminergic system and an
important part of the brain reward and motivation circuit.
The nucleus accumbens of rats comprises a central core
(AcC) surrounded on its medial, ventral and lateral sides by a
shell (AcS; Herkenham et al., 1984; Záborszky et al., 1985).
The core and shell were initially identified and distinguished
by the relative enrichment of the core in acetylcholinesterase
and the enrichment of the shell in argyrophilic fibers, respec-
tively (Záborszky et al., 1985). Subsequently, immunolabeling
showed differential labeling for substance P (Zahm, 1989)
and calbindin 28kD (Heimer et al., 1997; Martin et al., 1991;
Voorn et al., 1989; Zahm and Heimer, 1993) in rodents, with
the core poor in substance P-containing fibers and rich in
calbindinergic neuropil, and the shell moderate in substance
P-containing fibers and calbindinergic neuropil. Other
markers for the two sub-territories include calretininergic
and neurotensinergic perikarya, which are relatively enriched
in accumbens shell, and enkephalin and the GABA
A
receptor,
which are enriched in neurons of accumbens core (Zahm,
1999). The accumbens in mammals has also been shown to
have a third sub-territory, a rostral pole (Zahm, 2000; Zahm
and Brog, 1992; Zahm and Heimer, 1993).
Fig. 9 Transverse sections of chick brain showing nucleus
accumbens (Ac) core (AcC) and shell (AcS) subdivisions
processed with antibodies made to the following
neuropeptides/proteins: A. Substance P (SP), B. Neuropeptide
Y (NPY), C. Dopamine and cAMP-regulated phosphoprotein
(DARPP-32), and D. Calbindin (CB). Scale bars= 500 μm. The
lateral border of AcC is particularly well delineated in section
A, however, the border with the dorsomedial MSt is not
marked clearly. The NPY immunoreactivity (B) is useful for
demarcation of the shell. The putative core is dorsolateral to
the BSTl while the putative shell occurs ventrolateral to the
AcC and BSTl. Refer to list of abbreviations for names of other
structures identified.
83BRAIN RESEARCH 1424 (2011) 67101
The Nomenclature Forum suggested that the homologue of
mammalian nucleus accumbens resides in the ventromedial
striatum of birds (i.e. ventromedial MSt). Markers to confirm
this and better define the relative locations of the avian accum-
bens core and shell, however, were not available at that time.
More recently, developmental (Abellán and Medina, 2009)and
histochemical/hodological studies (Bálint and Csillag, 2007;
Bálint et al., 2011; Husband and Shimizu, 2011; Roberts et al.,
2002) have helped confirm the location of the avian nucleus
accumbens, and clarify the components that appear homolo-
gous to the mammalian accumbens shell and core (Fig. 9). Spe-
cifically, the accumbens in chickens can be identified as a
rostral ventromedial striatal territory that is derived from the
ventrobasal striatal progenitor zone (LGE4 in mammals), and
then its core subdivision can specifically be recognized as a re-
gion within accumbens at the base of the lateral ventricle that
is very rich in neurons expressing genes for SP,ENK,andNPY,
but moderate in cLmo4 (Abellán and Medina, 2009). By contrast,
the accumbens shell can be identified ventral and lateral to the
core by its enrichment in cLmo4. This shell region has a caudal
extension located medial to the intrapeduncular nucleus, with-
in what has been regarded as ventral MSt. Thus, this caudal ex-
tension appears to be part of accumbens shell in birds (Fig. 3).
There is good agreement among recent neurochemical and
hodological studies regarding the overall boundaries of the
entire nucleus accumbens in birds (Abellán and Medina,
2009; Bálint and Csillag, 2007; Bálint et al., 2011; Husband
and Shimizu, 2011; Roberts et al., 2002) and that a third
accumbens sub-territory, a rostral pole, occurs in birds (Bálint
and Csillag, 2007; Bálint et al., 2011; Husband and Shimizu,
2011) that may be comparable to that in mammals (Zahm
and Heimer, 1993). Projections into the avian nucleus accum-
bens resemble those in mammals, and to some extent define
its proposed subdivisions. For example, the nucleus tractus
solitarius projects prominently to accumbens shell but not
core in mammals. Bálint and Csillag (2007) showed by antero-
grade labeling in conjunction with immunolabeling that the
nucleus tractus solitarius (NTS) also projects to accumbens
in birds, and based on the differential projection within nucle-
us accumbens were able to distinguish an accumbens core-
like region (poor in NTS input) and an accumbens shell-like
region (rich in NTS input). Nucleus accumbens also receives
dopaminergic input from the ventral tegmental area (Kitt
and Brauth, 1981; Mezey and Csillag, 2002), and at least one
study in budgerigars (Roberts et al., 2002) has suggested
that the accumbens shell is especially high in tyrosine
hydroxylase-containing (i.e. dopaminergic) fibers. Finally, a
dense noradrenergic input from the A6 neurons of locus coer-
uleus to nucleus accumbens has been reported (Figs. 6F, G),
ending more heavily in accumbens shell (Bailhache and
Balthazart, 1993; Kitt and Brauth, 1986a; Mello et al., 1998;
Moons et al., 1995; Reiner et al., 1994; von Bartheld and
Bothwell, 1992). By contrast, somatic striatum receives only
modest noradrenergic innervation (Bailhache and Balthazart,
1993; Moons et al., 1995; Reiner et al., 1994).
Bálint et al. (2011) recently showed projections from the
chick AcS are similar to those reported in mammals, including
extensive projections to the lateral preoptic area, substantia
nigra pars compacta, ventral tegmental area, A8 dopaminer-
gic cell group, and moderate projections to the periaqueductal
gray, parabrachial complex and nucleus raphe (Groenewegen
and Russchen, 1984; Usuda et al., 1998; Zahm and Heimer,
1993). Although projections from the avian AcC have been
shown to be more extensive than reported for mammals, par-
ticularly in the caudal brainstem, significant projection areas
comparable to those in mammals have been noted, including
the substantia nigra pars compacta, the A8 dopaminergic cell
group, periaqueductal gray and the ventral tegmental area
(Groenewegen and Russchen, 1984; Usuda et al., 1998; Zahm
and Heimer, 1993).
As pointed out in a recent publication, however, identifica-
tion of subdivisions of avian nucleus accumbens as homologous
to particular accumbens subdivisions in mammals is not unam-
biguous and suggests caution needs to be applied (Husband and
Shimizu, 2011). They note that studies in budgerigars (Roberts
et al., 2002), chickens (Bálint and Csillag, 2007) and pigeons
(Husband and Shimizu, 2011) show that the two regions that
some have identified as avian accumbens core and shell do not
match accumbens core and shell in rats (Bubser et al., 2000;
Tan et al., 1999)orprimates(Brauer et al., 2000)particularly
in the localization of the calcium-binding protein calbindin.
Husband and Shimizu (2011) have, thus, used the designations
ventral (vAc) and dorsal (dAc) to subdivide the avian accumbens,
with their vAc and dAc being topographically comparable to the
accumbens shell (AcS) and accumbens core (AcC), respectively,
of others (Fig. 9). On the other hand, the above noted molecular
developmental, neurochemical, and hodological data and recent
immunohistochemical and tract-tracing studies, support the no-
tion that avian AcS and AcC resemble their similarly named
mammalian counterparts in many regards. While further stud-
ies are needed to evaluate the proposed boundaries and homol-
ogies of the avian accumbens subdivisions with respective
mammalian subdivisions, it may be that the counterparts will
not resemble one another by all neurochemical and hodological
criteria.
The nucleus accumbens in its entirety also receives input
from central components of the autonomic nervous system
such as the lateral hypothalamus (Berk and Hawkin, 1985),
lateral part of the bed nucleus of the stria terminalis (Atoji et
al. 2006), parabrachial nucleus (Wild et al., 1990), nucleus trac-
tus solitarius (Arends et al., 1988; Bálint and Csillag, 2007) and
dorsal motor nucleus of the vagus (Arends et al., 1988).
The avian accumbens also receives input from diverse tel-
encephalic and midbrain regions regarded as limbic struc-
tures, including the hippocampus (Atoji and Wild, 2004; Atoji
et al., 2002; Székely and Krebs, 1996; Veenman et al., 1995),
piriform cortex (Veenman et al., 1995), posterior pallial amyg-
dala (Atoji et al., 2006), and ventral tegmental area (Kitt and
Brauth, 1981; Mezey and Csillag, 2002; Moons et al., 1994).
Pallial input to avian accumbens is likely to be glutamatergic
(Csillag et al., 1997; Ding and Perkel, 2004; Ding et al., 2003;
Farries et al., 2005a; Reiner et al., 2001; Veenman and Reiner,
1996). Consistent with its input from limbic regions of brain,
the nucleus accumbens core and shell in birds are enriched
in the limbic system protein LAMP (Yamamoto and Reiner,
2005).
Alimbic loopinvolving medial regions of pallium (i.e.
mesopallium and nidopallium), subpallium and thalamus has
been described in birds (Husband and Shimizu, 2011). The cir-
cuit consists of a projection from a column of neurons in the
84 BRAIN RESEARCH 1424 (2011) 67101
medial pallium to the nucleus accumbens, from accumbens to
ventral pallidum (VP), from VP to the anterior dorsomedial tha-
lamic nucleus (DMA), and finally a return projection from DMA
to the medial pallium (Fig. 13 in Husband and Shimizu, 2011).
The restricted part of the medial pallium involved in this limbic
loophas been proposed to be equivalent to the mammalian
prefrontal cortex (Husband and Shimizu, 2011). This medial
pallial sector apparently corresponds to the medial pallial re-
gion previously shown to be involved in imprinting in chicks
(Bolhuis, 1999; Gruss and Braun, 1996; Horn, 1998; Maier and
Scheich, 1987; Metzger et al., 1998; Thode et al., 2005). Thus,
the pallial part of this loop, as well as the loop itself, is associat-
ed with learning and memory. Nonetheless there are behavior-
al, anatomical and electrophysiological data showing that
another brain region, the caudolateral nidopallium, may be
the equivalent of the avian prefrontal cortex (Güntürkün, 2005;
Waldmann and Güntürkün, 1993).
3.2.3. Ventral pallidum
The ventral pallidum (VP) of mammals is regarded as a medio-
ventral extension of globus pallidus and the viscerolimbic com-
ponent of the striatopallidal system. It occupies a relatively
extensive area in the basal forebrain of rats and primates,
extending ventrally and rostrally from the globus pallidus into
the olfactory tubercle (Alheid and Heimer, 1988). The VP pos-
sesses a number of pallidal traits by which it can be identified,
including enrichment in iron (Switzer et al., 1982), large aspiny
glutamate decarboxylase (GAD)-containing neurons, a dense
mat of terminals containing GAD, and dense mats of terminals
containing enkephalin and SP (Beach and McGeer, 1984; Haber
and Nauta, 1983; Haber and Watson, 1985; Switzer et al., 1982).
The terminals containing GAD, enkephalin and SP possess the
distinctive pallidal woolly fiber morphology of striatopallidal
terminals. The VP field in mammals contains two major neuro-
chemically distinct output neuron types, GABAergic neurons
with descending projections and cholinergic neurons with as-
cending projections within the telencephalon (Bengtson and
Osborne, 1999). The GABAergic VP neurons have their origin in
the pallidal progenitor zone of the developing subpallium, as
do globus pallidus neurons (Xu et al., 2008), while cholinergic
neurons have a preoptic origin (García-López et al., 2008;
Nóbrega-Pereira et al., 2010). The globus pallidus and ventral
pallidum in mammals, however, originate from different parts
of the MGE, with the globus pallidus derived from MGE2/MGE3
and the ventral pallidum arising from MGE4 (precommissural
part of VP) and MGE3 (commissural part of VP). This is similar
in birds, although the commissural part of VP (derived from
MGE3) appears to be missing (Abellán and Medina, 2009).
The VP of birds (Figs. 8C, D) was originally termed the ventral
paleostriatum, and the term was used for a large region that
contained both striatal and pallidal neuron types (Kitt and
Brauth, 1986a). The ventral paleostriatum was identified by
the abbreviation PVT (paleostriatum ventrale) in a chick brain
atlas (Kuenzel and Masson, 1988). Others used the term ventral
pallidum instead, and the Nomenclature Forum officially
renamed PVT as the ventral pallidum (VP). In doing so, the re-
gion definedas the ventral pallidum was narrowed to a territory
exclusively having pallidal neurochemistry (Reiner et al.,
2004b). Developmental studies in chicks have been instrumen-
tal in providing support for the identification of this region as
the ventral pallidum, and the homologue of part of the mam-
malian structure of the same name. Specifically, the VP displays
strong expression for Nkx2.1 (Puelles et al., 2000), Lhx6 and
Lhx7/8 (Abellán and Medina, 2009), which within subpallium
are uniquely expressed by pallidal neurons. Like the avian glo-
bus pallidus, the avian VP is rich in fibers and terminals con-
taining substance P, dynorphin, enkephalin and/or GABA
that represent the terminals of striatal neurons (Anderson and
Reiner, 1990a; Reiner et al., 1983, 1984; Veenman and Reiner,
1994; Veenman et al., 1995). Neurons in the avian VP also pos-
sess a similar glutamate receptor expression profile to that of
neurons in the avian globus pallidus (Wada et al., 2001), as
well as GABA receptors (Veenman et al., 1994). The GABAergic
projection neurons of VP in birds co-contain parvalbumin and
the neuropeptide LANT6 (Reiner and Anderson, 1993; Reiner
and Carraway, 1987; Veenman and Reiner, 1994). As in mam-
mals, the large GABAergic neurons of globus pallidus and ven-
tral pallidum arise from separate parts of the pallidal domain,
with the globus pallidus arising from the region homologous
to mammalian MGE2/MGE3, and the ventral pallidum arising
from the homologue of mammalian MGE4 (Abellán and Medina,
2009). The VP in birds only corresponds to the precommissural
VP of mammals (Abellán and Medina, 2009) and contains a
prominent group of cholinergic neurons, as it does in mammals
(Medina and Reiner, 1994). The cholinergic neurons of avian VP
(and those in globus pallidus and INP) derive from the POC sub-
pallial subdivision (Abellán and Medina, 2009), as also true of
the cholinergic neurons of mammalian VP (García-López et al.,
2008). They constitute a separate system that will be discussed
further in Section 3.4 (Basal telencephalic cholinergic and non-
cholinergic corticopetal system).
As true of globus pallidus, the avian ventral pallidum re-
ceives striatal input (in this case viscerolimbic) and it gives rise
to major descending output to targets very similar to those of
mammalian VP (Medina and Reiner, 1997). Viscerolimbic striatal
structures projecting to VP include the olfactory tubercle and
nucleus accumbens (Medina and Reiner, 1997). The VP in birds
also has been reported to receive pallial inputs from the
temporo-parieto-occipital area and diencephalic inputs from
the lateral hypothalamic area (Atoji and Wild, 2005; Berk and
Hawkin, 1985; Medina and Reiner, 1997; Veenman et al., 1995).
The avian VP, in turn, projects to a number of diencephalic
sites including, the subthalamic nucleus, paraventricular nucle-
us, dorsomedial thalamus, habenular nuclei, thalamic reticular
nucleus, and lateral hypothalamic area (Berk and Hawkin,
1985; Kitt and Brauth, 1981; Medina and Reiner, 1997; Veenman
et al., 1995). Several midbrain regions receive ventral pallidal
input as well, including the ventral tegmental area, substantia
nigra pars compacta, pedunculopontine tegmental nucleus,
central gray, and raphe (Kitt and Brauth, 1981; Medina and
Reiner, 1997).
3.2.4. Functional considerations for ventral basal ganglia
The dopaminergic input to viscerolimbic striatum in mammals,
especially nucleus accumbens, is widelythought to mediate the
rewarding effects of natural stimuli such as food, water and sex
(Salamone et al., 1997). The many autonomic circuit inputs and
outputs of the viscerolimbic basal ganglia are consistent
with such a function. Nucleus accumbens, in particular, plays
a major role in the initiation of reward-motivated behavior
85BRAIN RESEARCH 1424 (2011) 67101
(Roberts et al., 1980). Because reward provides the motivation
for learning, the nucleus accumbens has also been regarded as
critical for reward- and incentive-based learning (Salamone
et al., 1997). Moreover, drugs of abuse, such as cocaine, amphet-
amine, morphine and methamphetamine, produce their plea-
surable sensations by release of dopamine and increased
neuronal activation in nucleus accumbens, especially the
accumbens shell (Pierce and Kalivas, 1995; Pontieri, et al., 1994,
1995). The dopaminergic input from the posteromedial ventral
tegmental area to the medial olfactory tubercle and ventrome-
dial accumbens shell appears particularly involved in reward,
based on behavioral findings addressing the effects of intra-
accumbens cocaine or amphetamine administration (Ikemoto,
2007). The role of the ventral basal ganglia, however, is
more complex than a simple role in facilitating pursuit of re-
ward (Salamone et al., 1997). Nucleus accumbens also plays a
role in withholding responses so as to maximize reward, and le-
sions of accumbens core impair this ability and cause impulsive
behavior (Cardinal et al., 2001). The nucleus accumbens, espe-
cially the shell, is also responsive to stress (Kalivas and Duffy,
1995; King et al., 1997). Although neural circuit models (e.g. the
directindirect pathway model) have been developed to explain
how the somatic basal ganglia mediates motor control at the
neuronal level, how the striatopallidal circuitry of the viscero-
limbic basal ganglia mediates its role in reward-motivated be-
havior, choice and learning is yet to be elucidated at the
neuronal circuit level.
The viscerolimbic basal ganglia in birds appears to be in-
volved in reward, and reward-motivated learning. For example,
Delius et al. (1976) have shown that electrical stimulation of the
accumbens region has rewarding properties in birds, and medi-
al striatal neurons in chick show reward-related responses
(Yanagihara et al., 2001). Similarly, chicks with electrolytic le-
sions of the ventromedial striatum, including the accumbens
region, are impaired in their ability to associate color cues
with reward (Aoki et al., 2006; Izawa et al., 2003). Such lesions
also cause impulsivity in chicks. Additionally, the accumbens
region is associated with passive avoidance learning in chicks
(Stewart et al., 1996). One difficulty in relating the results of
these studies to specific parts of viscerolimbic basal ganglia is
that the stimulating, recording or lesion sites inthe above stud-
ies spanned the somatic and viscerolimbic medial striatum.
Thus, the relative contribution of limbic striatum and medial
MSt to these functions is not certain. As true of medial caudate
in mammals, medial MSt appears to be viscerolimbic in connec-
tivity, function and neurochemistry (Yamamoto and Reiner,
2005). As discussed in the previous section on the nucleus
accumbens, a limbic loopcircuit comprising prefrontal
cortex-like medial pallial neurons, medial MSt/nucleus accum-
bens, ventral pallidum anda mediodorsal-like thalamic nucleus
that may be involved in learning and memory has been demon-
strated in pigeons (Husband and Shimizu, 2011).
3.3. Subpallial amygdaloid nuclei: The extended
amygdala central and medial amygdala, and bed nuclei of
the stria terminalis
The amygdala in mammals has been recognized as consisting
of two major subpallial cell groups, the central and medial
amygdaloid nuclei (Swanson and Petrovich, 1998). Heimer
and coworkers recognized, however, that these two amygda-
loid nuclei were each confluent with neuronal corridors that
extended from them, through the territory below the globus
pallidus (i.e. sublenticular) to the bed nuclei of the stria termi-
nalis or BST (Alheid and Heimer, 1988; Alheid et al., 1995). The
distinctive feature of these neuronal corridors is that they
possess similar neurochemical features and connections as
the amygdaloid nuclei, with which they are confluent. The
central and medial amygdaloid nuclei, directly or by way of
the BST, are the major output nuclei of the amygdala (Paré
et al., 2004; Swanson, 2000). The central extended amygdala,
which we will use as a term to include the central amygdaloid
nucleus (CeA) and its corridor to the lateral BST (BSTL), is in-
volved in fear/anxiety and ingestive functions. The medial ex-
tended amygdala, which we will use as a term to include the
medial amygdaloid nucleus (MeA) and its corridor to the me-
dial BST (BSTM), is involved in reproduction and defense
(Alheid and Heimer, 1988; Alheid et al., 1995). We will refer
to the central extended amygdala together with the BSTL as
the central extended amygdala-BSTL complex (or simply cen-
tral extended amygdala complex), and the medial extended
amygdala together with the BSTM as the medial extended
amygdalaBSTM complex (or simply medial extended amyg-
dala complex).
