Download as pdf or txt
Download as pdf or txt
You are on page 1of 162

Progress in Neurobiology 70 (2003) 83–244

The neurobiology and control of anxious states


Mark J. Millan∗
Psychopharmacology Department, Centre de Rescherches de Croissy, Institut de Recherches (IDR) Servier,
125 Chemin de Ronde, 78290 Croissy-sur-Seine, Paris, France
Received 13 March 2003; accepted 21 May 2003

Abstract
Fear is an adaptive component of the acute “stress” response to potentially-dangerous (external and internal) stimuli which threaten to
perturb homeostasis. However, when disproportional in intensity, chronic and/or irreversible, or not associated with any genuine risk, it
may be symptomatic of a debilitating anxious state: for example, social phobia, panic attacks or generalized anxiety disorder. In view of
the importance of guaranteeing an appropriate emotional response to aversive events, it is not surprising that a diversity of mechanisms are
involved in the induction and inhibition of anxious states. Apart from conventional neurotransmitters, such as monoamines, ␥-amino-butyric
acid (GABA) and glutamate, many other modulators have been implicated, including: adenosine, cannabinoids, numerous neuropeptides,
hormones, neurotrophins, cytokines and several cellular mediators. Accordingly, though benzodiazepines (which reinforce transmission at
GABAA receptors), serotonin (5-HT)1A receptor agonists and 5-HT reuptake inhibitors are currently the principle drugs employed in the
management of anxiety disorders, there is considerable scope for the development of alternative therapies. In addition to cellular, anatomical
and neurochemical strategies, behavioral models are indispensable for the characterization of anxious states and their modulation. Amongst
diverse paradigms, conflict procedures—in which subjects experience opposing impulses of desire and fear—are of especial conceptual
and therapeutic pertinence. For example, in the Vogel Conflict Test (VCT), the ability of drugs to release punishment-suppressed drinking
behavior is evaluated. In reviewing the neurobiology of anxious states, the present article focuses in particular upon: the multifarious and
complex roles of individual modulators, often as a function of the specific receptor type and neuronal substrate involved in their actions;
novel targets for the management of anxiety disorders; the influence of neurotransmitters and other agents upon performance in the VCT;
data acquired from complementary pharmacological and genetic strategies and, finally, several open questions likely to orientate future
experimental- and clinical-research. In view of the recent proliferation of mechanisms implicated in the pathogenesis, modulation and,
potentially, treatment of anxiety disorders, this is an opportune moment to survey their functional and pathophysiological significance, and
to assess their influence upon performance in the VCT and other models of potential anxiolytic properties.
© 2003 Elsevier Ltd. All rights reserved.

Contents

1. General introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
1.1. Aims and scope of review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
1.2. Organization of review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
1.3. Anxiety disorders and experimental models: general considerations . . . . . . . . . . . . . . . . . . . 88
1.4. Conflict models for the detection of anxiolytic (or anxiogenic) activity . . . . . . . . . . . . . . . 88
1.4.1. Basic principles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
1.4.2. Validity and clinical pertinence of the VCT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

Abbreviations: Ach, acetylcholine; ACPC, aminocyclopropylcarboxylic acid; ACTH, adrenocorticotropic hormone; AR, adrenoceptor; AMPA,
␣-amino-2,3-dihydro-5-methyl-3-oxo-4-isoxazolepropanoic acid; AR, adrenoceptor; BDNF, brain-derived neurotrophic factor; BZP, benzodiazepine; CART,
cocaine- and amphetamine-related transcript; CCK, cholecystokinin; CRF, corticotrophin releasing factor; DA, dopamine; DRN, dorsal raphe nucleus;
FCX, frontal cortex; GABA, ␥-amino-butyric acid; GAD, general anxiety disorder; GIRK, G-protein-coupled K+ -receptor; GR, glucocorticoid receptor;
5-HT, serotonin; LC, locus coeruleus; MAO, monoamine oxidase A; MCH, melanin concentrating hormone; mCPP, meta-chlorophenylpiperazine; MR,
mineralocorticoid receptor; MRN, median raphe nucleus; MSH, melanocyte-stimulating hormone; NA, noradrenaline; NGF, nerve growth factor; NK,
neurokinin; NMDA, N-methyl-d-aspartate; NO, nitric oxide; NPY, neuropeptide Y; OFQ, orphaninFQ; PACAP, pituitary adenylyl cyclase activating pep-
tide; PAG, periaqueductal gray; SP, substance P; SSRI, selective serotonin reuptake inhibitor; Trk, tropomyosin-related kinase; VCT, Vogel Conflict Test;
VDCC, voltage-dependent calcium channel; VTA, ventrotegmental area
∗ Tel.: +33-1-55-72-24-25; fax: +33-1-55-72-24-70.

E-mail address: mark.millan@fr.netgrs.com (M.J. Millan).

0301-0082/$ – see front matter © 2003 Elsevier Ltd. All rights reserved.
doi:10.1016/S0301-0082(03)00087-X
84 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

1.5. The integration and expression of anxious states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89


1.5.1. Adaptive and non-adaptive features of anxiety: basic pathological
mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
1.5.2. “Trait” and “state” anxiety . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
1.5.3. Cerebral circuits for the modulation of anxious states . . . . . . . . . . . . . . . . . . . . . . . . 91
2. GABA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
2.1. GABAergic pathways . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
2.2. GABAA receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
2.2.1. Structure, localization and modulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
2.2.2. Benzodiazepines and related ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
2.2.3. Neurosteroids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
2.2.4. Propofol and other injectable general anesthetics . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
2.2.5. Ethanol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
2.3. GABAB receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
2.3.1. Coupling and localization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
2.3.2. Anxiolytic properties of GABAB receptor agonists . . . . . . . . . . . . . . . . . . . . . . . . . 107
3. Excitatory amino acids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
3.1. Glutamatergic pathways . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
3.2. NMDA receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
3.2.1. Cooperative operation of NMDA recognition and glycineB sites . . . . . . . . . . . . . 108
3.2.2. Anxiolytic actions of NMDA recognition site antagonists and open
channel blockers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
3.2.3. GlycineB site ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
3.2.4. Interaction with BZPs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
3.2.5. Modulatory sites on NR2B subunits: ifenprodil . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
3.3. AMPA receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
3.4. Kainate receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
3.5. Metabotropic receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
3.5.1. Multiple classes of functionally-heterogeneous metabotropic receptor . . . . . . . . 113
3.5.2. Anxiolytic profiles of ligands at metabotropic mGluR receptors . . . . . . . . . . . . . 113
3.6. Multiple glutamatergic mechanisms for anxiolysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
4. Monoamines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
4.1. Noradrenaline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
4.1.1. Noradrenergic pathways . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
4.1.2. α2 -AR agonists. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
4.1.3. Significance of α2 -AR subtypes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
4.1.4. α2 -AR antagonists . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
4.1.5. Role of α1 -ARs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
4.1.6. Role of β-ARs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
4.2. Dopamine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
4.2.1. Dopaminergic pathways . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
4.2.2. Multiple classes of dopamine receptor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
4.2.3. Dopamine “D2 -like” receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
4.2.4. Dopamine “D1 -like” receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
4.3. Serotonin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
4.3.1. Serotonergic pathways . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
4.3.2. 5-HT1A receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
4.3.3. 5-HT1B receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
4.3.4. Serotonin 5-HT2 receptor subtypes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
4.3.5. 5-HT3 receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
4.3.6. Other classes of 5-HT receptor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
4.4. Antidepressant agents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
4.4.1. Clinical actions of antidepressant agents in the treatment of anxious states . . . 137
4.4.2. SSRIs and mixed 5-HT/NA reuptake inhibitors: long-term, adaptive actions . . 138
4.4.3. Selective NA reuptake inhibitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
4.4.4. Antidepressant agents possessing 5-HT2A/2C antagonist properties . . . . . . . . . . . 139
4.4.5. Monoamine oxidase inhibitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
4.5. Antipsychotic agents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
5. Histamine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
5.1. Coupling, localization and interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
5.2. Control of anxious states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 85

6. Acetylcholine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
6.1. Cholinergic pathways and their modulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
6.2. Nicotinic receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
6.2.1. Localization and interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
6.2.2. Influence upon anxious states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
6.3. Muscarinic receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
7. Adenosine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
7.1. Generation, localization and coupling to multiple receptor subtypes . . . . . . . . . . . . . . . . . 143
7.2. Adenosine A1 receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
7.3. Adenosine A2A receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
7.4. Purinoceptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
8. Cannabinoids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
8.1. Generation and coupling to CB1 receptors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
8.2. Influence upon anxious states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
9. Neuropeptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
9.1. Neuropeptides/receptors previously examined by use of the VCT . . . . . . . . . . . . . . . . . . . 146
9.1.1. Cholecystokinin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
9.1.2. Corticotropin releasing factor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
9.1.3. Vasopressin and oxytocin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
9.1.4. Substance P and other tachykinins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
9.1.5. Neuropeptide Y . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
9.1.6. Glucagon-like peptide-1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
9.1.7. Galanin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
9.1.8. Neurotensin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
9.1.9. Multiple opioid peptides and melanocortins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
9.1.10. Melanin Concentrating Hormone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
9.1.11. Angiotensin/AT receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
9.2. Peptides/receptors as yet to be evaluated in the VCT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
9.2.1. Vasoactive intestinal peptide and pituitary adenlyl cyclase activating peptide . . 158
9.2.2. Somatostatin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
9.2.3. Natriuretic peptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
9.2.4. Bombesin-related peptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160
10. Melatonin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160
10.1. Coupling and localization. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160
10.2. Control of anxious states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160
11. Glucocorticoids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
11.1. Genomic and non-genomic actions in the CNS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
11.2. Interactions with neurotransmitters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
11.3. Control of anxious states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
11.4. The complex role of glucocorticoids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
12. Estrogen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
12.1. Genomic and non-genomic actions in the CNS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
12.2. Control of anxious states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
12.3. Clinical pertinence: gender and anxiety . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
13. Ion channels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
13.1. Calcium channels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
13.1.1. Operation and localization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
13.1.2. Control of anxious states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
13.1.3. Actions of GABApentin and pregabalin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
13.2. Sodium channels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
13.3. Potassium-channels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
14. Nitric oxide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
15. Neurotrophins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
15.1. Actions of neurotrophins at “Trk” receptors: cerebral localization of BDNF . . . . . . . . 167
15.2. Role of BDNF in the control of anxious states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
15.3. BDNF/TrkB receptors as a therapeutic target . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
15.4. Other neurotrophins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
16. Cytokines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
17. Sigma1 and sigma2 sites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
17.1. Sigma1 sites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
17.2. Sigma2 sites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
17.3. Perspectives for sigma ligands as anxiolytic agents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
86 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

18. General discussion: open questions, conceptual issues and future research directions . . . . . . . 171
18.1. Multiple substrates of action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
18.2. Cellular mechanisms of action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
18.3. Long-term actions of anxiolytic agents. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
18.4. Detection of anxiogenic and anxiolytic properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
18.5. Comparing actions in the VCT to other models. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
18.6. Complementary pharmacological and genetic strategies . . . . . . . . . . . . . . . . . . . . . . . . . . 175
18.7. Neuronal networks and anxious states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
18.8. Anxious states and cognitive function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
18.9. Redundancy in the control of anxious states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
18.10. Clinical validation of hypotheses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
19. Concluding comments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178

1. General introduction growth (Stork et al., 2000a,b). However, in these (and sev-
eral other) cases, little other compelling evidence is available
1.1. Aims and scope of review to suggest that such mechanisms directly and specifically
control anxious states. Further, in general, only a limited
Pharmacological studies, clinical investigations and, in range of models has been exploited in their behavioral
recent years, analyses of genetically-modified mice have characterization, and the influence of genetic background
implicated a remarkable diversity of mechanisms in the and developmental changes has not been systematically ad-
etiology, modulation and treatment of anxiety (Griebel, dressed. Such mechanisms were considered insufficiently
1999a,b; Hood et al., 2000; Belzung and Griebel, 2001; well characterized for inclusion herein, and these studies are
Blanchard et al., 2001a,b; Lesch, 2001; Moret and Briley, authoritatively discussed elsewhere (Picciotto, 1999; Weiss
2001; Wood and Toth, 2001; Clément et al., 2002; Kent et al., 2000; Belzung and Griebel, 2001; Holmes, 2001;
et al., 2002a,b). In this light, the major purpose of the Stephens et al., 2002; Wood and Toth, 2001; Clément et al.,
present article is to review the roles of specific neuro- 2002; Rodgers et al., 2002a; Wolfer et al., 2002; Groenink
transmitters, their receptors and other modulators in the et al., 2003). The present review does not, thus, comprise
control of anxious states. As a framework for discussion, an all-inclusive inventory of every mechanism potentially
the following aspects are afforded particular emphasis. involved in the control and management of anxious states.
First, their localization and actions in corticolimbic struc- In the characterization of anxious states and their modu-
tures involved in the pathogenesis of anxiety disorders, lation, behavioral tests for evaluation of potential anxiolytic
together with their influence upon intracellular transduc- and anxiogenic properties are indispensable. Numerous ex-
tion systems controlling neuronal excitability. Second, their perimental paradigms involving natural and trained behav-
modulation by exposure to fear-inducing and other stress- iors (“unconditioned” and “conditioned” responses) have
ful stimuli. Third, their interaction with ␥-amino-butyric been described (summarized in Table 1). Of particular in-
acid (GABA)-containing, serotonergic and noradrenergic terest are procedures incorporating an element of “conflict”
pathways, all of which play pivotal roles in the response to whereby the subject experiences opposing and concomitant
stress and the control of mood. Fourth, where known, the tendencies of desire (for example, to obtain a reward) and
phenotype of genetically-manipulated mice in which the of fear (avoidance of a potentially aversive stimulus). One
functional status of specific neurotransmitters, receptors or instructive and well-established model is the Vogel Conflict
mediators has been disrupted. Finally, their relevance as tar- Test (VCT) in which water-deprived rodents are exposed to a
gets (established or innovative) for the clinical management mild and intermittent electric shock via a water bottle (Vogel
of anxiety disorders. et al., 1971; Treit, 1985; Pollard and Howard, 1989; Millan
In fact, studies of genetically-modified mice have (of- and Brocco, 2003). To the best of the author’s knowledge,
ten serendipitously) revealed potentially interesting al- no article has previously exploited this procedure as a vehi-
terations in anxious behavior upon interference with a cle for a comprehensive survey of neuronal mechanisms in-
plethora of different genes (see also Section 18.9). These volved in the control of anxious behavior. Thus, the present
include protein kinases and intracellular messengers (Chen article focuses in particular upon the actions of agents in the
et al., 1994; Rondi-Reig et al., 1997; Bowers et al., 2000; VCT. Indeed, in view of the enormous amount of ground
Weeber et al., 2000), modulators of G-protein signaling covered, it would be impossible to systematically present
(Oliveira-Dos-Santos et al., 2000), serine proteases (Pawlak detailed information from all behavioral procedures. Never-
et al., 2003) and regulators of development and neuronal theless, data from complementary models for the detection
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 87

