ArticlePDF Available

The receptor architecture of the pigeons' nidopallium caudolaterale: An avian analogue to the mammalian prefrontal cortex

Authors:

Abstract and Figures

The avian nidopallium caudolaterale is a multimodal area in the caudal telencephalon that is apparently not homologous to the mammalian prefrontal cortex but serves comparable functions. Here we analyzed binding-site densities of glutamatergic AMPA, NMDA and kainate receptors, GABAergic GABA(A), muscarinic M(1), M(2) and nicotinic (nACh) receptors, noradrenergic α(1) and α(2), serotonergic 5-HT(1A) and dopaminergic D(1)-like receptors using quantitative in vitro receptor autoradiography. We compared the receptor architecture of the pigeons' nidopallial structures, in particular the NCL, with cortical areas Fr2 and Cg1 in rats and prefrontal area BA10 in humans. Our results confirmed that the relative ratios of multiple receptor densities across different nidopallial structures (their "receptor fingerprints") were very similar in shape; however, the absolute binding densities (the "size" of the fingerprints) differed significantly. This finding enables a delineation of the avian NCL from surrounding structures and a further parcellation into a medial and a lateral part as revealed by differences in densities of nACh, M(2), kainate, and 5-HT(1A) receptors. Comparisons of the NCL with the rat and human frontal structures showed differences in the receptor distribution, particularly of the glutamate receptors, but also revealed highly conserved features like the identical densities of GABA(A), M(2), nACh and D(1)-like receptors. Assuming a convergent evolution of avian and mammalian prefrontal areas, our results support the hypothesis that specific neurochemical traits provide the molecular background for higher order processes such as executive functions. The differences in glutamate receptor distributions may reflect species-specific adaptations.
Content may be subject to copyright.
ORIGINAL ARTICLE
The receptor architecture of the pigeons’ nidopallium
caudolaterale: an avian analogue to the mammalian
prefrontal cortex
Christina Herold Nicola Palomero-Gallagher
Burkhard Hellmann Sven Kro
¨ner Carsten Theiss
Onur Gu
¨ntu
¨rku
¨nKarl Zilles
Received: 26 October 2010 / Accepted: 12 January 2011 / Published online: 4 February 2011
ÓSpringer-Verlag 2011
Abstract The avian nidopallium caudolaterale is a mul-
timodal area in the caudal telencephalon that is apparently
not homologous to the mammalian prefrontal cortex but
serves comparable functions. Here we analyzed binding-
site densities of glutamatergic AMPA, NMDA and kainate
receptors, GABAergic GABA
A
, muscarinic M
1
,M
2
and
nicotinic (nACh) receptors, noradrenergic a
1
and a
2
, sero-
tonergic 5-HT
1A
and dopaminergic D
1
-like receptors using
quantitative in vitro receptor autoradiography. We com-
pared the receptor architecture of the pigeons’ nidopallial
structures, in particular the NCL, with cortical areas Fr2
and Cg1 in rats and prefrontal area BA10 in humans. Our
results confirmed that the relative ratios of multiple
receptor densities across different nidopallial structures
(their ‘receptor fingerprints’’) were very similar in shape;
however, the absolute binding densities (the ‘size’ of the
fingerprints) differed significantly. This finding enables a
delineation of the avian NCL from surrounding structures
and a further parcellation into a medial and a lateral part as
revealed by differences in densities of nACh, M
2
, kainate,
and 5-HT
1A
receptors. Comparisons of the NCL with the
rat and human frontal structures showed differences in the
receptor distribution, particularly of the glutamate recep-
tors, but also revealed highly conserved features like the
identical densities of GABA
A
,M
2
, nACh and D
1
-like
receptors. Assuming a convergent evolution of avian and
mammalian prefrontal areas, our results support the
hypothesis that specific neurochemical traits provide the
molecular background for higher order processes such as
executive functions. The differences in glutamate receptor
distributions may reflect species-specific adaptations.
Keywords Receptor autoradiography Prefrontal cortex
Nidopallium caudolaterale Rat Human Fr2 Cg1
BA10 Dopamine Glutamate GABA
Abbreviations
ACh Acetylcholine
AMPA a-Amino-3-hydroxy-5-methyl-4-isoxalone
propionic acid
Cg1 Cingulate cortex 1
CDL Dorsolateral corticoid area
EPSCs Excitatory postsynaptic currents
FR2 Frontal area 2
GABA c-Aminobutyric acid
GLI Gray level index
gluR1 Glutamate receptor subunit 1
HA Hyperpallium apicale
HVC Higher vocal center
IMM Intermediate and medial mesopallium ventrale
C. Herold (&)K. Zilles
C. and O. Vogt-Institute of Brain Research,
University of Du
¨sseldorf, 40225 Du
¨sseldorf, Germany
e-mail: christina.herold@uni-duesseldorf.de
N. Palomero-Gallagher K. Zilles
Institute of Neuroscience and Medicine INM-2,
Research Center Ju
¨lich, 52425 Ju
¨lich, Germany
B. Hellmann O. Gu
¨ntu
¨rku
¨n
Department of Biopsychology,
Institute of Cognitive Neuroscience,
Faculty of Psychology, Ruhr-University Bochum,
44780 Bochum, Germany
C. Theiss
Institute of Anatomy and Molecular Embryology,
Faculty of Medicine, Ruhr-University Bochum,
44780 Bochum, Germany
S. Kro
¨ner
School of Behavioral and Brain Sciences, The University
of Texas at Dallas, Richardson, TX 75080, USA
123
Brain Struct Funct (2011) 216:239–254
DOI 10.1007/s00429-011-0301-5
MNH Mediorostral nidopallium/hyperpallium
nACh Nicotinic acetylcholine
NCC Nidopallium caudocentrale
NCL Nidopallium caudolaterale
NCLl Nidopallium caudolaterale pars lateralis
NCLm Nidopallium caudolaterale pars medialis
NCM Nidopallium caudomediale
NFT Nidopallium fronto-trigeminale
NIM Nidopallium intermedium medialis
NMDA N-methyl-D-aspartate
PFC Prefrontal cortex
Introduction
The increasingly refined parcellation of the mammalian
cerebral cortex with anatomical methods enables various
analyses of its functional segregation (Uylings et al. 2000;
Amunts et al. 2004; Eickhoff et al. 2006, Amunts et al.
2007; Naito et al. 2008; Palomero-Gallagher et al. 2009;
Zilles and Amunts 2010). Similarly, in the last decades the
avian forebrain has been subdivided by various means.
These efforts have fostered a new understanding of the
avian telencephalic organization and the assumed homol-
ogies between avian and mammalian brain components
(Reiner et al. 2004). This new view, which is rooted in a
series of seminal studies over the last 40 years (Karten
1969), assumes that mammalian and avian pallia are
homologous in terms of shared pallial identity that derive
from common ancestry (Jarvis et al. 2005). This assump-
tion, however, does not imply that cortical or subcortical
pallial areas have to be one-to-one homologous to pallial
components in birds. Thus, pallial structures of birds and
mammals might be similar in terms of anatomical, physi-
ological and cognitive characteristics, but may still repre-
sent the result of convergent evolution.
The avian nidopallium caudolaterale (NCL) is such a
case. Numerous studies show that the mammalian pre-
frontal cortex (PFC) and the avian NCL share several
anatomical (Kro
¨ner and Gu
¨ntu
¨rku
¨n1999), neurochemical
(Bast et al. 2002; Karakuyu et al. 2007), electrophysio-
logical (Diekamp et al. 2002a; Kalenscher et al. 2005; Rose
and Colombo 2005), and functional (Gu
¨ntu
¨rku
¨n1997;
Diekamp et al. 2002b; Kalenscher et al. 2003; Lissek and
Gu
¨ntu
¨rku
¨n2005) characteristics; however, several genetic
(Puelles et al. 2000) and topological arguments (Medina
and Reiner 2000) make a homology between the PFC and
the NCL unlikely. Therefore, the similarities of these two
structures likely do not result from common ancestry but
represent the outcome of an evolutionary convergence.
Thus, a common selection pressure for an ‘executive’
behavioral repertoire possibly facilitates emergence of
non-homologous forebrain areas of mammals and birds that
share typical ‘prefrontal’ characteristics (Gu
¨ntu
¨rku
¨n2005a;
Kirsch et al. 2008).
The NCL displays a homogeneous cytoarchitecture and
does not differ considerably from neighboring portions of
the nidopallium either. The NCL was first defined by its
dopaminergic innervation and high tyrosine hydroxylase
density (Divac et al. 1985; Waldmann and Gu
¨ntu
¨rku
¨n
1993). To date, the outer borders and the internal structure
of the NCL have been analyzed with immunocytochemical
(Wynne and Gu
¨ntu
¨rku
¨n1995; Bock et al. 1997; Schnabel
et al. 1997; Durstewitz et al. 1998; Riters et al. 1999) and
ultrastructural methods (Metzger et al. 2002) as well as in
several tracing studies (Leutgeb et al. 1996; Metzger et al.
1998; Kro
¨ner and Gu
¨ntu
¨rku
¨n1999). Receptor autoradiog-
raphy is an additional powerful tool to define areal borders
and to derive region-specific receptor-density combinations
that define areas like ‘fingerprints’ (Zilles et al. 2002b).
Therefore, the first aim of this study was to map the
chemoarchitecture of the NCL. This approach is important
to define the areal borders between the nidopallium cau-
docentrale (NCC) and the NCL, since different studies
using tracing techniques or immunocytochemistry showed
discrepant delineations (Wynne and Gu
¨ntu
¨rku
¨n1995;
Waldmann and Gu
¨ntu
¨rku
¨n1993; Kro
¨ner and Gu
¨ntu
¨rku
¨n
1999; Atoji and Wild 2009). Although, the border to the
laterally and supraventricular located dorsolateral corticoid
area (CDL) and the NCL is easier to define, it was addi-
tionally included in the analysis here.
The second aim of this study was to investigate possible
subdivisions within the NCL because numerous studies in
mammals have implicated subdivisions of the PFC in the
processing of different stimulus domains (Levy and
Goldman-Rakic 1999). Similarly, there is also evidence for
a parcellation of the NCL based on functional, neuro-
chemical and hodological data (Leutgeb et al. 1996; Braun
et al. 1999; Kro
¨ner and Gu
¨ntu
¨rku
¨n1999; Riters et al. 1999;
Diekamp et al. 2002b). Riters et al. (1999) proposed a
dorsoventral distinction of the NCL based on the distri-
bution of tyrosine hydroxylase, choline acetyltransferase
and substance P labeled fibers and terminals. Accordingly,
lesions of the dorsal NCL result in delay-specific working
memory deficits (Diekamp et al. 2002b). The high density
of tyrosine hydroxylase positive fibers in the dorsal NCL
might be related to the important role of dopamine in
working memory functions as shown in primates (Goldman-
Rakic 1999) and birds (Karakuyu et al. 2007). Based on
connectivity data, however, Kro
¨ner and Gu
¨ntu
¨rku
¨n(1999)
assumed a frontocaudal distinction with the caudal portion
being tightly embedded within the limbic system.
The third aim was to compare the receptor fingerprints
of mammalian frontal and prefrontal areas with those of
the avian NCL in order to examine whether a common
240 Brain Struct Funct (2011) 216:239–254
123
functional repertoire is reflected by a similar pattern of
receptor architecture. For this purpose we studied the
receptor fingerprints of the medial and the lateral portions
of Brodmann’s human prefrontal cortex area BA10
(BA10m and BA10l, respectively) and the rat frontal area 2
(Fr2) as well as the rat prefrontal cingulate area 1 (Cg1)
(Brodmann 1909; Uylings et al. 2003). Owing to their
connectivity patterns with other neocortical areas, the
thalamus, the basal ganglia, and the amygdala, both Fr2
and Cg1 were structurally and functionally compared with
the dorsolateral prefrontal cortex in primates (Uylings et al.
2003; Van de Werd et al. 2010). However, it has to be
noted they there are still discrepancies in the delineations
of rat prefrontal and motor cortical structures; furthermore,
Fr2 is classified as the rodent’s motor cortex (Van Eden
et al. 1992; Zilles 1985).
Taken together, our receptor autoradiographic study was
aimed to constitute an independent approach to these open
questions.
Materials and methods
We examined a total of six pigeons (Columba livia)of
unknown sex and eight male rats (Long-Evans). Animals
were decapitated and the brains removed from the skull,
frozen immediately in isopentane at -40°C and stored at
-70°C. Serial coronal 10 lm sections were cut with a
cryostat microtome (2800 Frigocut E, Reichert-Jung).
Sections were thaw-mounted on gelatinized slides and
freeze-dried.
Post-mortem human brain tissue was studied from 2
control subjects (age 72 male and age 77 female, post-
mortem time 8 and 18 h) without a record of neurological
or psychiatric disorders and was obtained from the body
donor program of the Department of Anatomy, University
of Du
¨sseldorf, Germany. Causes of death were a heart
attack and carcinoma. Serial coronal cryosections (20 lm)
comprising the whole cross-section of unfixed brain blocks
were prepared at -20°C using a large-scale cryostat
microtome. Sections were thaw-mounted on gelatinized
slides, freeze-dried and stained with a modified cell
body staining for cytoarchitectonic analysis (Merker 1983;
Palomero-Gallagher et al. 2008) or processed for receptor
autoradiography.