Developmental, neurochemical, hodological and behavior-
al evidence suggests that territories corresponding to the
central and medial extended amygdalaBST complexes of
mammals are present in the avian subpallium as well (Aste
et al., 1998a; Jurkevich et al., 1997, 1999; Roberts et al., 2002; re-
views in Abellán and Medina, 2009; Reiner et al., 2004b; Xie et
al., 2010; Yamamoto et al., 2005). Moreover, both also appear
to be present in amphibian and reptilian subpallia (Martínez-
García et al., 2008; Morona and González, 2008), suggesting
that amygdaloid regions were present in the common tetra-
pod ancestor (Martínez-García et al., 2007). Evidence for cen-
tral and medial extended amygdalaBST complexes in birds
is presented in the following two sections below, with constit-
uents of the central extended amygdala and the BSTL, and
constituents of the medial extended amygdala and BSTM pre-
sented under separate subheadings, respectively.
3.3.1. Central extended amygdala and BSTL
In mammals, neurons of the central extended amygdala are
GABAergic and have descending projections to the lateral hy-
pothalamus, central gray, the parabrachial nucleus and the
nucleus of the solitary tract, which are autonomic centers
that regulate behaviors related to ingestion, fear, and stress/
anxiety (Alheid and Heimer, 1988; Alheid et al., 1995; de
Olmos et al., 2004; Swanson, 2000). GABAergic neurons of the
central extended amygdala projecting to these targets are typ-
ically enriched in any of several neuropeptides, including
corticotropin-releasing hormone (CRH), neurotensin, enkeph-
alin or somatostatin (Alheid et al., 1995; Moga and Gray,
1985; Panguluri et al., 2009; Paré and Smith, 1994; Poulin and
Timofeeva, 2008; Swanson and Petrovich, 1998). Consistent
with its role in autonomic functions, the central extended
amygdala receives input from the posterior intralaminar thal-
amus, the parabrachial nucleus, the nucleus of the solitary
tract, the insular cortex, and the pallial amygdala. It is distinc-
tively rich in calcitonin gene related peptide (CGRP) terminals,
86 BRAIN RESEARCH 1424 (2011) 67101
representing the parabrachial and posterior intralaminar tha-
lamic input (D'Hanis et al., 2007). The mammalian central
amygdala itself is primarily a striatal derivative (Bupesh et
al., 2011b; García-López et al., 2008; Puelles et al., 2000; Tole
et al., 2005; Waclaw et al., 2010), and recent fate mapping
data indicate that it contains Pax6-expressing cells derived
from dorsal LGE and Islet1-expressing cells derived from
ventral LGE (Bupesh et al., 2011b; Waclaw et al., 2010). Howev-
er, recent evidence also indicates that the medial part of the
central amygdala, and possibly the sublenticular corridor of
the central extended amygdala contain a mixture of neurons
of striatal and pallidal origin (Bupesh et al., 2011b). Pallidal
neurons invading the central amygdala and the corridor con-
tain somatostatin (Bupesh et al., 2011b; García-López et al.,
2008). The BSTL is primarily a pallidal derivative, but some
striatal cells expressing Pax6,Islet1 or Lmo4 invade the BSTL
(Bupesh et al., 2011b; García-López et al., 2008; reviewed in Me-
dina and Abellán, 2009). The central extended amygdaloid
complex in birds includes at least two components: the so-
called subpallial amygdaloid area (SpA; Reiner et al., 2004a;
Roberts et al., 2002; Wild et al., 1990; Yamamoto et al., 2005),
and the BSTL (Abellán and Medina, 2008, 2009; Aste et al.,
1998b; Jurkevich et al., 1999; Reiner et al., 2004b; Roberts et
al., 2002). Both of these were recognized by the Nomenclature
Forum, though not specifically in relationship to an entity
termed the extended central amygdala. The Nomenclature
Forum also did not recognize a central amygdaloid nucleus in
birds. The current understanding of the subdivisions of avian
central extended amygdala and the BSTL are discussed in the
following sections.
3.3.1.1. The subpallial amygdaloid area (SpA) and the central
amygdala. Abellán and Medina (2009) recognized that the
central extended amygdala in chick includes a territory
below the globus pallidus identified by the Nomenclature
Forum as the SpA and a more lateral territory below the cau-
dolateral striatum that the Forum did not recognize. Abellán
and Medina termed these two regions the medial and lateral
parts of the avian extended central amygdala (Fig. 10). They
noted that the medial portion of the central extended amyg-
dala, below the globus pallidus, may correspond to the
sublenticular corridor of the central extended amygdala of
mammals (Reiner et al., 2004a), and include striatal and pallidal
cells (Abellán and Medina, 2009). They proposed that the lateral
portion of the central extended amygdala may correspond to at
least part of the mammalian central amygdala, by the criteria
that it lies directly below the caudolateral striatum and is rich
in striatal, Pax6-expressing (Figs. 10E, F) neurons (Abellán and
Medina, 2009). They additionally recognized a caudolateral ven-
tral part of LSt (which they termed CLSt) as part of the central
extended amygdala as well, and suggested that this also was
part of the avian homologue of mammalian central amygdala
(Figs. 10E, F). Considerable neurochemical and hodological evi-
dence shows a strong similarity between the mammalian sub-
lenticular extended amygdala and the medial part of the avian
central extended amygdala of Abellán and Medina (2009),or
subpallial amygdala (SpA) as the Nomenclature Forum termed
it. As true of mammalian sublenticular central amygdala, neu-
rons of the avian central extended amygdala are GABAergic
(Abellán and Medina, 2009; Yamamoto et al., 2005). Moreover,
many of these neurons are striatal in neurochemistry, contain-
ing the characteristic striatal amygdaloid neuropeptides en-
kephalin, neurotensin and/or corticotropin releasing hormone
(Atoji et al., 1996; Molnar et al., 1994; Reiner et al., 2004a; Richard
et al., 2004; Roberts et al., 2002; Yamamoto et al., 2005). The in-
puts to the medial and lateral portions of the extended central
amygdala of birds also resemble those of the mammalian
Fig. 10 Major structures comprising the avian central
extended amygdala (EAce) complex, as seen in frontal (A,B) or
oblique-horizontal (CF) sections showing mRNA expression
of Lmo3 (A,B), SP (C,D) or Pax6 (E,F). Scale bars in A and C =1 mm
(applies to AF). Expressionof the gene Lmo3 helps to d elineate
the dorsal part of lateral bed nucleus of the stria terminalis
(BSTLd) and part of a lateral corridor that includes the
intrapeduncular nucleus (INP) and the EAce cell corridor,
located below the globus pallidus (GP). The striatal division is
rich in cells expressing substance P (SP) or Pax6, and both
markers are enriched in the striatal (lateral) part of the EAce
complex. In addition, many cells expressing Pax6 or SP also
invade (apparently by tangential migration) the pallidal (more
medial) parts of the EAce complex, including the EAcem and
part of the dorsal BSTL. Refer to list of abbreviations for names
of other structures identified.
From Abellán and Medina (2009).
87BRAIN RESEARCH 1424 (2011) 67101
central extended amygdala. In both, this regionreceives viscer-
olimbic input from the parabrachial nucleus (Wild et al., 1990),
the nucleus of the solitary tract, and the pallial amygdala
(Atoji et al., 2006; Veenman et al., 1995). Moreover, the avian
central extended amygdala, as well as the CLSt, is enriched in
CGRP terminals, (Lanuza etal., 2000; Reiner et al., 2004b; Roberts
et al., 2002; Yamamoto et al., 2005). Note, however, that neither
the lateral central extended amygdala nor the CLSt in birds is
nearly as rich in CGRP+ fibers as their suggested mammalian
homologue, the central amygdala. As in mammals, the GABAer-
gic/neuropeptidergic neurons of the central extended amygdala
in birds give rise to its outputs, notably to the BSTLand its sub-
nuclei (Fig. 10) and the nucleus of the solitary tract/dorsal vagal
nucleus (Abellán and Medina, 2009; Atoji et al., 2006; Berk, 1987;
Richard et al., 2004; Yamamoto et al., 2005), and may account for
some fibers in those regions containing enkephalin, neuroten-
sin and/or corticotropin releasing hormone. In summary, con-
siderable data support the homology of the medial part of the
avian central extended amygdala (i.e. the subpallial amygdala)
to the mammalian sublenticula r central extended amygdala, in-
cluding the presence of a neurochemically and hodologically
similar region in reptiles (Martínez-García et al., 2008). A homolo-
gy of the lateral part of the avian central extended amygdala
(CLSt) to the mammalian central amygdala also is supported by
considerable data, although differences between these two struc-
tures suggest further study of this issue is needed.
3.3.1.2. Striatal capsule. The intercalated cell masses of the
mammalian amygdala are subpallial neurons interposed be-
tween the central amygdala and the basal complex of the pallial
amygdala and the ventral endopiriform nucleus. The amygda-
loid intercalated cell masses appear to represent an integral
part of the central extended amygdala in mammals, and develop
from the dorsal part of the striatal subdivision of the developing
subpallium, that is the LGE (García-López et al., 2008; Kaoru et al.,
2010; Medina and Abellán, 2009; Waclaw et al., 2010). The amyg-
daloid intercalated cell masses have a GABAergic projection to
the central amygdala and the cholinergic corticopetal cell groups
of the basal telencephalon (including the basal magnocellular
complex), and they are involved in extinction of fear memories
(Paré et al., 2004). A distinctive set of subpallial neurons inter-
posed between the nidopallium and the lateral striatum has re-
cently been termed the avian striatal capsule (Puelles et al.,
2007), and been proposed to be comparable to the intercalated
cell masses of the mammalian amygdala (Abellán and Medina,
2009). Data supporting the proposal include similar develop-
mental origin from the dorsal part of the avian LGE homologue,
some molecular traits, and their position at the border between
the subpallium and what Abellán and Medina (2009) identify as
ventral pallium and thus regard as comparable to mammalian
ventral pallium, including part of the basal amygdalar complex
and the ventral endopiriform nucleus (Figs. 3CF). The connec-
tions of the avian striatal capsule are unknown and therefore
no hodological support exists at this time. Moreover, the avian
striatal capsule is not juxtaposed to the proposed avian central
amygdala (subpallial amygdalar area), unlike the intercalated
cell masses of the mammalian amygdala.
3.3.1.3. Lateral bed nucleus of the stria terminalis (BSTL). As
noted by the Nomenclature Forum, the BSTL in birds and
mammals is located at the base of the lateral ventricle near
the level of the septopallio-mesencephalic tract and anterior
commissure (Figs. 8C, D). It is characterized by a relative
abundance of neurotensinergic (Atoji et al., 1996; Reiner and
Carraway, 1987; Reiner et al., 2004b), enkephalinergic (Molnar
et al., 1994), and corticotropin-releasing hormone (CRH) neu-
rons (Panzica et al., 1986; Richard et al., 2004), and many calci-
tonin gene related peptide (CGRP; Lanuza et al., 2000) and
noradrenergic fibers (Reiner et al., 1994). In contrast a paucity
of cholinergic cells/fibers (Medina and Reiner, 1994), sparse
dopaminergic terminals (Bailhache and Balthazart, 1993; Reiner
et al., 1994, 2004b), and fewsubstance P-containing neurons and
fibers (Reiner et al., 1983, 2004b) have been reported. Like the
mammalian BSTL (Gray and Magnuson, 1987; Moga et al.,
1989; van der Kooy et al., 1984), the avian BSTL is reciprocally
connected with the hypothalamus, parabrachial nucleus, the
nucleus of the solitary tract, and the dorsal motor nucleus of
the vagus (Arends et al., 1988; Atoji et al., 2006; Bálint et al.,
2011; Berk, 1987; Wild et al., 1990). The BSTL in mammals and
birds, however, appears to be a complex territory with striatal
and pallidal cell subpopulations. In brief, the dorsal and medial
parts of the BSTL in birds and mammals are rich in cells that
develop from a comparable pallidal embryonic subdomain
(Abellán and Medina, 2009; García-López et al., 2008; Xu et al.,
2008), and in birds has been termed the dorsal BSTL by Abellán
and Medina (2009). Lateral to this, in a region termed by them
the dorsolateral BSTL, reside abundant neurons expressing
Pax6/Lmo4 that appear to derive from the striatal progenitor
zone (Figs. 10E, F; Abellán and Medina, 2008, 2009; García-
López et al., 2008; Xu et al., 2008). The BSTL in mammals also
contains a minor subpopulation of Pax6-expressing neurons
derived from dorsal LGE, and an abundant subpopulation of
Islet1-expressing neurons derived from ventral LGE (Bupesh et
al., 2011b). The dorsolateral BSTL of birds is the subdivision
rich in neurons possessing such striatal markers as corticotro-
pin releasing hormone, enkephalin and neurotensin. In adult
birds, however, even the dorsomedial BSTL contains some
enkephalinergic and neurotensinergic neurons, as also true in
mammals (Molnar et al., 1994; Reiner et al., 2004b).The dorsolat-
eral BSTL in birds is confluent with the subpallial amygdala, or
as Abellán and Medina termed it, the medial part of the central
extended amygdala. Note that since thestriatal central extend-
ed amygdala complex projects to the pallidal dorsal BSTL, this
system shows a striato-pallidal organization in birds, as also
noted in mammals (Alheid and Heimer, 1988; Swanson, 2000).
Given that reptiles too possess a BSTL possessing these same
various features, it seems likely that the BSTL is homologous
across amniotes (Martínez-García et al., 2008).
3.3.2. Functional considerations for the central extended
amygdala complex
The central extended amygdalaBSTL in mammals is involved
in food intake and fear/stress behaviors via its connections
with the central parts of the autonomic nervous system
(Luiten et al., 1987; Paré et al., 2004; van der Kooy et al.,
1984). The connections of the avian central extended amygda-
la and BSTL complex with the lateral hypothalamic area, para-
brachial nucleus, nucleus of the solitary tract, and dorsal
motor nucleus of the vagus are consistent with the view that
the homologous circuit in birds is involved with the same
88 BRAIN RESEARCH 1424 (2011) 67101
basic ingestive and visceral functions (Kuenzel, 1994, 2000;
Kuenzel and Blähser, 1993). In mammals, the central extend-
ed amygdalaBSTL complex projections involving neurons
containing corticotropin-releasing hormone (CRH) are partic-
ularly important for expression of stress and anxiety, and for
regulation of appetite (Clark and Kaiyala, 2003; Gallagher et
al., 2008; Heimer and Alheid, 1991; Krogh et al., 2008). This
peptide is also enriched in the projections of the central ex-
tended amygdala complex in birds and reptiles, particularly
in the BSTL projections (Martínez-García et al., 2008; Richard
et al., 2004), and the CRH+ amygdaloid neurons may thus
play a similar role in non-mammals as mammals (Crespi and
Denver, 2005; Meade andDenbow, 2003; Tachibana et al., 2006).
3.3.3. Medial extended amygdala and BSTM
In mammals, the medial extended amygdala consists of the
medial amygdala itself and a sublenticular neuronal corridor
leading from it to the BSTM. This medial amygdaloid corridor
lies inferior to the sublenticular central extended amygdala
corridor. The medial amygdala receives main olfactory and
vomeronasal input, is rich in neurons with receptors for go-
nadal steroids, and projects to medial preoptic and hypotha-
lamic regions involved in reproduction and defense (Alheid
et al., 1994, 1995; Swanson, 2000). Many of the projections of
the medial amygdala and BSTM are GABAergic (Swanson,
2000; Swanson and Petrovich, 1998), but some are glutamater-
gic (Choi et al., 2005). The complexity of these subpallial cell
groups is further evidenced by the multiple subdivisions de-
scribed for the medial amygdala and BSTM (Alheid et al.,
1995; de Olmos and Heimer, 1999; de Olmos et al., 1985, 2004;
Dong et al., 2001). Recent molecular and fate mapping data
in mouse aid understanding of the developmental basis of
this complexity. The medial amygdaloid nuclear complex
originates primarily from the caudoventral pallidal progenitor
subdivision (MGEcv, also sometimes called the anterior ento-
peduncular area, or AEP) and the commissural preoptic area
(POC) of the subpallium, as confirmed by the numerous
Nkx2.1-lineage neurons in the mature medial amygdala (Xu
et al., 2008) and by recent experimental fate mapping (Bupesh
et al., 2011a). MGEcv-derived neurons in the medial amygdala
can be distinguished from POC-derived neurons because the
former express the transcription factor Lhx6, while the latter
express Shh (García-López et al., 2008). The distribution of
neurons expressing Lhx6 and Shh indicates that MGEcv-
derived and POC-derived cells in mammals are segregated in
the posterior medial amygdala, but intermingled in the anteri-
or medial amygdala (Bupesh et al., 2011a; García-López et al.,
2008). The preoptic origin of part of the medial amygdala has
been confirmed by a fatemap of Dbx1-lineage cells, showing
that most of the nitrergic neurons (co-containing GABA) of
the medial amygdala originate in the preoptic subdivision
(Hirata et al., 2009). In addition, the medial amygdala appears
to include neurons that originate either in ventral pallium
(expressing the transcription factor Lhx9;Bupesh et al., 2011a;
García-López et al., 2008) or the supraopto-paraventricular do-
main (SPV) of the hypothalamus (expressing the transcription
factors Otp and Lhx5;Abellán et al., 2010; Bardet et al., 2008;
Bupesh et al., 2011a; García-Moreno et al., 2010). The ventral
pallial and SPV-derived neurons of medial amygdala are pre-
sumably glutamatergic (Abellán et al., 2010; García-López et
al., 2008), and possibly the source of the glutamatergic projec-
tions of the medial amygdala. The mammalian BSTM also in-
cludes cells derived from MGEcv, POC and extratelencephalic
sources (García-López et al., 2008), including the supraopto-
paraventricular domain (Abellán et al., 2010; Bupesh et al.,
2011a).
The Nomenclature Forum recognized the subpallial part of
nucleus taeniae in birds as a medial amygdala homologue
(Reiner et al., 2004b; Yamamoto et al., 2005). They also noted
the evidence for a BSTM in birds (Aste et al., 1998a; Jurkevich
et al., 1997, 1999; Reiner et al., 2004b; Roberts et al., 2002). In
addition to arginine vasotocin, the BSTM likewise was
shown to contain galanin (Klein et al., 2006). More recently,
Abellán and Medina (2009) termed the subpallial part of nucle-
us taeniae the subpallial medial amygdala (MeAs; Fig. 11).
They suggested that it formed a functional unit with the
BSTM (Fig. 11B) shown as BSTM1 (dorsolateral) and 2 (ventro-
medial) in chicks (Abellán and Medina, 2008, 2009; Jurkevich
et al., 1999), and possibly the ventral part of BSTL (BSTLv,
Fig. 11A; Abellán and Medina, 2008, 2009). A glutamatergic
Fig. 11 Major structures comprising the avian medial extended amygdala (EAme) complex, as seen in frontal sections
showing mRNA expression of the transcription factor Lhx6. These structures include the subpallial medial amygdala (MeAs)
and the BSTM (which show two subdivisions in chicken, called BSTM1 and BSTM2). Scale bar =1 mm. Note the expression of
Lhx6 in other pallidal structures of the telencephalon, including the globus pallidus (GP), the BSTLd and the medial EACe. Refer
to list of abbreviations for names of other structures identified.
Modified from Abellán and Medina (2009).
89BRAIN RESEARCH 1424 (2011) 67101
population of neurons has been identified in the avian medial
amygdala (Abellán and Medina, 2008, 2009; Abellán et al.,
2009; Bardet et al., 2008; Puelles et al., 2007), which include
neurons derived from the ventral pallium, and Otp-expressing
neurons of supraopto paraventricular origin. Additionally, a
sub-nucleus termed the amygdaloid taenial nucleus (ATn
shown in Fig. 17 of Puelles et al., 2007) has been identified in
chickens located dorsal and medial to the MeAs (Fig. 11 and
last cross-sectional plate of Fig. 1) suggesting that the avian
medial amygdala (MeAs) may comprise more than the tradi-
tional nucleus taeniae.
A medial amygdala and BSTM have also been described in
reptiles and amphibians (Martínez-García et al., 2008; Morona
and González, 2008), suggesting that a medial amygdala
BSTM corridor was present in the forebrain of the common
tetrapod ancestor.