Table 1 tection of anxiolytic properties, while Section 1.5 considers


Experimental models of anxiety widely used in rodents the adaptive significance of anxious states and outlines neu-
I. “Trait”, long-term anxious states ronal circuits implicated in their integration and expression.
(A) Rodent strains displaying high or low anxiety Section 2 focuses on the key inhibitory transmitter, GABA,
(B) Inter-individual differences within a defined strain
which exerts its actions via two major classes of ionotropic
(C) Chronic exposure to fear-provoking stimuli
(D) Genetic models: transgenic and knock-out mice (GABAA ) and metabotropic (GABAB ) receptors. GABAA
sites represent the molecular target of benzodiazepines
II. “State”, acute anxious states
(A) Unconditioned (BZP) such as diazepam (Valium® ) which are universally
(1) Exploration (avoidance, conflict) employed in the clinical management of anxious states. On
(i) Light–dark box (light chambers vs. dark chambers) the other hand, Section 3 focuses on the excitatory amino
(ii) Holeboard (nose pokes) acid, glutamate, which similarly acts via both ionotropic
(iii) Elevated plus-maze (open arms vs. closed arms)
and metabotropic receptors, yet which fulfils a radically
(iv) Open field (central squares vs. peripheral squares)
(v) Neophobia/emergence test (novel object) different role in the control of mood: drugs interacting with
glutamatergic mechanisms await therapeutic exploitation.
(2) Interaction based
(i) Active social interaction (unfamiliar rat pairs) Section 4 is devoted to the monoamines. Serotonergic and
(ii) Resident intruder noradrenergic mechanisms have long been privileged in the
(iii) Ultrasonic vocalization (separation induced) study and treatment (by antidepressant agents) of anxiety
(3) Acute response to aversive stimuli (environment or brain disorders. However, with the exception of 5-HT1A recep-
stimulation) tors, targeted by the agonist and anxiolytic agent, buspirone
(i) Freezing (Buspar® ), the clinical significance of individual subtypes
(ii) Ultrasonic vocalization
of 5-HT receptor—of which 14 have been cloned—remains
(iii) Startle
(iv) Autonomic-cardiovascular parameters (arterial pressure, heart nebulous. Likewise, the pathophysiological roles of specific
rate, endocrine secretion) subtypes of adrenoceptor (AR), of which nine are known,
(4) Defensive behavior to threatening stimuli are poorly characterized. Moreover, though often neglected,
(i) Fear/defence battery corticolimbic pools of dopamine, which acts via five sub-
(B) Conditioned types of receptor, are arguably of greater significance to
(1) Conflict procedures the induction and modulation of anxious states than gen-
(i) Geller–Seifter (operant, lever-pressing for reward) erally appreciated. Sections 5 and 6 consider two further,
(ii) Vogel Conflict Test “conventional” transmitters, histamine and acetylcholine,
(iii) Conditioned suppression (no punishment during test session)
for which studies of multiple receptor subtypes are begin-
(iv) Safety-signal withdrawal (no punishment during test session)
(v) Conditioned place aversion ning to provide insights into their complex roles in the
control of anxious states. Sections 7 and 8 deal with adeno-
(2) Non-conflict procedures
(i) Fear-induced freezing, startle and ultrasonic vocalizations sine and cannabinoids, respectively, both of which can be
(re-exposure to aversive environment) ubiquitously generated in the CNS, and both of which play
(ii) Shock-probe (burying of aversive object) major (but enigmatic) roles in the response to stress and
(3) Drug-discrimination fear. Section 9 discusses an array of neuropeptides vari-
(i) Anxiogenic agents ously implicated in the suppression and/or enhancement
Note that “conflict” and “non-conflict” might also be employed as a of anxious states. Though the significance of certain has
framework for classifying unconditioned models, e.g. exploration models not, as yet, been evaluated employing the VCT, they all
(conflict) vs. the acute response to aversive stimuli (non-conflict). Further, comprise attractive potential targets for the development of
actual exposure to aversive stimuli (usually punishment, as in the Vogel novel classes of anxiolytic agent. Section 10 focuses on
Conflict Test) as compared to potential exposure (as in unconditioned, ex-
ploratory models) would also offer an appropriate categorization scheme.
the pineal hormone, melatonin, while Sections 11 and 12,
Thus, no classification is definitive and all-inclusive (see text). Note also respectively, draw attention to the pertinence of genomic
that certain “state” models incorporate an element of chronicity inasmuch and non-genomic (cerebral) actions of the stress-sensitive
as they are repeatedly performed in the same population of subjects. For hormones, glucocorticoids and estrogen, in the modula-
example, the Geller–Seifter conflict test and drug-discrimination models. tion of anxious states. Section 13 outlines the relevance
“State” models are, of course, invariably utilized in the characterization
of anxious behavior in “trait” paradigms.
of voltage-dependent calcium—and, possibly, sodium and
potassium—channels as targets for the management of anx-
of anxiolytic (or anxiogenic) properties, in particular other iety disorders. Conversely, actions of the intra- and intercel-
conflict paradigms, are also carefully discussed. lular messenger, nitric oxide (NO) are evoked in Section 14.
Both rapid and longer-term actions of two further classes of
1.2. Organization of review modulator implicated in the pathogenesis and treatment of
anxiety disorders, neurotrophins and cytokines, are outlined
The review is structured as follows: in Sections 15 and 16, respectively. Section 17 focuses on
Sections 1.3 and 1.4 summarize general features of the the still perplexing roles of sigma1 and sigma2 sites. Fi-
VCT as compared to other procedures employed for the de- nally, Section 18 concludes with a discussion of several key
88 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

themes likely to orientate future research into mechanisms throughout this review, the genuine, clinical pertinence of
involved in the etiology and (improved) management of the VCT can only realistically be gauged within the light
anxious states. of controlled, therapeutic trials of mechanistically-diverse
The text is accompanied by tables summarizing the ac- agents—both active and inactive in this procedure—and it
tions of representative drugs in the VCT. In order to en- is salutary to reflect just how little clinical feedback is cur-
hance readability and to obviate encumbering the test with rently available in this regard.
superfluous detail, most specific information concerning the Table 1 offers a classification of models for the detection
actions of individual drugs and endogenous modulators is of anxiolytic properties, axed around the criteria of “con-
consigned to these tables. In addition, several figures are ditioned versus unconditioned” behaviors. The criteria of
provided in which the actions of certain key agents in the “punishment versus non-punishment” and of “conflict versus
VCT are illustrated, and in which their loci of action are non-conflict” provide alternative (and similarly valid) frame-
depicted. Though the accompanying citation list is by no works for categorizing such models. In any case, this table
means exhaustive, it is comprehensive. Other reviews cov- serves to emphasize that the VCT, while indubitably of im-
ering specific aspects of material discussed herein are gen- portance, is just one of many complementary models avail-
erously cited, though primary references are, in principle, able for the characterization of anxiolytic agents (Section
prioritized. Emphasis is also afforded to recent publications 18.5). Specific procedural features of the VCT (see in par-
incorporating novel findings not as yet considered or ref- ticular, Treit, 1985; Pollard and Howard, 1989; Millan and
erenced elsewhere. While the review attempts to minimize Brocco, 2003) as compared to other experimental paradigms
unnecessary duplication, individual sections are generally (Bignami, 1998; Thiébot et al., 1991; Treit, 1994; Rodgers,
self-sufficient in material in order to avoid incessant con- 1997; Rodgers et al., 1997; Clément and Chapoutier, 1998;
sultation of other parts. Numerous cross-references are also Belzung and Griebel, 2001; Blanchard et al., 2001a,b, 2003;
made to sections in which supplementary information con- Shekhar et al., 2001; Wall and Messier, 2001; Barros and
cerning specific issues under discussion can be found. Tomaz, 2002; Bourin and Hascoët, 2003; De Boer and
Koolhaas, 2003; File and Seth, 2003; Prut and Belzung,
1.3. Anxiety disorders and experimental models: 2003; Sanchez, 2003) employed for characterization of
general considerations anxiolytic drugs have been reviewed in detail elsewhere.
For individual models, these articles also provide additional
One major difficulty confronted in the experimental study information concerning the actions of drug interacting with
of anxious disorders is the absence of concrete parame- mechanisms discussed herein.
ters reflecting “anxiety” per se. Inasmuch as anxiety, like
other mood states, lies at the “interface” of neurobiology 1.4. Conflict models for the detection of anxiolytic (or
and psychology, the approach is essentially inferential. By anxiogenic) activity
analogy to depression and pain, and in contrast to hyper-
tension and diabetes, for example, no specific operational 1.4.1. Basic principles
measure of anxiety exists. “Surrogate” markers, such as an In conflict situations, the tendency of subjects to seek
increase in arterial pressure, tachycardia, hyperthermia and out a positive stimulus is countered by an opposite impulse
excessive secretion of glucocorticoids into the systemic cir- of avoidance founded upon either spontaneous (unlearned)
culation, can be instructive in the evaluation of anxiety. or conditioned (learned) fear. For example, in a procedure
However, these autonomic-somatic variables are generally in which a natural (exploratory) behavior is monitored, the
inappropriate—and impractical—for systematic characteri- plus-maze test (very much in vogue for the phenotyping of
zation of the role of cerebral mechanisms in the modula- transgenic mice), the urge to investigate an unfamiliar envi-
tion of anxious states. Accordingly, behavioral end-points ronment is opposed by the fear associated with its novelty
reflecting the emotive component of anxiety are generally (Rodgers and Johnson, 1995; Rodgers et al., 1997; Belzung
preferred: avoidance, escape, freezing and so forth—despite and Griebel, 2001). On the other hand, in the Geller and
the fact that “a motor withdrawal” response is clearly not a Seifter (1960) conflict test, rats trained to respond for a
feature of chronic, maladaptive, debilitating anxious states food reward are exposed to a modest electric shock dur-
in man. In behavioral paradigms employed for the evaluation ing a “conflict” component of the procedure: that is, the
of anxiolytic—and other classes of psychotropic—drugs, a response is both conditioned and punished. This original
motor response is, then, considered to express a mood state. conflict procedure remains of unquestionable value in the
In this respect, the VCT is no exception. evaluation of potential anxiolytic agents. Further, the inclu-
Partly for the above reasons, one must remain circum- sion of both punished and non-punished sessions facilitates
spect as concerns the exact “type” or “component” of anxi- the detection of specific anxiolytic properties of drugs as
ety which is measured with the VCT and other experimental compared to a generalized perturbation of motor function
models. Indeed, their precise relevance to the broad spec- resulting in an artifactual alteration in (punished) response
trum of anxiety disorders recognized in man is not easy to rates. However, the Geller–Seifter model requires a lengthy
decipher (Section 1.4.2). Correspondingly, as emphasized and daily period of training, and one drug may potentially
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 89

interfere with the actions of a second drug tested at a later afford a definitive response. Nevertheless, the VCT is of par-
date. ticular pertinence to generalized anxiety disorders (GAD)
(Barrett and Vanover, 1993; Shekhar et al., 2001) which are
1.4.2. Validity and clinical pertinence of the VCT characterized by excessive, persistent and unfounded anxi-
By contrast, the VCT, in which water-deprived rats re- ety and an exaggerated response to acute, anxiogenic situ-
ceive a mild shock from the spout of a test bottle, has ations. GAD is an under-recognized yet incapacitating and
the advantage of circumventing a prolonged training period chronic disorder which often reveals co-morbidity with other
and of avoiding potential drug interactions. Inasmuch as the psychiatric and neurological conditions (Culpepper, 2002;
VCT does not incorporate a non-punished session, caution Grachev et al., 2002; Keller, 2002; Kessler and Wittchen,
must be exercised in uncoupling the anxiolytic actions of 2002; Sramek et al., 2002; Wittchen et al., 2002). Arguably,
drugs from their (potential) influence upon motor perfor- positive actions of drugs in the VCT and other conflict
mance: their possible modification of nociceptive thresholds paradigms founded on conditioned fear are also relevant to
and water appetite must also be borne in mind (Treit, 1985; the therapy of social (and other) phobias, which similarly
Pollard and Howard, 1989; Millan and Brocco, 2003). As encompass a component of conditioned fear (Hidalgo et al.,
regards the cognitive nature of the VCT, responding has 2001). Their significance to other anxious states, such as
been variously considered as conditioned or unconditioned. panic attacks and obsessive–compulsive disorders, is less
This somewhat disconcerting interpretational ambivalence apparent (Gorman et al., 1989, 2000; Graeff and Deakin,
appears (as recently discussed in Millan and Brocco, 2003) 1991; Barrett and Vanover, 1993; Bourin, 1997; Bourin et al.,
to reflect the precise experimental design. Though animals 1998; Blanchard et al., 2001a,b; Shekhar et al., 2001).
are pre-exposed to the bottle in the absence of punishment Though the VCT has been widely employed in rats,
(that is, drinking behavior is learned), responding is punished the procedure has only recently been established in mice
only during the (non-signaled) test session. That is, there is (Umezu, 1999; Van Gaalen and Steckler, 2000). This is
no learned association of punishment with a response. an innovation to be welcomed inasmuch as transgenic and
The VCT has significant “face validity” inasmuch as the knock-out mice should ideally be phenotyped with a vari-
avoidance of an actual or threatened aversive stimulus, de- ety of models reflecting different facets of clinical anxiety,
spite a potential reward, clearly mimics human behavior— incorporating contrasting end-points, and reflecting both
though, as pointed out above (Section 1.3), avoidance is conditioned and unconditioned behaviors (Tarantino and
not necessarily an inherent feature of pathological, anx- Bucan, 2000; Weiss et al., 2000; Belzung and Griebel, 2001;
ious states. As concerns “construct validity”, psychologi- Wood and Toth, 2001; Clément et al., 2002; Van Gaalen
cal and biological processes underlying anxiety in the VCT et al., 2002).
likely resemble those in man, though this is more chal- To summarize, then, the VCT is a particularly important,
lenging to demonstrate. Indeed, more generally, the extent rigorous and clinically-relevant model for pharmacological
to which experimental models employed for the detection and genetic studies of the role of endogenous modulators and
of anxiolytic drugs mimic (heterogeneous classes of) clin- potential therapeutic agents in the control of anxious states.
ical anxiety disorder(s) remains a major question in need
of clarification. As intimated above (Section 1.3), the fun- 1.5. The integration and expression of anxious states
damental issue of predictive validity remains to be more
fully explored. Though the VCT is, for example, responsive 1.5.1. Adaptive and non-adaptive features of anxiety:
to BZPs (therapeutically-active) and refractory to cholecys- basic pathological mechanisms
tokinin (CCK)2 antagonists (essentially ineffective), insuffi- Fear and anxiety are crucial and adaptive components of
cient clinical information is available from other drug classes the overall behavioral and autonomic “stress” response to
to permit an unambiguous statement concerning either false dangerous situations which threaten to perturb homeosta-
positives or false negatives in the VCT. In this respect, of sis (Rosen and Schulkin, 1998; Haller, 2001; Mineka and
course, the VCT is no different from the Geller–Seifter pro- Öhman, 2002). Transient anxiety proportional to the chal-
cedure, the plus-maze test or any other paradigm currently lenge encountered elicits an appropriate response (often es-
employed for the detection of anxiolytic agents: sufficient cape or avoidance) and is of fundamental importance as a
clinical evidence is simply not yet available (Section 18.10). survival strategy for all higher animals, including man. Anx-
Moreover, for certain classes of drug (or individual agents), ious states are, thus, controlled by a highly complex sys-
experimental data from the VCT and other procedures re- tem of both inhibitory and facilitatory mechanisms: these
main limited, rendering generalized conclusions concern- are reciprocally regulated in a highly-dynamic fashion and
ing their pathophysiological roles and therapeutic pertinence display considerable redundancy. Indeed, a major message
somewhat hazardous. of this article (see also Section 18.9) is the existence of
These uncertainties raise the interrelated question as to a hierarchy of interactive processes which either counter
which type of clinical anxiety is most effectively modeled by or favor anxious states. The purpose of these homeostatic
the VCT. Likewise, in the absence of feedback from human controls is to: (1) maintain an appropriate degree of emo-
patients presenting diverse anxiety disorders, it is difficult to tionality (avoid pervasive anxiety) under non-threatening
90 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