Receptor autoradiography
Details of the autoradiographic labeling procedure have
been published elsewhere (Zilles et al. 2002b; Palomero-
Gallagher et al. 2009). Binding protocols are summarized
in Table 1. Three steps were performed in the following
sequence: (1) A preincubation step removed endogenous
ligand from the tissue. (2) During the main incubation step,
binding sites were labeled with triated ligand (total bind-
ing). Coincubation of the triated ligand and a 1,000 to
10,000-fold excess of an appropriate non-labeled ligand
(displacer) determined non-specific and thus non-dis-
placeable binding. Specific binding is the difference
between total and non-specific binding. (3) A final rinsing
step eliminated unbound radioactive ligand from the
sections.
The following binding sites were labeled according
to standardized protocols: a-amino-3-hydroxy-5-methyl-
4-isoxalone propionic acid (AMPA) with [
3
H] AMPA,
kainate with [
3
H]kainate, N-methyl-D-aspartate (NMDA)
with [
3
H]MK-801, c-aminobutyric acid A (GABA
A
) recep-
tor with [
3
H]muscimol, muscarinic cholinergic M
1
receptor
with [
3
H]pirenzepine, muscarinic cholinergic M
2
receptor
with [
3
H]oxotremorine-M, nicotinic cholinergic (nACh)
receptor with [
3
H]cytosine (pigeon) or [
3
H]epibatidine
(rat and human), noradrenergic a
1
adrenoreceptor with
[
3
H]prazosin, noradrenergic a
2
adrenoreceptor with [
3
H]
RX-821002, serotonergic 5-HT
1A
receptor with [
3
H]8-OH-
DPAT, and dopaminergic D
1
-like receptors with [
3
H]SCH
23390. Sections were air-dried overnight and subsequently
coexposed for 4–5 weeks against a tritium-sensitive film
(Hyperfilm, Amersham, Braunschweig, Germany) with
plastic [
3
H]-standards (Microscales, Amersham) of known
concentrations of radioactivity.
Image analysis
The resulting autoradiographs were subsequently processed
using densitometry with a video-based image analyzing
technique (Zilles et al. 2002b; Schleicher et al. 2005).
Autoradiographs were digitized by means of a KS-400
image analyzing system (Kontron, Germany) connected to
a CCD camera (Sony, Tokyo) equipped with a S-Ortho-
planar 60-mm macro lens (Zeiss, Germany). The images
were stored as binary files with a resolution of 512 9512
pixels and 8-bit gray value. The gray value images of the
coexposed microscales were used to compute a calibration
curve by non-linear, least-squares fitting, which defined the
relationship between gray values in the autoradiographs
and concentrations of radioactivity. This enabled the pixel-
wise conversion of the gray values of an autoradiograph
into the corresponding concentrations of radioactivity.
These concentrations of binding sites occupied by the
ligand under incubation conditions are transformed into
fmol binding site/mg protein at saturation conditions by
means of the equation: (K
D
?L)/A
S
9L, where K
D
is the
equilibrium dissociation constant of ligand-binding kinet-
ics, Lis the incubation concentration of ligand, and A
S
the
specific activity of the ligand. The mean of the gray values
contained in a specific region over a series of 4–5 sections
Brain Struct Funct (2011) 216:239–254 241
123
Table 1 Incubation conditions for receptor autoradiography
Receptor [
3
H] ligand
(incubation
concentration)
Displacer
(incubation
concentration)
Incubation buffer Preincubation
step
Main incubation step Rinsing step
AMPA [
3
H]AMPA (10 nM) Quisqualate
(10 lM)
50 mM Tris–acetate (pH 7.2) 3 910 min at
4°Cin
incubation
buffer
45 min at 4°C in incubation
buffer ?100 mM KSCN
494 s at 4°C in incubation
buffer ?292 s at 4°Cin
acetone/glutaraldehyde
Kainate [
3
H]kainate (8 nM) Kainate
(100 lM)
50 mM Tris–citrate (pH 7.1) 3 910 min at
4°Cin
incubation
buffer
45 min at 4°C in incubation
buffer ?10 mM Ca-acetate
494 s at 4°C in incubation
buffer ?292 s at 4°Cin
acetone/glutaraldehyde
NMDA [
3
H]MK-801
(5 nM)
MK-801
(100 lM)
50 mM Tris–HCl (pH 7.2) 15 min at 25°C
in incubation
buffer
60 min at 25°C in incubation
buffer ?30 lM
glycine ?50 lM spermidine
295 min at 4°C in incubation
buffer
Muscarinergic
cholinergic M
1
[
3
H]pirenzipine
(1nM)
Pirenzipine
(10 lM)
Modified Krebs–Ringer buffer (pH 7.4) 20 min at 25°C
in incubation
buffer
60 min at 25°C in incubation
buffer
295 min at 4°C in incubation
buffer
Muscarinergic
cholinergic M
2
[
3
H]oxotremonine-
M (0.8 nM)
Carbachol
(1 lM)
20 mM Hepes–Tris (pH 7.5)
?10 mM MgCl
2
20 min at 25°C
in incubation
buffer
60 min at 25°C in incubation
buffer
292 min at 4°C in incubation
buffer
Nicotinic
cholinergic
[
3
H]cytisine [1 nM] Nicotine
(10 lM)
50 mM Tris–HCl (pH7.4) ?120 mM
NaCl ?5 mM KCl ?1mM
MgCl
2
?2.5 mM CaCl
2
15 min at 22°C
in incubation
buffer
90 min at 4°C in incubation
buffer
292 min at 4°C in incubation
buffer
[
3
H]epibatidine
[0.5 nM]
Nicotine
(10 lM)
15 mM Hepes–Tris (pH 7.5) ?10 mM
NaCl ?5.4 mM KCl ?0.8 mM
MgCl
2
?1.8 mM CaCl
2
20 min at 22°C
in incubation
buffer
90 min at 4°C in incubation
buffer
195 min at 4°C in incubation
buffer
29up & down in distilled H
2
O
a
1
Adrenoreceptor
[
3
H]prazosin
(0.2 nM)
Phentolamine
(10 lM)
50 mM Tris–HCl (pH 7.4) 30 min at 37°C
in incubation
buffer
45 min at 30°C in incubation
buffer
295 min at 4°C in incubation
buffer
a
2
Adrenoreceptor
[
3
H]UK-14304
(1.4 nM)
Noradrenalin
(100 lM)
50 mM Tris–HCl (pH 7.7)
?100 lM MnCl
2
15 min at 22°C
in incubation
buffer
90 min at 22°C in incubation
buffer
5 min at 4°C in incubation buffer
GABA
A
[
3
H]muscimol
(6 nM:pigeon and
3 nM:human)
GABA
(10 lM)
50 mM Tris–citrate (pH 7.0) 3 95 min at
4°Cin
incubation
buffer
40 min at 4°C in incubation
buffer
393 s at 4°C in incubation buffer
Serotoninergic
5-HT
1A
[
3
H]8-OH-DPAT
(1 nM)
Serotonin
(10 lM)
170 mM Tris–HCl (pH 7.6) ?4mM
CaCl
2
?0.01% ascorbic acid
30 min at 22°C
in incubation
buffer
60 min at 22°C in incubation
buffer
195 min at 4°C in incubation
buffer
Dopaminergic
D
1
-like
[
3
H]SCH-23390
(0.5 nM)
SKF 83566
(1 lM)
50 mM Tris–HCl (pH 7.4) ?120 mM
NaCl ?5 mM KCl ?2 mM CaCl
2
?1 mM MgCl
2
?1lM mianserin
20 min at 22°C
in incubation
buffer
90 min at 22°C in incubation
buffer
2910 min at 4°C in incubation
buffer
242 Brain Struct Funct (2011) 216:239–254
123
from one animal is thus transformed into a receptor con-
centration (fmol/mg protein).
Anatomical identification
The borders of the NCL were identified based on previous
neurochemical (Waldmann and Gu
¨ntu
¨rku
¨n1993) and tract-
tracing studies (Kro
¨ner and Gu
¨ntu
¨rku
¨n 1998). The borders
of the NCL and surrounding structures as defined in the
atlas of Karten and Hodos (1967) were traced on prints of
the digitized autoradiographs. The borders of rat Fr2 and
Cg1 were anatomically identified based on a rat cortex atlas
(Zilles 1985). We decided to analyze these two regions
because they are assumed to be a part of the rat frontal and
prefrontal cortex (Uylings et al. 2003). The borders of
human BA10 were identified based on criteria defined by
Brodmann (Brodmann 1909). BA10m and BA10l were
additionally defined and traced onto digitized autoradio-
graphs (n=3 hemispheres). The mean of the gray values
contained in a specific region over a series of 4–5 sections
from one hemisphere is thus transformed into a receptor
concentration per unit protein (fmol/mg protein).
Statistical analysis
To investigate the chemoarchitectural differences between
the NCL and the surrounding structures, the binding site
concentrations of the NCL were compared with those of
the nidopallium caudomediale (NCM, located medial to
NCL) and the dorsolateral corticoid area (CDL, located
dorsolaterally to NCL, above the ventricle). First, a
Friedman ANOVA was conducted. If significant, pair wise
comparisons were run with the Wilcoxon rank test. Binding
site concentrations of the NCM were measured medial to
Field L. Differences between nidopallium caudolaterale
pars medialis (NCLm) and nidopallium caudolaterale pars
lateralis (NCLl) were further analyzed with Wilcoxon rank
tests.
Results
Receptor-binding site densities in avian pallial
structures
The most caudal portion of the avian nidopallium displays
a rather homogeneous cytoarchitecture. The only subven-
tricular cytoarchitectural feature that is clearly different
from the otherwise homogeneous pattern is Field L in the
most medial part. Within Field L, especially, the granular
layer L2 is readily visible. Ventrolaterally, the lamina
arcopallialis dorsalis defines the borderline between the
nidopallium and the arcopallial and the amygdalar
substructures. Dorsally, the caudal cap of the lateral ven-
tricle separates the nidopallium from the CDL and the
hippocampal and parahippocampal structures. The distri-
bution of different ligand-binding sites shows that the
cytoarchitectonically seemingly homogeneous caudal nid-
opallium is in fact comprised of several substructures.
Further, the examined receptor types not only enabled a
clear delineation of the NCL from the adjoining areas, but
also revealed the existence of two hitherto unknown sub-
entities. Stereotaxic coordinates, A 5.50 and A 6.75, were
chosen as exemplary levels for which all receptor types
were shown in Fig. 1a and b. Different binding-site den-
sities within the borders of the NCL could be followed up
to the most caudal aspect of the subventricular forebrain
where it constituted the most caudal tip of the nidopallium.
Frontally, the NCL was visible up to A 7.50 (anterior–
posterior coordinates according to the pigeon brain atlas of
Karten and Hodos 1967). Further, binding-site densities of
the different receptor ligands are presented relative to each
other in a 2-dimensional coordinate plot to construct
a receptor fingerprint for a given brain area (Fig. 2).
This allows us to compare the shape and the absolute size
of this receptor fingerprint across brain areas and between
species.
As illustrated in the autoradiographs and in the finger-
prints, glutamatergic AMPA and NMDA receptors show
the highest densities of all measured receptors, and were
followed by GABA
A
receptor densities. Conversely, lowest
values were found for nACh, a
1
and D
1
-like receptor
densities (Figs. 1a/b, 2a).
The mean density of AMPA receptors for the whole
NCL was 2,252 ±269 fmol/mg protein. A comparison of
binding densities between NCL, NCC and CDL using a
Friedman ANOVA showed no significant overall effects
[Chi Square (N=6, df =2) =1, n.s.].
Overall densities for kainate receptor-binding sites were
660 ±81 fmol/mg (Fig. 1a/b). Because kainate receptor-
binding sites were approximately fivefold lower than those
of AMPA receptors, and sixfold lower than those of
NMDA receptors, this resulted in a considerable indenta-
tion in the fingerprint (Fig. 2a). Binding of [
3
H]kainate was
highest in the most lateral portion of the NCL (Figs. 1a/b,
3). We labeled this area nidopallium caudolaterale pars
lateralis (NCLl) to differentiate it from the medial portion
of the NCL (NCLm). The Friedman ANOVA showed a
significant overall effect [Chi Square (N=6, df =2) =
10.33, p\0.01]. Binding was higher both in NCL and
NCC than in CDL (all N=6, T=0, p\0.05). Addi-
tionally, a significant higher concentration of kainate
receptors in the lateral than in the medial aspect of the NCL
was detected (N=6, T=1, p\0.05; Fig. 3).
Binding of [
3
H]MK-801 was very high throughout the
entire caudal nidopallium (Fig. 1a/b), indicating a high
Brain Struct Funct (2011) 216:239–254 243
123
density of NMDA receptors. Binding density reached
2,525 ±143 fmol/mg protein (Figs. 1a, 2a) in NCL. Like
for the AMPA receptors this results in a prominent peak in
the fingerprints (c.f Fig. 2a). The Friedman ANOVA
comparing NCL, NCC and CDL showed a significant
overall effect [Chi Square (N=6, df =2) =8.33,
p\0.05]. A subsequent Wilcoxon test revealed signifi-
cantly higher values for NCL and NCM over CDL (all
N=6, TB1, p\0.05).