3.3.3.1. Subpallial medial amygdala. In chickens, the subpal-
lial medial amygdala (called the subpallial amygdaloid nucleus
of the taeniae by the Nomenclature Forum) includes inter-
mingled neurons derived from pallidal/MGEcv and preoptic pro-
genitor zone subdivisions (expressing Lhx6 and Shh, respectively),
resembling in this respect the anterior subnucleus of the medial
amygdala of mammals (Abellán and Medina, 2009). Moreover,
the subpallial medial amygdala of birds (Fig. 11)contains
GABAergic neurons (Abellán and Medina, 2009; Sun et al., 2005;
Yamamoto et al., 2005) and nitrergic neurons (Balthazart et al.,
2003; Panzica et al., 1994), resembling the finding of GABAergic
and nitrergic neurons in the mammalian medial amygdala
(Swanson, 2000; Tanaka et al., 1997). In addition, the avian sub-
pallial medial amygdala receives olfactory input from the main
olfactory bulb (Reiner and Karten, 1985) and piriform cortex
(Bingman et al., 1994; Veenman et al., 1995), and projects to the
BST complex and medial preoptic region (Balthazart and Absil,
1997; Cheng et al., 1999). As true of the mamma lian medial amyg-
dala, the avian subpallial medial amygdala is reciprocally con-
nected with the hippocampal formation (Atoji and Wild, 2004;
Atoji et al., 2002). Importantly, the avian subpallial medial amyg-
dala is enriched in sex-steroid concentrating neurons possessing
estrogen and androgen receptors and the enzyme aromatase
(Balthazart et al., 1998; Foidart et al., 1999; Martinez-Vargas et
al., 1978). Their abundance is more striking in males (Watson
and Adkins-Regan, 1989), and reflects the role of the steroid re-
ceptors and the neurons that contain them in male sexual
behavior (Absil et al., 2002; Panzica et al., 1996; Thompson et
al., 1998). In summary, data in birds, reptiles and amphibians
(Martínez-García et al., 2008; Morona and González, 2008)
stronglysupport the homologyof avian subpallialmedial amyg-
dala and the medial amygdala of mammals, particularly its an-
terior subnucleus. A distinct sublenticular corridor from the
avian medial amygdala to the BSTM has not, however, been
clearly delineated.
3.3.3.2. Medial bed nucleus of the stria terminalis (BSTM). The
BSTM of chicken also includes neurons of pallidal/MGEcv
and preoptic origins, thus resembling the BSTM of mammals
(Abellán and Medina, 2008, 2009). In quail, a single BSTM nucle-
us has been documented by neurochemical criteria (Aste et al.,
1998a), while in chickens BSTM has been shown to consist of
two subnuclei (Fig. 11B), as evidenced by immunocytochemistry
and in situ hybridization histochemistry (Jurkevich et al., 1999).
The two BSTM nuclei of chickens were termed the BSTMdl and
BSTMvm by Jurkevich et al. (1999), and adopted as the BSTM1
and BSTM2, respectively, by the Nomenclature Forum (Reiner
et al., 2004b). Similar to the mammalian BSTM, steroid-
responsive neurons of the avian BSTM include cells containing
aromatase (Aste et al., 1998b; Balthazart et al., 1990; Roselli,
1991; Shinoda et al., 1994; Xie et al., 2011), and cells containing
vasotocin (Aste et al., 1998b; De Vries et al., 1994; Jurkevich et
al., 1999; Kiss et al., 1987; Viglietti-Panzica et al., 1992; Xie et al.,
2011). As also true of mammalian BSTM, the avian BSTM (and
potentially the embryologically-related ventral part of BSTL) re-
ceives olfactory input from the piriform cortex (Bingman et al.,
1994; Veenman et al., 1995), and possibly the subpallial medial
amygdala (Balthazart and Absil, 1997), and projects to the medi-
al preoptic nucleus and medial hypothalamus, which are in-
volved in male mating behavior (Absil et al., 2001, 2002; Xie et
al., 2010). In Japanese quail, AVT neurons of the BSTM are the
projection neurons targeting the POM and hypothalamus
(Absil et al., 2002) and some of their axons also reach the subpal-
lial medial amygdala (Balthazart and Absil, 1997).
3.3.4 Functional considerations for subpallial medial
amygdala and BSTM
In mammals, the medial extended amygdalaBSTM complex
plays a key role in mating, sexual, defensive, and aggressive be-
haviors, for which olfactory information to the medial amygda-
la is extremely important (Choi et al., 2005; Swanson, 2000). In
birds, the hodological and behavioral data indicate that the me-
dial amygdala and BSTM play a similar role (Absil et al., 2002;
Panzica et al., 1998; Thompson et al., 1998; Xie et al., 2010). Al-
though birds are relatively microsmatic, electrophysiological re-
sponses in the avian olfactory bulb to odorants are comparable
to those in mammals (McKeegan, 2002), and olfactory cues are
used by birds in social and sexual interactions (Caro and Baltha-
zart, 2010). The medial amygdala and BSTM, which receive ol-
factory information (Bingman et al., 1994; Veenman et al.,
1995), are involved in olfactory-related behaviors (Balthazart
and Schoffeniels, 1979). Numerous studies directly demonstrate
the role of the avian subpallial medial amygdala and BSTM in
reproductive behavior. For example BSTM1 and BSTM2 are sig-
nificantly larger in males, are steroid-responsive, and play a
role in male copulatory behavior (Aste et al., 1998b; Del Abril et
al., 1987; Guillamón and Segovia, 1997; Jurkevich et al., 1999;
Kiss et al., 1987; Panzica et al., 1998; Viglietti-Panzica et al.,
1992; Voorhuis et al., 1988). The avian BSTM2 is involved specif-
ically in male appetitive sexual behavior (Xie et al., 2010, 2011).
Arginine vasotocin (AVT) is the non-mammalian homologue
of vasopressin (Acher et al., 1993), and castration inquail elimi-
nates AVT-expressing neurons and decreases the number of
aromatase-expressing neurons in the BSTM and medial preop-
tic hypothalamus (Aste et al., 1998b; Panzica et al., 1999).
3.4. Basal telencephalic cholinergic and non-cholinergic
corticopetal system
In mammals, the basal telencephalic corticopetal system con-
sists of large cholinergic neurons dispersed over the pallidal-
substantia innominata region, the medial septum-diagonal
band nucleus, and the magnocellular preoptic nucleus (Gritti
90 BRAIN RESEARCH 1424 (2011) 67101
et al., 1993, 2003). These cells partly overlap the globus pallidus
and ventral pallidum of the basal ganglia and the extended
amygdala inthe so-called substantia innominata, but they rep-
resent functionally distinct neurons that typically project to cor-
tical regions. The cholinergic corticopetal projections are in
contrast to the major descending GABAergic projections typical
of the basal ganglia and BST-extended amygdala systems
(Alheid et al., 1995; Gritti et al., 1997). Corticopetal cholinergic
neurons play an important role in modulation of cortical activi-
ty, and in attentional and arousal processes (Záborszky et al.,
1999) that affect learning and memory (Cape and Jones, 2000;
Cape et al., 2000; Metherate et al., 1988, 1992). Similarly, the
avian corticopetal system consists of scattered large cholinergic
neurons that are dispersed over the pallidum-substantia inno-
minata, and septal-diagonal band regions (Medina and Reiner,
1994; Medina et al., 1995). In birds, telencephalic structures
containing cholinergic neurons (Reiner et al., 2004b)include
the basal magnocellular nucleus (NBM, Figs. 11Aand12), nucle-
us of the diagonal band, horizontal limb (NDBh, Figs. 11A
and 12A), nucleus of the diagonal band, vertical limb (NDBv,
Fig. 12B) and commissural nucleus of the septum (CoS, Fig. 11A).
3.4.1. Nucleus basalis magnocellularis
Cholinergic cells of nucleus basalis magnocellularis (NBM) in
birds take residence primarily in and about the lateral and medi-
al forebrain bundle but also dispersed in the globus pallidus,
ventral pallidum and intrapeduncular nucleus (Medina and Rei-
ner, 1994; Reiner et al., 2004b). This field of cholinergic neurons
thus overlaps GABAergic neurons in each region. The choliner-
gic neurons of the avian NBM (Figs. 11Aand12) are comparable
to those in the mammalian nucleus basalis of Meynert, which
overlaps the globus pallidus, ventral pallidum and substantia
innominata. Perikarya of the basal forebrain cholinergic system
of mammals projecttopographically to pallial and cortical areas,
including hippocampus, neocortex, and pallial amygdala
(Záborszky et al., 1999). Various pallial regions in birds also re-
ceive cholinergic innervation, including rostromedial nidopal-
lium, temporo-parieto-occipital area (TPO), hippocampal
complex and dorsal arcopallium (Medina and Reiner, 1994).
Tract-tracing and double-labeling data indicate that cholinergic
neurons of thebasal forebrain in birdsproject to the pallium in a
roughly topographic manner (Bagnoli et al., 1992; Krebs et al.,
1991; Medina et al., 1995). Since the intrapeduncular nucleus is
rich in cholinergic neuronsand projects to the TPO, its choliner-
gic neurons appear to be part of the NBM systemin birds (Brauth
et al., 1978). The cholinergic innervation of the pallium in birds
and mammalsis associated with their enrichment in muscarin-
ic cholinergic receptors (Brann et al., 1988; Dietl et al., 1988; Koh-
ler et al., 1995; Wächtler and Ebinger, 1989). In mammals, the
basal forebrain cholinergic corticopetal cell fields receive input
from the pallial amygdala, and the central and intercalated nu-
clei of the subpallial amygdala (Grove, 1988; Paré and Smith,
1994; Price and Amaral, 1981; Russchen, 1982; Russchen et al.,
1985a,b), and the brainstem reticular formation, including the
pedunculopontine nucleus (Datta and Prutzman, 2005; Pal and
Mallick, 2004). In birds, it appears that the region of the NBM
cholinergic neurons receives input from limbic striatum (in-
cluding nucleus accumbens) and from the arcopalliumamyg-
daloid complex (Medina and Reiner, 1997; Veenman et al.,
1995). Tract-tracing data also suggest that the NBM receives
input from a rostral rhombencephalic tegmental region (Kitt
and Brauth, 1986b) that contains the pedunculopontine nucle-
us. Thus, inputs, outputs, and neurochemistry of the NBM in
birds closely resemble those in mammals.
3.4.2. Nucleus of the diagonal band
Cholinergic neurons in the diagonal band and ventromedial
septum are also part of the mammalian basal forebrain cho-
linergic system (Woolf, 1991). Cholinergic neurons of the nu-
cleus of the diagonal band (NDB) occur as far caudally as the
nucleus of the septal commissure. Neurons from the vertical
limb of the diagonal band and medial septum provide major
cholinergic innervation of the hippocampus while those of
the horizontal limb of the diagonal band project to the olfacto-
ry bulbs. Similarly, the avian NDB (Figs. 11A and 12B) projects
heavily into the hippocampal and parahippocampal areas
(Atoji et al., 2002; Benowitz and Karten, 1976; Casini et al.,
1986; Montagnese et al., 2004). Tract-tracing and double-
Fig. 12 Major structures of the cholinergic corticopetal system seen with markers of ChAT (mRNA or protein), which label
cholinergic cells. A. Nucleus basalis magnocellularis (NBM). B. Nucleus of the diagonal band (NDB). C. Higher magnification of
the NBM of Fig. 12A showing more detail of ChAT expression in NBM, GP and low expression in LSt. Scale bars in A and
B= 1 mm.
Modified from Abellán and Medina (2009).
91BRAIN RESEARCH 1424 (2011) 67101
labeling data indicate that cholinergic neurons of the NDB are
at least partly the source of this projection to the hippocampal
complex, as well as to medial pallial territories of the Wulst
and dorsal ventricular ridge (Medina et al., 1995).
3.4.3. Nucleus commissuralis septi (Commissural septal
nucleus)
The avian commissural septal nucleus (CoS) can be observed
just dorsal to the anterior commissure and lateral to the nu-
cleus of the hippocampal commissure (Fig. 11). The nucleus
projects heavily to the hippocampal and parahippocampal
areas (Atoji et al., 2002; Benowitz and Karten, 1976; Casini et
al., 1986; Montagnese et al., 2004). In general, its inputs resem-
ble those to the NBM and NDB.
3.4.4. Functional considerations for corticopetal system
The basal forebrain system has been implicated in cortical
learning and memory in mammals (Záborszky et al., 1999),
and there is a strong correlation between the loss of choliner-
gic neurons and the loss of memory (Auld et al., 2002). Little is
known, however, about the role of these cholinergic neurons
in learning and memory in birds. The similarities between
mammals and birds in the inputs, outputs, and neurochemis-
try of this system of cholinergic neurons suggest that they
likely play a key role in learning and memory in birds as
well. Pharmacological blockade of muscarinic cholinergic re-
ceptors, in fact, has been shown to impair learning and mem-
ory in diverse avian species (Kohler et al., 1996; Mineau et al.,
1994; Patterson et al., 1990; Savage et al., 1994; Zhao et al.,
1997). Moreover, beta-amyloid toxicity, which is known to
damage the basal forebrain cholinergic system in mammals,
is known to impair memory in chicks (Gibbs et al., 2010).
4. Conclusions
The present goal was to provide current developmental, hodo-
logical, chemoarchitectonic and behavioral/functional data
that further refine understanding of the avian subpallium and
its relationship to that in mammals. We have organized the
cell groups of the lateral wall of the avian subpallium into four
distinct neural systems, and pointed out similarities and differ-
ences between mammals and birds in the constituent parts of
these systems. Noteworthy similarities that had not been previ-
ously recognized are apparent, as well as some differences that
advance understanding of the evolution and function of the
avian subpallium. A subsequent paper will address septal struc-
tures residing in the medial subpallial wall that comprise a fifth
neural system in the avian subpallium.
Acknowledgments
We wish to thank Lauren Kuenzel for her excellent technical
assistance for completing Fig. 1 and sizing, grouping and pro-
viding the appropriate magnification bars for some of the
photomicrographs presented in the paper. We also thank
Dr. Antonio Abellán for providing some of the images utilizing
in situ hybridization histochemistry. Supported in part by NSF
Grant # IOS-0842937 and Competitive USDA/AFRI/NIFA Grant
no. 2005-35203-15850 to W.J.K., NIH Grant # NS-19620, NS-28711
and NS-57722 to A.R., Grant OTKAT73219 (Hungary) to A.C.,
Spanish Ministry of Science and Innovation-FEDER Grant no.
BFU2009-07212/BFI to L.M. and NIH Grant RO1 MH066128 to D.P.
REFERENCES
Abellán, A., Medina, L., 2008. Expression of cLhx6 and cLhx7/8 suggests
a pallido-pedunculo-preoptic origin for the lateral and medial
parts of the avian bed nucleus of the stria terminalis. Brain Res.
Bull. 75, 299304.
Abellán, A., Medina, L., 2009. Subdivisions and derivatives of the
chicken subpallium based on expression of LIM regulatory genes
and markers of neuron subpopulations during development.
J. Comp. Neurol. 515, 465501.
Abellán, A., Legaz, I., Vernier, B., Rétaux, S., Medina, L., 2009. Olfactory
and amygdalar structures of the chicken ventral pallium based
on the combinatorial expression patterns of LIM and other
developmental regulatory genes. J. Comp. Neurol. 516, 166186.
Abellán, A., Vernier, B., Rétaux, S., Medina, L., 2010. Similarities
and differences in the forebrain expression of Lhx1 and Lhx5
between chicken and mouse: insights for understanding
telencephalic development and evolution. J. Comp. Neurol. 518,
35123528.
Absil, P., Foidart, A., Hemmings Jr., H.C., Steinbusch, H.W., Ball, G.F.,
Balthazart, J., 2001. Distribution of DARPP-32 immunoreactive
structures in the quail brain: anatomical relationship with
dopamine and aromatase. J. Chem. Neuroanat. 21, 2339.
Absil, P., Papello, M., Viglietti-Panzica, C., Balthazart, J., Panzica,
G.C., 2002. The medial preoptic nucleus receives vasotocinergic
inputs in male quail: a tract-tracing and immunocytochemical
study. J. Chem. Neuroanat. 24, 2739.
Acher, R., Chauvet, J., Chauvet, M.T., Michel, G., 1993. The avian
neurohypophysial hormone-neurophysin precursors: structure,
post-translational processing and evolution. In: Sharp, P.J. (Ed.),
Avian Endocrinology. Journal of Endocrinology, Ltd., Bristol,
pp. 149160.
Albin, R.L., Young, A.B., Penney, J.B., 1989. The functional anatomy
of basal ganglia disorders. Trends Neurosci. 12, 366375.
Alheid, G.F., Heimer, L., 1988. New perspectives in basal forebrain
organization of special relevance for neuropsychiatric disorders:
the striatopallidal, amygdaloid, and corticopetal components of
substantia innominata. Neuroscience 27, 139.
Alheid, G.F., Beltramino, C., Braun, A., Miselis, R.R., Francois, C., de
Olmos, J.S., 1994. Transition areas of the striatopallidal system
with the extended amygdala in the rat andprimate: observations
from histochemistry and experiments with mono- and
trans-synaptic tracers. In: Percheron, G., McKenzie, J.S., Feger, J.
(Eds.),The Basal Ganglia IV. NewIdeas and Data on Structure and
Function. Advances in Behavioral Biology, 41. Plenum Press, N.Y,
pp. 95107.
Alheid, G.F., de Olmos, J.S., Beltramino, C.A., 1995. Amygdala and
extended amygdala. In: Paxinos, G. (Ed.), The Rat Nervous
System. Academic Press, San Diego, pp. 495578.
Anderson, K.D., Reiner, A., 1990a. Extensive co-occurrence of
substance P and dynorphin in striatal projection neurons: an
evolutionarily conserved feature of basal ganglia organization.
J. Comp. Neurol. 295, 339369.
Anderson, K.D., Reiner, A., 1990b. Distribution and relative abundance
of neurons in the pigeon forebrain containing somatostatin,
neuropeptide Y, or both. J. Comp. Neurol. 299, 261282.
Anderson, K.D., Reiner, A., 1991. Striatonigral projection neurons:
a retrograde labeling study of the percentages that contain
substance P or enkephalin in pigeons. J. Comp. Neurol. 303,
658673.
92 BRAIN RESEARCH 1424 (2011) 67101
Anderson, K.D., Karle, E.J., Reiner, A., 1991. Ultrastructural single- and
double-label immunohistochemical studies of substance
P-containing terminals and dopaminergic neurons in the
substantia nigra in pigeons. J. Comp. Neurol. 309, 341362.
Aoki, N., Suzuki, R., Izawa, E.-I., Csillag, A., Matsushima, T., 2006.
Localized lesions of ventral striatum, but not arcopallium,
enhanced impulsiveness in choices based on anticipated
spatial proximity of food rewards in domestic chicks. Behav.
Brain Res. 168, 112.
Arends, J.J., Wild, J.M., Zeigler, H.P., 1988. Projections of the nucleus
of the tractus solitarius in the pigeon (Columba livia). J. Comp.
Neurol. 278, 405429.
Aste, N., Balthazart, J., Absil, P., Grossmann, R., Mülhbauer, E.,
Viglietti-Panzica, C., Panzica, G.C., 1998a. Anatomical and
neurochemical definition of the nucleus of the stria terminalis
in Japanese quail (Coturnix japonica). J. Comp. Neurol. 396,
141157.
Aste, N., Panzica, G.C., Viglietti-Panzica, C., Harada, N., Balthazart,
J., 1998b. Distribution and effects of testosterone on aromatase
mRNA in the quail forebrain: a non-radioactive in situ
hybridization study. J. Chem. Neuroanat. 14, 103115.
Atoji, Y., Wild, J.M., 2004. Fiber connections of the hippocampal
formation and septum and subdivisions of the hippocampal
formation in the pigeon as revealed by tract tracing and kainic
acid lesions. J. Comp. Neurol. 475, 426461.
Atoji, Y., Wild, J.M., 2005. Afferent and efferent connections of the
dorsolateral corticoid area and a comparison with connections
of the temporo-parieto-occipital area in the pigeon (Columbia
livia). J. Comp. Neurol. 485, 165182.
Atoji, Y., Shibata, N., Yamamoto, Y., Suzuki, Y., 1996. Distribution
of neurotensin-containing neurons in the central nervous
system of the pigeon and the chicken. J. Comp. Neurol. 375,
187211.
Atoji, Y., Wild, J.M., Yamamoto, Y., Suzuki, Y., 2002. Intratelencephalic
connections of the hippocampus in pigeons (Columba livia).