circumstances; (2) efficiently respond to potential threats Scherrer et al., 2000; Gratacos et al., 2001; Hettema et al.,
with a (transient) “fear” proportional to the danger encoun- 2001; Holmes, 2001; Stein et al., 2001a,b; Clément et al.,
tered; (3) permit adaptive behavioral responses, such as 2002; Freimer et al., 2002; Glatt and Freimer, 2002).
escape or avoidance; and (4) rapidly restore a “normal” emo-
tional status once the threat has passed. The aberrant oper- 1.5.2. “Trait” and “state” anxiety
ation of mechanisms controlling mood and the response to Certain authors distinguish personality-based (tempera-
stress—due to genetic, developmental and/or environmen- ment), sustained, “trait” anxiety to ephemeral, fear-induced
tal factors (Heim and Nemeroff, 1999; Holmes, 2001)— “state” anxiety (see Table 1). Trait and state anxiety may,
can provoke a perturbation of equilibrium, an abrupt or in principle, be engendered by mechanisms outlined in
gradual shift to a new “stasis” (“set-point”) and, in patholog- the preceding paragraph (Section 1.5.1) (Stewart et al.,
ical cases, a clinically-defined anxiety disorder. Correspond- 1993; Abadie et al., 1999; Kalin et al., 2001; Radu et al.,
ingly, if the anxiety experienced is excessive with respect to 2003). The VCT—like most other widely-used paradigms—
the trigger, persists following its withdrawal, is elicited in the conforms to “state” anxiety (Herman et al., 1986; Rägo
absence of potential threats, is associated with non-adaptive et al., 1988; Molewijk et al., 1995a,b; Telch et al., 1996;
behaviors, or is simply omnipresent for no apparent reason, Liebsch et al., 1998; Davis and Shi, 1999; Belzung and
then an “anxiety disorder” is generally diagnosed and treat- Griebel, 2001; Keogh and Cochrane, 2002; Ohl et al.,
ment advocated. 2002). Experimentally, “trait” anxiety may be modeled
Transient anxiety (fear) may, then, be compared to the (Table 1) by: (1) genetically-modified lines of mice (see
acute pain provoked by noxious (tissue-damaging) or po- Section 1.1 for references to reviews); (2) chronic stress ex-
tentially noxious stimuli. However, by analogy to extreme posure (citations throughout this article); (3) highly-anxious
or chronic pain, when disproportional or prolonged (even strains of mice and rats (Stewart et al., 1993; Tejani-Butt
irreversible), anxiety is maladaptive and, generally, re- et al., 1994; Kiaanman et al., 1995; Steimer et al., 1997;
flects an underlying pathology requiring treatment (Millan, Berton et al., 1998; Desousa et al., 1998; Dockstader and
1999, 2002a). Pursuing the analogy to pain (Millan, 1999), Van Der Kooy, 2001; Sudakov et al., 2001; Bouwknecht
excessive anxiety may reflect the exaggerated (in terms of and Paylor, 2002; Pardon et al., 2002; Rodgers et al.,
duration or intensity) activation of mechanisms normally 2002a,b; Avgustinovich et al., 2003); and (4) inherent dif-
fulfilling a physiological role in the induction of fear. How- ferences in anxious behavior between individual animals
ever, this is not necessarily the case and, by analogy to of a common strain (Colombo et al., 1985; Cools et al.,
differences between chronic versus acute pain, mechanisms 1993; Rodgers, 1997; Ho et al., 2002; Wunderlich et al.,
underlying persistent, pathological anxious states may be 2002).
qualitatively as well as quantitatively distinct to those In fact, considerable efforts have been valuably invested
involved in transitory (adaptive) anxious states. Several (op. cit.) in the characterization of strains displaying differ-
conceptual examples of pathological mechanisms underly- ential responsiveness to stress and contrasting patterns of
ing anxious states may be provided. First, anxiety may be anxious behavior. Nevertheless, it is somewhat disappoint-
evoked by the failure of an endogenous anxiolytic mecha- ing that individual intra-strain differences have received so
nism which either: (1) operates “constitutively” (tonically, in little attention inasmuch as neither animal nor human popu-
the resting state); and/or (2) is recruited following cessation lations are homogeneous in terms of “basal” levels of anxi-
of a fear-eliciting stimulus in order to restore equilibrium. ety and the response to anxiolytic agents. Indeed, individual
In the former instance, an enduring anxious state may arise “disparities” between animals are of obvious heuristic per-
spontaneously and, in the latter, anxiety may be anoma- tinence to the striking differences seen amongst (“healthy”)
lously protracted and severe relative to the transient threat humans as concerns the response to stress and anxious be-
encountered. Second, a transmitter which normally exerts havior. Even bearing in mind the more complex cognitive
no—or only a subliminal, positive—influence upon anxious and emotional dimension of anxious states in man, appro-
behavior may become a powerful inducer of anxious states priate analyses of inter-subject variability in experimental
(“break-through”). This may occur, for example, following studies should generate insights into those factors which
the initiation of adaptive changes in the amygdala, hip- predispose to, or otherwise modify, anxious behavior in
pocampus or other corticolimbic structures upon chronic man. Regrettably, then, for the majority of investigators,
exposure to fear-inducing situations or other stressors. This heterogeneity within experimental groups is considered
comprises yet another analogy to “plastic” events underly- as—at best—noise and, more seriously, as a profound
ing pathological pain which permit normally sub-threshold nuisance which can frustrate the inviolate objective of
(innocuous) inputs to become intensely aversive (Millan, statistically-significant inter-group comparisons. Few re-
1999). The influence of environmental events (stressors) ports in which variability is underlined have, therefore,
upon mechanisms inducing anxious states is discussed penetrated the literature and such enterprising studies, re-
throughout this article, while the significance of develop- quiring unusual analytical strategies, are to be encouraged.
mental and genetic factors—which lie outside its scope— Furthermore, of particular interest are studies of the in-
has been considered elsewhere (Heim and Nemeroff, 1999; fluence of drugs upon performance in the VCT and other
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 91

models of “state” anxiety undertaken in “high-anxious” as mitted to the integration of its emotional (and cognitive)
compared to “low-anxious” individuals or strains. components.
In this respect, numerous interconnected limbic and corti-
1.5.3. Cerebral circuits for the modulation of anxious cal structures have been implicated in the emotional and/or
states mnesic mechanisms which underlie and modulate anxious
states. A comprehensive discussion of their roles is beyond
1.5.3.1. Corticolimbic integration of anxious states. It is the scope of the present paper (Fig. 1) (for reviews, see Gray,
not unreasonabe to contend that the entire organism is in- 1987; Gorman et al., 1989; Reiman et al., 1989; Deakin
volved in the response to (and modulation of) stress and fear and Graeff, 1991; Fanselow, 1991; Davis, 1992a,b;
inasmuch as anxious states impact upon, and are modified Aggleton, 1993; Charney and Deutch, 1996; Risold and
by, virtually all major systems: motor, sensory, endocrine, Swanson, 1997; Davidson et al., 1999; Pikkarainen et al.,
immune, cardiovascular and, of course, neural. Further, it 1999; Groenewegen and Uylings, 2000; Lovick, 2000;
is important to acknowledge the therapeutic significance Grachev et al., 2002; Phan et al., 2002; File and Seth, 2003).
of interventions which palliate non-emotional symptoms For a historical perspective on this issue, which can only
known to provoke, aggravate or otherwise (reciprocally) be briefly chronicled in the following paragraph, articles by
interact with anxious states, such as chronic pain, cardio- Pratt (1992) and LeDoux (2000), are recommended.
vascular disease and any variety of somatic complaints Broca (1878) originally coined the term “limbic” for
(Millan, 1999; Di Piero et al., 2001; Keogh and Cochrane, a series of phylogenetically-conserved structures, but
2002; Rainville, 2002). Nevertheless, in the experimental the first conscious and enduring attempt to provide a
characterization of anxiety, its induction and alleviation, neuroanatomically-circumscribed, physiological founda-
we are generally concerned with cerebral circuits com- tion for the expression of emotion and fear was made by

Fig. 1. Schematic representation of the organization of structures involved in the integration and induction of anxious states: in particular conditioned
fear, an important characteristic of the Vogel Conflict Test. For simplicity, not all anatomical pathways are indicated: for example, those between various
regions of the cortex are not fully illustrated, and several other pathways of ascending nociceptive information are not shown. Note the central importance
of the amygdala which can be subdivided into essentially two sub-territories. First, the basolateral complex, the major region for direct (thalamus and
parabrachial nucleus (PBN)) and indirect (cortex) receipt of sensory and other modes of input. The basolateral complex displays reciprocal projections
with many structures from which it receives afferents. Second, the central nucleus, the principle region for projections to hindbrain regions coordinating
the behavioral, autonomic and endocrine response to fear, a role to which—according to the concept of an “extended amygdala”—the adjacent bed
nucleus of the stria terminals (not shown) also contributes. Other abbreviations are as follows (more or less clockwise). PF/F: prefrontal/frontal; ENT:
entorhinal; CX: cortex; MB (SMB): mammillary bodies (supramammillary bodies); CING: cingulate; ASSN: association; THAL: thalamus; INS: insular;
SS: somatosensory; STT: spinothalamic tract; DMV/NAMB—dorsal motor nucleus of the vagus/nucleus ambiguous; HR: heart rate; AP: arterial pressure;
LAT HYPOTH: lateral hypothalamus; PVN: paraventricular nucleus; ACTH: adrenocorticotropic hormone; NRPC: nucleus reticularis penduncocellularis;
PAG: periaqueductal gray; FMN: facial motor nucleus; VTA: ventrotegmental area; LC: locus coeruleus; DLTN: dorsolateral segmental nucleus; DA:
dopamine; NA: noradrenaline; ACh: acetylcholine and NACC: nucleus accumbens. This figure is to a large extent inspired by the work of Davis et al.
(1994) and LeDoux (2000). For further details, see text.
92 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Papez, 1937. His “Circle of Papez”’ embraces a projec- primarily derived from the central nucleus and the contigu-
tion (the fornix) from the hippocampus to the mammillary ous bed nucleus of the stria terminals, whereas the basolat-
bodies which are linked in turn to the anterior nucleus eral amygdaloid complex is principally responsible for the
of the thalamus: the anterior thalamus is connected to receipt and filtering of cortical and subcortical (mostly tha-
the cingulate cortex which closes the circuit in transmit- lamic) sensory input (Fig. 1) (Gray and Magnuson, 1992;
ting information back to the hippocampus. This nascent Pitkänen et al., 1997; Davis and Shi, 1999; Richter-Levin
limbic system matured in the work of MacLean (1949, and Akirav, 2000; Kozicz, 2002; Saldivar-Gonzalez et al.,
1952), who rectified the original omission of the amygdala 2003; Walker et al., 2003).
and elaborated a network encompassing the frontal cortex Synaptic plasticity in the amygdala has been convinc-
(FCX) and other subcortical nuclei, such as the septum ingly implicated in the induction, processing and extinction
and nucleus accumbens (Barili et al., 1998; Pralong et al., of conditioned fear, in the generation of anticipatory anxiety
2002). The septum was subsequently incorporated into a and in the coordination of the global response to threats.
septo-hippocampal “behavioral inhibition system” subject Such phenomena are highly relevant to behavior in the VCT
to modulation by cortical input (Gray, 1987; Gray and and several other experimental models in rodents (Table 1),
McNaughton, 2000). Amongst several hypothalamic nu- as well as to GAD (and phobias) in man (Deakin and
clei, the paraventricular nucleus deserves special mention Graeff, 1991; Davis, 1992a,b; Graeff et al., 1993, 2001;
in view of its critical role in integrating the adrenocortical Davis et al., 1994; Fanselow, 1994; Davis and Shi, 1999;
response to stress (Yadin et al., 1993; Herman et al., 2002; LeDoux, 2000; Blair et al., 2001; Kim et al., 2001;
Carrasco and Van de Kar, 2003). In addition to the mam- Boshuisen et al., 2002; McGaugh et al., 2002). Both
millary bodies themselves, the overlying supramammillary the acquisition and extinction of conditioned fear repre-
nucleus, which is interlocked with the septo-hippocampal sent forms of “learning” implying alterations in synaptic
system and amygdala, is likewise implicated in the control strength. Moreover, extinction (the progressive dissipation
of emotion—including fear-conditioning—and cognition of the conditioned response upon repetitive exposure to the
(Vertes, 1992; Kirk, 1998; Wirtshafter et al., 1998; Pan and initially-neutral, non-aversive conditioned stimulus), does
McNaughton, 2002). Moreover, together with the periaque- not merely reflect spontaneous recovery or passive decay
ductal gray (PAG), amygdala and medial hypothalamus, an of events underlying conditioned fear, but rather active
important role has been ascribed to the inferior colliculus inhibition by superimposition of novel adaptive processes
in the assimilation of aversive input and the coordination of integrated at different synapses (LeDoux, 2000; Blair et al.,
anxious and defensive behaviors: the latter structure trig- 2001; Myers and Davis, 2002; Vianna et al., 2003). In the
gers (conditioned and unconditioned) responses to auditory development of innovative anxiolytic agents it may, thus,
stimuli via its projections to the amygdala (Graeff, 1990; be possible to specifically target and strengthen processes
Brandao et al., 1994; Maisonnette et al., 2000; Macedo underlying the extinction of conditioned fear expressed in
et al., 2002; Brandao et al., 2003). Finally, virtually all the amygdala and other corticolimbic structures.
cortical regions (insular, orbital, entorhinal, temporal, as- Lesion and stimulation studies have convincingly im-
sociation, frontal, pre-frontal, cingulate and parietal) have plicated the amygdala in the control of behavior in the
been shown to play significant, yet contrasting, roles in VCT and other conflict models (Hodges et al., 1987; Yadin
the response to—and control of—fear and stress (Hurley et al., 1991; Kopchia et al., 1992; Möller et al., 1994, 1997;
et al., 1991; Pratt, 1992; Bechara et al., 2000; Groenewegen Shibata et al., 1989; Yamashita et al., 1999)—though these
and Uylings, 2000; Grachev et al., 2002; Pralong et al., approaches have also unveiled roles for other structures,
2002). such as the mammillary bodies, the septum, the frontal
cortex and raphe nuclei (Vogel et al., 1971; Thiébot et al.,
1.5.3.2. Key roles of the amygdala, the hippocampus and 1983; Yamashita et al., 1989a,b; Pratt, 1992; Yadin et al.,
the PAG. Over recent years, a burgeoning literature has fo- 1993; Wieronska et al., 2001).
cussed on the amygdala as an epicenter of events modulating Curiously, comparable findings with the VCT do not ap-
the response to fear in animals and in man, and as a major pear to have been acquired for the hippocampus which bidi-
site of action for anxiolytic agents (Fig. 1) (Pitkänen et al., rectionally communicates with the amygdala and numerous
1997; Swanson and Petrovich, 1998; Davis and Shi, 1999; other corticolimbic structures. It has been proposed that the
Van De Kar and Blair, 1999; LeDoux, 2000; Mineka and hippocampus is specifically required for representation of
Öhman, 2002; Weidenfeld et al., 2002). Indeed, the amyg- the context within which conditioning occurs: that is, the en-
dala possesses an extensive pattern of reciprocal connec- vironment in which the conditioned and the unconditioned
tions with cortical, limbic, monoaminergic and other struc- stimulus are paired, but this concept remains somewhat
tures implicated in the emotional, cognitive, autonomic and controversial (Kim and Fanselow, 1992; Fanselow, 2000;
endocrine response to stress (Gray and Magnuson, 1992; Cahill and McGaugh, 1998; Gewirtz et al., 2000; LeDoux,
Swanson and Petrovich, 1998; LeDoux, 2000; Richter-Levin 2000; Richter-Levin and Akirav, 2000; Burgess et al., 2002;
and Akirav, 2000; Kjelstrup et al., 2002; McGaugh et al., Grillon, 2002; Kjelstrup et al., 2002; Mineka and Öhman,
2002; Carrasco and Van de Kar, 2003). Output pathways are 2002; Pralong et al., 2002).
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 93