GABA
A
receptor-binding sites were labeled with
[
3
H]muscimol (mean density 1,810 ±188 fmol/mg pro-
tein). Since binding increased medial to NCC, the border of
the NCL could be easily visualized (Fig. 1a/b). The Fried-
man ANOVA showed a significant main effect [Chi Square
(N=6, df =2) =12, p\0.005]. Subsequent Wilcoxon
tests revealed a significantly stepwise decrease of binding
from NCC over NCL to CDL (all N=6, T=0, p\0.05)
that is particularly illustrated in the fingerprints (Fig. 2a).
The study of binding sites for the neurotransmitter
acetylcholine revealed low densities for all analyzed cho-
linergic receptors. Binding of [
3
H]pirenzepine to muscari-
nergic cholinergic receptors of the M
1
-type was very low in
the caudolateral nidopallium (NCL: 151 ±24 fmol/mg
protein; Figs. 1a/b, 2a). A further differentiation within
NCL was not visible. The Friedman ANOVA using the
data from NCL, NCC and CDL showed a significant
overall effect [Chi Square (N=6, df =2] =6.52,
p\0.05). Subsequent Wilcoxon tests revealed that the
concentration was significantly lower in NCL compared to
both NCC and CDL (N=6, TB1, p\0.05; Fig. 2).
M
2
-receptors presented the highest densities of all
determined cholinergic receptors in the nidopallial struc-
tures (269 ±39 fmol/mg protein; Fig. 2a). The Friedman
ANOVA comparing NCL, NCC and CDL showed a
significant overall effect [Chi Square (N=6, df =2) =
10.33, p\0.01). Subsequent Wilcoxon tests revealed a
Fig. 1 Color-coded autoradiographs showing the distribution of
AMPA, kainate, NMDA, GABA
A
,M
1
,M
2
, nicotinic cholinergic
(nACh), a
1
,a
2
, 5-HT
1A
and D
1
-like receptors in coronal sections
through the pigeon brain at rostrocaudal levels A 5.50 (a) and A 6.75
(b). Extent of the NCL at each of these levels is highlighted in gray in
the schematic drawing. Scale bars code for receptor densities in fmol/
mg protein
244 Brain Struct Funct (2011) 216:239–254
123
significantly stepwise increase of binding strength from
CDL over NCL to NCM (N=6, TB1, p\0.05) and a
parcellation of NCLm and NCLl (N=6, T=0, p\0.05;
Fig. 3).
Binding of [
3
H]cytisine to nicotinic receptors was very
low in the whole lateral aspect of the nidopallium
(144 ±12 fmol/mg protein, Fig. 1a/b and Fig. 2a), indi-
cating low densities of nACh receptors (Fig. 2a). The
Friedman ANOVA showed a significant overall effect [Chi
Square (N=6, df =2) =12, p\0.005]. Subsequent
Wilcoxon tests revealed a significantly stepwise decrease
of binding strength from CDL over NCL to NCC (N=6,
T=0, p\0.05). Further, binding densities between
NCLm and NCLl differed significantly (N=6, T=0,
p\0.05; Fig. 3).
The noradrenergic a
1
receptor was visualized by means
of [
3
H]prazosin (127 ±16 fmol/mg protein; Fig. 1a/b).
Although in few cases the ventral aspect of the NCL,
abutting the arcopallium, displayed some higher binding,
this was not consistently observed. A differentiation
between NCLl and NCLm was not evident. The Friedman
ANOVA comparing NCL, NCC and CDL showed a sig-
nificant overall effect [Chi Square (N=6, df =2) =12,
p\0.005]. Subsequent Wilcoxon tests revealed a signifi-
cantly stepwise decrease of binding strength from CDL
over NCL to NCC (all N=6, T=0, p\0.05; Fig. 2a).
[
3
H]RX821002 binds to noradrenergic a
2
receptor and
displayed moderate binding in NCL (308±27 fmol/mg
protein). Substructures within the NCL were not visible
(Fig. 1a/b). The Friedman ANOVA showed a significant
overall effect [Chi Square (N=6, df =2) =9.33,
p\0.01]. Subsequent Wilcoxon tests revealed that bind-
ing in NCC was significantly higher than both in NCL and
CDL (all N=6, T=0, p\0.05; Fig. 2).
Serotonergic 5-HT
1A
receptor-binding sites were visu-
alized with [
3
H]8-OH-DPAT. NCL revealed lower densi-
ties (374 ±67 fmol/mg protein) than the medially abutting
nidopallial areas, again providing the possibility to clearly
identify the medial wall of the NCL (Fig. 1a/b). The
Friedman ANOVA showed a significant overall effect
Fig. 1 continued
Brain Struct Funct (2011) 216:239–254 245
123
[Chi Square (N=6, df =2] =12, p\0.005). Sub-
sequent Wilcoxon tests revealed a significantly stepwise
increase of binding strength from CDL over NCL to NCC
(all N=6, T=0, p\0.05, Fig. 2a). Furthermore,
5-HT
1A
receptors were more abundant in NCLm than
NCLl (N=6, T=0, p\0.05; Fig. 3).
[
3
H]SCH23390 was used to reveal the location and
density of dopaminergic D
1
-like receptors. Ligand bind-
ing was mainly concentrated within the NCL without
showing a difference between the lateral and the medial
component (Fig. 1a/b). Although density in NCL was
rather low (92 ±12 fmol/mg protein), a Friedman
ANOVA comparing NCL, NCC, and CDL showed a
significant overall effect [Chi Square (N=6, df =2) =
12, p\0.01]. A subsequent Wilcoxon test revealed sig-
nificantly higher values for NCL and CDL over NCC
(N=6, T=0, p\0.05; Fig. 2) as well as significantly
higher values for NCL than for CDL (N=6, T=0,
p\0.05; Fig. 2).
Based on the different binding site densities for kai-
nate, NMDA, GABA
A
,M
1
,M
2
, nACh, a
1
,a
2
, 5-HT
1A
and
D
1
-like receptors a detailed outline of the NCL is depicted
in Fig. 4.
Comparison of receptor-binding site densities
in the avian NCL to mammalian prefrontal structures
In the rat (Fig. 2b) and human (Fig. 2c) prefrontal areas
examined, AMPA and GABA
A
receptors showed the
highest densities of all measured receptor types, and were
followed by NMDA receptor densities (Fig. 2b/c). Lowest
values were found for nACh, and D
1
-like receptor
densities.
Human and rat prefrontal areas differed considerably in
their relative balance of ionotropic glutamatergic receptors.
In human areas, BA10l and BA10m, kainate receptor
densities were comparable to those of AMPA receptors,
and only slightly lower than those of NMDA receptors
(Fig. 2c). In rat areas, Fr2 and Cg1, similar to the situation
described for the pigeon nidopallial areas, kainate receptor
Fig. 2 Receptor fingerprints for CDL, NCL, NCC of the pigeon
pallium (a), for Fr2 and Cg1 of the rat cortex (b) and for the BA10l
and BA10m of the human cortex (c). The mean densities (fmol/mg
protein) of glutamatergic (AMPA, kainate, NMDA), GABAergic
(GABA
A
), acetylcholinergic muscarinic (M
1
,M
2
) and nicotinic
(nACh), adrenergic (a
1
,a
2
), serotonergic (5-HT
1A
) and dopaminergic
(D
1
-like) receptors are displayed in polar coordinate plots. The lines
connecting the mean densities define the shape of the fingerprint
based on 11 different binding sites for each area. Note that the scales
in acare different. BA10l Brodmann area 10 lateral, BA10m
Brodmann area 10 medial, CDL area corticoidea dorsolateralis, NCL
nidopallium caudolaterale, NCC nidopallium caudocentrale, Fr2
frontal area 2, Cg1 cingulate cortex 1
b
246 Brain Struct Funct (2011) 216:239–254
123
densities were considerably lower than those of AMPA
(fourfold lower) or NMDA (five to sixfold lower) receptor
densities (Fig. 2b). Thus, the pigeon and rat, but not the
human fingerprints presented a conspicuous indentation at
the level of the kainate receptors.
The examined human and rat prefrontal areas presented
the same balance of cholinergic receptor densities, with
highest concentrations for the muscarinic M
1
cholinergic
type and lowest values for the nicotinic receptor (Fig. 2b/
c). This pattern differs however, from that of pigeons, since
nidopallial areas contain higher M
2
than M
1
receptor
densities (Fig. 2a).
In the group of monoaminergic receptors, noradrenergic
a
1
receptor densities were higher than those of a
2
receptors
in both human and rat prefrontal areas (Fig 2b/c). Con-
versely, a
1
receptor densities were lower than of a
2
receptor densities in the pigeon nidopallium (Fig. 2a).
Serotoninergic 5-HT
1A
receptor densities were higher than
those of a
1
receptors in human areas BA10l and BA10m,
whereas the opposite holds true for rat areas Fr2 and Cg1
(Fig. 2b/c). D
1
-like binding-site densities showed neither
differences between the analyzed prefrontal structures nor
the pigeon’s NCL (Fig. 2b/c).
Discussion
Using a quantitative analysis of 11 different receptor-
binding sites, the present study aimed to (1) analyze the
areal borders of the constituents of the caudolateral part of
the pigeons’ telencephalon, (2) to reveal possible subdivi-
sions within the NCL, (3) to compare the receptor finger-
prints of NCL and the surrounding NCC and CDL with
those of frontal areas in mammals.
Fig. 3 Histogram of the mean
receptor densities (fmol/mg
protein) of the pigeon’s areas
NCLm and the NCLl. Error
bars represent standard
deviations. Asterisks indicate
significant differences between
receptor densities
Fig. 4 Atlas of the NCL in serial frontal sections based on different receptor densities. The length of the bar represents 3 mm
Brain Struct Funct (2011) 216:239–254 247
123
Areal delineation in the pigeons’ caudolateral
telencephalon
Moving from centromedial to lateral, the avian caudolat-
eral telencephalon is constituted by the three areas: NCC,
NCL, and CDL. The NCC receives its input predominantly
from the dorsal intermediate mesopallium and projects to
arcopallial subfields. The arcopallial outflow to the medial
hypothalamus could imply that NCC is involved in neu-
roendocrine and autonomic functions and is limbic in
nature (Yamamoto and Reiner 2005; Atoji and Wild 2009).
The interconnectivity between NCC and NCL seems to be
surprisingly weak (Atoji and Wild 2009; Kro
¨ner and
Gu
¨ntu
¨rku
¨n1999). Further, the pattern of afferents and
efferents of NCC and NCL is considerably different
(Leutgeb et al. 1996; Metzger et al. 1998; Kro
¨ner and
Gu
¨ntu
¨rku
¨n1999; Atoji and Wild 2009). Thus, although
NCC and NCL cannot be delineated by cytoarchitectonic
means and were subsumed into area Ne16 in the quanti-
tative cytoarchitectonic study of Rehka
¨mper and Zilles
(1991), they show marked differences in hodology. The
study of Atoji and Wild (2009) placed the borderline
between NCC and NCL far more laterally than the
immunocytochemical and connectivity analyses conducted
on the NCL (Waldmann and Gu
¨ntu
¨rku
¨n1993; Leutgeb
et al. 1996; Kro
¨ner and Gu
¨ntu
¨rku
¨n1999; Riters et al.
1999). In fact, according to Atoji and Wild (2009), NCLm
would be part of NCC. Interestingly, the reconstruction of
the location of retrogradely labeled neurons in Atoji and
Wild (2009) reveals a border that is more close to that of
the present study and similar to the original delineation by
Waldmann and Gu
¨ntu
¨rku
¨n(1993) and this is reflected by
the distribution patterns of a
1
, 5-HT
1A
and D
1
-like recep-
tors. However, the caudal aspect of the avian nidopallium
is organized in clusters with fuzzy borders; in addition, not
all receptor-binding sites defined clear boundaries between
areas. Thus, the distribution patterns of the receptors con-
firm a smooth transition at the caudal site and both areas
probably do not have a clear boundary at that point.
Therefore, in the most caudal portion of the nidopallium,
the delineation between NCC and NCL becomes extremely
difficult and may have led to different findings in the past
(Atoji and Wild 2009).
Towards the lateral border, the distinction between NCL
and CDL is easy due to the ventricle that separates these
two areas. The CDL is considered to be mostly limbic in
nature and was hodologically compared to the mammalian
cingulate cortex (Yamamoto and Reiner 2005; Atoji and
Wild 2005; Csillag and Montagenese 2005). It shares
similarities with the receptor architecture of the hippo-
campal formation (data not shown) and nidopallial struc-
tures. CDL extents rostrally up to A 6.75 where NCL and
CDL are no longer separated by the lateral ventricle but
directly abut each other. At this point, the autoradiographic
data revealed a less fuzzy transition when compared to the
caudal aspects of NCL and NCC, depicting that NCL fol-
lows the outer curvature of the telencephalon but always
stays about 1 mm away from the pial surface. Similarly,
the rostral border of the NCL is easier to define as it tapers
up to A 7.50.