J. Comp. Neurol. 447, 177199.
Atoji, Y., Saito, S., Wild, J.M., 2006. Fiber connections of the compact
division of the posterior pallial amygdala and lateral part of the
bed nucleus of the stria terminalis in the pigeon (Columba livia).
J. Comp. Neurol. 499, 161182.
Auld,D.S.,Kornecook,T.J.,Bastianetto,S.,Quirion,R.,2002.
Alzheimer's disease and the basal forebrain cholinergic system:
relations to beta-amyloid peptides, cognition, and treatment
strategies. Prog. Neurobiol. 68, 209245.
Bachy, I., Berthon, J., Rétaux, S., 2002. Defining pallial and subpallial
divisions in the developing Xenopus forebrain. Mech. Dev. 117,
163172.
Bagnoli, P., Fontanesi, G., Alesci, R., Erichsen, J.T., 1992. Distribution
of neuropeptide Y, substance P, and choline acetyltransferase in
the developing visual system of the pigeon and effects of
unilateral retina removal. J. Comp. Neurol. 318, 392414.
Bailhache, T., Balthazart, J., 1993. The catecholaminergic system
of the quail brain: immunocytochemical studies of dopamine
beta-hydroxylase and tyrosine hydroxylase. J. Comp. Neurol.
329, 230256.
Bálint, E., Csillag, A., 2007. Nucleus accumbens subregions:
hodological and immunohistochemical study in the domestic
chick (Gallus domesticus). Cell Tissue Res. 327, 221230.
Bálint, E., Kitka, T., Zachar, G., Ádám, Á., Hemmings Jr., A.C., Csillag,
A., 2004. Abundance and location of DARPP-32 in
striato-tegmental circuits of domestic chicks. J. Chem.
Neuroanat. 28, 2736.
Bálint, E., Mezey, S., Csillag, A., 2011. Efferent connections of nucleus
accumbens subdivisions of the domestic chicken (Gallus
domesticus): an anterograde pathway tracing study. J.Comp.
Neurol. 519, 29222953.
Ball, G.F., Casto, J.M., Balthazart, J., 1995. Autoradiographic
localization of D1-like dopamine receptors in the forebrain of
male and female Japanese quail and their relationship with
immunoreactive tyrosine hydroxylase. J. Chem. Neuroanat. 9,
121133.
Balthazart, J., Absil, P., 1997. Identification of catecholaminergic
inputs to and outputs from aromatase-containing brain areas
of the Japanese quail by tract tracing combined with tyrosine
hydroxylase immunocytochemistry. J. Comp. Neurol. 382,
401428.
Balthazart, J., Schoffeniels, E., 1979. Pheromones are involved in
the control of sexual behaviour in birds. Naturwissenschaften
66, 5556.
Balthazart, J., Foidart, A., Surlemont, C., Vockel, A., Harada, N.,
1990. Distribution of aromatase in the brain of the Japanese
Quail, Ring Dove and Zebra Finch: an immunocytochemical
study. J. Comp. Neurol. 301, 276288.
Balthazart, J., Foidart, A., Baillien, M., Harada, N., Ball, G.F., 1998.
Anatomical relationships between aromatase and tyrosine
hydroxylase in the quail brain: double-label
immunocytochemical studies. J. Comp. Neurol. 391,
214226.
Balthazart, J., Panzica, G.C., Krohmer, R.W., 2003. Anatomical
relationships between aromatase-immunoreactive neurons
and nitric oxide synthase as evidenced by NOS
immunohistochemistry or NADPH diaphorase histochemistry
in the quail forebrain. J. Chem. Neuroanat. 25 (1), 3951.
Bardet, S.M., Martinez-de-la-Torre, M., Northcutt, R.G., Rubenstein,
J.L., Puelles, L., 2008. Conserved pattern of OTP-positive cells in
the paraventricular nucleus and other hypothalamic sites of
tetrapods. Brain Res. Bull. 75, 231235.
Beach, T.G., McGeer, E.G., 1984. The distribution of substance
P in the primate basal ganglia: an immunohistochemical
study of baboon and human brain. Neuroscience 13,
2952.
Bengtson, C.P., Osborne, P.B., 1999. Electrophysiological properties
of anatomically identified ventral pallidal neurons in rat brain
slices. Ann. N. Y. Acad. Sci. 877, 691694.
Bennett, B.D., Bolam, J.P., 1993. Characterization of
calretinin-immunoreactive structures in the striatum of the
rat. Brain Res. 609, 137148.
Benowitz, L.I., Karten, H.J., 1976. Organization of the tectofugal
visual pathway in the pigeon: a retrograde transport study.
J. Comp. Neurol. 167, 503520.
Berk, M.L., 1987. Projections of the lateral hypothalamus and bed
nucleus of the stria terminalis to the dorsal vagal complex in
the pigeon. J. Comp. Neurol. 260, 140156.
Berk, M.L., Hawkin, R.F., 1985. Ascending projections of the
mammillary region in the pigeon: emphasis on telencephalic
connections. J. Comp. Neurol. 239, 330340.
Bingman,V.P.,Casini,G.,Nocjar,C.,Jones,T.J.,1994.Connections
of the piriform cortex in homing pigeons (Columba livia)
studied with fast blue and WGA-HRP. Brain Behav. Evol. 43,
206218.
Bolhuis, J.J., 1999. Early learning and the development of filial
preferences in the chick. Behav. Brain Res. 98, 245252.
Bottjer, S.W., 1993. The distribution of tyrosine hydroxylase
immunoreactivity in the brains of male and female zebra
finches. J. Neurobiol. 24, 5169.
Bottjer, S.W., Halsema, K.A., Brown, S.A., Miesner, E.A., 1989.
Axonal connections of a forebrain nucleus involved with vocal
learning in zebra finches. J. Comp. Neurol. 279, 312326.
Brainard, M.S., Doupe, A.J., 2000. Interruption of a basal
gangliaforebrain circuit prevents plasticity of learned
vocalizations. Nature 404, 762766.
Brann, M.R., Buckley, N.J., Banner, T.I., 1988. The striatum and
cerebral cortex express different muscarinic receptor mRNAs.
Fed. Eur. Biochem. Soc. 230, 9094.
Brauer, K., Häußer, M., Härtig, W., Arendt, T., 2000. The core-shell
dichotomy of nucleus accumbens in the rhesus monkey as
revealed by double-immunofluorescence and morphology of
cholinergic interneurons. Brain Res. 858, 151162.
93BRAIN RESEARCH 1424 (2011) 67101
Brauth, S.E., Ferguson, J.L., Kitt, C.A., 1978. Prosencephalic
pathways related to the paleostriatum of the pigeon (Columba
livia). Brain Res. 147, 205221.
Brauth, S.E., Reiner, A., Kitt, C.A., Karten, H.J., 1983. The substance
P-containing striatotegmental path in reptiles: an
immunohistochemical study. J. Comp. Neurol. 219, 305327.
Brox, A., Puelles, L., Ferreiro, B., Medina, L., 2003. Expression of the
genes GAD67 and Distal-less-4 in the forebrain of Xenopus
laevis confirms a common pattern in tetrapods. J. Comp. Neurol.
461, 370393.
Brüning, G., 1993. Localization of NADPH-diaphorase in the brain
of the chicken. J. Comp. Neurol. 334, 192208.
Brüning, G., Funk, U., Mayer, B., 1994. Immunocytochemical
localization of nitric oxide synthase in the brain of the chicken.
Neuroreport 5, 24252428.
Bubser, M., Scruggs, J.L., Young, C.D., Deutch, A.Y., 2000. The
distribution and origin of the calretinin-containing innervation of
the nucleus accumbens of the rat. Eur. J. Neurosci. 12,15911598.
Bupesh, M., Legaz, I., Abellán, A., Medina, L., 2011a. Multiple
telencephalic and extratelencephalic embryonic domains
contribute neurons to the medial extended amygdala. J. Comp.
Neurol. 519, 15051525.
Bupesh, M., Abellán, A., Medina, L., 2011b. Genetic and
experimental evidence support the continuum of the central
extended amygdala and a multiple embryonic origin of its
principal neurons. J. Comp. Neurol. 519, 35073531.
Cape, E.G., Jones, B.E., 2000. Effects of glutamate agonist versus
procaine microinjections into the basal forebrain cholinergic
cell area upon gamma and theta EEG activity and sleepwake
state. Eur. J. Neurosci. 12, 21662184.
Cape, E.G., Manns, I.D., Alonso, A., Beaudet, A., Jones, B.E., 2000.
Neurotensin-induced bursting of cholinergic basal forebrain
neurons promotes gamma and theta cortical activity together
with waking and paradoxical sleep. J. Neurosci. 20, 84528461.
Cardinal, R.N., Pennicott, D.R., Sugathapala, C.L., Robbins, T.W.,
Everitt, B.J., 2001. Impulsive choice induced in rats by lesions of
the nucleus accumbens core. Science 292, 24992501.
Caro, S.P., Balthazart, J., 2010. Pheromones in birds:myth or reality?
J. Comp. Phys. A 196, 751766.
Carrillo, G.D., Doupe, A.J., 2004. Is the songbird Area X striatal,
pallidal, or both? An anatomical study. J. Comp. Neurol. 473,
415437.
Casini, G., Bingman, V.P., Bagnoli, P., 1986. Connections of the
pigeon dorsomedial forebrain studied with WGA-HRP and
3H-proline. J. Comp. Neurol. 245, 454470.
Casto, J.M., Ball, G.F., 1994. Characterization and localization of D1
dopamine receptors in the sexually dimorphic vocal control
nucleus, area X, and the basal ganglia of European starlings.
J. Neurobiol. 25, 767780.
Cheng, H.C., Long, J.P., 1974. Dopaminergic nature of
apomorphine-induced pecking in pigeons. Eur. J. Pharmacol.
26, 313320.
Cheng, M., Chaiken, M., Zuo, M., Miller, H., 1999. Nucleus taenia of
the amygdala of birds: anatomical and functional studies in
ring doves (Streptopelia risoria) and European starlings (Sturnus
vulgaris). Brain Behav. Evol. 53, 243270.
Choi,G.B.,Dong,H.-W.,Murphy,A.J.,Valenzuela,D.M.,Yancopoulos,
G.D., Swanson, L.W., Anderson, D.J., 2005. Lhx6 delineates a
pathway mediating innate reproductive behaviors from the
amygdala to the hypothalamus. Neuron 46, 647660.
Clark, M.S., Kaiyala, K.J., 2003. Role of corticotropin-releasing factor
family peptides and receptors in stress-related psychiatric
disorders. Semin. Clin. Neuropsychiatry 8, 119136.
Cobos, I., Puelles, L., Martinez, S., 2001a. The avian telencephalic
subpallium originates inhibitory neurons that invade tangentially
the pallium (dorsal ventricular ridge and cortical areas). Dev. Biol.
239, 3045.
Cobos, I., Shimamura, K., Rubenstein, J.L., Martinez, S., Puelles, L.,
2001b. Fate map of the avian anterior forebrain at the
four-somite stage, based on the analysis of quail-chick chimeras.
Dev. Biol. 239, 4667.
Cobos, I., Broccoli, V., Rubenstein, J.L.R., 2005. The vertebrate
ortholog of Aristaless is regulated by Dlx genes in the developing
forebrain. J. Comp. Neurol. 483, 292303.
Crespi, E.J., Denver, R.J., 2005. Roles of stress hormones in food
intake regulation in anuran amphibians throughout the life
cycle. Comp. Biochem. Physiol. A 141, 381390.
Csillag, A., Székely, A.D., Stewart, M.G., 1997. Synaptic terminals
immunolabeled against glutamate in the lobus parolfactorius
of domestic chicks (Gallus domesticus) in relation to afferents
from the archistriatum. Brain Res. 750, 171179.
D'Hanis, W., Linke, R., Yilmazer-Hanke, D.M., 2007. Topography of
thalamic and parabrachial calcitonin gene-related peptide
(CGRP) immunoreactive neurons projecting to subnuclei of the
amygdala and extended amygdala. J. Comp. Neurol. 505,
268291.
Datta, S., Prutzman, S.L., 2005. Novel role of brain stem
pedunculopontine tegmental adenylyl cyclase in the regulation of
spontaneous REM sleep in the freely moving rat. J. Neurophysiol.
94, 19281937.
de Olmos, J.S., Heimer, L., 1999. The concepts of the ventral
striatopallidal system and extended amygdala. Ann. N. Y.
Acad. Sci. 877, 132.
de Olmos, J.S., Alheid, G.F., Beltramino, C.A., 1985. Amygdala. In:
Paxinos, G. (Ed.), The Rat Nervous System. Academic Press,
Sydney, pp. 223334.
de Olmos, J.S., Beltramino, C.A., Alheid, G., 2004. Amygdala and
extended amygdala of the rat: a cytoarchitectonical,
fibroarchitectonical, and chemoarchitectonical survey, In:
Paxinos, G. (Ed.), The Rat Nervous System, Third Edition.
Elsevier Academic Press, San Diego, pp. 509603.
De Vries, G.J., Al Shamma, H.A., Zhou, L., 1994. The sexually
dimorphic vasopressin innervation of the brain as a model for
steroid modulation of neuropeptide transmission. In: Luine,
V.N., Harding, C.F. (Eds.), Hormonal Restructuring of the Adult
Brain. Basic and Clinical Perspectives, 473. Ann. N. Y. Acad. Sci,
New York, pp. 95120.
Del Abril, A., Segovia, S., Guillamón, A., 1987. The bed nucleus of
the stria terminalis in the rat: regional sex differences controlled
by gonadal steroids early after birth. Brain Res. 429, 295300.
Delius, J.D., Williams, A.S., Wooton, R.J., 1976. Motivation
dependence of brain self-stimulation in the pigeon. Behav.
Proc. 1, 1527.
DeLong, M.R., 1990. Primate models of movement disorders of
basal ganglia origin. Trends Neurosci. 13, 281285.
den Boer-Visser, A.M., Dubbeldam, J.L., 2002. The distribution of
dopamine, substance P, vasoactive intestinal polypeptide and
neuropeptide Y immunoreactivity in the brain of the collared
dove, Streptopelia decaocto. J. Chem. Neuroanat. 23, 127.
Dietl, M.M., Palacios, J.M., 1988. Neurotransmitter receptors in
the avian brain. I. Dopamine receptors. Brain Res. 439,
354359.
Dietl, M.M., Cortés, R., Palacios, J.M., 1988. Neurotransmitter
receptors in the avian brain. II. Muscarinic cholinergic
receptors. Brain Res. 439, 360365.
Ding, L., Perkel, D.J., 2004. Long-term potentiation in an avian
basal ganglia nucleus essential for vocal learning. J. Neurosci.
24, 488494.
Ding, L., Perkel, D.J., Farries, M.A., 2003. Presynaptic depression of
glutamatergic synaptic transmission by D1-like dopamine
receptor activation in the avian basal ganglia. J. Neurosci. 23,
60866095.
Dong, H.-W., Petrovich, G.D., Swanson, L.W., 2001. Topography of
projections from amygdala to bed nuclei of the stria terminalis.
Brain Res. Rev. 38, 192246.
Doupe, A.L., Perkel, D.J., Reiner, A., Stern, E.A., 2005. Birdbrains
could teach basal ganglia research a new song. Trends
Neurosci. 28, 353363.
94 BRAIN RESEARCH 1424 (2011) 67101
Durstewitz, D., Kroner, S., Hemmings Jr., H.C., Güntürkün, O., 1998.
The dopaminergic innervation of the pigeon telencephalon:
distribution of DARPP-32 and co-occurrence with glutamate
decarboxylase and tyrosine hydroxylase. Neuroscience 83,
763779.
Durstewitz, D., Kroner, S., Güntürkün, O., 1999. The dopaminergic
innervation of the avian telencephalon. Prog. Neurobiol. 59,
161195.
Edinger, L., 1908. The relations of comparative anatomy to
comparative psychology. J. Comp. Neurol. Psychol. 18, 437457.
Elliot-Smith, G., 1901. Notes upon the natural subdivisions of the
cerebral hemisphere. J. Anat. Physiol. 35, 431454.
Elshatory, Y., Gan, L., 2008. The LIM-homeobox gene Islet-1 is
required for the development of restricted forebrain cholinergic
neurons. J. Neurosci. 28, 32913297.
Farries, M.A., Perkel, D.J., 2000. Electrophysiological properties of
avian basal ganglia neurons recorded in vitro. J. Neurophysiol.
84, 25022513.
Farries, M.A., Perkel, D.J., 2002. A telencephalic nucleus essential
for song learning contains neurons with physiological
characteristics of both striatum and globus pallidus. J. Neurosci.
22, 37763787.
Farries, M.A., Ding, L., Perkel, D.J., 2005a. Evidence for Directand
Indirectpathways through the song system basal ganglia.
J. Comp. Neurol. 484, 93104.
Farries, M.A., Meitzen, J., Perkel, D.J., 2005b. Electrophysiological
properties of neurons in the basal ganglia of the domestic
chick: conservation and divergence in the evolution of the
avian basal ganglia. J. Neurophysiol. 94, 454467.
Fernández, A.S., Pieau, C., Repérant, J., Boncinelli, E., Wasserf, M.,
1998. Expression of the Emx-1 and Dlx-1 homeobox genes
define three molecularly distinct domains in the telencephalon
of mouse, chick, turtle and frog embryos: implications for the
evolution of telencephalic subdivisions in amniotes.
Development 125, 20992111.
Figueredo-Cárdenas, G., Medina, L., Reiner, A., 1996. Calretinin is
largely localized to a unique population of striatal interneurons
in rats. Brain Res. 709, 145150.
Fisher, S.E., Scharff, C., 2009. FOXP2 as a molecular window into
speech and language. Trends Genet. 25, 166177.
Flames, N., Pla, R., Gelman, D.M., Rubenstein, J.L., Puelles, L.,
Marín, O., 2007. Delineation of multiple subpallial progenitor
domains by the combinatorial expression of transcriptional
codes. J. Neurosci. 27, 96829695.
Foidart, A., Lakaye, B., Grisar, T., Ball, G.F., Balthazart, J., 1999.
Estrogen receptor-beta in quail: cloning, tissue expression and
neuroanatomical distribution. J. Neurobiol. 40, 327342.
Gale, S.D., Perkel, D.J., 2005. Properties of dopamine release and
uptake in the songbird basal ganglia. J. Neurophysiol. 93,
18711879.
Gale, S.D., Perkel, D.J., 2010. A basal ganglia pathway drives
selective auditory responses in songbird dopaminergic neurons
via disinhibition. J. Neurosci. 30, 10271037.
Gale, S.D., Person, A.L., Perkel, D.J., 2008. A novel basal ganglia
pathway forms a loop linking a vocal learning circuit and its
dopaminergic input. J. Comp. Neurol. 508, 824839.
Gallagher, J.P., Orozco-Cabal, L.F., Liu, J., Shinnick-Gallagher, P., 2008.
Synaptic physiology of central CRH system. Eur. J. Pharmacol.
583, 215225.
García-Calero, E., Puelles, L., 2009. Enc1 expression in the chick
telencephalon at intermediate and late stages of development.
J. Comp. Neurol. 517, 564580.
García-López, M., Abellán, A., Legaz, I., Rubenstein, J.L.R., Puelles,
L., Medina, L., 2008. Histogenetic compartments of the mouse
centromedial and extended amygdala based on gene expression
patterns during development. J. Comp. Neurol. 506, 4674.
García-Moreno, F., Pedraza, M., Di Giovannantonio, L.G., Di Salvio,
M., López-Mascaraque, L., Simeone, A., De Carlos, J.A., 2010. A
neuronal migratory pathway crossing from diencephalon to
telencephalon populates amygdala nuclei. Nat. Neurosci. 13,
680689.
Garda, A.L., Puelles, L., Rubenstein, J.L.,Medina, L., 2002. Expression
patterns of Wnt8b and Wnt7b in the chicken embryonic brain
suggest a correlation with forebrain patterning centers and
morphogenesis. Neuroscience 113, 689698.
Gerfen, C.R., 1992. The neostriatal mosaic: multiple levels of
compartmental organization. Trends Neurosci. 15, 133139.
Gibbs, M.E., Maksel, D., Gibbs, Z., Hou, X., Summers, R.J., Small, D.H.,
2010. Memory loss caused by β-amyloid protein is rescued
by a β
3
-adrenoceptor agonist. Neurobiol. Aging 31, 614624.