A major component of cerebral circuits involved in co- attributed to any single locus is highly misleading, if not
ordinating the defensive and aversive response to fear and frankly erroneous. In analogy to other higher functions,
stress, the PAG, also justifies brief comment. In distinc- such as pain and memory (see Damasio, 1989; Crick and
tion to the amygdala (with which it is interlinked), several Koch, 1998; Millan, 1999, 2002a; Bechara et al., 2000;
authors have implicated the PAG in elaboration of the Manns et al., 2003), extended and diffuse networks rather
stereotyped, reflexive, autonomic and behavioral fight/flight than any well-defined entity, embody the experience of
response to unconditioned fear, a function related to clin- anxiety and other emotions (Section 18.7).
ical panic attacks. Correspondingly, stimulation of the Adding a further level of complexity, many neurotrans-
PAG is aversive both in animals and in human subjects mitters and receptors fulfil multiple roles in the modulation
(Deakin, 1988, 1991; Deakin and Graeff, 1991; Pratt, 1992; of anxious states, acting in contrasting or similar fashions as
Graeff et al., 1993, 2001; Lovick, 2000; Bandler et al., 2000; a function of the precise cerebral circuits with which they
Blanchard et al., 2001a,b). As discussed in Section 4.3.1, interact, and the precise timing and conditions of their en-
these differential roles of the amygdala and the PAG have gagement (see ensuing sections).
been evoked in an attempt to reconcile a superficially The above points have important ramifications both for
contradictory body of data indicating that 5-HT exerts a studies of mechanisms underlying the induction and con-
complex and receptor-dependent inhibitory and facilitatory trol of anxious states, and for the assignment of functions
influence upon specific anxious states. to specific brain structures, neurotransmitters and receptors.
Thus, numerous interlinked and interacting corticolimbic Indeed, it would be naı̈ve to impose a unitary function to any
centers orchestrate the overall response to fear and stress. As of the structures, neurotransmitters or receptors discussed
indicated above for the amygdala, these regions are hetero- in this review. This assertion has (as further discussed in
geneous in that they incorporate functionally-distinct nuclei Section 18) evident implications for therapeutic strategies
(Graeff, 1990; Pratt, 1992; Pitkanen et al., 1997; Fanselow, designed to improve the management of anxiety disorders.
2000; Gray and McNaughton, 2000; LeDoux, 2000; Lovick,
2000; Levita et al., 2002). However, for the sake of sim-
plicity, and since studies aimed at the localization of drug 2. GABA
actions have often not differentiated specific sub-territories,
the present article usually adopts the “generic” term. 2.1. GABAergic pathways
Finally, in light of the crucial role of corticolimbic no-
radrenergic, serotonergic and dopaminergic pathways in the GABAergic neurones constitute the major mode of in-
expression of anxious states, the significance of their cell hibitory transmission throughout the CNS. Corticolimbic
clusters of origin, the locus coeruleus (LC), the raphe nuclei structures involved in the modulation of anxious states,
(of which the dorsal raphe nucleus (DRN) is embedded in such as the hippocampus, lateral septum, PAG and amyg-
the ventral PAG), and the ventrotegmental area (VTA), re- dala, contain major networks of GABAergic interneurones
spectively, should also be emphasized (Sections 4.1.1, 4.3.1 as well as, in certain cases, GABAergic projection neu-
and 4.2.1, respectively). rones (Scheel-Krüger and Petersen, 1982; Sanger, 1985;
Shephard, 1986, 1987; Cherubini and Conti, 2001; Mody,
1.5.3.3. Diffuse neuronal circuits and anxious states. The 2001).
preceding paragraphs briefly evoked the key roles of several GABAergic pathways exert an inhibitory influence upon
corticolimbic structures in mechanisms integrating the re- the release of many neurotransmitters known to mediate
sponse to external (and internal) fear-inducing stimuli, and anxiogenic actions (see below). Their suppressive influence
in the induction and modulation of anxious states. Amongst (expressed at both the cell body and terminal level) upon
these, the amygdala has emerged as a nexus of primordial corticolimbic noradrenergic and serotonergic projections,
importance in much the same way as the dorsal horn of the of which a hyperactivity is implicated in the induction of
spinal cord and the mediobasal hypothalamus fulfil pivotal anxious states, should be accentuated in the present context
roles in the assimilation and processing of information con- (Sections 4.1.1 and 4.3.1, respectively). Accordingly, condi-
trolling nociception and endocrine secretion, respectively tioned fear is accompanied by an activation of GABAergic
(Millan, 1999, 2002a; Hermann et al., 2002; Carrasco and interneurones in both noradrenergic (LC) and serotonergic
Van de Kar, 2003). (raphe nuclei) cell clusters, consistent with the notion that
Nonetheless, in referring to the amygdala (Section their activity is subject to a GABAergic “brake” under condi-
1.5.3.2) and other—ostensibly—discrete structures men- tions of stress (Wang et al., 1992; Gobert et al., 2000; Millan
tioned above, it should not be forgotten that they exten- et al., 2000d; Somogyi and Llewellyn-Smith, 2001; Ishida
sively and reciprocally interact via numerous classes of et al., 2002). Nevertheless, the degree to which a reduction
neuronal pathway and transmitter. Thus, as encapsulated of noradrenergic and serotonergic transmission participates
by the historical concept that a “Circle of Papez” underlies in the anxiolytic properties of GABAergic agents—and the
the expression of emotion (Section 1.5.3.1), the notion that extent to which an acceleration of NA and 5-HT release
the generation, inhibition or experience of anxiety can be mediates the anxiety precipitated by BZP withdrawal—is
94 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

still disputed (Scheel-Krüger and Petersen, 1982; Sanger, 2.2. GABAA receptors
1985; Shephard, 1986, 1987; Pan and Williams, 1989;
File et al., 1992; Plaznik et al., 1994a; Tao et al., 1996; 2.2.1. Structure, localization and modulation
Dekeyne et al., 2000b; Millan et al., 2001a; Mohler et al., Pentameric GABAA receptors are enriched throughout
2002; Shekhar et al., 2002). Indeed, actions of GABAergic corticolimbic structures and are well represented in raphe
neurones in the hippocampus and other limbic regions nuclei, the LC and the VTA, the nuclei of origin of ascending
postsynaptic to noradrenergic and serotonergic neurones serotonergic, noradrenergic and mesocortical/mesolimbic
are likewise implicated in the expression of their anxiolytic dopaminergic pathways, respectively. Accordingly, GABAA
properties (Oleskewich and Lacaille, 1992; Söderpalm et al., receptors contribute to the inhibitory influence of GABA
1997; Nazar et al., 1999b). Further, GABAergic neurones upon serotonergic, noradrenergic and dopaminergic neu-
are inhibitory to the stress-induced release of dopamine rones, actions exerted both at the level of their perikarya
(DA), glutamate, CCK, corticotrophin-releasing factor and upon their terminals in the cortex, hippocampus, amyg-
(CRF) and several other anxiogenic mediators discussed dala and other limbic regions (Davis et al., 1994; Fritschy
in the present article (Owens et al., 1989, 1991; Karreman and Möhler, 1995; Tao et al., 1996; Liu and Glowa, 1999;
and Moghaddam, 1996; Keim and Shekhar, 1996; Beaufour Gobert et al., 2000; Millan et al., 2000d, 2001a; Pirker et al.,
et al., 2001a; Miklös and Kovacs, 2002). The global con- 2000; Celada et al., 2001; Shekhar et al., 2002). Of partic-
trol of anxious states by GABAergic mechanisms entails ular note in the present context is the suppressive influence
thus: (1) modulation of monoaminergic transmission; (2) of GABAA receptors upon hypothalamic and corticolimbic
interactions with monoaminergic receptors postsynaptic to secretion of CRF (Section 9.1.2) (Cullinan, 2000; Skelton
monoaminergic projections; and (3) actions independent of et al., 2000).
monoaminergic pathways (Sections 2.2.2.2 and 2.2.2.3). Consistent with a role in the control of anxious states,
There is curiously little functional evidence for a pertur- levels of GABAA receptors and their subunits are modified
bation in GABAergic transmission in patients with anxiety by exposure to stress (Biggio et al., 1984; Giovannini et al.,
disorders (Section 2.2.1) (Goddard et al., 2001, 2002). Nev- 2001). Alterations in the binding properties and density
ertheless, in line with neurochemical studies indicating that of GABAA receptors—principally, as monitored by stud-
interruption of GABAergic transmission can trigger anxiety ies with radioligands of BZP sites (Section 2.2.2)—have,
(Martijena et al., 2002; Stork et al., 2002a), mice deficient further, been specifically associated with: (1) experimental
in glutamate decarboxylase, the principle synthetic route to models of anxiety; (2) rat strains displaying differential
GABA in the CNS (Millan, 2002b), show enhanced anxiety sensitivity to stress; and (3) anxiety disorders in man (Rägo
(Kash et al., 1999; Stork et al., 2000a, 2003). On the con- et al., 1988; Harro et al., 1990; Kang et al., 1991; Kaschka
trary, anxiolytic actions of drugs which interfere with neu- et al., 1995; Tiihonen et al., 1997b; Abadie et al., 1999;
ronal GABA uptake have been demonstrated in the VCT and Malizia, 2002). Long-term administration of BZPs them-
other conflict models (Table 2) (Ågmo et al., 1991; Gadea selves leads to changes in corticolimbic populations of
and Lopez-Colomé, 2001; Schmitt et al., 2002). Further, GABAA receptors which are related to the progressive de-
though anxiolytic properties of inhibitors of GABA transam- velopment of tolerance and dependence to their actions.
inase (a rate-limiting enzyme of GABA transformation— Similarly, abrupt discontinuation of treatment with BZPs or
ultimately—into glutamate (Millan, 2002b)) such as vi- other GABAergic modulators provokes marked alterations
gabatrin have received scant attention, they have likewise in the relative densities of GABAA receptor subunits (Zhao
been documented in the VCT and several other procedures et al., 1994; Impagnatiello et al., 1996; Fahey et al., 1999;
(Liljequist and Engel, 1984a,b; Ågmo et al., 1991; Sherif Liu and Glowa, 1999; Nutt and Malizia, 2001; Tanay et al.,
and Ahmad, 1995; Sherif and Oreland, 1995; Shors and 2001; Malizia, 2002; Mascia et al., 2002).
Oreland, 1995; Dalvi and Rodgers, 1996). Recently, in GABAA receptors are assembled from several fami-
small-scale clinical trials, vigabatrin was found to display lies of subunit, of which at least 16 occur in the CNS
efficacy in the treatment of panic disorders (Zwanzger et al., (Sieghart, 1995; Barnard et al., 1998; Rudolph et al., 1999;
2001a,b). Whiting et al., 1999; Kittler et al., 2002; Möhler et al.,
The inhibitory influence of GABA upon neuronal activ- 2002). Though their stoichiometry remains under discus-
ity is expressed via chloride-permeable, ionotropic GABAA sion, the majority of central sites appear to be “ternary”
receptors and via metabotropic GABAB receptors, both of associations of two ␣-subunits, two ␤-subunits and a single
which, as outlined below (Sections 2.2 and 2.3, respectively), (intervening) ␥-subunit, surrounding a central ion channel.
are implicated in the control of anxious states (Barnard et al., As concerns ␥2 sites, the ␥2S isoform predominates over
1998; Bowery and Enna, 2000). Despite their occurrence in its—less BZP-responsive—␥2L homologue in the limbic
the hippocampus and other limbic structures, heteromeric, system and cortex. Individual ␣-subunits, which confer
chloride-permeable “GABAC ” receptors are not discussed specific functional properties to GABAA receptors, are dif-
further herein as information specifically pertaining to a role ferentially distributed in the CNS. Notably ␣1 -, ␣2 - and
in the modulation of anxious states has not been forwarded ␣3 -subtypes are concentrated in the cortex, hippocampus
(Zhang et al., 2001a; Möhler et al., 2002). and various limbic structures, ␣4 - and ␣5 -subunits are
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 95

Table 2
Influence of drugs modulating GABA availability, and of GABAA receptor ligands, upon behavior in the Vogel Conflict Test
Drug Class Species Route Effect Active dose Maximum Reference
(strain) (structure) range change (%)
␥-Acetylene- GABA transaminase Rat (W) i.p. + 50 mg/kg +115 Ågmo et al. (1991)
GABA inhibitor
Valproate GABA transaminase Rat (W) i.p. IA 200–400 mg/kg IA Ågmo et al. (1991)
inhibitor Rat (SD) i.p. + 400 mg/kg +340 Liljequist and Engel
(1984a,b)
SKF 100330 GABA reuptake Rat (W) i.p. + 12 mg/kg +86 Ågmo et al. (1991)
inhibitor
Muscimol GABAA agonist Rat (SD) i.c. (lateral + 5 ng +170 Drugan et al. (1986)
septum)
Rat (LH) i.c. (DRN) + 5 and 10 ng +212 Higgins et al. (1988)
Rat (SD) i.c.v. + 150–300 ng +620 Canzani et al. (1980)
THIP GABAA agonist Rat (SD) i.c.v. + 2000 ng +340 Canzani et al. (1980)

Picrotoxin GABAA Ant Rat (W) i.p. − 2 mg/kg −95 Ågmo et al. (1991)
Isoniazide GABAA Ant Rat (W) i.p. − 200 mg/kg −87 Ågmo et al. (1991)
Drug actions are expressed as a percentage of control values (defined as 0%). Thus, +100%, for example, represents a two-fold increase in
punished responses; (+) significant increase in responses; (−) significant decrease in responses and IA, no significant change (inactive). THIP:
4,5,6,7-tetrahydroisoxyzolo [5,4-c]-pyridin-3-ol; W: Wistar; SD: Sprague–Dawley; LH: Lister Hooded; DRN: dorsal raphe nucleus; Ant: antagonist.

largely restricted to the hippocampus, and the ␣6 -subunit is postsynaptic to serotonergic pathways (Drugan et al., 1986;
confined to the cerebellum (Wisden et al., 1992; Marksitzer Higgins et al., 1988; Dalvi and Rodgers, 1996) in the control
et al., 1993; Gutierrez et al., 1994; Fritschy and Möhler, of anxious states. Further, upon intraventricular injection,
1995; Nusser et al., 1996; Bernard et al., 1998; Fritschy muscimol was also found to elicit anxiolytic actions in the
et al., 1998; Pirker et al., 2000; Nyiri et al., 2001; Korpi VCT, actions reproduced by its analogue, THIP (Cananzi
et al., 2002; Möhler et al., 2001, 2002). et al., 1980). In distinction to agonists, selective antagonists
Activation of GABAA receptors by GABA (which acts at GABAA receptors manifest anxiogenic properties in the
at the ␣/␤-subunit interface) results in neuronal hyperpo- VCT and related models (Getova et al., 1990; Ågmo et al.,
larization via the opening of chloride-permeable ion chan- 1991; Thiébot et al., 1991).
nels. This action is modulated by a remarkable diversity
of allosteric sites differentially located on various subunits 2.2.2. Benzodiazepines and related ligands
(MacDonald and Olsen, 1994; Hevers and Lüddens, 1998;
Meadows et al., 1998; Rudolph et al., 1999; Whiting et al., 2.2.2.1. Influence upon anxious states: diverse classes of
1999; Kittler et al., 2002; Korpi et al., 2002; Stell and Mody, ligand. In distinction to GABA, for which the recognition
2002). Of these, sites recognized by BZPs and related agents site is located at the junction between ␣- and ␤-subunits,
(Rudolph et al., 1999; Sigel et al., 2001; Möhler et al., BZPs act at the interface between ␣ and ␥2 -subunits of
2002), by neurosteroids (Hevers and Lüddens, 1998; GABA receptors. While inactive alone, they potentiate the
Falkenstein et al., 2000), by barbiturates (Birnir et al., potency, duration and, at certain synapses, amplitude of the
1997), by a further injectable anesthetic, propofol (Davis actions of GABA (MacDonald and Olsen, 1994; Günther
et al., 1997; Krasowski et al., 1998), and by ethanol (Kittler et al., 1995; Sieghart, 1995; Hajos et al., 2000b; Walters
et al., 2002; Korpi et al., 2002; Miczek et al., 2002) are et al., 2000; Stell and Mody, 2002).
of particular pertinence herein since they have been shown General features of the anxiolytic profiles of BZPs (Fig. 2)
to: (1) transduce the actions of drugs eliciting anxiolytic do not necessitate detailed consideration herein inasmuch as
actions in the VCT and other experimental models; and (2) they are well-established both clinically and experimentally
ameliorate anxious states in man. (Sanger, 1985; Shephard, 1986, 1987; Pratt, 1992; Kasper
Surprisingly, comparatively few data are available con- and Resinger, 2001; Blanchard et al., 2003; File and Seth,
cerning the anxiolytic and anxiogenic actions of GABAA 2003; Prut and Belzung, 2003). Indeed, as for most other
receptor agonists and antagonists, respectively, in the VCT models deemed predictive of therapeutic activity, the demon-
(Table 2). Nevertheless, anxiolytic properties of the proto- stration of responsiveness to BZPs was an ineluctable step
typical agonist, muscimol, upon introduction either into the in the initial characterization of the VCT (Vogel et al., 1971;
lateral septum or into the DRN support a role of GABAA Pollard and Howard, 1989). Thus, rather than a detailed
sites both presynaptic (Tao et al., 1996; Celada et al., 2001; discussion, the reader is referred to Table 3 which com-
Wirtshafter and Sheppard, 2001; Adell et al., 2002) and piles positive findings with a cohort of familiar (and several
96 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