Subdivisions of the NCL
Our findings reveal a clear parcellation of the avian nid-
opallium that is in line with tracing studies (Rehka
¨mper
and Zilles 1991; Leutgeb et al. 1996; Kro
¨ner and Gu
¨ntu
¨r-
ku
¨n1999; Atoji and Wild 2009). Earlier studies have
shown functional and neurochemical subdivisions of the
NCL (Leutgeb et al. 1996; Braun et al. 1999; Kro
¨ner and
Gu
¨ntu
¨rku
¨n1999; Riters et al. 1999). Here, a new subdi-
vision into a medial and a lateral part is proposed by the
differences of the mean receptor densities of nACh, M
2
,
kainate, and 5-HT
1A
receptors. Some earlier tracing and
neurochemical studies revealed a possible dorsal and ven-
tral component (Leutgeb et al. 1996; Braun et al. 1999;
Riters et al. 1999). The neurochemical subdivision into a
dorsal and a ventral component also coincides with hod-
ological data showing that only dorsal NCL receives
afferents from multimodal thalamic nuclei (Korzeniewska
and Gu
¨ntu
¨rku
¨n1990;Gu
¨ntu
¨rku
¨n and Kro
¨ner 1999) and
contributes more significantly to working memory perfor-
mance (Diekamp et al. 2002a,b). Dorsal, but not ventral
NCL, is connected with a complex of association structures
in the rostromedial nidopallium and ventral hyperpallium
in different species of birds. In domestic chicken two
extensively overlapping structures, the mediorostral nid-
opallium/hyperpallium (MNH) and the intermediate and
medial mesopallium ventrale (IMM), play a pivotal role in
auditory and visual filial imprinting, respectively (Horn
1981; Braun et al. 1999). These areas are activated during
imprinting and lesions cause deficits in recognizing the
imprinting stimulus (Horn 1981; Horn et al. 1985). In
chicken, IMM is also a nodal point of initial memory
formation in one-trial passive avoidance learning with
gustatory cues (Rose 2000). Both MNH and IMM project
to dorsomedial NCL as shown in chicken (Metzger et al.
1998) and pigeons (Kro
¨ner and Gu
¨ntu
¨rku
¨n1999). How-
ever, we could not confirm a border between dorsal and
ventral NCL based on the receptor-density profiles. On the
other hand, Kro
¨ner and Gu
¨ntu
¨rku
¨n(1999) demonstrated
that the component labeled NCLl in our preparations
receives input from secondary areas of sensory represen-
tation and projects back to these structures. Furthermore, a
large number of neurons from NCL projects to the arco-
pallium and these output neurons are close to the densest
catecholaminergic innervations that are located in the
248 Brain Struct Funct (2011) 216:239–254
123
lateral part of the NCL (Waldmann and Gu
¨ntu
¨rku
¨n1993;
Kro
¨ner and Gu
¨ntu
¨rku
¨n1999). In addition, a large number
of medial NCL neurons project to the basal ganglia in
pigeons (Veenman et al. 1995; Kro
¨ner and Gu
¨ntu
¨rku
¨n
1999). Therefore, NCLl displayed a different connectivity
pattern from NCLm. Due to the curvature of the NCL,
NCLl is positioned more dorsally than NCLm. Thus, a
dorsoventral subdivision of the NCL could mistakenly be
concluded from the lateromedial differentiation of a
semilunar structure.
The neurochemistry of the caudolateral avian forebrain
In NCL, NCC, and CDL the highest receptor densities were
detected for glutamatergic and GABA
A
receptors. This is
in line with earlier studies that determined receptor levels
in the nidopallium of various bird species (Dietl and
Palacios 1988; Stewart et al. 1988,1999; Mitsacos et al.
1990; Aamodt et al. 1992; Veenman et al. 1994; Ben-Ari
et al. 1997; Salvatierra et al. 1997). Pigeons showed higher
AMPA and NMDA receptor concentrations in the nid-
opallium when compared to other birds, while the amount
of GABA
A
receptor densities seemed to be similar in
pigeons, chicks and zebra finches (Stewart et al. 1988;
Henley and Barnard 1990; Veenman et al. 1994; Martinez
de la Torre et al. 1998; Stewart et al. 1999; Pinaud and
Mello 2007). The present study reports for the first time
kainate receptor densities in the pigeon’s pallium. If
compared to AMPA and NMDA receptors, kainate binding
was about four times lower in all of the above-mentioned
structures. However, like for the NMDA receptors, kainate
binding differed between the CDL and the nidopallial
structures, showing a clear segregation. This is in line with
an immunohistochemical study in quails, showing that
AMPA and NMDA receptors have higher densities than
kainate receptors in the nidopallium. In addition, kainate
and NMDA binding is lower in the CDL while the AMPA
receptor subunit GluR1 was intensely labeled in the CDL
(Cornil et al. 2000). Binding of the GABA
A
receptor also
increased from the surface to the deeper nidopallial areas,
confirming earlier immunohistochemical und receptor
autoradiographic studies (Rehka
¨mper and Zilles 1991;
Veenman et al. 1994). In the nidopallium, cholinergic
muscarinic and nicotinic receptors showed an intermediate
to low density, which is in line with results from other
studies of muscarinic or nicotinic binding sites in the tel-
encephalon of pigeons, chicks, quails, sparrows, and star-
lings (Dietl et al. 1988; Ball et al. 1990; Sorenson and
Chiappinelli 1992). As described for the GABA
A
receptor,
the M
2
receptor density increases from the superficial CDL
over the NCL to the NCM while the nACh receptor den-
sities decreases. The boundaries of the NCL were revealed
by all cholinergic receptors.
The monoaminergic receptors were differentially dis-
tributed. Their densities ranged from very low (D
1
-like
receptors) to moderate (5-HT
1A
receptors). Densities of the
a
2
receptors varied across different bird species in the CDL
and in the nidopallium (Balthazart and Ball 1989; Ball
et al. 1995; Diez-Alarcia et al. 2006). To our knowledge to
date no specific information about the densities of 5-HT
1A
receptor densities is available on the avian pallium,
although it was shown in a competition assay with [
3
H]5-
HT binding that 5-HT
1A
receptors were abundant in the
pigeon’s telencephalon (Waeber et al. 1989). Comparable
results were reported for the D
1
-like receptor in the nid-
opallium of pigeons (Dietl and Palacios 1988).
Comparison to mammals and functional considerations
As first shown by lesion experiments (Mogensen and
Divac 1982), the NCL is involved in executive functions.
More recent studies have confirmed that the NCL shares
many similarities with the mammalian prefrontal cortex
(Gu
¨ntu
¨rku
¨n, 2005a,b; Kirsch et al. 2008). These findings
can be seen in parallel to observations in corvids and
parrots which possess cognitive abilities that are compa-
rable to those of monkeys and apes (Bird and Emery 2010;
Hunt and Gray 2003; Emery and Clayton 2004; Kenward
et al. 2005; Seed et al. 2006; Prior et al. 2008; Taylor et al.
2009; Pollok et al., 2000). As observed for other mammals
(Harvey and Krebs 1990) this is accompanied by an
increased encephalization (Cnotka et al. 2008) and a rela-
tive growth of associative forebrain areas (Mehlhorn et al.
2010). Based on topographical and genetic arguments both
the NCL and the prefrontal cortex seem to be a case of
homoplasy (Puelles et al. 2000). Additionally, the mor-
phological organization of avian and mammalian fore-
brains differs importantly, with the avian pallium having a
nuclear organization while the mammalian dorsal pallium
assumes a laminar structure. Thus, a layered cortical
structure appears not to be a prerequisite for higher cog-
nitive functions (Kirsch et al. 2008). In contrast to the
NCL, less is known about the CDL and its functions. The
connections of the avian CDL share similarities with those
of the mammalian cingulate cortex (Vogt and Pandya
1987; Atoji and Wild 2005). Neurobehavioral studies in
which the CDL was lesioned as part of larger lesions to the
lateral nidopallium or the hippocampal formation indicate
a role for the CDL in spatial memory (Hartmann and
Gu
¨ntu
¨rku
¨n1998; Bingman et al. 1985; Colombo et al.
2001; Gagliardo et al. 2001). Only one study showed that
CDL lesions did not impair performance in simultaneous
pattern or delayed alternation discrimination tasks
(Gagliardo et al. 1996). Receptor autoradiography and
receptor fingerprints of brain regions provide a tool to
compare the chemoarchitecture between different species.
Brain Struct Funct (2011) 216:239–254 249
123
Therefore, our results will be further discussed in the light
of comparative studies in birds, primates and rats.
As in the pigeon’s NCL and CDL, high receptor den-
sities for glutamatergic and GABAergic receptors were
found in the prefrontal regions investigated here, as well as
in other cortical regions of rats, monkeys and humans
(Gebhard et al. 1995; Geyer et al. 1998; Zilles et al. 2002a,
b; Palomero-Gallagher and Zilles 2004). However, there
were differences in the amount of distinct glutamate
receptors between species. AMPA and NMDA receptors
showed high concentrations in the NCL and the CDL of
pigeons and chicks (Bock et al. 1997) if compared to
frontal structures of mammals. Kainate receptors seemed to
be very low in rat FR2 and Cg1, while they did not differ
substantially between human BA10 and the NCL, and
between the CDL and the human cingulate cortex (Palo-
mero-Gallagher et al. 2009). By contrast, the amounts of
GABA
A
receptors were equally distributed in the prefrontal
areas of all the investigated species here and also in the
NCL of pigeons and chicks (Stewart et al. 1988). The same
is true for the CDL and the human as well as the macaque
cingulate cortex (Bozkurt et al. 2005; Palomero-Gallagher
et al. 2009). Thus, there seems to be a shift towards higher
densities of glutamate receptors in avian nidopallial
structures. Therefore, the top right quadrant of the finger-
prints for the birds’ nidopallial structures differs in size
when compared to the rodent frontal areas, and differ in
shape for both species, if compared to human BA10.
Cholinergic M
1
receptors were highest in human if
compared to macaque monkey, rhesus monkey, rat and
pigeon, while M
2
and nicotinic receptors showed equal
densities (Bozkurt et al. 2005; Lidow et al. 1989). How-
ever, pigeons showed an inverted pattern of M
1
/M
2
binding
if compared to other species. ACh is an essential regulator
of cortical excitability and plays important roles for arou-
sal, attention, and cognitive processes (Sarter and Bruno
2000; Hasselmo and Stern 2006; Briand et al. 2007; Sarter
et al. 2009). These functions are mediated by muscarinic
and nicotinic ACh receptors. In the cerebral cortex the M
1
receptor is preferentially expressed in pyramidal cells and
enriched on the extrasynaptic membrane of their dendrites
and spines (Yamasaki et al. 2010). The M
2
receptor is the
primary muscarinic autoreceptor presynaptically regulating
ACh release in the forebrain of rodents and primates
including humans (Mrzljak et al. 1995; Zhang et al. 2002).
Both receptor subtypes are metabotropic. M
1
couples to a
stimulatory G-protein whereas M
2
couples to an inhibitory
G-protein. Genetic variation of the CHRM2 gene encoding
the M
2
receptor selectively influence muscarinic presyn-
aptic inhibition (Comings et al. 2003). The nACh receptors
are fast-acting ligand-gated ion channels producing EPSPs.
A recent genetic approach showed that both, fast-acting
nicotinic receptors and slow-acting muscarinic receptors
influence in a synergistic system the efficiency of shifting
visuospatial attention in the PFC (Greenwood et al. 2009).
In pigeons, central cholinergic systems are important for
temporal memory processes and spatial orientation during
homing, two processes that also involve the NCL
(Gagliardo and Divac 1993; Santi and Weise 1995; Kohler
et al. 1996; Riters and Bingman 1999).
Like for the muscarinic cholinergic receptors, the same
inverted ratio was detected in the NCL and in the CDL for
the noradrenergic a
1
and a
2
receptors if compared to pre-
frontal or cingulate structures in mammals. In humans,
macaque monkeys and rats higher amounts of a
1
than of a
2
receptors were described (Goldman-Rakic et al. 1990;
Bozkurt et al. 2005; Palomero-Gallagher et al. 2009). Both
receptor types are metabotropic and a
1
receptors are cou-
pled to stimulatory G-proteins, while a
2
receptors are
coupled to inhibitory G-proteins. In the PFC of monkeys,
a
2
receptors are located postsynaptically at the dendritic
spines of pyramidal neurons where glutamate receptors are
concentrated (Aoki et al. 1998). Behavioral pharmacolog-
ical studies in rodents, monkeys, and humans demonstrated
that systemically or locally administered a
2
receptor ago-
nists could improve PFC cognitive performances (Robbins
and Arnsten 2009). Further, it was shown that stimulation
of a
2
receptors suppresses glutamate synaptic transmission
in the PFC and tunes the synaptic output to an optimal state
for working memory function (Wang et al. 2007; Ji et al.
2008). In songbirds noradrenalin is involved in song
learning at different developmental stages by controlling
local circuits in the higher vocal center (HVC) (Fortune
and Margoliash 1995) and modulation of auditory
responses through attention processes (Castelino and
Schmidt 2010). The HVC could be an oscine specialization
of the dorsal NCL (Farries 2001). Because both the M
1
/M
2
and the a
1
/a
2
ratio show an inverted pattern in the NCL
resulting in an increased inhibitory control on local circuits
this may be a compensating mechanism for the shift to
glutamatergic processing.
The densities of 5-HT
1A
receptors were equal in the
prefrontal areas of humans, monkeys and pigeons, while
rats showed lower densities (Goldman-Rakic et al. 1990).
The 5-HT
1A
subtype is of particular interest, since it is one
of the main mediators of 5-HT and contributes to a lot of
prefrontal functions (Sakaue et al. 2000; de Almeida et al.