Goldberg, J.H., Fee, M.S., 2010. Singing-related neural activity
distinguishes four classes of putative striatal neurons in the
songbird basal ganglia. J. Neurophysiol. 103, 20022014.
Goldberg, J.H., Adler, A.,Bergman, H., Fee, M.S., 2010. Singing-related
neural activity distinguishes two putative pallidal cell types
in the songbird basal ganglia: comparison to the primate
internal and external pallidal segments. J. Neurosci. 30,
70887098.
González, A., López, J.M., Marín, O., 2002a. Expression pattern of
the homeobox protein NKX2-1 in the developing Xenopus
forebrain. Brain Res. Gene Expr. Patterns 1, 181185.
González, A., López, J.M., Sánchez-Camacho, C., Marín, O., 2002b.
Regional expression of the homeobox gene NKX 2-1defines
pallidal and interneuronal populations in the basal ganglia of
amphibians. Neuroscience 114, 567575.
Goodman, I.J., Stitzel, R.E., 1977. Corpus striatal (paleostriatal
complex) interventions and stereotyped behavior in pigeons.
Soc. Neurosci. Abstr. 3, 37.
Goodman, I., Zacny, J., Osman, A., Azzaro, A., Donovan, C., 1982.
Lesion-produced telencephalic catecholamine imbalances and
altered operant pecking rates in pigeons. Physiol. Behav. 29,
10451050.
Goodman, I., Zacny, J., Osman, A., Azzaro, A., Donovan, C., 1983.
Dopaminergic nature of feeding-induced behavioral
stereotypes in stressed pigeons. Pharmacol. Biochem. Behav.
18, 153158.
Gray, T.S., Magnuson, D.J., 1987. Galanin-like immunoreactivity
within amygdaloid and hypothalamic neurons that project to
the midbrain central grey in rat. Neurosci. Lett. 83, 264268.
Gritti, I., Mainville, L., Jones, B.E., 1993. Codistribution of
GABA- with acetylcholine-synthesizing neurons in the basal
forebrain of the rat. J. Comp. Neurol. 329, 438457.
Gritti, I., Mainville, L., Mancia, M., Jones, B.E., 1997. GABAergic and
other noncholinergic basal forebrain neurons, together with
cholinergic neurons, project to the mesocortex and isocortex
in the rat. J. Comp. Neurol. 383, 163177.
Gritti, I., Manns, I.D., Mainville, L., Jones, B.E., 2003. Parvalbumin,
calbindin, or calretinin in cortically projecting and GABAergic,
cholinergic, or glutamatergic basal forebrain neurons of the
rat. J. Comp. Neurol. 458, 1131.
Groenewegen, H.J., Russchen, F.T., 1984. Organization of the
efferent projections of the nucleus accumbens to pallidal,
hypothalamic, and mesencephalic structures: a tracing and
immunohistochemical study in the cat. J. Comp. Neurol. 223,
347367.
Groenewegen, H.J., Uylings, H.B.M., 2000. The prefrontal cortex
and the integration of sensory, limbic and autonomic
information. Progr. Brain Res. 126, 328.
Grove, E.A., 1988. Neural associations of the substantia innominata
in the rat: afferent connections. J. Comp. Neurol. 277, 315346.
Gruss, M., Braun, K., 1996. Distinct activation of monoaminergic
pathways in chick brain in relation to auditory imprinting and
stressful situations: a microdialysis study. Neuroscience 76,
891899.
Guillamón, A., Segovia, S., 1997. Sex differences in the vomeronasal
system. Brain Res. Bull. 44, 377382.
Güntürkün, O., 2005. The avian prefrontal cortexand cognition.
Curr. Opin. Neurobiol. 15, 686693.
95BRAIN RESEARCH 1424 (2011) 67101
Haber, S.N., Nauta, W.J.H., 1983. Ramifications of the globus pallidus
in the rat as indicated by patterns of immunohistochemistry.
Neuroscience 9, 245260.
Haber, S.N., Watson, S.J., 1985. The comparative distribution of
enkephalin, dynorphin and substance P in the human globus
pallidus and basal forebrain. Neuroscience 14, 10111024.
Haesler, S., Wada, K., Nshdejan, A., Morrisey, E.E., Lints, T., Jarvis,
E.D., Scharff, C., 2004. FoxP2 expression in avian vocal learners
and non-learners. J. Neurosci. 24, 31643175.
Haesler, S., Rochefort, C., Georgi, B., Licznerski, P., Osten, P.,
Scharff, C., 2007. Incomplete and inaccurate vocal imitation
after knockdown of FoxP2 in songbird basal ganglia nucleus
area X. PLoS Biol. 5, e321. doi:10.1371/journal.pbio.0050321.
Heimer, L., Alheid, G.F., 1991. Piecing together the puzzle of basal
forebrain anatomy. In: Napier, T.C., Kalivas, P.W., Hanin, I.
(Eds.), The Basal Forebrain: Anatomy to Function. Plenum
Press, New York, pp. 142.
Heimer, L., Wilson, R.D., 1975. The subcortical projections of
allocortex: similarities in the neuronal associations of the
hippocampus, the piriform cortex and the neocortex. In:
Santini, M. (Ed.), Golgi Centennial Symposium Proceedings.
Raven Press, New York, pp. 173193.
Heimer, L., Roberts, M., Switzer, R., 1976. The olfactory tubercle: a
structure of varying stratifica tion. Exp. Brain Res. Suppl. 1, 141147.
Heimer, L., Alheid, G.F., de Olmos, J.S., Groenewegen, H.J., Haber,
S.N., Harlan, R.E., Zahm, D.S., 1997. The accumbens: beyond the
core-shell dichotomy. J. Neuropsychiatry Clin. Neurosci. 9,
354381.
Herkenham, M., Moon-Edley, S., Stuart, J., 1984. Cell clusters in the
nucleus accumbens of the rat and the mosaic relationship of
opiate receptors, acetylcholinesterase and subcortical afferent
terminations. Neuroscience 11, 561593.
Herrick, C.J., 1956. The Evolution of Human Nature. University of
Texas Press, Austin.
Hirata, T., Li, P., Lanuza, G.M., Cocas, L.A., Huntsman, M.M., Corbin,
J.C., 2009. Identification of distinct telencephalic progenitor
pools for neuronal diversity in the amygdala. Nat. Neurosci. 12,
141149.
Holmgren,N., 1925. Points of view concerning forebrain morphology
in higher vertebrates. Acta Zool. 6, 413477.
Hoover, B.R., Marshall, J.F., 1999. Population characteristics of
preproenkephalin mRNA-containing neurons in the globus
pallidus of the rat. Neurosci. Lett. 265, 199202.
Hoover, B.R., Marshall, J.F., 2002. Further characterization of
preproenkephalin mRNA-containing cells in the rodent globus
pallidus. Neuroscience 111, 111125.
Horn, G., 1998. Visual imprinting and the neural mechanisms of
recognition memory. Trends Neurosci. 21, 300305.
Huot, P., Parent, A., 2007. Dopaminergic neurons intrinsic to the
striatum. J. Neurochem. 101, 14411447.
Husband, S.A., Shimizu, T., 2011. Calcium-binding protein
distributions and fiber connections of the nucleus accumbens
in the pigeon (Columba livia). J. Comp. Neurol. 519, 13711394.
Ikemoto, S., 2007. Dopamine reward circuitry: two projection
systems from the ventral midbrain to the nucleus
accumbens-olfactory tubercle complex. Brain Res. Rev. 56 (1),
2778.
Izawa, E.-I., Zachar, G., Yanagihara, S., Matsushima, T., 2003.
Localized lesion of caudal part of lobus parolfactorius caused
impulsivechoice in the domestic chick: evolutionarily conserved
function of ventral striatum. J. Neurosci. 23, 18941902.
Jarvis, E.D., Gunturkun, O., Bruce, L., Csillag, A., Karten, H., Kuenzel,
W., Medina, L., Paxinos, G., Perkel, D.J., Shimizu, T., Striedter, G.,
Wild, J.M., Ball, G.F., Dugas-Ford, J., Durand, S.E., Hough, G.E.,
Husband, S., Kubikova, L., Lee, D.W., Mello, C.V., Powers, A.,
Siang, C., Smulders, T.V., Wada, K., White, S.A., Yamamoto, K.,
Yu, J., Reiner, A., Butler, A.B., 2005. Avian brains and a new
understanding of vertebrate brain evolution. Nat. Rev. Neurosci.
6, 151159.
Jiao, Y., Medina, L., Veenman, C.L., Toledo, C., Puelles, L., Reiner,
A., 2000. Identification of the anterior nucleus of the ansa
lenticularis in birds as the homolog of the mammalian
subthalamic nucleus. J. Neurosci. 20, 69987010.
Jozsa,R.,Korf,H.W.,Csernus,V.,Mess,B.,1988.Thyrotropin-releasing
hormone (TRH)-immunoreactive structures in the brain of the
domestic mallard. Cell Tissue Res. 251, 441449.
Juorio, A.V., Vogt, M., 1967. Monoamines and their metabolites in
the avian brain. J. Physiol. 189, 489518.
Jurkevich, A., Barth, S.W., Grossmann, R., 1997. Sexual
dimorphism of arg-vasotocin gene expressing neurons in the
telencephalon and dorsal diencephalon of the domestic fowl.
An immunocytochemical and in situ hybridization study. Cell
Tissue Res. 287, 6977.
Jurkevich, A., Barth, S.W., Kuenzel, W.J., Kohler, A., Grossman, R.,
1999. Development of sexually dimorphic vasotocinergic
system in the bed nucleus of stria terminalis in chickens.
J. Comp. Neurol. 408, 4660.
Kalivas, P.W., Duffy, P., 1995. Selective activation of dopamine
transmission in the shell of the nucleus accumbens by stress.
Brain Res. 675, 325328.
Källén, B., 1951. On the ontogeny of the reptilian forebrain. Nuclear
structures and ventricular sulci. J. Comp. Neurol. 95, 397447.
Källén, B., 1953. On the nuclear differentiation during ontogenesis
in the avian forebrain and some notes on the amniote
strio-amygdaloid complex. Acta Anat. (Basel) 17, 7284.
Källén, B., 1962. Embryogenesis of brain nuclei in the chick
telencephalon. Ergeb. Anat. Entwicklungsgesch 36, 6282.
Kao, M.H., Doupe, A.J., Brainard, M.S., 2005. Contributions of an
avian basal gangliaforebrain circuit to real-time modulation
of song. Nature 433, 638643.
Kaoru, T., Liu, F.-C., Ishida, M., Oishi, T., Hayashi, M., Kitagawa, M.,
Shimoda, K., Takahashi, H., 2010. Molecular characterization
of the intercalated cell masses of the amygdala: implications
for the relationship with the striatum. Neuroscience 166,
220230.
Karle, E.J., Anderson, K.D., Medina, L., Reiner, A., 1996. Light and
electron microscopic immunohistochemical study of the
dopaminergic terminals in the pigeon straitum using antisera
against dopamine and tyrosine hydroxylase. J. Comp. Neurol.
369, 109124.
Karten, H.J., 1969. The organization of the avian telencephalon
and some speculations on the phylogeny of the amniote
telencephalon. Ann. N. Y. Acad. Sci. 167, 146179.
Karten, H.J., Dubbeldam, J.L., 1973. The organization and projections
of the paleostriatal complex in the pigeon (Columba livia).
J. Comp. Neurol. 148, 6189.
Kawai, Y., Takami, K., Shiosaka, S., Emson, P.C., Hillyard, C.J., Girgis,
S., MacIntyre, I., Tohyama, M., 1985. Topographic localization
of calcitonin gene related peptide in the rat brain: an
immunohistochemical analysis. Neuroscience 15, 747763.
Kelley, A.E., 1999.Neural integrative activities of nucleus accumbens
subregions in relation to motivation andlearning. Psychobiology
27, 198213.
King, D., Zigmond, M.J., Finlay, J.M., 1997. Effects of dopamine
depletion in the medial prefrontal cortex on the stress-induced
increase in extracellular dopamine in the nucleus accumbens
core and shell. Neuroscience 77, 141153.
Kiss, J.Z., Voorhuis, T.A., van Eekelen, J.A., de Kloet, E.R., de Wied,
D., 1987. Organization of vasotocin-immunoreactive cells and
fibers in the canary brain. J. Comp. Neurol. 263, 347364.
Kita, H., Kita, T., 2001. Number, origins, and chemical types of rat
pallidostriatal projection neurons. J. Comp. Neurol. 437,
438448.
Kitt, C.A., Brauth, S.E., 1981. Projections of the paleostriatum upon the
midbrain tegmentum in the pigeon. Neuroscience 6, 15511566.
Kitt, C.A., Brauth, S.E., 1986a. Telencephalic projections from
midbrain and isthmal cell groups in the pigeon. I. Locus
coeruleus and subcoeruleus. J. Comp. Neurol. 247, 6991.
96 BRAIN RESEARCH 1424 (2011) 67101
Kitt, C.A., Brauth, S.E., 1986b. Telencephalic projections from
midbrain and isthmal cell groups in the pigeon. II. The nigral
complex. J. Comp. Neurol. 247, 92110.
Klein, S., Jurkevich, A., Grossmann, R., 2006. Sexually dimorphic
immunoreactivity of galanin and colocalization with arginine
vasotocin in the chicken brain (Gallus gallus domesticus).
J. Comp. Neurol. 499, 828839.
Kohler, E.C., Messer Jr., W.S., Bingman, V.P., 1995. Evidence for
muscarinic acetylcholine receptor subtypes in the pigeon
telencephalon. J. Comp. Neurol. 362, 271282.
Kohler, E.C., Riters, L.V., Chaves, L., Bingman, V.P., 1996. The
muscarinic acetylcholine receptor antagonist scopoloamine
impairs short-distance homing pigeon navigation. Physiol.
Behav. 60, 10571061.
Koster, R., 1957. Comparative studies of emesis in pigeons and
dogs. J. Pharmacol. Exp. Ther. 119, 406417.
Krebs, J.R., Erichsen, J.T., Bingman, V.P., 1991. The distribution of
neurotransmitters and neurotransmitter-related enzymes in
the dorsomedial telencephalon of the pigeon (Columba livia).
J. Comp. Neurol. 314, 467477.
Kreitzer, A.C., Malenka, R.C., 2008. Striatal plasticity and basal
ganglia circuit function. Neuron 60, 543554.
Krogh, K., Ostergaard, K., Sabroe, S., Laurberg, S., 2008. Clinical
aspects of bowel symptoms in Parkinson's disease. Acta
Neurol. Scand. 117, 6064.
Krützfeldt,N.O.E., Wild, J.M., 2005. Definition and novel connections
of the entopallium of the pigeon(Columba livia). J. Comp. Neurol.
490, 4056.
Kubikova, L., Wada, K., Jarvis, E.D., 2010. Dopamine receptors in a
songbird brain. J. Comp. Neurol. 518, 741769.
Kuenzel, W.J., 1994. Central neuroanatomical systems involved in
the regulation of food intake in birds and mammals. J. Nutr.
124, 1355S1370S.
Kuenzel, W.J., 2000. The autonomic nervous system of birds. In:
Whittow, G.C. (Ed.), Sturkie's Avian Physiology. Academic
Press, San Diego, pp. 101122.
Kuenzel, W.J., Blähser, S., 1993. The visceral forebrain system in
birds: its proposed anatomical components and functions.
Poult. Avian Biol. Rev. 5, 2936.
Kuenzel, W.J., Masson, M., 1988. A Stereotaxic Atlas of the Brain of
the Chick (Gallus domesticus). Johns Hopkins University Press,
Baltimore.
Lanuza, E., Davies, D.C., Landete, J.M., Novejarque, A.,
Martínez-García, F., 2000. Distribution of CGRP-like
immunoreactivity in the chick and quail brain. J. Comp.
Neurol. 421, 515532.
Laverghetta, A.V., Toledo, C., Veenman, C.L., Reiner, A., 2005. Cellular
localization of AMPA-type glutamate receptor subunits in the
basal ganglia of pigeons (Columba livia). Brain Behav. Evol. 67,
1038.
Leblois, A., Person, A.L., Bodor, A., Perkel, D.J., 2009. Millisecond
timescale disinhibition mediates fast information
transmission through an avian basal ganglia loop.
J. Neurosci. 29, 1542015433.
Leblois, A., Wendel, B.J., Perkel, D.J., 2010. Striatal dopamine
modulates basal ganglia output and regulates social
context-dependent behavioral variability through D1
receptors. J. Neurosci. 30, 57305743.
Lévesque, M., Parent, A., 2005. The striatfugal fiber system in
primates:a reevaluation of itsorganization basedon single-axon
tracing studies. Proc. Natl. Acad. Sci. U. S. A. 102, 1188811893.
Long, J.E., Cobos, I., Potter, G.B., Rubenstein, J.L.R., 2009a. Dlx1&2
and Mash1 transcription factors control MGE and CGE
patterning and differentiation through parallel and overlapping
pathways. Cereb. Cortex 19, i96i106.
Long, J.E., Swan, C., Liang, W., Cobos, I., Potter, G.B., Rubenstein,
J.L.R., 2009b. Dlx1&2 and Mash1 transcription factors control
striatal patterning and differentiation through parallel and
overlapping pathways. J. Comp. Neurol. 512, 556572.
Luiten, P.G., Ter Horst, G.J., Steffens, A.B., 1987. The hypothalamus,
intrinsic connections and outflow pathways to the endocrine
system in relation to the control of feeding and metabolism.
Prog. Neurobiol. 28, 154.
Luo, M., Perkel, D.J., 1999. Long-range GABAergic projection in a
circuit essential for vocal learning. J. Comp. Neurol. 403,
6884.
Maier, V., Scheich, H., 1987. Acoustic imprinting in guinea fowl
chicks: age dependence of 2 deoxyglucose uptake in relevant
forebrain areas. Dev. Brain Res. 31, 1527.
Manabe, T., Tatsumi, K., Inoue, M., Makinodan, M., Yamauchi, T.,
Makinodan, E., Yokoyama, S., Sakumura, R., Wanaka, A., 2007.
L3/Lhx8 is a pivotal factor for cholinergic differentiation of
murine embryonic stem cells. Cell Death Differ. 14, 10801085.
Marín, O., Rubenstein, J.L., 2001. A long, remarkable journey:
tangential migration in the telencephalon. Nat. Rev. Neurosci. 2,
780790.
Marín, O., Rubenstein, J.L., 2003. Cell migration in the forebrain.
Annu. Rev. Neurosci. 26, 441448.
Marín, O., Anderson, S.A., Rubenstein, J.L., 2000. Origin and
molecular specification of striatal interneurons. J. Neurosci. 20,
60636076.
Marín, F., Herrero, M.T., Vyas, S., Puelles, L., 2005. Ontogeny of
tyrosine hydroxylase mRNA expression in mid- and forebrain:
neuromeric pattern and novel positive regions. Dev. Dyn. 234,
709717.
Martin, L.J., Spicer, D.M., Lewis, M.H., Gluck, J.P., Cork, L.C., 1991.
Social deprivation of infant rhesus monkeys alters the
chemoarchitecture of the brain: I. Subcortical regions. J. Neurosci.
11, 33443358.
Martínez-García, F., Novejarque, A., Lanuza, E., 2007. Evolution of
the amygdala in vertebrates. In: Kaas, J., Bullock, T.H. (Eds.),
Evolution of Nervous Systems: A Comprehensive Reference.
Non-mammalian Vertebrates, 2. Academic Press-Elsevier,
Amsterdam, pp. 255334.
Martínez-García, F., Novejarque, A., Lanuza, E., 2008. Two
interconnected functional systems in the amygdala of amniote
vertebrates. Brain Res. Bull. 75, 206213.
Martinez-Vargas, M.C., Keefer, D.A., Stumpf, W.E., 1978. Estrogen
localization in the brain of the lizard, Anolis carolinensis (1).
J. Exp. Zool. 205, 141147.
McKeegan, D.E., 2002. Spontaneous and odour evoked activity in
single avian olfactory bulb neurones. Brain Res. 929, 4858.
Meade, S., Denbow, D.M., 2003. The interaction of bombesin and
corticotropin-releasing hormone on ingestive behavior in the
domestic fowl. Physiol. Behav. 78, 611614.
Medina, L., Abellán, A., 2009. Development and evolution of the
pallium. Semin. Cell Dev. Biol. 20, 698711.