750
LICKS/3 MIN

500

250

0
VEH VEH 10.0 15.0 20.0 MG/KG, I.P.
CLORAZEPATE

NO SHOCK SHOCK

750
LICKS/3 MIN

500

250

0
VEH VEH 0.16 0.63 1.25 2.5 MG/KG, I.P.
TRIAZOLAM

NO SHOCK SHOCK

Fig. 2. Actions of the benzodiazepine, clorazepate, and of the triazalo-benzodiazepine, triazolam, in the Vogel Conflict Test. Data are means ± S.E.M.s.
Star indicates significance of differences to corresponding vehicle values. ( ) P < 0.05. Data for clorazepate are from Millan et al. (1997) and for
triazolam from Brocco and Millan (unpublished observations).

lesser-known) agents. For example, the conventional (and and the non-subtype-selective partial agonists, imidazenil,
␣-subunit, non-selective) BZPs, diazepam, chlordiazepoxide bretazenil and pagoclone (Giusti et al., 1993; Griebel et al.,
and clorazepate (Barrett and Gleeson, 1991; Griebel et al., 1999a,c; Sorbera et al., 2001; Dubinsky et al., 2002).
1999a; Dekeyne et al., 2000b; Flores and Pellón, 2000); In a fashion opposite to these positive modulators, neg-
the chemically-distinct and likewise, non-␣ subunit selec- ative modulators (or inverse agonists) at the BZP site on
tive triazolo-BZPs, alprazolam and triazolam (Söderpalm GABAA receptors compromise operation of the associated
et al., 1989; Giusti et al., 1993; Millan et al., 1997, 2001a; ion channel and, correspondingly, express anxiogenic prop-
Nazar et al., 1997; Rowlett et al., 2001; Sanna et al., 2002; erties (Dorow et al., 1983; Thiébot et al., 1988, 1991; Gentil
Brocco, unpublished observations); the cyclopyrrolone, zo- et al., 1989; Cole et al., 1995a,b; Sanger and Cohen, 1995;
plicone (Ueki et al., 1987; Benavides et al., 1990; Carlson Paronis et al., 2001; Sarter et al., 2001; Bourin and Hascoët,
et al., 2001), the preferential (full) ␣1 -subunit agonists and 2003; Prut and Belzung, 2003). Neutral antagonists (devoid
hypnotics, zolpidem (an imidazopyridine) and zaleplon (a of intrinsic activity), such as flumazenil, interfere with the
pyrazolopyrimidine) (Depoortere et al., 1986; Hadingham actions of both positive and negative modulators but are
et al., 1993; Kleven and Koek, 1999; Sanger, 1995; Sanger bereft of anxiolytic or anxiogenic activity in experimental
et al., 1996; Rush, 1998; Damgen and Lüddens, 1999; models (Corda et al., 1983; Liljequist and Engel, 1984a,b;
Paconis et al., 2001; McLeod et al., 2002); abecarnil, a Mizoule et al., 1985; Paconis et al., 2001; Wieronska et al.,
preferential (full) agonist at ␣1 and ␣3 as compared to 2001; Dubinsky et al., 2002; Bourin and Hascoët, 2003; Prut
␣2 -subunits (Stephens et al., 1990; Griebel et al., 1999a,c) and Belzung, 2003). On the other hand, robust panicogenic
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 97

Table 3
Influence of benzodiazepines and related ligands upon behavior in the Vogel Conflict Test
Drug Class Species Route (structure) Effect Active dose Maximum Reference
(strain) range change (%)
Diazepam BZP agonist Rat (W) i.p. + 10 mg/kg +186 Dekeyne et al. (2000a,b)
Rat (SD) i.p. + 0.25–4.0 mg/kg +200 Söderpalm and Engel (1989)
Rat (LE) i.p. + 2–10 mg/kg +125 Carlson et al. (2001)
Rat (H) i.p. + 2–4 mg/kg +500 Vogel et al. (1980)
Rat (W) i.c. (amygdala) + 20 ␮g +255 Shibata et al. (1989)
Rat (SD) i.c. (hippocampus) + 40 ␮g/side +330 Kataoka et al. (1991)
Mouse s.c. + 1 mg/kg +133 Umezu et al. (1999)
Chlordiazepoxide BZP agonist Rat (W) i.p. + 10–20 mg/kg +370 Millan et al. (2001a)
Rat (H) i.p. + 6–8 mg/kg +667 Vogel et al. (1980)
Rat (SD) i.p. + 4–27 mg/kg +214 Vogel et al. (1980)
Lormetazepam BZP agonist Rat (W) i.c. (amygdala) + 1 ␮g +380 Shibata et al. (1989)
Flurazepam BZP agonist Rat (SD) i.p. + 4–27 mg/kg +214 Vogel et al. (1980)
Rat (H) i.p. + 6–8 mg/kg +667 Vogel et al. (1980)
Rat (W) i.c. (amygdala) + 40–80 ␮g +190 Shibata et al. (1989)
Rat (LH) i.c. (amygdala) + 5 ␮g +154 Higgins et al. (1991)
Alprazolam Triazolo-BZP Rat (SD) i.p. + 0.5–16 mg/kg +220 Söderpalm and Engel (1989)
agonist Rat (LE) i.p. + 2–5 mg/kg +125 Carlson et al. (2001)
Triazolam Triazolo-BZP Rat (SD) i.p. + 0.01 and 0.1 mg/kg +220 Depoortere et al. (1986)
agonist Rat (W) i.p. + 1.25 mg/kg +380 Brocco (unpublished
observations)
Midazolam Triazolo-BZP Rat (W) i.p. + 1–4 mg/kg +94 Stefanski et al. (1992)
agonist Rat (W) i.c. (hippocampus) + 10 and 20 ␮g +180 Nazar et al. (1999b)
Rat (W) i.c. (hippocampus) + 10 and 20 ␮g +300 Stefanski et al. (1993a)
Rat (W) i.c. (N. accumbens) IA 5–20 ␮g IA Stefanski et al. (1993a)
Rat (?) i.c. (amygdala) + 1 ␮g +1000 Schell-Krüger and Petersen
(1982)
Zolpidem ␣1 > ␣2 ,␣3 -agonist Rat (SD) i.p. + 1.0–10.0 mg/kg +200 Depoortere et al. (1986)
Rat (W) i.c. (hippocampus) IA 0.1–10 ␮g IA Nazar et al. (1999b)
Abecarnil ␣1 > ␣2 ,␣3 -agonist Rat (W) i.p. + 0.3–10.0 mg/kg +825 Stephens et al. (1990)
Rat (W) i.p. + 30.0 mg/kg +250 Griebel et al. (1999c)
Zopiclone ␣1 ,␣2 ,␣3 -agonist Rat (W) i.p. + 10 mg/kg +282 Shibata et al. (1989)
Rat (W) i.c. (amygdala) + 1 ␮g +185 Shibata et al. (1989)
SL 651,498 ␣2 ,␣3 > ␣1 -agonist Rat (W i.p. + 10 mg/kg +122 Griebel et al. (2001b)
and SD)
Bretazenil BZP partial agonist Rat (W) i.p. + 1 and 10 mg/kg +310 Griebel et al. (1999c)
Rat (W) i.c. (hippocampus) IA 0.1–10 ␮g/site IA Nazar et al. (1999b)
Imidazenil BZP partial agonist Rat (W) i.p. + 30.0 mg/kg +225 Griebel et al. (1999c)
RWJ-51204 BZP partial agonist Rat (LE) p.o. + 0.1–10 mg/kg NI Dubinsky et al. (2002)
Flumazenil BZP Ant Rat (W) i.p. IA >10 mg/kg IA Wieroñska et al. (2001)
Rat (SD) p.o. IA 10–50 mg/kg IA Liljequist and Engel (1984a,b)
Ethyl-␤-carboline BZP inv. agonist Rat (W) i.p. − 10 mg/kg −50 Mizoule et al. (1985)
Methyl-␤-carboline BZP inv. agonist Rat (SD) i.v. − 0.15 mg/kg −60 Corda et al. (1983)
FG 7142 BZP inv. agonist Rat (W) i.p. − 5 mg/kg −50 Mizoule et al. (1985)
Rat (SD) i.v. − 4 mg/kg −55 Corda et al. (1983)
Abbreviations—N. accumbens: nucleus accumbens; LE: Long–Evans; H: Holtzman; BZP: benzodiazepine; IA: no significant change (inactive) and NI:
not indicated. For other abbreviations, see Table 2. For ␣-subunits, relative activities refers to efficacies rather than affinities.

effects of flumanezil have been documented in man (Nutt, underlie the distinctively powerful and broad-based anxi-
1996). olytic profile of BZPs—though also contributing to their
panoply of side-effects (Section 18.1) (Petersen et al., 1985;
2.2.2.2. Influence upon anxious states: multiple mechanisms Pratt, 1992; Menard and Treit, 1999; Sramek et al., 2002).
of action. BZPs exert their actions at multiple supraspinal Employing the VCT, anxiolytic actions of diazepam and
loci and, as a function of the structure involved, their actions other BZPs were seen upon their application into the dorsal
are probably expressed upon different components of anx- hippocampus (Katoaka et al., 1991; Stefański et al., 1993a;
ious states. This involvement of diverse sites of action may Plaznik et al., 1994a), observations mimicked by studies
98 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

of exploratory behavior in a novel environment (Stefański 2.2.2.3. Influence upon anxious states: monoaminergic and
et al., 1993a) though, curiously, few other data are available non-monoaminergic mechanisms. The anxiolytic actions
for this structure. By analogy, only a single study (Katoaka of BZPs in limbic structures involve both pre- and post-
et al., 1982), employing a Geller conflict procedure, has lo- synaptic interactions with serotonergic terminals (Section
calized anxiolytic actions of BZPs to the mammillary bod- 4.3.1) (Pratt, 1992; Tao et al., 1996; Dekeyne et al., 2000b;
ies. On the other hand, by use of the VCT (Scheel-Kruger Millan et al., 2001b; Olivier et al., 2001; Adell et al., 2002).
and Petersen, 1982; Petersen et al., 1985; Shibata et al., There is also evidence that actions of BZPs upon serotoner-
1989) and other conflict procedures, numerous investiga- gic perikarya participate in their anxiolytic actions (Fig. 4).
tors have pinpointed the amygdala as a crucial structure for Unambiguous evidence that BZPs act at sites in raphe nu-
anxiolytic actions of BZPs, a conclusion underscored by clei to enhance responding in conventional conflict models
models involving untrained behaviors (Menard and Treit, has not, in fact, been presented but, employing a modified
1999). The majority of these studies suggest that the ba- conflict procedure (Section 18.5), a role for the DRN was
solateral complex of the amygdala is a more prominent deduced (Thiébot et al., 1980, 1982). Moreover, employ-
locus of action for BZPs than the central nucleus, pos- ing models of unconditioned behaviors, both the DRN and
sibly reflecting the greater concentration of BZP sites in the MRN have been identified as substrates for anxiolytic
the former region. However, as implied above, it is possi- actions of BZPs (Costall et al., 1989; Menard and Treit,
ble that these two sub-territories fulfil complementary roles 1999).
in mediating different elements of the anxiolytic profile of An interruption of the stress-induced activation of nora-
BZPs (Scheel-Kruger and Petersen, 1982; Menard and Treit, drenergic neurones by actions in the LC also contributes to
1999). The lateral septum has also been identified as an im- the anxiolytic properties of BZPs (Fig. 4) (Owens et al.,
portant structure for expression of the anxiolytic properties 1989, 1991; Shekhar et al., 2002). In addition, the ability of
of BZPs, though its role differs to that of amygdalar nuclei BZPs to brake the engagement of mesolimbic-mesocortical
(Menard and Treit, 1999). While data for actions of BZPs in dopaminergic projections originating in the VTA may par-
conflict procedures appear to be unavailable for this struc- ticipate in their relief of anxious states (Fig. 4) (Millan et al.,
ture, evidence for anxiolytic actions of GABAA agonists is 2000d; Morrow et al., 2000; Beaufour et al., 2001a; Dazzi
available (Drugan et al., 1986). There is, finally, evidence et al., 2001; Gifkins et al., 2002).
for a role of the dorsal PAG in mediating the anxiolytic Notwithstanding the significance of this generalized,
actions of BZPs in the open-field procedure (Menard and inhibitory influence of BZPs upon the response of cor-
Treit, 1999). Thus, contrasting neuronal substrates are im- ticolimbic monoaminergic pathways to fear and stress, a
plicated in (various components) of the anxiolytic actions of reduction of monoamine release is neither necessary nor
BZPs, of which several regions, notably the amygdala and sufficient to account for their anxiolytic actions under all
the hippocampus, are involved in their effects in the VCT conditions—an axiom applicable to other classes of anxi-
(Fig. 3). olytic agent (see Section 18.1). Indeed, as explained below

Fig. 3. Schematic illustration of the sites of action of agents interacting with GABA receptors, NMDA receptors and metabotropic glutamatergic receptors
in the Vogel Conflict Test. The loci of action indicated are based upon results published for the Vogel Conflict Test (see tables). Additional sites of
anxiolytic action, revealed by other procedures, should be borne in mind. Abbreviations—BZP: benzodiazepine; T-BZP: triazolo-benzodiazepine; GlyB:
glycine B; mGlu: metabotropic glutamatergic; BARB: barbiturate; AGO: agonist; ANT: antagonist; D/MRN: dorsal/median raphe nucleus and PAG,
periaqueductal gray.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 99

Fig. 4. Influence of the benzodiazepines, diazepam and chlordiapoxide, upon extracellular levels of monoamines in the frontal cortex of freely-moving
rats. Drugs were administered at doses which elicit robust increases in punished responses in the Vogel Conflict Test. Data are means ± S.E.M.s. Dialysis
levels are expressed as a percentage of baseline, pre-injection values. Data are from Millan et al. (1997, 2001a). For the Vogel Conflict Test, ( ) P < 0.05
to vehicle values and, for dialysis data, ( ) P < 0.05 to basal pre-injection values.