2008). In the human cingulate cortex the density of the
5-HT
1A
subtype is slightly higher than in the CDL (Palo-
mero-Gallagher et al. 2009). In birds less is known about
the serotonergic contribution to executive functions, but it
was shown that serotonin release was increased in the NCL
during a working memory task (Karakuyu et al. 2007).
D
1
-like receptors showed the lowest densities of all
measured receptor types in the assumed prefrontal and
cingulate regions of pigeons, rats, monkeys, cats, tree
250 Brain Struct Funct (2011) 216:239–254
123
shrews and humans (Richfield et al. 1989; Goldman-Rakic
et al. 1990; Palomero-Gallagher et al. 2009). In mammals,
low densities of D
1
-like receptors in frontal areas are
associated with volume transmission of dopamine and a
diffuse action of dopamine on multiple components of
cortical networks (reviewed in Gonzalez-Burgos et al.
2007). These results also reveal that the dopaminergic
system seems to be highly conserved across species,
although prefrontal structures evolved independently
(Callier et al. 2003). Thus, the dopaminergic system and its
interactions with other systems might constitute a key
element for our understanding of the anatomical/chemical
traits that are necessary for proper executive functions. The
low density of D
1
-like receptors might also explain why
species share similar deficits if signaling via this receptor-
type is disturbed (Zahrt et al. 1997; Williams and Castner
2006; Herold et al. 2008; McNab and Klingberg 2008;
Rose et al. 2010).
In summary, it appears that the GABAergic and dopa-
minergic systems are highly conserved across the species
studied here, which have a long history of separate evo-
lution (Jarvis et al. 2005). This could result from a common
selection pressure for a structure that serves executive
functions, i.e., the control of higher order processes. This
includes the integration and manipulation of information
from all modalities in order to generate a proper behavior
in a given situation. These functions rely on specific con-
nections to other brain structures and the modulation of
information flow through these circuits. Thus, similar
evolutionary pressures on information processing in birds
might result in a comparable or analogue pattern of specific
receptor compositions that would resemble those in the
neocortex of mammals. Future studies need to examine
differences between various bird species, as well as
between different mammalian species to confirm these
conclusions.
Acknowledgments Supported by a grant from the BMBF through
the Bernstein Focus Group ‘Varying Tunes’’ to O.G.
Conflict of interest The authors declare that they have no conflict
of interest.
References
Aamodt SM, Kozlowski MR, Nordeen EJ, Nordeen KW (1992)
Distribution and developmental change in [3H]MK-801 binding
within zebra finch song nuclei. J Neurobiol 23:997–1005
Amunts K, Weiss PH, Mohlberg H, Pieperhoff P, Eickhoff S, Gurd
JM, Marshall JC, Shah JN, Fink GR, Zilles K (2004) Analysis of
the neural mechanisms underlying verbal fluency in cytoarchi-
tectonically defined stereotaxic space—the roles of Brodmann
areas 44 and 45. NeuroImage 22:42–56
Amunts K, Schleicher A, Zilles K (2007) Cytoarchitecture of the
cerebral cortex—more than localization. NeuroImage 37:1061–
1065
Aoki C, Venkatesan C, Go CG, Forman R, Kurose H (1998) Cellular
and subcellular sites for noradrenergic action in the monkey
dorsolateral prefrontal cortex as revealed by the immunocyto-
chemical localization of noradrenergic receptors and axons.
Cereb Cortex 8:269–277
Atoji Y, Wild JM (2005) Afferent and efferent connections of the
dorsolateral corticoid area and a comparison with connections of
the temporo-parieto-occipital area in the pigeon (Columba livia).
J Comp Neurol 485:165–182
Atoji Y, Wild JM (2009) Afferent and efferent projections of the
central caudal nidopallium in the pigeon (Columba livia).
J Comp Neurol 517:350–370
Ball GF, Nock B, Wingfield JC, McEwen BS, Balthazart J (1990)
Muscarinic cholinergic receptors in the songbird and quail brain:
a quantitative autoradiographic study. J Comp Neurol
298:431–442
Ball GF, Casto JM, Balthazart J (1995) Autoradiographic localization
of D1-like dopamine receptors in the forebrain of male and
female Japanese quail and their relationship with immunoreac-
tive tyrosine hydroxylase. J Chem Neuroanat 9:121–133
Balthazart J, Ball GF (1989) Effects of the noradrenergic neurotoxin
DSP-4 on luteinizing hormone levels, catecholamine concentra-
tions, alpha 2-adrenergic receptor binding, and aromatase
activity in the brain of the Japanese quail. Brain Res
492:163–175
Bast T, Diekamp B, Thiel C, Schwarting RK, Gu
¨ntu
¨rku
¨n O (2002)
Functional aspects of dopamine metabolism in the putative
prefrontal cortex analogue and striatum of pigeons (Columba
livia). J Comp Neurol 446:58–67
Ben-Ari Y, Khazipov R, Leinekugel X, Caillard O, Gaiarsa JL (1997)
GABAA, NMDA and AMPA receptors: a developmentally
regulated ‘menage a trois’. Trends Neurosci 20:523–529
Bingman VP, Ioale
`P, Casini G, Bagnoli P (1985) Dorsomedial
forebrain ablations and home loft association behavior in homing
pigeons. Brain Behav Evol 26:1–9
Bird CD, Emery NJ (2010) Rooks perceive support relations similar
to six-month-old babies. Proc Biol Sci 277:147–151
Bock J, Schnabel R, Braun K (1997) Role of the dorso-caudal
neostriatum in filial imprinting of the domestic chick: a
pharmacological and autoradiographical approach focused on
the involvement of NMDA-receptors. Eur J Neurosci 9:1262–
1272
Bozkurt A, Zilles K, Schleicher A, Kamper L, Arigita ES, Uylings
HB, Ko
¨tter R (2005) Distributions of transmitter receptors in the
macaque cingulate cortex. Neuroimage 25:219–229
Braun K, Bock J, Metzger M, Jiang S, Schnabel R (1999) The
dorsocaudal neostriatum of the domestic chick: a structure
serving higher associative functions. Behav Brain Res 98:211–
218
Briand LA, Gritton H, Howe WM, Young DA, Sarter M (2007)
Modulators in concert for cognition: modulator interactions in
the prefrontal cortex. Prog Neurobiol 83:69–91
Brodmann K (1909) Vergleichende Lokalisationslehre der Großhirnr-
inde in ihren Prinzipien dargestellt auf Grund des Zellenbaues,
Barth, Leipzig; English translation available in Garey, L. J.
Brodmann’s Localization in the Cerebral Cortex (Smith Gordon,
London, 1994)
Callier S, Snapyan M, Le Crom S, Prou D, Vincent JD, Vernier P
(2003) Evolution and cell biology of dopamine receptors in
vertebrates. Biol Cell 95:489–502
Castelino CB, Schmidt MF (2010) What birdsong can teach us about
the central noradrenergic system. J Chem Neuroanat 39:96–111
Brain Struct Funct (2011) 216:239–254 251
123
Cnotka J, Gu
¨ntu
¨rku
¨n O, Rehka
¨mper G, Gray RD, Hunt GR (2008)
Extraordinary large brains in tool-using New Caledonian crows
(Corvus moneduloides). Neurosci Lett 433:241–245
Colombo M, Broadbent NJ, Taylor CS, Frost N (2001) The role of the
avian hippocampus in orientation in space and time. Brain Res
919:292–301
Comings DE, Wu S, Rostamkhani M, McGue M, Lacono WG, Cheng
LS, MacMurray JP (2003) Role of the cholinergic muscarinic 2
receptor (CHRM2) gene in cognition. Mol Psychiatry 8:10–11
Cornil C, Foidart A, Minet A, Balthazart J (2000) Immunocytochem-
ical localization of ionotropic glutamate receptors subunits in the
adult quail forebrain. J Comp Neurol 428:577–608
Csillag A, Montagnese CM (2005) Thalamotelencephalic organiza-
tion in birds. Brain Res Bull 66:303–310
de Almeida J, Palacios JM, Mengod G (2008) Distribution of 5-HT
and DA receptors in primate prefrontal cortex: implications for
pathophysiology and treatment. Prog Brain Res 172:101–115
Diekamp B, Kalt T, Gu
¨ntu
¨rku
¨n O (2002a) Working memory neurons
in pigeons. J Neurosci 22:RC210
Diekamp B, Gagliardo A, Gu
¨ntu
¨rku
¨n O (2002b) Nonspatial and
subdivision-specific working memory deficits after selective
lesions of the avian prefrontal cortex. J Neurosci 22:9573–9580
Dietl MM, Palacios JM (1988) Neurotransmitter receptors in the
avian brain. I. Dopamine receptors. Brain Res 439:354–359
Dietl MM, Cortes R, Palacios JM (1988) Neurotransmitter receptors
in the avian brain. II. Muscarinic cholinergic receptors. Brain
Res 439:360–365
Diez-Alarcia R, Pilar-Cuellar F, Paniagua MA, Meana JJ, Fernandez-
Lopez A (2006) Pharmacological characterization and autora-
diographic distribution of alpha2-adrenoceptor antagonist
[3H]RX 821002 binding sites in the chicken brain. Neuroscience
141:357–369
Divac I, Mogensen J, Bjorklund A (1985) The prefrontal ‘cortex’ in
the pigeon. Biochemical evidence. Brain Res 332:365–368
Durstewitz D, Kro
¨ner S, Hemmings HC Jr, Gu
¨ntu
¨rku
¨n O (1998) The
dopaminergic innervation of the pigeon telencephalon: distribu-
tion of DARPP-32 and co-occurrence with glutamate decarbox-
ylase and tyrosine hydroxylase. Neuroscience 83:763–779
Eickhoff S, Amunts K, Mohlberg H, Zilles K (2006) The human
parietal operculum. II. Stereotaxic maps and correlation with
functional imaging results. Cereb Cortex 16:268–279
Emery NJ, Clayton NS (2004) The mentality of crows: convergent
evolution of intelligence in corvids and apes. Science
306:1903–1907
Farries MA (2001) The oscine song system considered in the context
of the avian brain: lessons learned from comparative neurobi-
ology. Brain Behav Evol 58:80–100
Fortune ES, Margoliash D (1995) Parallel pathways and convergence
onto HVc and adjacent neostriatum of adult zebra finches
(Taeniopygia guttata). J Comp Neurol 360:413–441
Gagliardo A, Divac I (1993) Effects of ablation of the presumed
equivalent of the mammalian prefrontal cortex on pigeon
homing. Behav Neurosci 107:280–288
Gagliardo A, Bonadonna F, Divac I (1996) Behavioural effects of
ablations of the presumed ‘prefrontal cortex’ or the corticoid in
pigeons. Behav Brain Res 78:155–162
Gagliardo A, Ioale
`P, Odetti F, Bingman VP, Siegel JJ, Vallortigara G
(2001) Hippocampus and homing in pigeons: left and right
hemispheric differences in navigational map learning. Eur J
Neurosci 13:1617–1624
Gebhard R, Zilles K, Schleicher A, Everitt BJ, Robbins TW, Divac I
(1995) Parcellation of the frontal cortex of the New World
monkey Callithrix jacchus by eight neurotransmitter-binding
sites. Anat Embryol (Berl) 191:509–517
Geyer S, Matelli M, Luppino G, Schleicher A, Jansen Y, Palomero-
Gallagher N, Zilles K (1998) Receptor autoradiographic
mapping of the mesial motor and premotor cortex of the
macaque monkey. J Comp Neurol 397:231–250
Goldman-Rakic PS (1999) The ‘psychic’’ neuron of the cerebral
cortex. Ann N Y Acad Sci 868:13–26
Goldman-Rakic PS, Lidow MS, Gallager DW (1990) Overlap of
dopaminergic, adrenergic, and serotoninergic receptors and
complementarity of their subtypes in primate prefrontal cortex.
J Neurosci 10:2125–2138
Gonzalez-Burgos G, Kro
¨ner S, Seamans JK (2007) Cellular mech-
anisms of working memory and its modulation by dopamine in
the prefrontal cortex of primates and rats. In: Tseng KY, Atzori
M (eds) Monoaminergic Modulation of Cortical Excitability.
Springer, Berlin, pp 125–152
Greenwood PM, Lin MK, Sundararajan R, Fryxell KJ, Parasuraman R
(2009) Synergistic effects of genetic variation in nicotinic and
muscarinic receptors on visual attention but not working
memory. Proc Natl Acad Sci USA 106:3633–3638
Gu
¨ntu
¨rku
¨n O (1997) Cognitive impairments after lesions of the
neostriatum caudolaterale and its thalamic afferent: functional
similarities to the mammalian prefrontal system? J Brain Res
38:133–143
Gu
¨ntu
¨rku
¨n O (2005a) Avian and mammalian ‘prefrontal cortices’’:
limited degrees of freedom in the evolution of the neural mecha-
nisms of goal-state maintenance. Brain Res Bull 66:311–316
Gu
¨ntu
¨rku
¨n O (2005b) The avian ‘prefrontal cortex’ and cognition.