Medina, L., Reiner, A., 1994. Distribution of cholineacetyltransferase
immunoreactivity in the pigeon brain. J. Comp. Neurol. 342,
497537.
Medina, L., Reiner, A., 1995. Neurotransmitter organization and
connectivity of the basal ganglia in vertebrates: implications
for the evolution of basal ganglia. Brain Behav. Evol. 46,
235258.
Medina, L., Reiner, A., 1997. The efferent projections of the
dorsal and ventral pallidal parts of the pigeon basal ganglia,
studied with biotinylated dextran amine. Neuroscience 81,
773802.
Medina, L., Anderson, K.D., Karle, E.J., Reiner, A., 1995. An
ultrastructural double-label immunohistochemical study of
the enkephalinergic input to dopaminergic neurons of the
substantia nigra in pigeons. J. Comp. Neurol. 357, 408432.
Mello, C.V., Pinaud, R., Ribeiro, S., 1998. Noradrenergic system of
the zebra finch brain: immunocytochemical study of
dopamine-beta-hydroxylase. J. Comp. Neurol. 400, 207228.
Metherate, R., Tremblay, N., Dykes, R.W., 1988. Transient and
prolonged effects of acetylcholine on responsiveness of cat
somatosensory cortical neurons. J. Neurophysiol. 59, 12531276.
97BRAIN RESEARCH 1424 (2011) 67101
Metherate, R., Cox, C.L., Ashe, J.H., 1992. Cellular bases of
neocortical activation: modulation of neural oscillations by the
nucleus basalis and endogenous acetylcholine. J. Neurosci. 12,
47014711.
Metzger, M., Jiang, S., Wang, J., Braun, K., 1996. Organization of the
dopaminergic innervation of forebrain areas relevant to
learning: a combined immunohistochemical/retrograde tracing
study in the domestic chick. J. Comp. Neurol. 376, 127.
Metzger,M.,Jiang,S.,Braun,K.,1998.Organizationofthedorsocaudal
neostriatal complex: a retrograde and anterograde tracing study
in the domestic chick withspecial emphasis on pathways
relevant to imprinting. J. Comp. Neurol. 395, 380404.
Mezey, S., Csillag, A., 2002. Selective striatalconnections of midbrain
dopaminergic nuclei in the chick (Gallus domesticus). Cell Tissue
Res. 308, 3546.
Mineau, P., Boag, P.T., Beninger, R.J., 1994. The effects of
physostigmine and scopolamine on memory for food caches in
the black-capped chickadee. Pharmacol. Biochem. Behav. 49,
363370.
Moga, M.M., Gray, T.S., 1985. Peptidergic efferents from the
intercalated nuclei of the amygdala to the parabrachial nucleus
in the rat. Neurosci. Lett. 61, 1318.
Moga, M.M., Saper, C.B., Gray, T.S., 1989. Bed nucleus of the stria
terminalis: cytoarchitecture, immunohistochemistry, and
projection to the parabrachial nucleus in the rat. J. Comp.
Neurol. 283, 315332.
Molnar, M., Casini, G., Davis, B.M., Bagnoli, P., Brecha, N., 1994.
Distribution of proenkephalin mRNA in the chicken and
pigeon telencephalon. J. Comp. Neurol. 348, 419432.
Montagnese, C.M., Szekely, A.D., Adam, A., Csillag, A., 2004. Efferent
connections of septal nuclei of the domestic chick (Gallus
domesticus): an anterograde pathwaytracing study with a bearing
on functional circuits. J. Comp. Neurol. 469, 437456.
Moons, L., Van, G.J., Ghijsels, E., Vandesande, F., 1994.
Immunocytochemical localization of L-dopa and dopamine in
the brain of the chicken (Gallus domesticus). J. Comp. Neurol.
346, 97118.
Moons, L., D'Hondt, E., Pijcke, K., Vandesande, F., 1995. Noradrenergic
system in the chicken brain: immunocytochemical study with
antibodies to noradrenaline and dopamine-beta-hydroxylase.
J. Comp. Neurol. 360, 331348.
Moreno, N., Bachy, I., Retaux, S., González, A., 2004.
LIM-homeodomain genes as developmental and adult genetic
markers of Xenopus forebrain functional subdivisions. J. Comp.
Neurol. 472, 5272.
Moreno, N., Domínguez, L., Rétaux, S., González, A., 2008a. Islet1
as a marker of subdivisions and cell types in the developing
forebrain of Xenopus. Neuroscience 154, 14231439.
Moreno, N., González, A., Rétaux, S., 2008b. Evidences for tangential
migrations in Xenopus telencephalon: developmental patterns
and cell tracking experiments. Dev. Neurobiol. 68, 504520.
Moreno, N., Rétaux, S., González, A., 2008c. Spatio-temporal
expression of Pax6 in Xenopus forebrain. Brain Res. 1239, 9299.
Moreno, N., González, A., Rétaux, S., 2009. Development and
evolution of the subpallium. Semin. Cell Dev. Biol. 20, 735743.
Morona, R., González, A., 2008. Calbindin-D28k and calretinin
expression in the forebrain of anuran and urodele amphibians:
further support for newly identified subdivisions. J. Comp.
Neurol. 511, 187220.
Nauta, W.J.H., Karten, H.J., 1970. A general profile of the vertebrate
brain, with sidelights on the ancestry of the cerebral cortex. In:
Schmitt, F.O. (Ed.), The Neurosciences, Second Study Program.
Rockefeller University Press, New York, pp. 726.
Nery, S., Fishell, G., Corbin, J.G., 2002. The caudal ganglionic
eminence is a source of distinct cortical and subcortical cell
populations. Nat. Neurosci. 5, 12791287.
Nóbrega-Pereira, S., Gelman, D., Bartolini, G., Pla, R., Pierani, A.,
Marín, O., 2010. Origin and molecular specification of globus
pallidus neurons. J. Neurosci. 30, 28242834.
Nottebohm, F., Stokes, T.M., Leonard, C.M., 1976. Central control of
song in the canary, Serinus canaria. J. Comp. Neurol. 165,
457486.
Okuhata, S., Saito, N., 1987. Synaptic connections of thalamo-cerebral
vocal nuclei of the canary. Brain Res. Bull. 18, 3544.
Ölveczky, B.P., Andalman, A.S., Fee, M.S., 2005. Vocal experimentation
in the juvenile songbird requires a basal ganglia circuit. PLoS Biol.
3 (5), e153.
Pal, D., Mallick, B.N., 2004. GABA in pedunculo pontine tegmentum
regulates spontaneous rapid eye movementsleep by acting on
GABA
A
receptors in freely moving rats. Neurosci. Lett. 365,
200204.
Panguluri, S., Saggu, S., Lundy, R., 2009. Comparison ofsomatostatin
and corticotrophin-releasing hormone immunoreactivity in
forebrain neurons projecting to taste-responsive and
non-responsive regions of the parabrachial nucleus in rat. Brain
Res. 1298, 5769.
Panzica, G.C., Viglietti-Panzica, C., Fasolo, A., Vandesande, F.,
1986. CRF-like immunoreactive system in the quail brain.
J. Hirnforsch. 27, 539547.
Panzica, G.C., Arévalo, R., Sánchez, F., Alonso, J.R., Aste, N.,
Viglietti-Panzica, C., Aijón, J., Vázquez, R., 1994. Topographical
distribution of reduced nicotinamide adenine dinucleotide
phosphate-diaphorase in the brain of the Japanese quail.
J. Comp. Neurol. 342, 97114.
Panzica, G.C., Garzino, A., García-Ojeda, E., 1996. Coexistence of
NADPH-diaphorase and tyrosine hydroxylase in the
mesencephalic catecholaminergic system of the Japanese
quail. J. Chem. Neuroanat. 11, 3747.
Panzica, G.C., Castagna,C., Viglietti-Panzica, C., Russo, C., Tlemçani,
O., Balthazart, J., 1998. Organizational effects of estrogens on
brain vasotocin and sexual behavior in quail. J. Neurobiol. 37,
684699.
Panzica, G., Pessatti, M., Viglietti-Panzica, C., Grossmann, R.,
Balthazart,J., 1999. Effects of testosterone on sexually dimorphic
parvocellular neurons expressing vasotocin mRNA in the male
quail brain. Brain Res. 850, 5562.
Paré, D., Smith, Y., 1994. GABAergic projection from the intercalated
cell masses of the amygdala to the basal forebrain in cats.
J. Comp. Neurol. 344, 3349.
Paré, D., Quirk, G.J., Ledoux, J.E., 2004. New vistas on amygdala
networks in conditioned fear. J. Neurophysiol. 92, 19.
Parent, A., Hazrati, L.-N., 1995. Functional anatomy of the basal
ganglia. I. The cortico-basal ganglia thalamo-cortical loop.
Brain Res. Rev. 20, 91127.
Parent, A., Charara, A., Pinault, D., 1995. Single striatofugal axons
arborizing in both pallidal segments and in the substantia
nigra in primates. Brain Res. 698, 280284.
Patterson, T.A., Lipton, J.R., Bennett, E.L., Rozenzweig, M.R., 1990.
Cholinergic receptor antagonists impair formation of
intermediate-term memory in the chick. Behav. Neural Biol.
54, 6374.
Person, A.L., Gale, S.D., Farries, M.A., Perkel, D.J., 2008. Organization
of songbird basal ganglia, including Area X. J. Comp. Neurol. 508,
840866.
Pierce, R.C., Kalivas, P.W., 1995. Amphetamine produces
sensitized increases in locomotion and extracellular dopamine
preferentially in the nucleus accumbens shell of rats
administered repeated cocaine. J. Pharmacol. Exp. Ther. 275,
10191029.
Pombero, A., Martínez, S., 2009. Telencephalic morphogenesis
during the process of neurulation. An experimental study
using quail-chick chimeras. J. Comp. Neurol. 512, 784797.
Pontieri, F.E., Colangelo, V., La Riccia, M., Pozzilla, C., Passarelli, F.,
Orzi, F., 1994. Psychostimulant drugs increase glucose
utilization in the shell of the rat nucleus accumbens.
Neuroreport 5, 25612564.
Pontieri, F.E., Tanda, G., Di Chiara, G., 1995. Intravenous cocaine,
morphine, and amphetamine preferentially increase
98 BRAIN RESEARCH 1424 (2011) 67101
extracellular dopamine in the shellas compared with the
coreof the rat nucleus accumbens. Proc. Natl. Acad. Sci. U. S. A.
92, 1230412308.
Poulin, A.-M., Timofeeva, E., 2008. The dynamics of neuronal
activation during food anticipation and feeding in the brain of
food-entrained rats. Brain Res. 1227, 128141.
Price, J.L., Amaral, D.G., 1981. An autoradiographic study of the
projections of the central nucleus of the monkey amygdala.
J. Neurosci. 1, 12421259.
Puelles, L., Medina, L., 1994. Development of neurons expressing
tyrosine hydroxylase and dopamine in the chicken brain: a
comparative segmental analysis. In: Smeets, W.J.A.J., Reiner, A.
(Eds.), Phylogeny and Development of Catecholamine Systems
in the CNS of Vertebrates. Cambridge Univ. Press, Cambridge,
pp. 381404.
Puelles, L., Kuwana, E., Puelles, E., Bulfone, A., Shimamura, K.,
Keleher, J., Smiga, S., Rubenstein, J.L., 2000. Pallial and subpallial
derivatives in the embryonic chick and mouse telencephalon,
traced by the expression of the genes Dlx-2, Emx-1, Nkx-2.1,
Pax-6, and Tbr-1. J. Comp. Neurol. 424, 409438.
Puelles, L., Martínez-de-la-Torre, M., Paxinos, G., Watson, C.,
Martínez, S., 2007. The Chick Brain in Stereotaxic Coordinates.
Academic Press, San Diego.
Redies, C., Medina, L., Puelles, L., 2001. Cadherin expression by
embryonic subdivisions and derived gray matter structures in
the telencephalon of the chicken. J. Comp. Neurol. 438,
253285.
Reiner, A., 2005. A new avian brain nomenclature: why, how and
what. Brain Res. Bull. 66, 317331.
Reiner, A., Anderson, K.D., 1990. The patterns of neurotransmitter
and neuropeptide co-occurrence among striatal projection
neurons: conclusions based on recent findings. Brain Res. Rev.
15 (3), 251265.
Reiner, A., Anderson, K.D., 1993. Co-occurrence of
gamma-aminobutyric acid, parvalbumin and the
neurotensin-related neuropeptide LANT6 in pallidal, nigral
and striatal neurons in pigeons and monkeys. Brain Res. 624,
317325.
Reiner, A., Carraway, R.E., 1987. Immunohistochemical and
biochemical studies on Lys8-Asn9-neurotensin 813
(LANT6)-related peptides in the basal ganglia of pigeons,
turtles, and hamsters. J. Comp. Neurol. 257, 453476.
Reiner, A., Karten, H.J., 1985. Comparison of olfactory bulb
projections in pigeons and turtles. Brain Behav. Evol. 27,
1127.
Reiner, A., Brecha, N.C., Karten, H.J., 1982a. Basal ganglia pathways
to the tectum: the afferent and efferent connections of the
lateral spiriform nucleus of pigeon. J. Comp. Neurol. 208, 1636.
Reiner, A., Karten, H.J., Brecha, N.C., 1982b. Enkephalin-mediated
basal ganglia influences over the optic tectum: immunohisto-
chemistry of the tectum and the lateral spiriform nucleus in
pigeon. J. Comp. Neurol. 208, 3753.
Reiner, A., Karten, H.J., Solina, A.R., 1983. Substance P: localization
within paleostriatal-tegmental pathways in the pigeon.
Neuroscience 9, 6185.
Reiner, A., Davis, B.M., Brecha, N.C., Karten, H.J., 1984. The distribution
of enkephalinlike immunoreactivity in the telencephalon of the
adult and developing domestic chicken. J. Comp. Neurol. 228,
245262.
Reiner, A., Karle, E.J., Anderson, K.D., Medina, L., 1994.
Catecholaminergic perikarya and fibers in the avian nervous
system. In: Smeets, W.J.A.J., Reiner, A. (Eds.), Phylogeny and
Development of Catecholaminergic Systems in the CNS of
Vertebrates. Cambridge University Press, Cambridge, pp. 135181.
Reiner, A., Medina, L., Veenman, C.L., 1998a. Structural and
functional evolution of the basal ganglia in vertebrates. Brain
Res. Rev. 28, 235285.
Reiner, A., Perera, M., Paullus, R., Medina, L., 1998b.
Immunohistochemical localization of DARPP-32 in striatal
projection neurons and striatal interneurons in pigeons.
J. Chem. Neuroanat. 16, 1733.
Reiner, A., Medina, L., Haber, S.N., 1999. The distribution of
dynorphinergic terminals in striatal target regions in comparison
to the distribution of substance P-containing and enkephalinergic
terminals in monkeys and humans. Neuroscience 88 (3), 775793.
Reiner, A., Stern, E.A., Wilson, C.J., 2001. Physiology and morphology
of intratelencephalically projecting corticostriatal-type neurons
in pigeons as revealed by intracellular recording and cell filling.
Brain Behav. Evol. 58, 101114.
Reiner, A., Laverghetta, A.V., Meade, C.A., Cuthbertson, S.L., Bottjer,
S.W., 2004a. An immunohistochemical and pathway tracing
study on the striatopallidal organization of Area X in the male
zebra finch. J. Comp. Neurol. 469, 239261.
Reiner, A., Perkel, D.J., Bruce, L.L., Butler, A.B., Csillag, A., Kuenzel,
W., Medina, L., Paxinos, G., Shimizu, T., Striedter, G., Wild, M.,
Ball, G.F., Durand, S., Güntürkün, O., Lee, D.W., Mello, C.V.,
Powers, A., White, S.A., Hough, G., Kubikova, L., Smulders, T.V.,
Wada, K., Dugas-Ford, J., Husband, S., Yamamoto, K., Yu, J.,
Siang, C., Jarvis, E.D., 2004b. Revised nomenclature for avian
telencephalon and some related brainstem nuclei. J. Comp.
Neurol. 473, 377414.
Richard, S., Martinez-Garcia, F., Lanuza, E., Davies, D.C., 2004.
Distribution of corticotropin-releasing factor-immunoreactive
neurons in the central nervous system of the domestic chicken
and Japanese quail. J. Comp. Neurol. 469, 559580.
Richfield, E.K., Young, A.B., Penney, J.B., 1987. Comparative
distribution of dopamine D-1 and D-2 receptors in the basal
ganglia of turtles, pigeons, rats, cats, and monkeys. J. Comp.
Neurol. 262, 446463.
Rieke, G.K., 1980. Kainic acid lesions of pigeon paleostriatum: a
model for study of movement disorders. Physiol. Behav. 24,
683687.
Rieke, G.K., 1981. Movement disorders and lesions of pigeon
brain stem analogues of basal ganglia. Physiol. Behav. 26,
379384.
Rieke, G.K., 1982. The TPc, the avian substantia nigra: pharmacology
and behavior. Physiol. Behav. 28, 755763.
Roberts, D.C.S., Koob, G.F., Klonoff, P., Fibiger, H.C., 1980. Extinction
and recovery of cocaine self-administration following
6-hydroxydopamine lesions of the nucleus accumbens.
Pharmacol. Biochem. Behav. 12, 781787.
Roberts, T.F., Hall, W.S., Brauth, S.E., 2002. Organization of theavian
basal forebrain: chemical anatomy in the parrot (Melopsittacus
undulatus). J. Comp. Neurol. 454, 383408.
Roselli, C.E., 1991. Synergistic induction of aromatase activity in
rat brain by estradiol and 5 alpha-dihydrotestosterone.
Neuroendocrinology 53, 7984.
Russchen, F.T., 1982. Amygdalopetal projections in the cat. II.
Subcortical afferent connections. A study with retrograde
tracing techniques. J. Comp. Neurol. 207, 157176.
Russchen, F.T., Amaral, D.G., Price, J.L., 1985a. The afferent
connections of the substantia innominata in the monkey,
Macaca fascicularis. J. Comp. Neurol. 242, 127.
Russchen, F.T., Bakst, I., Amaral, D.G., Price, J.L., 1985b. The
amygdalostriatal projections in the monkey. An anterograde
tracing study. Brain Res. 329, 241257.
Salamone, J.D., Cousins, M.S., Snyder, B.J., 1997. Behavioral
functions of nucleus accumbens dopamine: empirical and
conceptual problems with the Anhedonia Hypothesis.
Neurosci. Biobehav. Rev. 21, 341359.
Sanberg, P.R., Mark, R.F., 1983. The effect of striatal lesions in the
chick on haloperidol-potentiated tonic immobility.
Neuropharmacology 22, 253257.
Savage, L.M., Stanchfield, M.A., Overmeier, J.B., 1994. The effects
of scopolamine, diazepam, and lorazepam on working
memory in pigeons: an analysis of reinforcement procedures
and sample problem type. Pharmacol. Biochem. Behav. 48,
183191.
99BRAIN RESEARCH 1424 (2011) 67101
Scharff, C., Nottebohm, F., 1991. A comparative study of the
behavioral deficits following lesions of various parts of the
zebra finch song system: implications for vocal learning.
J. Neurosci. 11, 28962913.
Schnabel, R., Braun, K., 1996. Development of dopamine receptors
in the forebrain of the domestic chick in relation to auditory
imprinting. An autoradiographic study. Brain Res. 720,
120130.
Schnabel, R., Metzger, M., Jiang, S., Hemmings Jr., H.C., Greengard,
P., Braun, K., 1997. Localization of dopamine D1 receptors and
dopaminoceptive neurons in the chick forebrain. J. Comp.
Neurol. 388, 146168.
Schwarcz, R., Fuxe, K., Agnati, L.F., Hökfelt, T., Coyle, J.T., 1979.
Rotational behaviour in rats with unilateral striatal kainic acid
lesions: a behavioural model for studies on intact dopamine
receptors. Brain Res. 170 (3), 485495.
Shinoda, K., Nagano, M., Osawa, Y., 1994. Neuronal aromatase
expression in preoptic, striatal, and amygdaloid regions during
late prenatal and early postnatal development in the rat.
J. Comp. Neurol. 343, 113129.
Smeets, W.J., Reiner, A., 1994. Catecholamines in the CNS of
vertebrates: current concepts of evolution and functional
significance. In: Smeets, W.J., Reiner, A. (Eds.), Phylogeny and
Development of Catecholamine Systems in the CNS of
Vertebrates. Cambridge Univ. Press, Cambridge, pp. 463481.