(Section 2.2.2.3), one component of the anxiolytic pro- Boonyanaruthee, 1997; Millan et al., 2000d, 2001b;
file of BZPs may be exerted via populations of GABAA Beaufour et al., 2001a).
receptors which do not interact (at least presynaptically at It is, therefore, critical to recognize that BZPs exert their
the perikarya level) with monoaminergic pathways. Fur- distinctive anxiolytic (and undesirable) properties in ex-
ther, in contrast to traditional BZPs (such as diazepam), perimental models—and, presumably, in man—via actions
which exert a pronounced, dose-dependent and robust in- at a diversity of loci involving both monoaminergic and
hibitory influence upon hippocampal and frontocortical non-monoaminergic mechanisms.
release of 5-HT, NA and DA at anxiolytic doses (Fig. 4),
triazolo-BZPs (such as alprazolam) display a complex pat- 2.2.2.4. Functional roles of GABAA receptor subunits.
tern of structure-dependent increases and decreases in ex- Genetic disruption of individual isoforms of the ␣-subunit
tracellular levels of monoamines (Section 18.1) (Lima et al., of GABAA receptors by point mutations at position 101
1995; Broderick, 1997; Broderick et al., 1998; Millan et al., (substitution of a histidine by an arginine residue) suggests
2000b,d; Bentué-Ferrer et al., 2001). Fig. 5 exemplifies the that ␣1 -, ␣2 -, ␣3 - and ␣5 -subunits may each be associated
marked anxiolytic properties of alprazolam at a dose which, with contrasting facets of the functional profile of BZPs
in distinction to conventional BZPs (Fig. 4), decreases (Table 4) (Sieghart, 1995; Möhler et al., 2002). In interpret-
dialysis levels of 5-HT but not NA or DA in the FCX. ing the results discussed below, however, it should be borne
Indeed, preservation of catecholaminergic transmission in mind that the deletion of one individual subunit may
may account for the utility of triazolo-BZPs—as compared result in adaptive changes of others (Sur et al., 2001; Korpi
to conventional BZPs—in the management of depressive et al., 2002; Kralic et al., 2002a,b; Thompson et al., 2002).
states (Jonas and Cohon, 1993; Petty et al., 1995; Broderick Though mice overexpressing ␥2S -subunits are virtually in-
et al., 1998; Broderick, 1997; Srisurapanon and distinguishable from wild-type conspecifics, populations de-
100 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Fig. 5. Influence of the triazolo-benzodiazepines, triazolam and alprazolam, upon extracellular levels of monoamines in the frontal cortex of freely-moving
rats. Drugs were administered at doses which elicit robust increase in punished responses in the Vogel Conflict Test. Data are means ± S.E.M.s.
Dialysis levels are expressed as a percentage of baseline, pre-injection values. Data are from Millan et al. (2001d), and Gobert and Millan (unpublished
observations). ( ) P < 0.05 to vehicle values for the Vogel Conflict Test and, for dialysis data, ( ) P < 0.05 to basal pre-injection values.

ficient in ␥2S -subunits are refractory to BZPs (Günther et al., resistant to the anxiolytic actions of diazepam. This effect
1995; Wick et al., 2000). Further, heterozygous (±) ␥2S mice was specific inasmuch as pentobarbital was still effective,
present an anxiogenic phenotype (Crestani et al., 1999). As consistent with its action at a separate site on GABAA
mentioned above (Section 2.2.2.1), BZPs act at the interface receptors (Section 2.2.4) (Löw et al., 2000). A similar ap-
between ␣ and ␥2 -subunits. The probable explanation for the proach suggested that the myorelaxant actions of diazepam
anxious phenotype of ␥2S (±) mice is a reduction in the clus- are mediated by ␣2 -subunits, presumably reflecting their ex-
tering of GABAA receptors in somatic and dendritic regions pression in spinal motoneurones (Löw et al., 2000; Crestani
of neurones in the hippocampus, amygdala and/or cortex, et al., 2001; Möhler et al., 2001, 2002).
structures in which ␥2S -subunits are concentrated (Gutierrez ␣1 -Subunits are found in GABAergic neurones them-
et al., 1994; Crestani et al., 1999; Pirker et al., 2000; Nyiri selves and are highly expressed in the cerebellum, basal
et al., 2001). By analogy to the ␥2S -isoform, mice overex- ganglia, thalamus and cortex (Section 2.2.1). In distinc-
pressing ␥2L -subunits are unremarkable, while populations tion to ␣2 -subunits, inactivation of ␣1 -subunits eliminated
with deleted ␥2L -sites are anxiogenic (Günther et al., 1995; the sedative and amnesic actions of diazepam whereas its
Homanics et al., 1999; Wick et al., 2000). Intriguingly, mice anxiolytic (and myorelexant) properties were maintained.
lacking ␥2L -subunits are more sensitive to BZPs due to sub- The sedative actions of zolpidem were also abolished in
stitution of absent ␥2L -subunits by their ␥2S counterparts mice lacking ␣1 -subunits, and these effects were selective
which favor an “agonist-preferring”, BZP-responsive con- inasmuch as pentobarbital continued to compromise motor
figuration of GABAA receptors (Quinlan et al., 2000). function (Gao et al., 1993; Rudolph et al., 1999; Crestani
Introduction of a point mutation into the ␣2 -subunit, et al., 2000, 2002a; McKernan et al., 2000; Reynolds
which is enriched in corticolimbic structures, rendered mice et al., 2001; Kralic et al., 2002a). (These observations con-
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 101

Table 4 trasting patterns of data as concerns the phenotype of mice


Differential functional roles of GABAA receptor ␣-subunits as evaluated with dysfunctional ␣3 -subunits. In line with this supposition,
by genetic studies in mice
in an independent series of studies adopting a similar genetic
Activity ␣1 ␣2 ␣3 ␣5 approach but different experimental conditions, it was re-
Anxiolytic No Yes (Yes) No cently found that selective elimination of ␣3 -subunits did ab-
Amnesic Yes ? ? Yes rogate the anxiolytic properties of BZPs (McKernan, 2002).
Myorelaxant No Yes Yesa No
␣3 -Subunits, which are localized on noradrenergic neurones
Hypnotic (EEG) No Yes Yes No
Sedative Yes No No No facilitatory to motor function (Crestani et al., 2001), may
Anticonvulsant Yes Yes?b Yes?b Yes?b also play a role complementary to ␣1 -subunits in mediating
Yes: involved; No: not involved; (Yes): to be confirmed as data cur-
the myorelaxant properties of (high doses of) BZPs (Möhler
rently inconsistent and EEG: electroencephalography; a At high doses of et al., 2002).
diazepam; b Though the role of ␣1 -subunits in anticonvulsant properties Targeted disruption of ␣5 -subunits, which are principally
is confirmed, it remains to be elucidated which additional site(s) are in- found at extrasynaptic sites in the hippocampus (Section
volved. Note that the indication “No” should be considered as provisional 2.2.1), failed to impair the anxiolytic actions of diazepam.
since the involvement of specific subunits is dependent upon experimental
conditions, including the choice of benzodiazepine employed to validate
This result questions a major role of ␣5 -subunits in the
function. Further, pharmacological studies of selective agents at specific control of anxious states by BZPs despite reports that in-
␣-subunits have, in certain cases, generated data indicating roles which verse agonists at ␣5 -subunits manifest anxiogenic properties
have not, as yet, been clearly revealed by studies of genetically-modified (Crestani et al., 2002b; Garcia et al., 2002; McKernan, 2002;
mice. See text for further details. Data summarized from Crestani et al. Navarro et al., 2002). Interestingly, cognitive performance—
(1999, 2000, 2001, 2002a,b), Rudoph et al. (1999), Löw et al. (2000),
McKernan et al. (2000), Kralic et al. (2002a,b), McKernan (2002) and
including association learning and fear conditioning—was
Möhler et al. (2002). improved in mice lacking ␣5 -subunits (Collinson et al., 2002;
Crestani et al., 2002b; Garcia et al., 2002) though equivalent
findings with selective inverse agonists at ␣5 -subunits await
form, incidentally, to the somewhat anachronistic notion confirmation (McKernan, 2002).
that GABAA receptors incorporating ␣1 -subunits may be Recombinant GABAA receptors incorporating ␣4 -subunits
equated with pharmacologically-defined “GABAA /BZ1 ” are essentially refractory to the actions of BZPs. Accord-
sites thought to transduce the sedative and hypnotic actions ingly, it has been posited that an up-regulation of hip-
of BZPs, as compared to “GABA/BZ2 ” sites—containing pocampal ␣4 -subunits by sudden termination of long-term
other subunits—which mediate their anxiolytic properties BZP administration participates in the negative emotional
(Dennis et al., 1988; McLeod et al., 2002).) In complemen- symptoms of the accompanying withdrawal syndrome—
tary studies, the use of ␣1 -subunit preferring antagonists though this hypothesis implicitly assumes the existence
has similarly suggested that ␣1 -subunits—and other (as yet of tonically-active, endogenous BZP-like ligands (Follesa
unidentified) mechanisms—mediate the sedative actions et al., 2001, 2002).
of BZPs in primates (Platt et al., 2002). Nevertheless, a It should, finally, be noted that substitution of ␥-subunits
possible incoherence between genetic and pharmacologi- by ␦-subunits (present in the hippocampus, but not the
cal approaches is provided by findings that a preferential amygdala), ε-subunits (enriched in the hypothalamus)
␣1 -subunit antagonist, ␣-carboline-3-carboxylate-t-butyl or ␪-subunits (confined to peripheral tissues) generates
ester, did interfere with the anxiolytic (and myorelax- GABAA receptors insensitive to BZPs (Barnard et al., 1998;
ant) effects of BZPs in rodents (Griebel et al., 1999a,c; Möhler et al., 2002).
Belzung, 2001). Interestingly, further, on the basis of
pharmacological studies, ␣1 -subunits were implicated in 2.2.2.5. Functional profiles of subunit-selective agents.
the “interoceptive” (discriminative stimulus), subjective— Notwithstanding the need for more extensive analysis, these
though not anxiolytic—effects of BZPs in primates (Lelas elegant studies of the functional consequences of point
et al., 2002; Rowlett et al., 2002, 2003). mutations in specific ␣-subunits of GABAA receptors pro-
Counterintuitively, in view of the fact that serotonergic vide an instructive framework for improved interpretation
cell bodies express ␣3 but not ␣1 - (or ␣2 ) subunits, it was re- of the actions of drugs differentially recognizing ␣1 -, ␣2 -
ported that deletion of ␣3 -subunits did not interfere with the and ␣3 -subunits. (That is, drugs which display contrasting
anxiolytic actions of diazepam (Gao et al., 1993; Fritschy efficacies at individual subunits inasmuch as the synthe-
and Möhler, 1995; Pirker et al., 2000; Möhler et al., 2001, sis of agents possessing absolute selectively—in terms of
2002). However, ␣2 -subunits exist on afferent terminals in affinity—has not, to date, proven feasible.) Thus, the novel
raphe nuclei (Pirker et al., 2000), and several other GABAA pyridoindole derivative, SL651,498, which behaves as a
receptor subunits occur both in the DRN and in the LC full agonist at ␣2 - and ␣3 as compared to ␣1 -subunits,
(Moragues et al., 2002). Thus, subunits other than ␣3 may displays robust anxiolytic actions in the VCT and other
modify monoaminergic transmission. Further, no informa- models in the absence of sedative (and amnesic) properties
tion from conflict models is available, and studies with other (Griebel et al., 2001b). Moreover, the recently-described
drugs and/or anxiolytic procedures may have yielded con- ligand, L838,417, reveals potent anxiolytic properties in
102 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

the relative absence of motor disruption, a profile which anxiety responsive to currently-available BZPs in the ab-
may be ascribed to its combined agonist actions at ␣2 - and sence of sedative and cognitive side-effects.
␣3 -subunits as compared to ␣1 -subunits (McKernan et al., Sixth, it has become almost a platitude to cite the amnesic
2000). Contrariwise, the potent hypnotic versus anxiolytic properties of BZPs as a troubling complication of their uti-
actions of zolpidem and zaleplon are congruent with their lization. It should not, however, be ignored that for certain
preferential (high efficacy) activation of ␣1 versus ␣2 , ␣3 categories of anxiety disorder, an inability to forget is a car-
and (even less potently) ␣5 -subunits (Benavides et al., 1990; dinal and diagnostic feature. Correspondingly, caution is im-
Meadows et al., 1998; Damgen and Lüddens, 1999; Doble, perative in the exploitation of drugs possessing, for example,
1999; Rush et al., 1999; Platt et al., 2002; Sanna et al., inverse agonist properties at ␣5 -subunits, and consequently
2002). Finally, the relatively modest myorelexant effects anticipated to “improve” mnesic function (Section 18.8).
of abecarnil are consistent with lower efficacy at ␣2 as Seventh, though elimination of sedative properties may, at
compared to ␣1 -subunits (Pribilla et al., 1993). first sight, also appear to represent a major advantage, this
supposition does not fully reflect the reality of therapeutic
2.2.2.6. Clinical relevance of subunit-selective drugs: open trials and common clinical usage. Indeed, introduction of
questions. Clearly, then, the assignment of specific func- the 5-HT1A receptor agonist, buspirone (Section 4.3.2.3),
tional correlates to distinct subunits of GABAA receptors encountered major difficulties precisely due to its lack of
offers the perspective of novel classes of anxiolytic agent sedative properties in patients accustomed to BZPs.
possessing an improved therapeutic window between doses Eighth, as for all neuropsychiatric indications, a consid-
exerting beneficial actions as compared to those provoking eration of paramount importance in the treatment of anxi-
deleterious side-effects. However, there remain several ques- ety disorders is the influence of drugs upon sleep (Benca,
tions and caveats. 1994a,b; Roth, 2001; Smith et al., 2002a). Sleep is almost
First, inasmuch as myorelaxant properties are primarily invariably disturbed in anxious patients—while, recipro-
mediated by the same ␣2 -subunit as that which transduces cally, patients with primary insomnia frequently reveal
the anxiolytic actions of BZPs, drugs exerting a “pure” anx- high state anxiety (Benca, 1994a; Roth, 2001; Smith et al.,
iolytic profile may be difficult to obtain—though the clini- 2002a,b). Indeed, patients suffering from panic attacks or
cal benefits of myorelaxation in the management of anxious GAD are characterized by a decline of sleep continuity, re-
states should not be neglected, notably in patients suffering duced total time and efficiency of sleep, high sleep latency
from discomfort of muscular origin (Millan, 1999). and early morning waking. These features are often encoun-
Second, development of tolerance and dependence upon tered in depressive subjects though, in contrast to the latter,
protracted administration of agents acting exclusively (as ag- anxious patients reveal a suppression of “REM” sleep dura-
onists or antagonists) at specific ␣-subunits warrants careful tion and no change “in REM” latency (Papadimitriou et al.,
examination. 1988a; Haurie et al., 1989; Mellman and Uhde, 1989a;
Third, as intimated above, employing diverse drugs and Benca, 1994a,b). Thus, successful (subunit-selective and
experimental approaches, it will be important to corroborate other) anxiolytic agents should favor high-quality, regular
and amplify insights provided by genetic studies into the and prolonged sleep without provoking a deterioration of
roles of specific GABAA receptor subunits (Reynolds et al., “morning-after” activities (Haefely et al., 1992; Gieschke
2001). et al., 1994; Patat et al., 1994; Borbély, 1995; Ashton,
Fourth, possibly reflecting species differences, though 1994; Noble et al., 1998; Mitler, 2000; Mailliet et al., 2001;
partial BZP agonists display functional profiles distinct Platt et al., 2002; Smith et al., 2002a,b). In this light, it has
from those of conventional (full agonist) BZPs in many been suggested that the hypnotic qualities of BZPs involve
experimental models (Stephens et al., 1990; Martin et al., ␣2 - and ␣3 -subunits, though this awaits confirmation (Platt
1993; Potokar and Nutt, 1994; Nazar et al., 1997; Dubinsky et al., 2002)
et al., 2002; Platt et al., 2002; McMahon and France, 2003),
they have not proven to be markedly different in the clinic 2.2.3. Neurosteroids
in terms of their therapeutic index between anxiolytic and
adverse effects (Haefely et al., 1992; Busto et al., 1994; 2.2.3.1. Multiple central actions. The term “neurosteroids”
Guldner et al., 1995; Aufdembrinke, 1998; Rush, 1998; (a contraction of neuroactive steroids) refers to a range
Sandford et al., 2001). Thus, clinical studies will be decisive of steroids synthesized in glial cells and neurones from
in assessing the advantages of subunit-selective anxiolytic cholesterol and other precursors. Neurosteroids exert rapid
agents as compared to traditional drugs which interact (non-genomic) actions in the brain leading to alterations
indiscriminately with all ␣-subunits of GABAA receptors. in neuronal excitability (Lambert et al., 1995; Beaulieu,
Fifth, the precise relationship between discrete ␣-subunits 1997; Maurice et al., 1999; Rupprecht and Holsboer, 1999;
and the heterogeneity of anxious disorders in man remains Falkenstein et al., 2000; Rupprecht et al., 2001; Miksys and
to be established. It should not, for example, be assumed Tyndale, 2002; Van Broekhoven and Verkes, 2003). In this
on the basis of extant data that the selective recruitment of regard, innumerable central actions of neurosteroids have
␣2 -subunits will suffice for effective relief of all forms of been described, including effects at the following sites:
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 103