Curr Opin Neurobiol 15:686–693
Gu
¨ntu
¨rku
¨n O, Kro
¨ner S (1999) A polysensory pathway to the
forebrain of the pigeon: the ascending projections of the nucleus
dorsolateralis posterior thalami (DLP). Eur J Morphol 37:185–
189
Hartmann B, Gu
¨ntu
¨rku
¨n O (1998) Selective deficits in reversal
learning after neostriatum caudolaterale lesions in pigeons:
possible behavioral equivalencies to the mammalian prefrontal
system. Behav Brain Res 96:125–133
Harvey PH, Krebs JR (1990) Comparing brains. Science 249:140–146
Hasselmo ME, Stern CE (2006) Mechanisms underlying working
memory for novel information. Trends Cogn Sci 10:487–493
Henley JM, Barnard EA (1990) Autoradiographic distribution of
binding sites for the non-NMDA receptor antagonist CNQX in
chick brain. Neurosci Lett 116:17–22
Herold C, Diekamp B, Gu
¨ntu
¨rku
¨n O (2008) Stimulation of dopamine
D1 receptors in the avian fronto-striatal system adjusts daily
cognitive fluctuations. Behav Brain Res 194:223–229
Horn G (1981) Neural mechanisms of learning: an analysis of
imprinting in the domestic chick. Proc R Soc Lond B Biol Sci
213:101–137
Horn G, Bradley P, McCabe BJ (1985) Changes in the structure of
synapses associated with learning. J Neurosci 5:3161–3168
Hunt GR, Gray RD (2003) Diversification and cumulative evolution
in New Caledonian crow tool manufacture. Proc Biol Sci
270:867–874
Jarvis ED, Gu
¨ntu
¨rku
¨n O, Bruce L, Csillag A, Karten H, Kuenzel W,
Medina L, Paxinos G, Perkel DJ, Shimizu T, Striedter G, Wild
JM, Ball GF, Dugas-Ford J, Durand SE, Hough GE, Husband S,
Kubikova L, Lee DW, Mello CV, Powers A, Siang C, Smulders
TV, Wada K, White SA, Yamamoto K, Yu J, Reiner A, Butler
AB (2005) Avian brains and a new understanding of vertebrate
brain evolution. Nat Rev Neurosci 6:151–159
Ji XH, Cao XH, Zhang CL, Feng ZJ, Zhang XH, Ma L, Li BM (2008)
Pre- and postsynaptic beta-adrenergic activation enhances
excitatory synaptic transmission in layer V/VI pyramidal neu-
rons of the medial prefrontal cortex of rats. Cereb Cortex
18:1506–1520
Kalenscher T, Diekamp B, Gu
¨ntu
¨rku
¨n O (2003) Neural architecture of
choice behaviour in a concurrent interval schedule. Eur J
Neurosci 18:2627–2637
252 Brain Struct Funct (2011) 216:239–254
123
Kalenscher T, Gu
¨ntu
¨rku
¨n O, Calabrese P, Gehlen W, Kalt T,
Diekamp B (2005) Neural correlates of a default response in a
delayed go/no-go task. J Exp Anal Behav 84:521–535
Karakuyu D, Herold C, Gu
¨ntu
¨rku
¨n O, Diekamp B (2007) Differential
increase of extracellular dopamine and serotonin in the
‘prefrontal cortex’ and striatum of pigeons during working
memory. Eur J Neurosci 26:2293–2302
Karten HJ (1969) The ascending auditory pathway in the pigeon
(Columba livia). II. Telencephalic projections of the nucleus
ovoidalis thalami. Brain Res 11:134–53
Karten H, Hodos W (1967) A stereotaxic atlas of the brain of the
pigeon (Columba livia). The Johns Hopkins University Press,
Baltimore
Kenward B, Weir AA, Rutz C, Kacelnik A (2005) Behavioural ecology:
tool manufacture by naive juvenile crows. Nature 433:121
Kirsch JA, Gu
¨ntu
¨rku
¨n O, Rose J (2008) Insight without cortex:
lessons from the avian brain. Conscious Cogn 17:475–483
Kohler EC, Riters LV, Chaves L, Bingman VP (1996) The muscarinic
acetylcholine antagonist scopolamine impairs short-distance
homing pigeon navigation. Physiol Behav 60:1057–1061
Korzeniewska E, Gu
¨ntu
¨rku
¨n O (1990) Sensory properties and
afferents of the N. dorsolateralis posterior thalami of the pigeon.
J Comp Neurol 292:457–479
Kro
¨ner S, Gu
¨ntu
¨rku
¨n O (1999) Afferent and efferent connections of
the caudolateral neostriatum in the pigeon (Columba livia): a
retro- and anterograde pathway tracing study. J Comp Neurol
407:228–260
Leutgeb S, Husband S, Riters LV, Shimizu T, Bingman VP (1996)
Telencephalic afferents to the caudolateral neostriatum of the
pigeon. Brain Res 730:173–181
Levy R, Goldman-Rakic PS (1999) Association of storage and
processing functions in the dorsolateral prefrontal cortex of the
nonhuman primate. J Neurosci 19:5149–5158
Lidow MS, Gallager DW, Rakic P, Goldman-Rakic PS (1989)
Regional differences in the distribution of muscarinic choliner-
gic receptors in the macaque cerebral cortex. J Comp Neurol
289:247–259
Lissek S, Gu
¨ntu
¨rku
¨n O (2005) Out of context: NMDA receptor
antagonism in the avian ‘prefrontal cortex’ impairs context
processing in a conditional discrimination task. Behav Neurosci
119:797–805
Martinez de la Torre M, Mitsacos A, Kouvelas ED, Zavitsanou K,
Balthazart J (1998) Pharmacological characterization, anatomi-
cal distribution and sex differences of the non-NMDA excitatory
amino acid receptors in the quail brain as identified by CNQX
binding. J Chem Neuroanat 15:187–200
McNab F, Klingberg T (2008) Prefrontal cortex and basal ganglia
control access to working memory. Nat Neurosci 11:103–107
Medina L, Reiner A (2000) Do birds possess homologues of
mammalian primary visual, somatosensory and motor cortices?
Trends Neurosci 23:1–12
Mehlhorn J, Hunt GR, Gray RD, Rehka
¨mper G, Gu
¨ntu
¨rku
¨n O (2010)
Tool-making new caledonian crows have large associative brain
areas. Brain Behav Evol 75:63–70
Merker B (1983) Silver staining of cell bodies by means of physical
development. J Neurosci Methods 9:235–241
Metzger M, Jiang S, Braun K (1998) Organization of the dorsocaudal
neostriatal complex: a retrograde and anterograde tracing study
in the domestic chick with special emphasis on pathways
relevant to imprinting. J Comp Neurol 395:380–404
Metzger M, Jiang S, Braun K (2002) A quantitative immuno-electron
microscopic study of dopamine terminals in forebrain regions of
the domestic chick involved in filial imprinting. Neuroscience
111:611–623
Mitsacos A, Dermon CR, Stassi K, Kouvelas ED (1990) Localization
of L-glutamate binding sites in chick brain by quantitative
autoradiography. Brain Res 513:348–352
Mogensen J, Divac I (1982) The prefrontal ‘cortex’ in the pigeon.
Behavioral evidence. Brain Behav Evol 21:60–66
Mrzljak L, Pappy M, Leranth C, Goldman-Rakic PS (1995) Cholin-
ergic synaptic circuitry in the macaque prefrontal cortex. J Comp
Neurol 357:603–617
Naito E, Scheperjans F, Eickhoff SB, Amunts K, Roland P, Zilles K,
Ehrsson HH (2008) Cytoarchitectonic areas in human superior
parietal lobule are functionally implicated by an illusion of
bimanual interaction with an external object. J Neurophysiol
99:695–703
Palomero-Gallagher N, Zilles K (2004) Isocortex. In: Paxinos G (ed)
The rat nervous system, 3rd edn edn. Acadamic Press, San
Diego, pp 729–757
Palomero-Gallagher N, Mohlberg H, Zilles K, Vogt B (2008)
Cytology and receptor architecture of human anterior cingulate
cortex. J Comp Neurol 508:906–926
Palomero-Gallagher N, Vogt B, Mayberg HS, Schleicher A, Zilles K
(2009) Receptor architecture of human cingulate cortex: insights
into the four-region neurobiological model. Hum Brain Mapp
30:2336–2355
Pinaud R, Mello CV (2007) GABA immunoreactivity in auditory and
song control brain areas of zebra finches. J Chem Neuroanat
34:1–21
Pollok B, Prior H, Gu
¨ntu
¨rku
¨n O (2000) Development of object
permancence in the food storing magpie (Pica pica). J Comp
Psychol 114:148–157
Prior H, Schwarz A, Gu
¨ntu
¨rku
¨n O (2008) Mirror-induced behavior in
the magpie (Pica pica): evidence of self-recognition. PLoS Biol
6:e202
Puelles L, Kuwana E, Puelles E, Bulfone A, Shimamura K, Keleher J,
Smiga S, Rubenstein JL (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 424:409–438
Rehka
¨mper G, Zilles K (1991) Parallel evolution in mammalian and
avian brains: comparative cytoarchitectonic and cytochemical
analysis. Cell Tissue Res 263:3–28
Reiner A, Perkel DJ, Bruce LL, Butler AB, Csillag A, Kuenzel W,
Medina L, Paxinos G, Shimizu T, Striedter G, Wild M, Ball GF,
Durand S, Gu
¨ntu
¨rku
¨n O, Lee DW, Mello CV, Powers A, White
SA, Hough G, Kubikova L, Smulders TV, Wada K, Dugas-Ford
J, Husband S, Yamamoto K, Yu J, Siang C, Jarvis ED (2004)
Revised nomenclature for avian telencephalon and some related
brainstem nuclei. J Comp Neurol 473:377–414
Richfield EK, Young AB, Penney JB (1989) Comparative distribu-
tions of dopamine D-1 and D-2 receptors in the cerebral cortex
of rats, cats, and monkeys. J Comp Neurol 286:409–426
Riters LV, Bingman VP (1999) The effects of lesions to the
caudolateral neostriatum on sun compass based spatial learning
in homing pigeons. Behav Brain Res 98:1–15
Riters LV, Erichsen JT, Krebs JR, Bingman VP (1999) Neurochem-
ical evidence for at least two regional subdivisions within the
homing pigeon (Columba livia) caudolateral neostriatum.
J Comp Neurol 412:469–487
Robbins TW, Arnsten AF (2009) The neuropsychopharmacology of
fronto-executive function: monoaminergic modulation. Annu
Rev Neurosci 32:267–287
Rose SP (2000) God’s organism? The chick as a model system for
memory studies. Learn Mem 7:1–17
Rose J, Colombo M (2005) Neural correlates of executive control in
the avian brain. PLoS Biol 3:e190
Brain Struct Funct (2011) 216:239–254 253
123
Rose J, Schiffer AM, Dittrich L, Gu
¨ntu
¨rku
¨n O (2010) The role of
dopamine in maintenance and distractability of attention in the
‘prefrontal cortex’ of pigeons. Neuroscience 167:232–237
Sakaue M, Somboonthum P, Nishihara B, Koyama Y, Hashimoto H,
Baba A, Matsuda T (2000) Postsynaptic 5-hydroxytryptamine
(1A) receptor activation increases in vivo dopamine release in rat
prefrontal cortex. Br J Pharmacol 129:1028–1034
Salvatierra NA, Torre RB, Arce A (1997) Learning and novelty
induced increase of central benzodiazepine receptors from chick
forebrain, in a food discrimination task. Brain Res 757:79–84
Santi A, Weise L (1995) The effects of scopolamine on memory for
time in rats and pigeons. Pharmacol Biochem Behav 51:271–277
Sarter M, Bruno JP (2000) Cortical cholinergic inputs mediating
arousal, attentional processing and dreaming: differential affer-
ent regulation of the basal forebrain by telencephalic and
brainstem afferents. Neuroscience 95:933–952
Sarter M, Parikh V, Howe WM (2009) nAChR agonist-induced
cognition enhancement: integration of cognitive and neuronal
mechanisms. Biochem Pharmacol 78:658–667
Schleicher A, Palomero-Gallagher N, Morosan P, Eickhoff SB,
Kowalski T, de Vos K, Amunts K, Zilles K (2005) Quantitative
architectural analysis: a new approach to cortical mapping. Anat
Embryol (Berl) 210:373–386
Schnabel R, Metzger M, Jiang S, Hemmings HC Jr, Greengard P,
Braun K (1997) Localization of dopamine D1 receptors and
dopaminoceptive neurons in the chick forebrain. J Comp Neurol
388:146–168
Seed AM, Tebbich S, Emery NJ, Clayton NS (2006) Investigating
physical cognition in rooks, Corvus frugilegus. Curr Biol
16:697–701
Sorenson EM, Chiappinelli VA (1992) Localization of 3H-nicotine,
125I-kappa-bungarotoxin, and 125I-alpha-bungarotoxin binding
to nicotinic sites in the chicken forebrain and midbrain. J Comp
Neurol 323:1–12
Stewart MG, Bourne RC, Chmielowska J, Kalman M, Csillag A,
Stanford D (1988) Quantitative autoradiographic analysis of the
distribution of [3H]muscimol binding to GABA receptors in
chick brain. Brain Res 456:387–391
Stewart MG, Cristol D, Philips R, Steele RJ, Stamatakis A, Harrison
E, Clayton N (1999) A quantitative autoradiographic comparison
of binding to glutamate receptor sub-types in hippocampus and
forebrain regions of a food-storing and a non-food-storing bird.