Soha, J.A., Shimizu, T., Doupe, A.J., 1996. Development of the
catecholaminergic innervation of the song system of the male
zebra finch. J. Neurobiol. 29, 473489.
Sohrabji,F.,Nordeen,E.J.,Nordeen,K.W.,1990.Selectiveimpairment
of song learning following lesions of a forebrain nucleus in the
juvenile zebra finch. Behav. Neural Biol. 53, 5163.
Spooner, C.E., Winters, W.D., 1966. Neuropharmacological profile
of the young chick. Int. J. Neuropharmacol. 5, 217236.
Stenman, J., Yu, R.T., Evans, R.M., Campbell, K., 2003. Tlx and Pax6
co-operate genetically to establish the pallio-subpallial
boundary in the embryonic mouse telencephalon. Development
130, 11131122.
Stewart,M.G.,Kabai,P.,Harrison,E.,Steele,R.J.,Kossut,M.,Gierdalski,
M., Csillag, A., 1996. The involvement of dopamine in the
striatum in passive avoidance training in the chick. Neuroscience
70, 714.
Striedter, G.F., 1997. The telencephalon of tetrapods in evolution.
Brain Behav. Evol. 49, 179213.
Striedter, G.F., Marchant, T.A., Beydler, S., 1998. The neostriatum
develops as part of the lateral pallium in birds. J. Neurosci. 18,
58395849.
Stühmer, T., Anderson, S.A., Ekker, M., Rubenstein, J.L.R., 2002a.
Ectopic expression of the Dlx genes induces glutamic acid
decarboxylase and Dlx expression. Development 129,
245252.
Stühmer, T., Puelles, L., Ekker, M., Rubenstein, J.L., 2002b. Expression
from a Dlx gene enhancer marks adult mouse cortical GABAergic
neurons. Cereb. Cortex 12, 7585.
Sun, Z., Reiner, A., 2000. Localization of dopamine D1A and D1B
receptor mRNAs in the forebrain and midbrain of the domestic
chick. J. Chem. Neuroanat. 19, 211224.
Sun, Z., Wang, H.B., Laverghetta, A.V., Yamamoto, K., Reiner, A.,
2005. The distribution and cellular localization of glutamic acid
decarboxylase-65 (GAD65) mRNA in the forebrain and
midbrain of domestic chick. J. Chem. Neuroanat. 29, 265281.
Sussel, L., Marín, O., Kimura, S., Rubenstein, J.L., 1999. Loss of
Nkx2.1 homeobox gene function results in a ventral to dorsal
molecular respecification within the basal telencephalon:
evidence for a transformation of the pallidum into the striatum.
Development 126, 33593370.
Swanson, L.W., 2000. Cerebral hemisphere regulation of motivated
behavior. Brain Res. 886, 113164.
Swanson, L.W., Petrovich, G.D., 1998. What is the amygdala?
Trends Neurosci. 21, 323331.
Switzer, R.C., Hill, J., Heimer, L., 1982. The globus pallidus and its
rostro-ventral extension intotheolfactorytubercleofthe
rat: a cyto- and chemoarchitectural study. Neuroscience 7,
18911904.
Székely, A.D., Krebs, J.R., 1996. Efferent connectivity of the
hippocampal formation of the zebra finch (Taenopygia guttata):
an anterograde pathway tracing study using Phaseolus vulgaris
leucoagglutinin. J. Comp. Neurol. 368, 198214.
Székely, A.D., Boxer, M.I., Stewart, M.G., Csillag, A., 1994. Connectivity
of the lobus parolfactorius of the domestic chicken (Gallus
domesticus): an anterograde and retrograde pathway tracing
study. J. Comp. Neurol. 348, 374393.
Tachibana, T., Sato, M., Oikawa, D., Furuse, M., 2006. Involvement
of CRF on the anorexic effect of GLP-1 in layer chicks. Comp.
Biochem. Physiol. A 143, 112117.
Tan, Y., Williams, E.S., Zahm, D.S., 1999. Calbindin-D 28kD
immunofluorescence in ventral mesencephalic neurons labeled
following injections of Fluoro-Gold in nucleus accumbens
subterritories: inverse relationship relative to known neurotoxin
vulnerabilities. Brain Res. 844, 6777.
Tanaka, M., Ikeda, T., Hayashi, S., Iijima, N., Amaya, F., Hisa, Y.,
Ibata, Y., 1997. Nitrergic neurons in the medial amygdala
project to the hypothalamic paraventricular nucleus of the rat.
Brain Res. 777, 1321.
Teramitsu,I.,Kudo,L.C.,London,S.E.,Geschwind,D.H.,White,
S.A., 2004. Parallel FoxP1 and FoxP2 expressionin songbird and
human brain predicts functional interactions. J. Neurosci. 24,
31523163.
Thode, C., Bock, J., Braun, K., Darlison, M.G., 2005. The chicken
immediate-early gene ZENK is expressed in the medio-rostral
neostriatum/hyperstriatum ventrale, a brain region involved
in acoustic imprinting, and is up-regulated after exposure to
an auditory stimulus. Neuroscience 130, 611617.
Thompson, R.R., Goodson, J.L., Ruscio, M.G., Adkins-Regan, E.,
1998. Role of the archistriatal nucleus taeniae in the sexual
behavior of male Japanese quail (Coturnix japonica): a comparison
of function with the medial nucleus of the amygdala in
mammals. Brain Behav. Evol. 51, 215229.
Tole, S., Remedios, R., Saha, B., Stoykova, A., 2005. Selective
requirement of Pax6, but not Emx2, in the specification and
development of several nuclei of the amygdaloid complex.
J. Neurosci. 25, 27532760.
Toresson, H., Campbell, K., 2001. A role for Gsh1 in the developing
striatum and olfactory bulb of Gsh2 mutant mice. Development
128, 47694780.
Toresson, H., Parmar, M., Campbell, K., 2000. Expression of Meis
and Pbx genes and their protein products in the developing
telencephalon: implications for regional differentiation. Mech.
Dev. 94, 183187.
Usuda, I., Tanaka, K., Chiba, T., 1998. Efferent projections of the
nucleus accumbens in the rat with special reference to
subdivision of the nucleus: biotinylated dextran amine study.
Brain Res. 797, 7393.
van der Akker, W.M., Brox, A., Puelles, L., Durston, A.J., Medina, L.,
2008. Comparative functional analysis provides evidence for a
crucial role for the homeobox gene Nkx2.1/Titf-1 in forebrain
evolution. J. Comp. Neurol. 506, 211223.
van der Kooy, D., Koda, L.Y., McGinty, J.F., Gerfen, C.R., Bloom, F.E.,
1984. The organization of projections from the cortex,
amygdala, and hypothalamus to the nucleus of the solitary
tract in rat. J. Comp. Neurol. 224, 124.
Veenman, C.L., Reiner, A., 1994. The distribution of
GABA-containing perikarya, fibers, and terminals in the
forebrain and midbrain of pigeons, with particular reference to
the basal ganglia and its projection targets. J. Comp. Neurol.
339, 209250.
Veenman, C.L., Reiner, A., 1996. Ultrastructural morphology of
synapses formed by corticostriatal terminals in the avian
striatum. Brain Res. 707, 112.
100 BRAIN RESEARCH 1424 (2011) 67101
Veenman, C.L., Albin, R.L., Richfield, E.K., Reiner, A., 1994.
Distributions of GABA
A
, GABA
B
, and benzodiazepine receptors
in the forebrain and midbrain of pigeons. J. Comp. Neurol. 344,
161189.
Veenman, C.L., Wild, J.M., Reiner, A., 1995. Organization of the avian
corticostriatalprojection system: a retrograde and anterograde
pathwaytracingstudyinpigeons.J.Comp.Neurol.354,87126.
Veenman, C.L., Medina, L., Reiner, A., 1997. Avian homologues of
the mammalian intralaminar, mediodorsal and midline
thalamic nuclei: Immunohistochemical and hodological
evidence. Brain Behav. Evol. 49, 7898.
Viglietti-Panzica,C., Anselmetti, G.C., Balthazart, J., Aste,N., Panzica,
G.C., 1992. Vasotocinergic innervation of the septal region in
the Japanese quail: sexual differences and the influence of
testosterone. Cell Tissue Res. 267, 261265.
Vincent, S.R., Johansson, O., Hökfelt, T., Skirboll, L., Elde, R.P.,
Terenius, L., Kimmel, J., Goldstein, M., 1983. NADPH-diaphorase:
a selective histochemical marker for striatal neurons containing
both somatostatin- and avian pancreatic polypeptide
(APP)-like immunoreactivities. J. Comp. Neurol. 217, 252263.
von Bartheld,C.S., Bothwell, M., 1992.Development and distribution
of noradrenergic and cholinergic neurons and their trophic
phenotypes in the avian ceruleus complex and midbrain
tegmentum. J. Comp. Neurol. 320, 479500.
Voorhuis, T.A.M., Kiss, J.Z., de Kloet, E.R., de Wied, D., 1988.
Testosterone-sensitive vasotocin-immunoreactive cells and
fibers in the canary brain. Brain Res. 442, 139146.
Voorn, P., Gerfen, C.R., Groenewegen, H.J., 1989. Compartmental
organization of the ventral striatum of the rat:
immunohistochemical distribution of enkephalin, substance
P, dopamine, and calcium-binding protein. J. Comp. Neurol.
289, 189201.
Wächtler, K., Ebinger, P., 1989. The pattern of muscarinic
acetylcholine receptor binding in the avian forebrain.
J. Hirnforsch. 30, 409414.
Waclaw, R.R., Ehrman, L.A., Pierani, A., Campbell, K., 2010.
Developmental origin of the neuronal subtypes that comprise
the amygdalar fear circuit in the mouse. J. Neurosci. 30,
69446953.
Wada, K., Hagiwara, M., Jarvis, E.D., 2001. Brain evolution revealed
through glutamate receptor expression profiles. Abstr. Soc.
Neurosci. 27, 1425.
Wada, K., Sakaguchi, H., Jarvis, E.D., Hagiwara, M., 2004. Differential
expression of glutamate receptors in avian neural pathways for
learned vocalization. J. Comp. Neurol. 476, 4464.
Waldmann, C., Güntürkün, O., 1993. The dopaminergic innervation
of the pigeon caudolateral forebrain: immunocytochemical
evidence for a prefrontal cortexin birds? Brain Res. 600 (2),
225234.
Watson, J.T., Adkins-Regan, E., 1989. Neuroanatomical localization
of sex steroid concentrating cells in the Japanese quail (Coturnix
japonica): Autoradiography with (
3
H)-estradiol, (
3
H)-testosterone
and (
3
H)-dihydrotestosterone. Neuroendocrinology 49, 5164.
White, S.A., Fisher, S.E., Geschwing, D.H., Scharff, C., Holy, T.E.,
2006. Singing mice, songbirds and more: models for FOXP2
function and dysfunction in human speech and language.
J. Neurosci. 26, 1037610379.
Wild, J.M., 1987. Thalamic projections to the paleostriatum and
neostriatum in the pigeon (Columba livia). Neuroscience 20,
305327.
Wild, J.M., Arends, J.J., Zeigler, H.P., 1990. Projections of the
parabrachial nucleus in the pigeon (Columba livia). J. Comp.
Neurol. 293, 499523.
Wild, J.M., Karten, H.J., Frost, B.J., 1993. Connections of the
auditory forebrain in the pigeon (Columba livia). J. Comp.
Neurol. 337, 3262.
Williams, H., Mehta, N., 1999. Changes in adult zebra finch song
require a forebrain nucleus that is not necessary for song
production. J. Neurobiol. 39, 1428.
Woolf, N.J., 1991. Cholinergic systems in mammalian brain and
spinal cord. Prog. Neurobiol. 37, 475524.
Wullimann, M.F., Mueller, T., 2004. Teleostean and mammalian
forebrains contrasted: evidence from genes to behavior.
J. Comp. Neurol. 475, 143162.
Xie, J., Kuenzel, W.J., Anthony, N.B., Jurkevich, A., 2010. Subpallial
and hypothalamic areas activated following sexual and
agonistic encounters in male chickens. Physiol. Behav. 101,
344359.
Xie, J., Kuenzel, W.J., Sharp, P.J., Jurkevich, A., 2011. Appetitive and
consummatory sexual and agonistic behavior elicits FOS
expression in aromatase and vasotocin neurons within the
preoptic area and bed nucleus of the stria terminalis of male
domestic chickens. J. Neuroendocrinol. 23, 232243.
Xu, Q., Tam, M., Anderson, S.A., 2008. Fate mapping Nkx2.1-lineage
cells in the mouse telencephalon. J. Comp. Neurol. 506, 1629.
Yamamoto, K., Reiner, A., 2005. Distribution of the limbic
system-associated membrane protein (LAMP) in pigeon
forebrain and midbrain. J. Comp. Neurol. 486, 221242.
Yamamoto, K., Sun, Z., Wang, H.B., Reiner, A., 2005. Subpallial
amygdala and nucleus taeniae in birds resemble extended
amygdala and medial amygdala in mammals in their
expression of markers of regional identity. Brain Res. Bull. 66,
341347.
Yanagihara, S., Izawa, E.I., Koga, K., Matsushima, T., 2001.
Reward-related neuronal activities in basal ganglia of
domestic chicks. Neuroreport 12, 14311453.
Yun, K., Potter, S., Rubenstein, J.L.R., 2001. Gsh2 and Pax6 play
complementary roles in dorsoventral patterning of the
mammalian telencephalon. Development 128, 193205.
Yun, K., Garel, S., Fischman, S., Rubenstein, J.L., 2003. Patterning of
the lateral ganglionic eminence by the Gsh1 and Gsh2
homeobox genes regulates striatal and olfactory bulb
histogenesis and the growth of axons through the basal
ganglia. J. Comp. Neurol. 461, 151165.
Záborszky, L., Alheid, G.F., Beinfeld, M.C., Eiden, L.E., Heimer, L.,
Palkovits, M., 1985. Cholecystokinin innervation of the ventral
striatum: a morphological and radioimmunological study.
Neuroscience 14, 427453.
Záborszky, L., Pang, K., Somogyi, J., Nadasdy, Z., Kallo, I., 1999. The
basal forebrain corticopetal system revisited. Ann. N. Y. Acad.
Sci. 877, 339367.
Zahm, D.S., 1989. Evidence for a morphologically distinct
subpopulation of striatopetal axons following injections of
WGA-HRP into the ventral tegmental area in the rat. Brain Res.
482, 145154.
Zahm, D.S., 1999. Functionalanatomical implications of the
nucleus accumbens core and shell subterritories. Ann. N. Y.
Acad. Sci. 877, 113128.
Zahm, D.S., 2000. An integrative neuroanatomical perspective on
some subcortical substrates of adaptive responding with
emphasis on the nucleus accumbens. Neurosci. Biobehav. Rev.
24, 85105.
Zahm, D.S., Brog, J.S., 1992. On the significance of subterritories in
the accumbenspart of the rat ventral striatum. Neuroscience
50, 751767.
Zahm, D.S., Heimer, L., 1993. Specificity in the efferent projections
of the nucleus accumbens in the rat: comparison of the rostral
pole projection patterns with those of the core and shell.
J. Comp. Neurol. 327, 220232.
Zhao, W.Q., Feng, H., Bennett, P., Ng, K.T., 1997. Inhibition of
intermediate-term memory following passive avoidance
training in neonate chicks by a presynaptic cholinergic
blocker. Neurobiol. Learn. Mem. 67, 207213.
Zhao, Y., Marin, O., Hermesz, E., Powell, A., Flames, N., Palkovits,
M., Rubenstein, J.L.R., Westphal, H., 2003. The LIM-homeobox
gene Lhx8 is required for the development of many cholinergic
neurons in the mouse forebrain. Proc. Natl. Acad. Sci. U. S. A.
100, 90059010.
101BRAIN RESEARCH 1424 (2011) 67101
... Regulation of the fear/stress response by the extended amygdala in nonmammals is still poorly understood. In birds, this is in part due to the high divergence of the telencephalon in evolution, which has made it challenging to identify the amygdala in sauropsids (reviewed by Kuenzel et al., 2011;Martínez-García et al., 2002Medina et al., 2011Medina et al., , 2017Medina et al., , 2023Reiner et al., 2004). The avian BSTL was identified in the pallidal territory of the subpallium, adjacent to the lateral ventricle (Aste et al., 1998;Kuenzel et al., 2011;Puelles et al., 2000;Reiner et al., 2004). ...
... In birds, this is in part due to the high divergence of the telencephalon in evolution, which has made it challenging to identify the amygdala in sauropsids (reviewed by Kuenzel et al., 2011;Martínez-García et al., 2002Medina et al., 2011Medina et al., , 2017Medina et al., , 2023Reiner et al., 2004). The avian BSTL was identified in the pallidal territory of the subpallium, adjacent to the lateral ventricle (Aste et al., 1998;Kuenzel et al., 2011;Puelles et al., 2000;Reiner et al., 2004). This nucleus has descending projections to hypothalamic and brainstem centers similar to those of mammalian BSTL involved in the regulation of the endocrine, autonomic, and behavioral aspects of the stress response (Atoji et al., 2006;Hanics et al., 2017). ...
... In mammals, CRF and ENK cells of the central amygdala play different roles: while CRF cells are involved in long-term components of fear learning and recall (related to anxiety; Asok et al., 2018;Davis et al., 2010;Gafford & Ressler, 2015;Pitts et al., 2009;Pomrenze et al., 2019), ENK/PKCδ cells promote anxiolysis and analgesia (Douceau et al., 2022;Paretkar & Dimitrov, 2019). Like in mammals, ENK cells and CRF cells of the avian central amygdala have subpallial origin Vicario et al., 2014Vicario et al., , 2015Vicario et al., , 2017present study) and are likely GABAergic, as typical for cells in the subpallium Kuenzel et al., 2011;Medina et al., 2011Medina et al., , 2017Medina et al., , 2023. Thus, projections from these cells to BSTL would lead to the inhibition of BSTL cells (Figure 10). ...
Article
Full-text available
In mammals, the central extended amygdala is critical for the regulation of the stress response. This regulation is extremely complex, involving multiple subpopulations of GABAergic neurons and complex networks of internal and external connections. Two neuron subpopulations expressing corticotropin‐releasing factor (CRF), located in the central amygdala and the lateral bed nucleus of the stria terminalis (BSTL), play a key role in the long‐term component of fear learning and in sustained fear responses akin to anxiety. Very little is known about the regulation of stress by the amygdala in nonmammals, hindering efforts for trying to improve animal welfare. In birds, one of the major problems relates to the high evolutionary divergence of the telencephalon, where the amygdala is located. In the present study, we aimed to investigate the presence of CRF neurons of the central extended amygdala in chicken and the local connections within this region. We found two major subpopulations of CRF cells in BSTL and the medial capsular central amygdala of chicken. Based on multiple labeling of CRF mRNA with different developmental transcription factors, all CRF neurons seem to originate within the telencephalon since they express Foxg1, and there are two subtypes with different embryonic origins that express Islet1 or Pax6. In addition, we demonstrated direct projections from Pax6 cells of the capsular central amygdala to BSTL and the oval central amygdala. We also found projections from Islet1 cells of the oval central amygdala to BSTL, which may constitute an indirect pathway for the regulation of BSTL output cells. Part of these projections may be mediated by CRF cells, in agreement with the expression of CRF receptors in both Ceov and BSTL. Our results show a complex organization of the central extended amygdala in chicken and open new venues for studying how different cells and circuits regulate stress in these animals.
... The pallium, especially the dorsal pallium, largely differs in each lineage of vertebrates: it is absent as a morphological entity in amphioxus 2 and is present as only a layered area in fishes and amphibians, while it is a simply layered region in reptiles and aves, but a large intricate multilayered cortex in mammals 3 . In contrast, the subpallial regions or the subpallium, the deep-seated basal ganglia (striatum and pallidum), are highly conserved in vertebrates with respect to the expression patterns of transcription factors, neuronal types and some neural connections, despite varieties in shapes and sizes [4][5][6] . Thus, from an evolutionarily point of view, the pallium has undergone divergent trajectories in different vertebrates, and the dorsal pallium has experienced the greatest development in mammals 7,8 . ...