5-HT3 receptors, glycineA receptors, nicotinic and mus- ing levels of allopregnanolone has been detected in patients
carinic receptors, oxytocin receptors, N-methyl-d-aspartate with GAD (Semeniuk et al., 2001), as well as patients suf-
(NMDA), kainate and ␣-amino-2,3-dihydro-5-methyl-3- fering from panic disorders and other anxious states (Purdy
oxo-4-isoxazolepropanoic acid (AMPA) receptors, and et al., 1991; Steimer et al., 1997; Rupprecht and Holsboer,
sigma1 binding sites (Section 17). All might, in theory, be 1999; Vallee et al., 2000; Girdler et al., 2001; Ströhle et al.,
involved in the influence of neurosteroids upon anxious 2002, 2003). This may represent a negative-feedback re-
states though little if any concrete information is available sponse counteracting the accompanying anxiety, though a
in this regard (Park-Chung et al., 1997; Maurice et al., 1999; very recent report suggested that neurosteroid levels may ac-
Zinder and Dar, 1999; Falkenstein et al., 2000; Meyer et al., tually be transiently reduced during a panic attack (Ströhle
2002; Van Broekhoven and Verkes, 2003). Indeed, the prin- et al., 2003).
ciple mode of action implicated in the anxiolytic properties It has been hypothesized that the anxiolytic effects
of neurosteroids is their positive, allosteric modulation of of (long-term) administration of SSRIs (Section 4.4) re-
GABAA receptors. The term “epalon” has been specifically flect (independently of 5-HT transporters), the progres-
applied to (endogenous and synthetic) steroids which exert sive (5␣-reductase and 3␣-hydroxysteroid oxidoreductase-
such actions (Lan and Gee, 1997), but it is not employed in catalysed) formation of allopregnanolone (Uzunov et al.,
the present article. 1996; Guidotti and Costa, 1998; Uzunova et al., 1998;
Ströhle et al., 1999; Griffin and Mellon, 1999; Serra et al.,
2.2.3.2. Generation and modulation. Cholesterol in the 2001; Roca et al., 2002). Clinical support for this con-
brain is sequentially converted (primarily in the mito- tention is provided by observations that cerebral levels of
chondria of glial cells) via pregnenolone and progesterone neurosteroids—and GABA—are elevated by fluoxetine in
into reduced metabolites of the latter: allopregnanolone premenstrual females, in parallel with an alleviation of their
(or 3␣-hydroxy-5␣-pregnan-20-one) and pregnanolone (or emotional symptoms (Halbreich et al., 1996; Uzunov et al.,
3␣-hydroxy-5␤-pregnan-20-one), respectively (Lambert 1996; Wang et al., 1996a; Bixo et al., 1997; Dimmock et al.,
et al., 1995; Beaulieu, 1997; Maurice et al., 1999; Rupprecht 2000; Girdler et al., 2001; Epperson et al., 2002; Sanacora
and Holsboer, 1999; Falkenstein et al., 2000; Rupprecht et al., 2002). However, there are also indications for a rapid
et al., 2001; Miksys and Tyndale, 2002; Van Broekhoven and facilitatory influence of fluoxetine upon neurosteroid levels
Verkes, 2003). Circulating sources of progesterone are also in rodents (Griffin and Mellon, 1999; Serra et al., 2001),
(peripherally and centrally) converted into allopregnanolone. despite its anxiogenic properties upon acute administration
These neurosteroids enhance the frequency and duration (Section 4.4.2). Thus, the temporal and causal relationship
of opening of GABAA receptor-coupled chloride channels, between the influence of SSRIs and other antidepressant
properties shared by: (1) the deoxycorticosterone deriva- agents upon corticolimbic generation of neurosteroids as
tives, 3␣,5␣- and 3␣,5␤-tetrahydrodeoxycorticosterone compared to their control of anxious states requires further
(3␣,5␣- and 3␣,5␤-THDOC) and (2) the testosterone clarification (Van Broekhoven and Verkes, 2003). Moreover,
metabolites, androsterone and 3␣-androstanediol, which it was recently found that fluoxetine can directly facilitate
may also be derived from circulating, endocrine sources the actions of GABA at GABAA receptors, further compli-
(Section 12) (Lambert et al., 1995; Beaulieu, 1997; Maurice cating analysis of the role of GABAergic transmission in
et al., 1999; Rupprecht and Holsboer, 1999; Falkenstein its influence upon anxious states (Robinson et al., 2003).
et al., 2000; Rupprecht et al., 2001; Miksys and Tyndale,
2002; Van Broekhoven and Verkes, 2003). In contrast to 2.2.3.3. Anxiolytic properties: roles of GABAA subunits.
these reduced steroids, other products of cholesterol trans- Synthetic and endogenous neurosteroids (positive modula-
formation may act as antagonists at GABAA sites, includ- tors) suppress the anxiogenic actions and gene expression
ing progesterone itself, pregnenolone sulfate—which can of corticotrophin-releasing factor (CRF) (Section 9.1.2)
also activate NMDA receptors—and its processing product, (Britton et al., 1992; Patchev et al., 1994) and evoke an
dehydroepiandrosterone sulfate (op. cit.). impressive spectrum of anxiolytic actions at non-sedative
The first synthetic drug shown to enhance activity at doses in diverse experimental models (Britton et al., 1991;
GABAA receptors via an action at “neurosteroid” sites was Wieland et al., 1995, 1997; Akwa et al., 1999; Gasior et al.,
the general anaesthetic, alphaxalone. Numerous other exoge- 1999; Rupprecht and Holsboer, 1999; Gomez et al., 2002;
nous neurosteroids have since been proposed as therapeutic Van Broekhoven and Verkes, 2003; De Boer and Koolhaas,
agents for the treatment of BZP-sensitive anxiety disorders 2003). Correspondingly, central administration of neuros-
in the hope of achieving improved efficacy while avoiding teroids is associated with anxiolytic properties in the VCT
motor and cognitive side-effects, though this remain to be and other conflict paradigms (Table 5) (Crawley et al., 1986;
proven clinically (Lan and Gee, 1997; Gasior et al., 1999; Wieland et al., 1991; Carboni et al., 1996; Czlonkowska
Vanover et al., 2000; Visser et al., 2002). et al., 1999). Concordant with studies of BZP receptor
The synthesis of neurosteroids is accelerated by stress ligands (Section 2.2.2.2), the amygdala has been identi-
(Purdy et al., 1991; Steimer et al., 1997; Rupprecht and fied as an important locus for expression of the anxiolytic
Holsboer, 1999) and, intriguingly, an elevation in circulat- properties of neurosteroids possibly in interaction with neu-
104 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Table 5
Influence of neurosteroids, of injectable anesthetics and of ethanol upon behavior in the Vogel Conflict Test
Drug Class Species Route Effect Active dose Maximum Reference
(strain) (structure) range change (%)
Allopregnanolone GABAA NST agonist Mouse (SW) i.p. + 20 mg/kg +140 Wieland et al. (1991)
Rat (W) i.c.v. + 10 ␮g +290 Czlonkowska et al. (1999)
Rat (SD) i.c.v. + 2.5 and 10 ␮g +430 Carboni et al. (1999)
3␣,5␣-THDOC GABAA NST agonist Rat (SD) i.p. + 10 mg/kg +1000 Crawley et al. (1986)
Rat (W) i.c.v. + 10–30 ␮g +200 Czlonkowska et al. (1999)
3␣,5␤-THDOC GABAA NST agonist Rat (W) i.c.v. + 20 and 30 ␮g +250 Czlonkowska et al. (1999)
3␣,5␤-P-20-one GABAA NST agonist Rat (SD) i.c.v. + 2.5–20 ␮g +200 Carboni et al. (1999)
3␣,5␣-P-20␣-diol GABAA NST agonist Rat (SD) i.c.v. + 10–20 ␮g +275 Carboni et al. (1999)
3␣,5␣-P-20␤-diol GABAA NST agonist Rat (SD) i.c.v. + 60 ␮g +160 Carboni et al. (1999)
3␣,5␤-P-20␣-diol GABAA NST agonist Rat (SD) i.c.v. + 10 and 20 ␮g +620 Carboni et al. (1999)
Pregnenolone GABAA NST Ant Rat (W) i.c.v. IA 10–30 ␮g IA Czlonkowska et al. (1999)
3␤,5␣-P-20-one GABAA NST Ant Rat (SD) i.c.v. IA 2.5–5 ␮g IA Carboni et al. (1999)
Phenobarbital Barbiturate Rat (?) i.p. + 15 mg/kg +41 Getova et al. (1990)
Anesthetic Rat (W) i.c. (amygdala) + 40 ␮g +220 Shibata et al. (1989)
Pentobarbital Barbiturate Rat (H) i.p. + 5 and 10 mg/kg +450 Vogel et al. (1980)
Anesthetic Mouse (ICR) s.c. + 20 mg/kg +51 Umezu (1999)
Propofol Anesthetic Rat (W) i.p. + 20 and 40 mg/kg +340 Matsuo et al. (1997)
4-I-Pro Anesthetic Rat (SD) i.p. + 60 mg/kg +330 Sanna et al. (1999)
Ethanol Modulator Rat (SD) i.p. + 2000 mg/kg +290 Liljequist and Engel (1984a,b)
Abbreviations—H: Holtzman; SW: Swiss; ICR: Swiss CD; NST: neurosteroid; IA: no significant change (inactive); 3␣,5␣- and 3␣,5␤-THDOC: 3␣,5␣- and
3␣,5␤-tetrahydrodeoxycorticosterone; P: pregnan and 4-I-Pro: 4-iodo-2,6-diisopropylphenone (propofol analogue). For other abbreviations, see Table 2.

ropeptides Y (NPY) (Section 9.1.5)—though the septum lation by protein kinase-mediated phosphorylation (Hodges
and hippocampus may likewise be implicated (Akwa et al., et al., 1991, 2002; Krishek et al., 1994; Lin et al., 1996;
1999; Bitran et al., 1999; Ferrara et al., 2001). Negative McDonald et al., 1998; Millan, 1999; Bowers et al., 2000)
modulators at neurosteroid sites are generally devoid of ac- and, in this regard, the actions of neurosteroids are differ-
tivity though anxiogenic (and even anxiolytic) actions have entially modified as compared to those of BZPs and GABA
been detected under certain conditions (Purdy et al., 1992; itself. Of particular interest, mice lacking the “ε” isoform
Picazo and Fernandez-Guasti, 1995; Akwa et al., 1999). of protein kinase C (which indirectly targets GABAA re-
The significance of individual GABAA receptor subunits ceptors) displayed both behavioral and endocrine—low
to the actions of neurosteroids is less apparent than for adrenocorticotropic hormone (ACTH) release—signs of
their BZP counterparts, though it is well-established that blunted anxiety, together with an enhanced sensitivity to
␤-subunits are of comparatively minor significance in this neurosteroids. In line with these findings, the actions of
regard (Section 2.2.2.4) (Lambert et al., 1995; Smith et al., neurosteroids in wild-type populations could be mimicked
1998b; Gulinello et al., 2001; Belelli et al., 2002; Bianchi by specific inhibitors of protein kinase C-ε (Hodges et al.,
et al., 2002). Interestingly, in contrast to BZPs, genetic 1999, 2002).
disruption of ␦-subunits, which are well represented in Following prolonged exposure to neurosteroids, precipi-
the hippocampus and nucleus accumbens, interfered with tation of withdrawal is accompanied by a complex pattern
the anxiolytic and sedative (though not cognitive) actions of regionally-specific alterations in GABAA receptor sub-
of neurosteroids. This finding coincides with studies in unit expression—these effects are partially mediated ge-
recombinant systems of GABAA receptors incorporating nomically by oxidized derivatives of neurosteroids acting at
␦-subunits which are highly sensitive to neurosteroids estrogen receptors (Section 12) (Biggio et al., 2001). The ac-
(Laurie et al., 1992; Wisden et al., 1992; Zhu et al., 1996; companying behavioral syndrome, which includes rebound
Mihalek et al., 1999; Adkins et al., 2001; Belelli et al., anxiety may—by analogy to BZPs themselves (Section
2002; Vicini et al., 2002; Wohlfarth et al., 2002). 2.2.2.4)—be partly attributable to an up-regulation of hip-
The above observations underscore the notion that neu- pocampal GABAA receptors incorporating ␣4 -subunits
rosteroids act at sites different to those recognized by (Smith et al., 1998b; Follesa et al., 2001, 2002; Gulinello
BZPs (Section 2.2.2.4). In support of contrasting (and com- et al., 2001, 2002; Mascia et al., 2002; Möhler et al., 2002;
plementary) mechanisms of action of neurosteroids and Hsu and Smith, 2003).
BZPs, they exert their anxiolytic actions synergistically in
a flumazenil-resistant and flumazenil-sensitive manner, re- 2.2.3.4. Other properties of neurosteroids. Notwithstand-
spectively (Van Broekhoven and Verkes, 2003). Moreover, ing the above observations, much remains to be learned
various subunits of GABAA receptors are subject to modu- concerning the interrelationship between neurosteroids,
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 105

GABAA receptors and anxious states. Notably, mechanisms of anxious states, though the mechanistic bases remain un-
controlling the local synthesis, release and accumulation of clear (MacIver, 1997; Schulte et al., 2000; Pain et al., 2002;
neurosteroids still elude identification. Further, processes Pashkov and Hemmings, 2002; Westphalen and Hemmings,
regulating the balance between positive allosteric modula- 2003).
tors, such as pregnanolone, as compared to negative modu- A recent report raised the possibility that potentiation of
lators, such as pregnenolone, remain to be clarified, and the transmission at GABAB receptors (Section 2.3) participates
precise contribution of GABAA receptors to the global in- in the anxiolytic (and other functional) actions of propofol
fluence of neurosteroids upon anxious states requires further (Schweiler et al., 2003).
evaluation (Falkenstein et al., 2000; Van Broekhoven and
Verkes, 2003). From a therapeutic perspective, the potential 2.2.5. Ethanol
advantages (and disadvantages) of neurosteroids as alterna-
tive (or adjunctive) agents to BZPs in the treatment of anxi- 2.2.5.1. Anxiolytic properties. It has long been held (the
ety disorders should become clearer as results from clinical “tension reduction hypothesis”) that a predisposition to
trials begin to be disclosed (see also comments in Section voluntary ethanol consumption reflects its anxiolytic prop-
2.2.2.6). This is a crucial issue since it cannot, a priori, be erties: that is, anxious patients are both more sensitive to
guaranteed that neurosteroids will be devoid of the detrimen- the reinforcing effects of ethanol and more vulnerable to
tal effects of BZPs. For example, recent studies have revealed its pernicious, long-term addictive potential (Cappell and
reinforcing properties of allopregnanolone in rats, while Herman, 1972; Altman et al., 1996; Silvestre et al., 2002)
neurosteroids substitute for BZPs in drug-discrimination (see also Section 4.2.1). as regards the implication of
procedures. These data raise the possibility of a similar sus- dopaminergic mechanisms in coping strategies in response
ceptibility for abuse in man (Engel et al., 2001; McMahon to anxiety and stress. Though a causal relationship be-
et al., 2001; Fish et al., 2002; Miczek et al., 2002; Sinnott tween “trait” anxiety (tension reduction) and ethanol abuse
et al., 2002; McMahon and France, 2003). has been vigorously repudiated—and alcoholic patients
are clearly a heterogeneous population (Cloninger, 1987;
2.2.4. Propofol and other injectable general anesthetics Schuckit and Hesselbrock, 1994; Stephens et al., 2002;
Injectable general anesthetics such as propofol and the Van Erp et al., 2001)—both epidemiological and experimen-
barbiturate, pentobarbital, both potentiate responsiveness tal support for this hypothesis has been advanced. For exam-
to GABA and directly activate GABAA receptors in its ple, anxiety (and stress in general) is predictive of recidivism
absence. Mutational analyses of specific GABAA recep- in ethanol addicts and provokes reinstatement of ethanol-
tor subunits have shown that, in contrast to BZPs, the seeking behavior in rodents (Lê et al., 1998; Brady and
␣2 -subunit is not critical for expression of the anxiolytic Sonne, 1999). Intriguingly, the amygdala appears to fulfil
properties of pentobarbitol (Löw et al., 2000) which—like a critical role both in the self-administration of ethanol and
other barbiturates—displays robust activity in the rat and in the expression of its anxiolytic properties (Stewart et al.,
mouse VCT (Table 5) (Vogel et al., 1980; Shibata et al., 1993; Colombo et al., 1995; Hyytia and Koob, 1995; Spa-
1989; Getova et al., 1990; Umezu, 1999; Prut and Belzung, nagel et al., 1995; Roberts et al., 1996; Möller et al., 1997).
2003). Further, the GABAA receptor subunits recognized Thus, in the amygdala and, possibly, other limbic struc-
by this and other intravenous anesthetics differ to those tures, the facilitatory influence of ethanol upon GABA
engaged by their volatile counterparts, such as isoflurane receptor-mediated ionic currents likely accounts for its
(Puia et al., 1991; Harris et al., 1995; Birnir et al., 1997; broad influence upon emotion and cognitive function
Krasowski et al., 1998; Amin, 1999; Chen et al., 1999; (Pohorecky, 1990; Mihic, 1999; Korpi et al., 2002; Miczek
Haseneder et al., 2002; Jenkins et al., 2002). Binding to ␤- et al., 2002) including its anxiolytic properties in the VCT
(but not ␥-) subunits appears to be of particular importance and other procedures—though anti-conflict actions are best
for the functional actions of pentobarbital and propofol, demonstrated where parameters are adjusted to yield modest
though this remains to be corroborated (Concas et al., 1991; response suppression (Vogel et al., 1971; Petersen and Buus
Hara et al., 1993; Cestari et al., 1996; Hevers and Lüddens, Lassen, 1981; Liljequist and Engel, 1984a,b; Criswell and
1998; Krasowski et al., 1998; Sanna et al., 1999; Haseneder Breese, 1989; Koob et al., 1987; Möller et al., 1997; Silvestre
et al., 2002; Korpi et al., 2002; Williams and Akabas, 2002). et al., 2002; De Boer and Koolhaas, 2003). Though recurrent
In analogy to BZPs, it has been shown that barbiturates, ethanol administration elicits tissue-dependent alterations
propofol and propofol derivatives all exert pronounced anx- in the gene expression of various ␣-, ␤- and ␥-GABAA
iolytic activity in the VCT in parallel with a reduction in the receptor subunits (Grobin et al., 2000), its binding domain
accompanying induction of 5-HT release (Table 5) (Matsuo remains ill-defined. The contention that ␥2 -subunits (the
et al., 1996, 1997; Chen et al., 1999; Sanna et al., 1999). long splice variant) are a prerequisite for its behavioral and
General anesthetics also exert a complex pattern of (in gen- other actions remains controversial (Wafford and Whiting,
eral, inhibitory) effects upon corticolimbic noradrenergic, 1992; Homanics et al., 1999; Cagetti et al., 2003). Fur-
dopaminergic and, probably, glutamatergic transmission: it ther, though certain actions of ethanol may, like those of
is plausible that these effects contribute to their modulation neurosteroids, be dependent on the integrity of ␦-subunits,
106 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