Behav Brain Res 98:89–94
Taylor AH, Hunt GR, Medina FS, Gray RD (2009) Do New
Caledonian crows solve physical problems through causal
reasoning? Proc Biol Sci 276:247–254
Uylings HBM, Sanz-Arigita E, de Vos K, Smeets WJAJ, Pool CW,
Amunts K, Rajkowska G, Zilles K (2000) The importance of a
human 3D database and atlas for studies of prefrontal and
thalamic functions. Progr Brain Res 126:357–368
Uylings HB, Groenewegen HJ, Kolb B (2003) Do rats have a
prefrontal cortex? Behav Brain Res 146:3–17
Van De Werd HJ, Rajkowska G, Evers P, Uylings HB (2010)
Cytoarchitectonic and chemoarchitectonic characterization of
the prefrontal cortical areas in the mouse. Brain Struct Funct
214:339–353
Van Eden CG, Lamme VA, Uylings HB (1992) Heterotopic cortical
afferents to the medial prefrontal cortex in the rat. A combined
retrograde and anterograde tracer study. Eur J Neurosci 4:77–97
Veenman CL, Albin RL, Richfield EK, Reiner A (1994) Distributions
of GABAA, GABAB, and benzodiazepine receptors in the
forebrain and midbrain of pigeons. J Comp Neurol 344:161–189
Veenman CL, Wild JM, Reiner A (1995) Organization of the avian
‘corticostriatal’ projection system: a retrograde and anterograde
pathway tracing study in pigeons. J Comp Neurol 354:87–126
Vogt BA, Pandya DN (1987) Cingulate cortex of the rhesus monkey:
II. Cortical afferents. J Comp Neurol 262:271–289
Waeber C, Dietl MM, Hoyer D, Palacios JM (1989) 5.HT1 receptors
in the vertebrate brain. Regional distribution examined by
autoradiography. Naunyn Schmiedebergs Arch Pharmacol
340:486–494
Waldmann C, Gu
¨ntu
¨rku
¨n O (1993) The dopaminergic innervation of
the pigeon caudolateral forebrain: immunocytochemical evi-
dence for a ‘prefrontal cortex’ in birds? Brain Res 600:225–234
Wang M, Ramos BP, Paspalas CD, Shu Y, Simen A, Duque A,
Vijayraghavan S, Brennan A, Dudley A, Nou E, Mazer JA,
McCormick DA, Arnsten AF (2007) Alpha2A-adrenoceptors
strengthen working memory networks by inhibiting cAMP-HCN
channel signaling in prefrontal cortex. Cell 129:397–410
Williams GV, Castner SA (2006) Under the curve: critical issues for
elucidating D1 receptor function in working memory. Neurosci-
ence 139:263–276
Wynne B, Gu
¨ntu
¨rku
¨n O (1995) Dopaminergic innervation of the
telencephalon of the pigeon (Columba livia): a study with
antibodies against tyrosine hydroxylase and dopamine. J Comp
Neurol 357:446–464
Yamamoto K, Reiner A (2005) Distribution of the limbic-system
associated membran protein (LAMP) in pigeon forebrain and
midbrain. J Comp Neurol 486:221–242
Yamasaki M, Matsui M, Watanabe M (2010) Preferential localization
of muscarinic M1 receptor on dendritic shaft and spine of
cortical pyramidal cells and its anatomical evidence for volume
transmission. J Neurosci 30:4408–4418
Zahrt J, Taylor JR, Mathew RG, Arnsten AF (1997) Supranormal
stimulation of D1 dopamine receptors in the rodent prefrontal
cortex impairs spatial working memory performance. J Neurosci
17:8528–8535
Zhang W, Yamada M, Gomeza J, Basile AS, Wess J (2002) Multiple
muscarinic acetylcholine receptor subtypes modulate striatal
dopamine release, as studied with M1–M5 muscarinic receptor
knock-out mice. J Neurosci 22:6347–6352
Zilles K (1985) The cortex of the rat, a stereotaxic atlas. Springer
Verlag, Berlin
Zilles K, Amunts K (2010) Centenary of Brodmann’s map—
conception and fate. Nat Rev Neurosci 11:139–145
Zilles K, Schleicher A, Palomero-Gallagher N, Amunts K (2002a)
Quantitative analysis of cyto- and receptor architecture of the
human brain. In: Mazziotta JC, Toga A (eds) Brain mapping: the
methods. Elsevier, Amsterdam, pp 573–602
Zilles K, Palomero-Gallagher N, Grefkes C, Scheperjans F, Boy C,
Amunts K, Schleicher A (2002b) Architectonics of the human
cerebral cortex and transmitter receptor fingerprints: reconciling
functional neuroanatomy and neurochemistry. Eur Neuropsy-
chopharmacol 12:587–599
254 Brain Struct Funct (2011) 216:239–254
123
... In altricial birds like zebra finches, a fundamental part of the early developmental conditions is determined by the parents, because nestlings completely rely on their parents for nutrition, warmth and protection. In adulthood, these offspring were sacrificed to measure GR expression in blood and five brain regions, including regions previously showed to be involved in HPA regulation in mammals (hypothalamus, amygdala, hippocampus, and nidopallium caudolateraleanalogous to the mammalian prefrontal cortex; [50,51]) and early life effects (amygdala, ventral striatumalso hippocampus -). We tested for the effects of developmental treatment on GR expression across tissues, to investigate whether harsher conditions during early life led to reduced GR expression in this species. ...
... Hemispheres were split sagittally with a razor blade. Regions were then isolated in a fixed manner, and for both hemispheres simultaneously, in the following order: 1) hippocampus (HP), 2) (ventral) striatum (VS), 3) hypothalamus (HYP), 4) nidopallium caudolaterale (NCL) -analogous to the mammalian prefrontal cortex [50,51] and 5) amygdala (AMY). The locations of these brains regions were based on the Zebra finch Expression Brain Atlas (ZEBrA) [55]. ...
Article
Full-text available
Early-life environment can affect organisms for life on many levels. The glucocorticoid receptor (GR) gene has a pivotal role mediating organismal physiological and behavioral responses to environmental change, and is sensitive to early-life environmental conditions and epigenetic programming. Longitudinal studies require non-lethal sampling of peripheral tissues (e.g. blood), but this approach is dependent on the extent to which GR expression in peripheral tissues covaries with GR expression in central tissues. To test for the long-term effects of early life adversity on GR expression across brain and peripheral tissues, we manipulated developmental conditions of captive zebra finches (n=45), rearing them in either benign or harsh conditions through manipulation of parental foraging costs. We measured relative GR mRNA expression in blood and five brain regions in adulthood: hippocampus, hypothalamus, amygdala, ventral striatum, and the nidopallium caudolaterale (analogous to the mammalian prefrontal cortex), using qPCR. We further tested whether GR expression was modulated by natal brood size (which affected growth), age at sampling, and sex. GR expression correlations between tissues varied widely in magnitude and direction, ranging from -0.27 to +0.80, indicating that our understanding of developmental effects on GR expression and associated phenotypes needs to be region specific rather than organism wide. A more consistent pattern was that GR expression increased with age in blood, ventral striatum and hippocampus; GR expression was independent of age in other tissues. Developmental treatment did not affect GR expression in any of the tissues measured directly, but in blood and ventral striatum of adult females we found a positive correlation between nestling mass and GR expression. Thus, GR expression in blood reflected early life conditions as reflected in growth in adult females, showing patterns in one brain tissue, but not ubiquitous across brain regions. These results point at sex-dependent physiological constraints during development, shaping early life effects on GR expression in females only. Further study is required to investigate whether these tissue-dependent effects more generally reflect tissue-dependent long-term effects of early life adversity. This, together with investigating the physiological consequences of GR expression levels on individual performance and coping abilities, will be fundamental towards understanding the mechanisms mediating long-term impacts of early life, and the extent to which these can be quantified through non-lethal sampling.
... This aligns with previous research that has found dim ALAN alters the neuroarchitecture of the nidopallium caudolateral, the avian equivalent of the prefrontal cortex (Gunturkun, 2005;Gunturkun and Bugnyar, 2016;Taufique et al., 2019). The nidopallium caudolateral has been implemented in mimicking prefrontal area structures by having the same receptor architecture as the Brodmann Area 10 in humans, which is involved in many processes including reward and conflict, working memory, and pain (Herold et al., 2011;Peng et al., 2018). IEG activation in areas associated with memory support previous findings that ALAN impairs learning and memory (Liu et al., 2022). ...
... Although not much is known about the avian caudal striatum, this area is related to anxiety and pain in mice (Jin et al., 2020). Additionally, the nidopallium caudolateral has been associated with the Brodmann Area 10 in humans, also involved in pain reception (Herold et al., 2011;Peng et al., 2018). ...
Article
Full-text available
Artificial light at night (ALAN) is a pervasive pollutant that alters physiology and behavior. However, the underlying mechanisms triggering these alterations are unknown, as previous work shows that dim levels of ALAN may have a masking effect, bypassing the central clock. Light stimulates neuronal activity in numerous brain regions which could in turn activate downstream effectors regulating physiological response. In the present study, taking advantage of immediate early gene (IEG) expression as a proxy for neuronal activity, we determined the brain regions activated in response to ALAN. We exposed zebra finches to dim ALAN (1.5 lux) and analyzed 24 regions throughout the brain. We found that the overall expression of two different IEGs, cFos and ZENK, in birds exposed to ALAN were significantly different from birds inactive at night. Additionally, we found that ALAN-exposed birds had significantly different IEG expression from birds inactive at night and active during the day in several brain areas associated with vision, movement, learning and memory, pain processing, and hormone regulation. These results give insight into the mechanistic pathways responding to ALAN that underlie downstream, well-documented behavioral and physiological changes.
... This means that birds are so far the only animal model for studying the development and processing of speech information in the brain, which has greatly stimulated research within the field of comparative neuroanatomy and pallial evolution (Brenowitz et al., 1997;Brainard & Doupe, 2002;Jarvis, 2004;Nottebohm, 2005;Jarvis, 2019). Further, after more than 365 million years of separate evolution birds have evolved a different pallial (neocortical) brain organisation compared to mammals but show similar connectivity between relevant brain areas, neurochemical features, neuron numbers and gene expression profiles of cells that are functionally related to cognition (Herold et al., 2011;Shanahan et al., 2013;Herold et al., 2014;Colquitt et al., 2021;Kverková et al., 2022;Ströckens et al., 2022). Such comparisons can yield basic insights into the links between brain structure and function and offer the unprecedented chance of gaining deep conceptual insights into fundamental brain functions. ...
Article
Full-text available
In recent years, brain research has indisputably entered a new epoch, driven by substantial methodological advances and digitally enabled data integration and modelling at multiple scales— from molecules to the whole brain. Major advances are emerging at the intersection of neuroscience with technology and computing. This new science of the brain combines high-quality research, data integration across multiple scales, a new culture of multidisciplinary large-scale collaboration and translation into applications. As pioneered in Europe’s Human Brain Project (HBP), a systematic approach will be essential for meeting the coming decade’s pressing medical and technological challenges. The aims of this paper are to: develop a concept for the coming decade of digital brain research, discuss this new concept with the research community at large, to identify points of convergence, and derive therefrom scientific common goals; provide a scientific framework for the current and future development of EBRAINS, a research infrastructure resulting from the HBP’s work; inform and engage stakeholders, funding organisations and research institutions regarding future digital brain research; identify and address the transformational potential of comprehensive brain models for artificial intelligence, including machine learning and deep learning; outline a collaborative approach that integrates reflection, dialogues and societal engagement on ethical and societal opportunities and challenges as part of future neuroscience research.
... The mesopallium is an associative forebrain region that does not receive direct sensory information from the thalamus and has been suggested to be correlated with behavioral innovation and flexibility 32 . A sub-region within the nidopallium, the NCL in particular, has been compared to the mammalian prefrontal cortex in controlling executive function, working memory, planning, flexible thinking, and attending objects of interest, particularly if the object is associated with a reward 29,30,33 . However, much of our observed nidopallial activity occurred more rostrally (Fig. 2, Supplementary Fig. 1 in SI) in areas associated with processing trigeminal, visual, and auditory sensory information 34 . ...
Article
Full-text available
Tools enable animals to exploit and command new resources. However, the neural circuits underpinning tool use and how neural activity varies with an animal’s tool proficiency, are only known for humans and some other primates. We use 18F-fluorodeoxyglucose positron emission tomography to image the brain activity of naïve vs trained American crows (Corvus brachyrhynchos) when presented with a task requiring the use of stone tools. As in humans, talent affects the neural circuits activated by crows as they prepare to execute the task. Naïve and less proficient crows use neural circuits associated with sensory- and higher-order processing centers (the mesopallium and nidopallium), while highly proficient individuals increase activity in circuits associated with motor learning and tactile control (hippocampus, tegmentum, nucleus basorostralis, and cerebellum). Greater proficiency is found primarily in adult female crows and may reflect their need to use more cognitively complex strategies, like tool use, to obtain food.
... Cholinergic M1 receptors were highest in humans if compared to macaque monkey, rhesus monkey, rat and pigeon, while M2 and nicotinic receptors showed equal densities [53]. However, pigeons showed an inverted pattern of M1/M2 binding in the NCL compared to other species which suggests an increased inhibitory control on local circuits, this may be a compensating mechanism for the shift to glutamatergic processing which was at highest concentration in the avian nidopallium [54]. ...