Article
Full-text available
The mechanisms underlying the organization and evolution of the telencephalic pallium are not yet clear.. To address this issue, we first performed comparative analysis of genes critical for the development of the pallium (Emx1/2 and Pax6) and subpallium (Dlx2 and Nkx1/2) among 500 vertebrate species. We found that these genes have no obvious variations in chromosomal duplication/loss, gene locus synteny or Darwinian selection. However, there is an additional fragment of approximately 20 amino acids in mammalian Emx1 and a poly-(Ala)6–7 in Emx2. Lentiviruses expressing mouse or chick Emx2 (m-Emx2 or c-Emx2 Lv) were injected into the ventricle of the chick telencephalon at embryonic Day 3 (E3), and the embryos were allowed to develop to E12–14 or to posthatchling. After transfection with m-Emx2 Lv, the cells expressing Reelin, Vimentin or GABA increased, and neurogenesis of calbindin cells changed towards the mammalian inside-out pattern in the dorsal pallium and mesopallium. In addition, a behavior test for posthatched chicks indicated that the passive avoidance ratio increased significantly. The study suggests that the acquisition of an additional fragment in mammalian Emx2 is associated with the organization and evolution of the mammalian pallium.
... The latter pathway is also present in birds. Connectivity and chemoarchitecture of these regions, homologous with mammalian progenitor zones, basal ganglia, and extended amygdala (Bupesh et al., 2011;Kuenzel et al., 2011;Vicario et al., 2014Vicario et al., , 2015Vicario et al., , 2017Martínez-Cerdeño et al., 2018), have been extensively investigated also by our research group (Bálint et al., 2004;Montagnese et al., 2004Montagnese et al., , 2008Hanics et al., 2012Hanics et al., , 2017 found to impair socially motivated vocalization of domestic chicks (Zachar et al., 2017). A great deal of studies has tackled the pre-and postnatal development of dopaminergic neurons and pathways (see below), the developmental dynamics of the ventrotegmentalaccumbens (mesolimbic) pathway, including the chemorepellent or chemoattractant factors specifically involved in the segregation of distinct dopaminergic pathways/connections (see Bissonette and Roesch, 2016). ...
Article
Full-text available
Gestational exposure of mice to valproic acid (VPA) is one currently used experimental model for the investigation of typical failure symptoms associated with autism spectrum disorder (ASD). In the present study we hypothesized that the reduction of dopaminergic source neurons of the VTA, followed by perturbed growth of the mesotelencephalic dopamine pathway (MT), should also modify pattern formation in the dopaminoceptive target regions (particularly its mesoaccumbens/mesolimbic portion). Here, we investigated VPA-evoked cellular morphological (apoptosis-frequency detected by Caspase-3, abundance of Ca-binding proteins, CaBP), as well as synaptic proteomic (western blotting) changes, in selected dopaminoceptive subpallial, as compared to pallial, regions of mice, born to mothers treated with 500 mg/kg VPA on day 13.5 of pregnancy. We observed a surge of apoptosis on VPA treatment in nearly all investigated subpallial and pallial regions; with a non-significant trend of similar increase the nucleus accumbens (NAc) at P7, the age at which the MT pathway reduction has been reported (also supplemented by current findings). Of the CaBPs, calretinin (CR) expression was decreased in pallial regions, most prominently in retrosplenial cortex, but not in the subpallium of P7 mice. Calbindin-D 28K (CB) was selectively reduced in the caudate-putamen (CPu) of VPA exposed animals at P7 but no longer at P60, pointing to a potency of repairment. The VPA-associated overall increase in apoptosis at P7 did not correlate with the abundance and distribution of CaBPs, except in CPu, in which the marked drop of CB was negatively correlated with increased apoptosis. Abundance of parvalbumin (PV) at P60 showed no significant response to VPA treatment in any of the observed regions we did not find colocalization of apoptotic (Casp3+) cells with CaBP-immunoreactive neurons. The proteomic findings suggest reduction of tyrosine hydroxylase in the crude synaptosome fraction of NAc, but not in the CPu, without simultaneous decrease of the synaptic protein, synaptophysin, indicating selective impairment of dopaminergic synapses. The morpho-functional changes found in forebrain regions of VPA-exposed mice may signify dendritic and synaptic reorganization in dopaminergic target regions, with potential translational value to similar impairments in the pathogenesis of human ASD.
... These cells were previously observed in chicken and quail using immunohistochemistry, but were proposed to belong to the extended amygdala (Richard et al., 2004). Like in mammals (Alheid & Heimer, 1988), corticopetal subpallial cells in chicken partially overlap the extended amygdala, the pallidal parts of the basal ganglia, and the lateral and medial forebrain bundles, but they belong to a different functional system characterized by projections to the pallium (Leutgeb et al., 1996;discussed by Abellán & Medina, 2009;Kuenzel et al., 2011). Corticopetal cells include cholinergic, GABAergic and glutamatergic neuronal subpopulations Medina & Reiner, 1994), and it would be interesting to investigate coexpression of these neurotransmitters with CRF. ...
Article
Full-text available
Understanding the neural mechanisms that regulate the stress response is critical to know how animals adapt to a changing world and is one of the key factors to be considered for improving animal welfare. Corticotropin-releasing factor (CRF) is crucial for regulating physiological and endocrine responses, triggering the activation of the sympathetic nervous system and the hypothalamo-pituitary-adrenal axis (HPA) during stress. In mammals, several telencephalic areas, such as the amygdala and the hippocampus, regulate the autonomic system and the HPA responses. These centers include subpopulations of CRF containing neurons that, by way of CRF receptors, play modulatory roles in the emotional and cognitive aspects of stress. CRF binding protein also plays a role, buffering extracellular CRF and regulating its availability. CRF role in activation of the HPA is evolutionary conserved in vertebrates, highlighting the relevance of this system to help animals cope with adversity. However, knowledge on CRF systems in the avian telencephalon is very limited, and no information exists on detailed expression of CRF receptors and binding protein. Knowing that the stress response changes with age, with important variations during the first week posthatching, the aim of this study was to analyze mRNA expression of CRF, CRF receptors 1 and 2, and CRF binding protein in chicken telencephalon throughout embryonic and early posthatching development, using in situ hybridization. Our results demonstrate an early expression of CRF and its receptors in pallial areas regulating sensory processing, sensorimotor integration and cognition, and a late expression in subpallial areas regulating the stress response. However, CRF buffering system develops earlier in the subpallium than in the pallium. These results help to understand the mechanisms underlying the negative effects of noise and light during prehatching stages in chicken, and suggest that stress regulation becomes more sophisticated with age.
... Rostral is approximately right, dorsal up, and bars = 1 mm. Abbreviations: HVC (proper name), lMAN (lateral magnocellular nucleus of the anterior nidopallium), RA (robust nucleus of the arcopallium), Area X (Area X of striatum, note that songbird striatum also contains pallidal projection neurons and interneurons that anatomically distinguish it from mammalian striatum that separates these features 20 ). Brain regions from both vocal-motor (HVC and RA) and -learning circuits (Area X) were micropunch dissected, total RNA extracted, cDNA synthesized, and PCR amplified. ...
Article
Full-text available
The non-euphorigenic phytocannabinoid cannabidiol (CBD) has been used successfully to treat childhood-onset epilepsies. These conditions are associated with developmental delays that often include vocal learning. Zebra finch song, like language, is a complex behavior learned during a sensitive period of development. Song quality is maintained through continuous sensorimotor refinement involving circuits that control learning and production. Within the vocal motor circuit, HVC is a cortical-like region that when partially lesioned temporarily disrupts song structure. We previously found CBD (10 mg/kg/day) improves post-lesion vocal recovery. The present studies were done to begin to understand mechanisms possibly responsible for CBD vocal protection. We found CBD markedly reduced expression of inflammatory mediators and oxidative stress markers. These effects were associated with regionally-reduced expression of the microglial marker TMEM119. As microglia are key regulators of synaptic reorganization, we measured synapse densities, finding significant lesion-induced circuit-wide decreases that were largely reversed by CBD. Synaptic protection was accompanied by NRF2 activation and BDNF/ARC/ARG3.1/MSK1 expression implicating mechanisms important to song circuit node mitigation of oxidative stress and promotion of synaptic homeostasis. Our findings demonstrate that CBD promotes an array of neuroprotective processes consistent with modulation of multiple cell signaling systems, and suggest these mechanisms are important to post-lesion recovery of a complex learned behavior.
... These cells were previously observed in chicken and quail using immunohistochemistry, but were proposed to belong to the extended amygdala (Richard et al., 2004). Like in mammals (Alheid and Heimer, 1988), corticopetal subpallial cells in chicken partially overlap the extended amygdala, the pallidal parts of the basal ganglia, and the lateral and medial forebrain bundles, but they belong to a different functional system characterized by projections to the pallium (Leutgeb et al., 1996;discussed by Abellán and Medina, 2009;Kuenzel et al., 2011). Corticopetal cells include cholinergic, GABAergic and glutamatergic neuronal subpopulations (Medina and Reiner, 1994;, and it would be interesting to investigate coexpression of these neurotransmitters with CRF. ...
Preprint
Full-text available
Understanding the neural mechanisms that regulate the stress response is critical to know how animals adapt to a changing world and is one of the key factors to be considered for improving animal welfare. Corticotropin releasing factor (CRF) is crucial for regulating physiological and endocrine responses, triggering the activation of the sympathetic nervous system and the hypothalamo - pituitary - adrenal axis (HPA) during stress. In mammals, several telencephalic areas, such as the amygdala and the hippocampus, regulate the autonomic system and the HPA responses. These centers include subpopulations of CRF containing neurons that, by way of CRF receptors, play modulatory roles in the emotional and cognitive aspects of stress. CRF binding protein also plays a role, buffering extracellular CRF and regulating its availability. CRF role in activation of the HPA is evolutionary conserved in vertebrates, highlighting the relevance of this system to help animals cope with adversity. However, knowledge on CRF systems in the avian telencephalon is very limited, and no information exists on detailed expression of CRF receptors and binding protein. Knowing that the stress response changes with age, with important variations during the first week posthatching, the aim of this study was to analyze mRNA expression of CRF, CRF receptors 1 and 2, and CRF binding protein in chicken telencephalon throughout embryonic and early posthatching development, using in situ hybridization. Our results demonstrate an early expression of CRF and its receptors in pallial areas regulating sensory processing, sensorimotor integration and cognition, and a late expression in subpallial areas regulating the stress response. However, CRF buffering system develops earlier in the subpallium than in the pallium. These results help to understand the mechanisms underlying the negative effects of noise and light during prehatching stages in chicken, and suggest that stress regulation becomes more sophisticated with age.
Article
A core component of the avian pallial cognitive network is the multimodal nidopallium caudolaterale (NCL) that is considered to be analogous to the mammalian prefrontal cortex (PFC). The NCL plays a key role in a multitude of executive tasks such as working memory, decision‐making during navigation, and extinction learning in complex learning environments. Like the PFC, the NCL is positioned at the transition from ascending sensory to descending motor systems. For the latter, it sends descending premotor projections to the intermediate arcopallium (AI) and the medial striatum (MSt). To gain detailed insight into the organization of these projections, we conducted several retrograde and anterograde tracing experiments. First, we tested whether NCL neurons projecting to AI (NCL arco neurons) and MSt (NCL MSt neurons) are constituted by a single neuronal population with bifurcating neurons, or whether they form two distinct populations. Here, we found two distinct projection patterns to both target areas that were associated with different morphologies. Second, we revealed a weak topographic projection toward the medial and lateral striatum and a strong topographic projection toward AI with clearly distinguishable sensory termination fields. Third, we investigated the relationship between the descending NCL pathways to the arcopallium with those from the hyperpallium apicale, which harbors a second major descending pathway of the avian pallium. We embed our findings within a system of parallel pallio‐motor loops that carry information from separate sensory modalities to different subpallial systems. Our results also provide insights into the evolution of the avian motor system from which, possibly, the song system has emerged.
Article
Full-text available
The globus pallidus (GP) of primates is divided conventionally into distinct internal and external parts. The literature repeats since 1930 the opinion that the homolog of the primate internal pallidum in rodents is the hypothalamic entopeduncular nucleus (embedded within fiber tracts of the cerebral peduncle). To test this idea, we explored its historic fundaments, checked the development and genoarchitecture of mouse entopeduncular and pallidal neurons, and examined relevant comparative connectivity data. We found that the extratelencephalic mouse entopeduncular structure consists of four different components arrayed along a dorsoventral sequence in the alar hypothalamus. The ventral entopeduncular nucleus (EPV), with GABAergic neurons expressing Dlx5&6 and Nkx2-1, lies within the hypothalamic peduncular subparaventricular area. Three other formations-the dorsal entopeduncular nucleus (EPD), the prereticular entopeduncular nucleus (EPPRt ), and the preeminential entopeduncular nucleus (EPPEm )-lie within the overlying paraventricular area, under the subpallium. EPD contains glutamatergic neurons expressing Tbr1, Otp, and Pax6. The EPPRt has GABAergic cells expressing Isl1 and Meis2, whereas the EPPEm population expresses Foxg1 and may be glutamatergic. Genoarchitectonic observations on relevant areas of the mouse pallidal/diagonal subpallium suggest that the GP of rodents is constituted as in primates by two adjacent but molecularly and hodologically differentiable telencephalic portions (both expressing Foxg1). These and other reported data oppose the notion that the rodent extratelencephalic entopeduncular nucleus is homologous to the primate internal pallidum. We suggest instead that all mammals, including rodents, have dual subpallial GP components, whereas primates probably also have a comparable set of hypothalamic entopeduncular nuclei. Remarkably, there is close similarity in some gene expression properties of the telencephalic internal GP and the hypothalamic EPV. This apparently underlies their notable functional analogy, sharing GABAergic neurons and thalamopetal connectivity.
Article
Using a songbird, zebra finches, as a developmental drug abuse model we found previously that cannabinoid agonists administered during the sensorimotor period of vocal learning (50-75 days of age) persistently alter song patterns and cocaine responsiveness in adulthood. However, these effects were not produced in adults exposed to similar treatment regimens. Currently, we have used the MAGL inhibitor, JZL184, to test whether enhanced endocannabinoid signaling may similarly alter cocaine responsiveness. We found that, as expected and consistent with prior results, repeated developmental (but not adult) treatments with Δ9-tetrahydrocannabinol (THC, 3 mg/kg QD IM) resulted in increased time spent in cocaine-paired chambers. Unexpectedly and in contrast, repeated developmental JZL184 (4 mg/kg QD IM) treatments decreased time spent in cocaine-conditioned chambers. That is, young finches repeatedly treated with JZL184 avoided cocaine-paired chambers later in adulthood, while similar development treatments with THC had the opposite effect. To begin to identify brain regions that may underly this differential responsiveness we used c-Fos expression as a marker of neuronal activity. Differences in c-Fos expression patterns following placement of cocaine-conditioned finches into vehicle- vs. cocaine-paired chambers suggest distinct involvement of circuits through striatal and amygdaloid regions in respective effects of THC and JZL184. Results demonstrate that, like exogenous cannabinoid exposure, inhibition of MAGL activity during late post-natal development persistently alters behavior in adulthood. Contrasting effects of THC vs. MAGL inhibition with JZL184 suggests the latter alters development of brain regions to favor promotion of aversive rather than appetitive cocaine responsiveness later in adulthood.
Article
Full-text available
The extrathalamic relay from the brainstem reticular formation to the cerebral cortex in the basal forebrain has been thought to be constituted predominantly, if not exclusively, by cholinergic neurons. In contrast, the septohippocampal projection has been shown to contain an important contingent of g-aminobutyric acid (GABA)ergic neurons. In the present study, we investigated whether GABAergic neurons also contribute to the projection from the basal forebrain to neocortical regions, including the mesocortex (limbic) and the isocortex in the rat. For this purpose, retrograde transport of cholera toxin (CT) was examined from the medial prefrontal cortex for the mesocortex and from the parietal cortex for the isocortex and was combined with dual-immunohistochemical staining for either choline acetyltransferase (ChAT) or glutamic acid decarboxylase (GAD) in adjacent series of sections. Retrogradely labelled GAD 1 neurons were codistributed with retrogradely labelled ChAT 1 neurons through the basal forebrain from both the prefrontal and the parietal cortex, suggesting parallel, widespread cortical projections. The GAD 1 cortically projecting cells were similar in size to the ChAT 1 cells, thereby indicating that they comprise a contingent of the magnocellular basal cell complex. The proportions of retrogradely labelled neurons that were GAD 1 (approximately one-third) were equal to or greater than those that were ChAT 1 from both the prefrontal cortex and the parietal cortex. In addition, the total of GAD 1 and ChAT 1 neurons did not account for the total number of cortically projecting cells, indicating that another equivalent proportion of chemically unidentified noncholinergic neurons also contributes to the basalocortical projection. Accordingly, as in the allocortex, GABAergic, cholinergic, and other unidentified noncholinergic neurons may have the capacity to modulate activity in the mesocortex (limbic) and the isocortex through parallel, widespread projections.
Article
The basal ganglia in modern mammals, birds and reptiles (i.e. modern amniotes) are very similar in connections and neurotransmitters, suggesting that the evolution of the basal ganglia in amniotes has been very conservative. For example, the basal ganglia in all amniotes possess a dorsal striatum containing two main populations of projection neurons, substance P-containing (SP+) and enkephalin-containing (ENK+) neurons, which have major projections to the dorsal pallidum and the tegmentum (ventral tegmental area and substantia nigra, or VTA/SN). The VTA/SN, in turn, has a major dopaminergic (DA+) projection to the striatum in all amniotes. In this paper, we review these data on the basal ganglia in amniotes and note points of similarity and difference in the functional circuitry of the basal ganglia among amniotes. In addition, we review recent findings on the neurotransmitter organization and connectivity of the basal ganglia in amphibians and fishes, with the goal of assessing whether a basal ganglia showing the same basic features as in amniotes is observed in anamniotes. Published data indicate that in at least two groups of fishes (cartilaginous fishes and lungfishes) and apparently in amphibians, the basal ganglia is present and consists of a distinct striatum and pallidum. The striatum of amphibians, cartilaginous fishes, and lungfishes contain SP+ and ENK+ neurons that seem to project to the pallidum as well as to a brainstem cell group that appears comparable to the VTA/SN of amniotes. Data for ray-finned fishes also suggest the presence of a striatum containing SP+ and ENK+ neurons that projects to VTA/SN-like brainstem cell group. In the basal ganglia of ray-finned fishes, however, a distinct pallidum had not been identified. Finally, the brainstem cell group receiving striatal input in all anamniotes contains DA+ neurons that seem to project to the striatum. The present analysis suggests that a rudimentary basal ganglia was already present in the brain of the ancestral jawed vertebrates. This rudimentary basal ganglia likely consisted of a striatum and a pallidum, and the striatum probably already possessed the same basic connections and some of the same basic cell types as the basal ganglia of modern jawed vertebrates.
Article
Most zebrafish neurodevelopmental studies have focused on the embryo, which is characterized by primary neurogenesis of mostly transient neurons. Secondary neurogenesis becomes dominant in the hatching larva, when major brain parts are established and begin to differentiate. This developmental period allows for a comparative analysis of zebrafish brain organization with amniotes at equivalent stages of neurogenesis. Within a particular time window, the early forebrains of mice (Embyronic stage [E] 12.5/13.5 days [d]) and zebrafish (3 d) reveal highly comparable expression patterns of genes involved in neurogenesis, for example proneural and other transcription factors (Neurogenin1, NeuroD, Mash1/Zash1a and Pax6). Further topological correspondences are seen in the expression of LIM and homeobox genes, such as Lhx6/7, Tbr2 and Dlx2a. When this analysis is extended to gamma-aminobutyric acid/glutamic acid decarboxylase (GABA/GAD) cell patterns during this critical time window, an astonishing degree of similarity between the two species is again seen, for example regarding the presence of GABA/GAD cells in the subpallium, with the pallium only starting to be invaded by such cells from the subpallium. Furthermore, the expression of proneural and other genes correlates with GABA cell patterns (e.g. Mash1/Zash1a gene expression in GABA-positive and Neurogenin1/NeuroD in GABA-negative telencephalic regions) in mice and zebrafish. Data from additional vertebrates, such as Xenopus, are also highly consistent with this analysis. Therefore, the vertebrate forebrain appears to undergo a phylotypic stage of secondary neurogenesis, characterized by regionally separated GABAergic (inhibitory) versus glutamatergic (excitatory) cell production sites, which are obscured later in development by tangential migration. This period is highly advantageous for molecular neuroanatomical cross-species comparisons.