Chapter
Full-text available
Muscarinic acetylcholine receptors (mAChRs) are widely expressed in both the central nervous system and peripheral nervous system and play a crucial role in modulating cellular activity and function. While these receptors have been extensively studied in mammals, their presence and role in avian species remain a relatively unexplored area of research. Nonetheless, several studies have suggested the existence of multiple functional muscarinic receptors in various avian species, including the vestibular periphery of pigeons, retinal cells, intestinal smooth muscles, dorsal root ganglia, developing hearts in chickens, and avian salt glands. Despite this, only the M2-M5 subtypes have been characterized, except for some studies that suggest the existence of functional M1 receptors in avian species, such as in the dorsal root ganglia, retina, heart, and vestibular periphery. In this paper, we review the distribution of avian muscarinic receptor subtypes, the characterization of muscarinic acetylcholine receptors in various organs and organ systems, and the sequence similarity of mAChR 2 and mAChR 3 between various birds and animals. Given the current gaps in our understanding, more research is needed to investigate further the function and expression of mAChRs in avian species.
... This means that birds are so far the only animal model for studying the development and processing of speech information in the brain, which has greatly stimulated research within the field of comparative neuroanatomy and pallial evolution (Brenowitz et al., 1997;Brainard & Doupe, 2002;Jarvis, 2004;Nottebohm, 2005;Jarvis, 2019). Further, after more than 365 million years of separate evolution birds have evolved a different pallial (neocortical) brain organisation compared to mammals but show similar connectivity between relevant brain areas, neurochemical features, neuron numbers and gene expression profiles of cells that are functionally related to cognition (Herold et al., 2011;Shanahan et al., 2013;Herold et al., 2014;Colquitt et al., 2021;Kverková et al., 2022;Ströckens et al., 2022). Such comparisons can yield basic insights into the links between brain structure and function and offer the unprecedented chance of gaining deep conceptual insights into fundamental brain functions. ...
Preprint
Full-text available
Brain research has in recent years indisputably entered a new epoch, driven by substantial methodological advances and digitally enabled data integration and modeling at multiple scales – from molecules to the whole system. Major advances are emerging at the intersection of neuroscience with technology and computing. This new science of the brain integrates high-quality basic research, systematic data integration across multiple scales, a new culture of large-scale collaboration and translation into applications. A systematic approach, as pioneered in Europe’s Human Brain Project (HBP), will be essential in meeting the pressing medical and technological challenges of the coming decade. The aims of this paper are To develop a concept for the coming decade of digital brain research To discuss it with the research community at large, with the aim of identifying points of convergence and common goals To provide a scientific framework for current and future development of EBRAINS To inform and engage stakeholders, funding organizations and research institutions regarding future digital brain research To identify and address key ethical and societal issues While we do not claim that there is a ‘one size fits all’ approach to addressing these aspects, we are convinced that discussions around the theme of digital brain research will help drive progress in the broader field of neuroscience.
Article
Anesthetic-induced brain activity study is crucial in avian cognitive-, consciousness-, and sleep-related research. However, the neurobiological mechanisms underlying the generation of brain rhythms and specific connectivity of birds during anesthesia are poorly understood. Although different kinds of anesthetics can be used to induce an anesthesia state, a comparison study of these drugs focusing on the neural pattern evolution during anesthesia is lacking. Here, we recorded local field potentials (LFPs) using a multi-channel micro-electrode array inserted into the nidopallium caudolateral (NCL) of adult pigeons (Columba livia) anesthetized with chloral hydrate, pelltobarbitalum natricum or urethane. Power spectral density (PSD) and functional connectivity analyses were used to measure the dynamic temporal neural patterns in NCL during anesthesia. Neural decoding analysis was adopted to calculate the probability of the pigeon’s brain state and the kind of injected anesthetic. In the NCL during anesthesia, we found elevated power activity and functional connectivity at low-frequency bands and depressed power activity and connectivity at high-frequency bands. Decoding results based on the spectral and functional connectivity features indicated that the pigeon’s brain states during anesthesia and the injected anesthetics can be effectively decoded. These findings provide an important foundation for future investigations on how different anesthetics induce the generation of specific neural patterns.
Article
The involvement of the prefrontal cortex (PFC) in consciousness is an ongoing focus of intense investigation. An important question is whether representations of conscious contents and experiences in the PFC are confounded by post-perceptual processes related to cognitive functions. Here, I review recent findings suggesting that neuronal representations of consciously perceived contents—in the absence of post-perceptual processes—can indeed be observed in the PFC. Slower ongoing fluctuations in the electrophysiological state of the PFC seem to control the stability and updates of these prefrontal representations of conscious awareness. In addition to conscious perception, the PFC has been shown to play a critical role in controlling the levels of consciousness as observed during anesthesia, while prefrontal lesions can result in severe loss of perceptual awareness. Together, the convergence of these processes in the PFC suggests its integrative role in consciousness and highlights the complex nature of consciousness itself.
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
Auditory perception can be significantly disrupted by noise. To discriminate sounds from noise, auditory scene analysis (ASA) extracts the functionally relevant sounds from acoustic input. The zebra finch communicates in noisy environments. Neurons in their secondary auditory pallial cortex (caudomedial nidopallium, NCM) can encode song from background chorus, or scenes, and this capacity may aid behavioral ASA. Furthermore, song processing is modulated by the rapid synthesis of neuroestrogens when hearing conspecific song. To examine whether neuroestrogens support neural and behavioral ASA in both sexes, we retrodialyzed fadrozole (aromatase inhibitor, FAD) and recorded in vivo awake extracellular NCM responses to songs and scenes. We found that FAD affected neural encoding of songs by decreasing responsiveness and timing reliability in inhibitory (narrow-spiking), but not in excitatory (broad-spiking) neurons. Congruently, FAD decreased neural encoding of songs in scenes for both cell types, particularly in females. Behaviorally, we trained birds using operant conditioning and tested their ability to detect songs in scenes after administering FAD orally or injected bilaterally into NCM. Oral FAD increased response bias and decreased correct rejections in females, but not in males. FAD in NCM did not affect performance. Thus, FAD in the NCM impaired neuronal ASA but that did not lead to behavioral disruption suggesting the existence of resilience or compensatory responses. Moreover, impaired performance after systemic FAD suggests involvement of other aromatase-rich networks outside the auditory pathway in ASA. This work highlights how transient estrogen synthesis disruption can modulate higher-order processing in an animal model of vocal communication.
Article
Full-text available
Although previous research has emphasized the beneficial effects of dopamine (DA) on functions of the prefrontal cortex (PFC), recent studies of animals exposed to mild stress indicate that excessive DA receptor stimulation may be detrimental to the spatial working memory functions of the PFC (Arnsten and Goldman-Rakic, 1990; Murphy et al., 1994, 1996a,b, 1997). In particular, these studies have suggested that supranormal stimulation of D 1 receptors may contribute to the detrimental actions of DA in the PFC (Murphy et al., 1994, 1996a). The current study directly tested this hypothesis by examining the effects of infusing a full D 1 receptor agonist, SKF 81297, into the PFC of rats performing a spatial working memory task, delayed alternation. SKF 81297 produced a dose-related impairment in delayed-alternation performance. The impairment was reversed by pretreatment with a D 1 receptor antagonist, SCH 23390, consistent with drug actions at D 1 receptors. SCH 23390 by itself had no effect on performance, although slightly higher doses impaired performance (Murphy et al., 1994, 1996a). There was a significant relationship between infusion location and drug efficacy; animals with cannulae anterior to the PFC were not impaired by SKF 81297 infusions. Taken together, these results demonstrate that supranormal D 1 receptor stimulation in the PFC is sufficient to impair PFC working memory function. These cognitive data are consistent with recent electrophysiological studies of D 1 receptor mechanisms affecting the PFC (Williams and Goldman-Rakic, 1995; Yang and Seamans, 1996). Increased D 1 receptor stimulation during stress may serve to take the PFC “off-line” to allow posterior cortical and subcortical structures to regulate behavior, but may contribute to the vulnerability of the PFC in many neuropsychiatric disorders.
Article
Full-text available
A series of electron microscopic immunocytochemical studies was performed to analyze subcellular sites for noradrenergic modulation in monkey prefrontal cortex. One out of 12 noradrenergic varicosities, identified by dopamine β-hydroxylase immunocytochemistry within single ultrathin sections, forms morphologically identifiable junctions with small dendrites and spines. Accordingly, α2-adrenergic receptors, almost all of which are of the A-subtype, that occur in spines are localized discretely over postsynaptic membranes. α2-Adrenergic receptors are also found at sites along axons, dendritic shafts and astrocytic processes lacking morphologically identifiable synaptic junctions, suggesting that these receptors are activated by volume transmission. In particular, axonal α2-adrenergic receptors occur mostly at pre-terminal regions, suggesting that axo-axonic interactions may mediate reduction of neurotransmitter release at sites other than axo-spinous junctions by closing voltage-dependent calcium channels. These results indicate that noradrenergic modulation of prefrontal cortex involves synaptic interactions at spines of pyramidal neurons and nonsynaptic volume transmission to glia, dendritic shafts and axons.
Chapter
This chapter discusses the principles of classical architectonic mapping in the context of recent imaging techniques. It presents an observer-independent approach for a quantitative analysis of cortical areas and their borders, which is based on a multivariate statistical analysis of the cytoarchitecture and illustrates the application of this approach for the cytoarchitectonic mapping of the human visual cortex. Important criteria for cytoarchitectonic mapping are the absolute thickness of cortical layers, the proportionate thickness of a layer relative to the other cortical layers and to the total cortical depth, the presence of clearly recognizable laminar borders and vertical columns, the packing density and size of neuronal cell bodies, the homogeneous or clustered distribution of cell bodies throughout the layers, and the presence of special cell types such as Betz cells. Understanding the regional distributions of neurotransmitter receptors is likely to provide a crucial intermediary level of description between function and structure, since different cytoarchitectonic and functional areas have different mean receptor densities as well as distinct laminar distribution patterns. Another way for a better understanding of brain function and the underlying anatomy is to compare architectonic maps obtained in postmortem brains with activation maps obtained in functional imaging studies in a common spatial reference system. Since these two kinds of maps stem from different subsets of brains, such a comparison must be performed on a probabilistic basis.
Article
Visual evoked potentials (VEPs) in the associative neostriatum caudolaterale (NCL) have shorter latencies than those recorded in other visual forebrain areas. Therefore visual input into NCL probably stems from a subtelencephalic relay. Tracing experiments revealed a projection of the nucleus dorsolateralis posterior thalami (DLP) into those portions of NCL in which visual, auditory, and somatosensory afferents from intratelencephalic parasensory areas terminate. Since VEPs in NCL are abolished after DLP-lesions, this structure has to be the critical relay. However, DLP also projects to other associative forebrain areas and parts of the basal ganglia. Previous experiments had furthermore revealed that DLP-neurons integrate visual, auditory, and somatosensory inputs. Thus, the DLP-projection onto various associative forebrain areas represents a true polysensory thalamotelencephalic system.
Article
Abstract Newly hatched domestic chicks were either acoustically imprinted on 400 Hz tone pulses or visually imprinted on a rotating red light. Compared to naive control animals, both groups of imprinted chicks expressed significantly enhanced stimulus evoked 2-fluoro-2-deoxyglucose (2-FDG) uptake in circumscribed areas of the dorso-caudal neostriatum (Ndc). This enhanced excitability after imprinting seems not to be related to changes of NMDA-receptor densities as measured by quantitative receptor autoradiography. However, pharmacological blockade of NMDA-receptors in the dorso-caudal neostriatum leads to a marked suppression of stimulus-evoked 2-FDG uptake in the dorso-caudal neostriatum and also in the interconnected imprinting relevant forebrain area, medio-rostral neostriatum/hyperstriatum ventrale (MNH). Furthermore, chicks which received bilateral Ndc injections of the competitive NMDA antagonist DL-2-amino-5-phosphono valeric acid (APV) during the imprinting experiments showed a dose-dependent decrease of imprinting success compared to vehicle-injected controls. These results indicate that the dorso-caudal neostriatum may represent a polysensory associative brain region in which visual and acoustic features of imprinting objects may be integrated. The activation in this area evoked by the imprinting stimulus during and after imprinting is critically dependent on NMDA-receptor activation, which appears to be required for this learning process.
Article
Learning is a complex set of processes involving the acquisition and storage of information. Imprinting in the domestic chick was studied to analyse the neural basis of storage. The recently hatched chick learns the characteristics of a visually conspicuous object by being exposed to it. When a chick is trained in this way, biochemical changes can be detecte in the dorsal part of the forebrain. Through a series of experiments it was shown that these changes are unlikely to be non-specific consequences of training, but more probably reflect some aspect of the storage process. By using a radioautographic technique to localize the brain region more precisely, part of the hyperstriatum ventrale was implicated in this process. Bilateral destruction of the region before imprinting prevented acquisition, and bilateral destruction after imprinting impaired retention. After exposure for 140 min to an imprinting stimulus there was an increase in the area of contact between presynaptic and postsynaptic elements in the region. This effect was found on the left side only. Sequential lesions to left and right sides confirmed that there is a hemispheric asymmetry in the role of the region in the storage of information. The area receives input from the visual pathways and possibly from other sensory pathways, and projects to regions that are thought to be involved in the control of locomotor and viscero-endocrine functions. The results afford an opportunity for the further analysis both of storage and of the whole set of neural processes that underlie imprinting in the domestic chick.