their genetic elimination did not modify the anxiolytic prop- tissues, while abrupt suspension of its long-term administra-
erties of ethanol (Mihalek et al., 2001; Sundstrom-Poromaa tion elicits a opposite elevation in limbic glutamate release
et al., 2002). On the other hand, ␣1 - and ␤2 -subunits are together with a negative affective state characterized by
implicated in the hypnotic effects of ethanol (Blednov pronounced anxiety (Rossetti and Carboni, 1995; Shimuzu
et al., 2003). Further, ␣4 -subunits are of particular signif- et al., 1998; Yan et al., 1998; Piepponen et al., 2002). Direct
icance to the GABAA receptor-mediated enhancement of (inhibitory) interactions of ethanol with NMDA and, possi-
chloride-currents by low concentrations of ethanol (Korpi bly, kainate receptors may also participate in its functional
et al., 2002; Sundstrom-Poromaa et al., 2002), and hip- actions, such as (vide supra) an enhancement of mesolim-
pocampal (extrasynaptic) populations of ␣4 -subunits play bic dopaminergic transmission (Allgaier et al., 1999;
a prominent role in its influence upon mood, including its Woodward, 2000; Crowder et al., 2002). In this light, evi-
anxiolytic and stress-attenuating properties (Smith et al., dence has been forwarded that a component of the anxiolytic
1998b; Sundstrom-Poromaa et al., 2002). Mimicking re- properties of ethanol is mediated via its (non-competitive)
sults obtained both with BZPs themselves (Section 2.2.2.4) antagonist properties at NMDA receptors, though the pre-
and with neurosteroids (Section 2.2.3.2), discontinuation of cise locus of action of ethanol remains unclear. NR1/NR2B
protracted ethanol treatment elicits an up-regulation of hyp- receptor complexes, which are enriched in the forebrain,
pocampal ␣4 -subunits. In view of their resistance to BPZs, may be particularly sensitive to inhibition by ethanol
this surfeit of ␣4 -subunits may contribute to the accompa- (Section 3.2.2) (Moraes-Ferreira and Morato, 1997; Harrison
nying withdrawal syndrome (Devaud et al., 1997; Papadeas et al., 1998; Wirkner et al., 1999; Woodward, 2000; Criswell
et al., 2001; Möhler et al., 2002; Cagetti et al., 2003). et al., 2003).
A further parallel to neurosteroids—which may be in- In view of the above-discussed (Section 2.2.5.1) concor-
volved in the actions of ethanol (Mihalek et al., 2001; dance of experimental and clinical evidence in favor of a
Morrow et al., 2001; Cagetti et al., 2003; Tokunaga et al., close interrelationship amongst stress, anxiety and ethanol
2003)—is provided by the supersensitivity to its motor ac- abuse in man, many studies have focussed on the facilitatory
tions in mice genetically deficient in protein kinase C-ε, influence of acute ethanol administration upon the activ-
reproducing cellular findings that inhibitors of protein ki- ity of the hypothalamo-corticotropic axis. The transitory
nase C-ε potentiate actions of ethanol at GABAA receptors pulse of ACTH release provoked by ethanol is accompa-
(Hodges et al., 1999; Olive et al., 2000; Choi et al., 2002). nied by an enhancement in hypothalamic gene expression
In mice lacking this enzyme, functional “rescue” of pro- of CRF. Contrariwise, it was recently shown that repeti-
tein kinase C-ε by conditioned induction of its expression tive exposure to ethanol is characterized by a reversible
in the amygdala and other forebrain structures normalized (restored upon withdrawal) depression in CRF production
responsiveness to ethanol (Choi et al., 2002). It would be in the hypothalamic paraventicular nucleus (Spencer and
informative to undertake comparable studies specifically McEwen, 1990; Madeira and Paula-Barbosa, 1999; Silva
focussing on the anxiolytic properties of ethanol in light of et al., 2002). Extrapolating from these results, one may rea-
the implication of GABAA receptors in the amygdala in the sonably conjecture that ethanol treatment (irrespective of
control of anxious states (Section 2.2) and in the functional its duration) should have marked effects on corticolimbic
actions of ethanol (see above). CRF-containing pathways and, underpinning this asser-
tion, CRF-deficient mice revealed excessive ingestion of
2.2.5.2. Other actions of ethanol. It is convenient to evoke ethanol, apparently due to the blunting of its rewarding and
here several other mechanisms by which ethanol may mod- motor-stimulant properties (Olive et al., 2003) It is, thus,
ulate anxious states—though they are not necessarily ex- likely that both endocrine and neuronal actions of CRF
pressed via its allosteric interaction with GABAA receptors. contribute to the influence of ethanol upon anxious states,
Protein kinase C-ε (see Section 2.2.5.1) modulates the an issue awaiting experimental confirmation by use of the
induction by ethanol of DA release in the amygdala and VCT and other paradigms.
other limbic structures, though the precise underlying mech- Finally, the facilitation of 5-HT3 receptor function by et-
anisms remain obscure and actions of ethanol independent of hanol via an increase of channel activation may also contr-
GABAA receptors may also be involved (Olive et al., 2000; ibute to its influence upon anxious states (Yang et al., 2003).
Sher, 2003). Indeed, while the marked influence of ethanol Thus, though ethanol is considered here under the rubric
upon corticolimbic release of monoamines is clearly related of GABAA receptors, it exemplifies the principle that mul-
to its modulation of emotionality, the precise significance for tiple mechanisms are often involved in the (generally com-
its control of anxious states awaits clarification (Campbell plex) influence of specific drug classes upon anxious states.
and McBride, 1995; Kiianman et al., 1995; Carboni et al.,
2000; Köhnke et al., 2002). 2.3. GABAB receptors
It has also been suggested that alterations in the release of
glutamate contribute to the influence of ethanol upon anx- 2.3.1. Coupling and localization
ious states. Thus, ethanol has been shown to diminish gluta- In contrast to the rapid actions of GABA trans-
matergic transmission in the hippocampus and other limbic duced by ionotropic GABAA receptors, metabotropic, G
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 107

protein-coupled GABAB receptors are responsible for its mate, as well as CRF, CCK and other neuropeptides, likely
slower component of neuronal inhibition. Though they participates in their control of anxious states (Bonanno and
modulate the activity of adenylyl cyclase, a suppression Raiteri, 1994; Waldmeier et al., 1994; Tanaka et al., 2003).
of Ca2+ -conductance (P/Q and N-type voltage-dependent Furthermore, the high density of GABAB sites in the dor-
calcium channels (VDCCs)) is implicated in the autorecep- sal and medial raphe nuclei suggests a role synergistic to
tor/heteroceptor actions of GABAB sites at GABAergic/ GABAA receptors in modulating the activity of serotonergic
non-GABAergic classes of terminal, respectively (Section pathways. However, in raphe nuclei, the presence of GABAB
13.1). In addition, engagement of K+ -channels (“GIRKs”) sites on GABAergic interneurones themselves, as well as on
underlies their postsynaptic, hyperpolarizing properties at afferent terminals, should not be neglected. Consequently,
both soma and dendrites (Section 13.3) (Kaupmann et al., it would be simplistic to assume that all functional effects
1997; Bowery and Enna, 2000; Billinton et al., 2001; of GABAB agonists introduced into raphe nuclei can be at-
Bowery et al., 2002; Murakami et al., 2002). tributed to a direct suppression of serotonergic transmission
Virtually all GABAB receptors are heterodimers of (Wirtshafter et al., 1993; Tao et al., 1996; Abellan et al.,
two proteins termed GABAB1 and GABAB2 . Though 2000; Wirtshafter and Sheppard, 2001; Adell et al., 2002;
the GABAB1 subunit is primarily responsible for ligand Varga et al., 2002; Serrats et al., 2003).
recognition, the GABAB2 component is essential for cell
surface insertion of the receptor and its appropriate func- 2.3.2. Anxiolytic properties of GABAB receptor
tional operation (Kuner et al., 1999; Bowery and Enna, agonists
2000; Devi, 2001; Galvez et al., 2001; Bowery et al., Mice lacking GABAB receptors have not been reported
2002; Kniazeff et al., 2002). Several splice variants of to display alterations in anxiety (Schuler et al., 2001) and,
GABAB1 and GABAB2 proteins are known. N-terminal rather surprisingly, the GABAB receptor agonist, baclofen,
GABAB1␣ and GABAB1␤ isoforms of the former are the has not invariably elicited anxiolytic actions in rodents,
most prominent, possibly predominating as post and presy- though it blocks the anxiety elicited by stress, attenuates the
naptic sites, respectively, in relationship to GABAergic withdrawal syndrome provoked by discontinuation of BZP
neurones in the hippocampus, though their neuronal local- administration, and was reported to decrease state anxiety
ization varies in a tissue-dependent fashion (Benke et al., in alcohol-dependent patients (Andrews and File, 1993;
1999; Billinton et al., 1999; Bischoff et al., 1999; Lu Sanders and Shekhar, 1995; Dalvi and Rodgers, 1996;
et al., 1999; Margeta-Mitrovic et al., 1999; Ng et al., 2001; Addolorato et al., 2002). One possible explanation for the
Bowery et al., 2002). Despite the tendency of GABAB1 and variable influence of baclofen upon anxious states appeals
GABAB2 subunits to associate with alien proteins (see also to its actions at GABAB autoreceptors. Thus, occasional re-
Section 18.2), no GABAB receptor “subtypes” as such are ports of anxiogenic actions of baclofen—and an isolated re-
currently accepted (Bowery and Enna, 2000). port of anxiolytic properties of the antagonist, CGP35348—
GABAB receptors are broadly distributed throughout cor- may reflect an indirect, GABAB autoreceptor-mediated
ticolimbic structures, and GABAB1␣ and GABAB1␤ proteins suppression—enhancement—of GABA release onto post-
show distinctive and complementary patterns of localization. synaptic, anxiolytic GABAA receptors (Dalvi and Rodgers,
Both are enriched in the hippocampus, the hypothalamus and 1996; Zarrindast et al., 2001). It is also conceivable that
the amygdala. On the other hand, GABAB1␤ sites predomi- (postsynaptic) anxiolytic GABAB receptors are saturated
nate in the septum and nucleus accumbens, while GABAB1␣ under certain conditions of high stress, so agonists would
sites are more conspicuously expressed in the PAG, mam- achieve little if any further stimulation. In any event, while
millary bodies, LC and raphe nuclei (Benke et al., 1999; baclofen (like several other classes of putative anxiolytic
Bischoff et al., 1999; Lu et al., 1999; Margeta-Mitrovic et al., agent) was inactive in the mouse VCT, it generally elic-
1999; Liang et al., 2000). In line with their anatomical lo- its robust increases in punished responses in the rat VCT
calization, the moderating influence of GABAB receptors (Table 6) (Ketelaars et al., 1988; Shephard et al., 1992;
upon corticolimbic release of noradrenaline (NA) and gluta- Umezu, 1999).

Table 6
Influence of GABAB receptor ligands upon behavior in the Vogel Conflict Test
Drug Class Species Route Effect Active dose Maximum Reference
(strain) (structure) range (mg/kg) change (%)
Baclofen GABAB agonist Rat (W) i.p. IA 1.25–2.5 IA Ågmo et al. (1991)
Rat (W) i.p. + 0.46 and 1.0 +185 Ketelaars et al. (1988)
Rat (SD) i.p. + 0.5–2.0 +190 Shephard et al. (1992)
Mouse (ICR) s.c. IA 0.1–3.0 IA Umezu (1999)
␦-Amino-valeric acid GABAB Ant Rat (SD) i.p. IA >10 IA Shephard et al. (1992)
Abbreviations—ICR: Swiss CD and IA: no significant change (inactive). For other abbreviations, see Table 2.
108 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

As discussed elsewhere in this article (Section 13.1.3), 1999). Moreover, a perturbation of glutamatergic transmis-
GABApentin, which directly binds to ␣2␦ subunits of sion is implicated in the affective and mnesic symptoms
VDCCs, displays pronounced anxiolytic actions. Activ- of several neurological and psychiatric disorders, includ-
ity at P/Q-type (␣2␦ , ␤1 , ␣1A ) and N-type (␣2␦ , ␤1 , ␣1B ) ing anxious states (Carlsson and Carlsson, 1990; Shors
VDCCs controlling transmitter release can be indirectly et al., 1997; Schwendt and Jezová, 2000; Millan, 2002b;
suppressed by engagement of presynaptic GABAB recep- Moghaddam, 2002). Extracellular pools of glutamate are
tors (Wu and Saggau, 1997; Bowery and Enna, 2000), rapidly sequestered by at least five classes of transporter,
while activation of postsynaptic GABAB sites coupled including the key glial subtypes, “EAAT1” and “EAAT2”.
to GIRKs elicits neuronal hyperpolarization in the hip- These transporters play a vital role in the highly complex
pocampus and other structures (Section 2.3.1). Intriguingly, pattern of communication amongst glutamatergic terminals
then, GABApentin was claimed to behave as an agonist and astrocytes (Danbolt, 2001; Amara and Fontana, 2002;
at cloned and postsynaptic (hippocampal) populations of Millan, 2002b; Nedergaard et al., 2002).
GABAB1␣ /GABAB2 heterodimers positively and nega- A fascinating new angle on the role of glutamate in
tively coupled to GIRKs and VDCCs, respectively (Taylor cerebral neurotransmission and—it may be deduced—
et al., 1998; Bertrand et al., 2001; Ng et al., 2001). This the control of anxious states was recently unveiled by
action was postulated to be involved in its anti-epileptic the discovery of a third vesicular glutamate transporter
actions and, inasmuch as GABApentin did not affect in “non-glutamatergic” neurones, including VTA-derived
presynaptic GABAB1␤ /GABAB2 heterodimers, postsynaptic dopaminergic projections, raphe-localized serotonergic
GABAB1␣ /GABAB2 heterodimers might also be involved fibers, striatal cholinergic interneurones and, quite extraor-
in its anxiolytic properties (Ng et al., 2001; Patel et al., dinarily, a subset of hippocampal and cortical GABAergic
2001). However, the contention that GABApentin displays interneurones (Takamori et al., 2001; Fremeau et al., 2002;
agonist properties at GABAB receptors has been vigorously Gras et al., 2002). By extrapolation from studies of types I
challenged (Stringer and Lorenzo, 1999; Lanneau et al., and II transporters, it may be inferred that glutamate can be
2001; Bowery et al., 2002). released from these pathways. Though this remains to be
Irrespective of the potential role of GABAB receptors in formally proven, in line with this assertion, previous studies
the anxiolytic actions of gabapentin, allosteric modulators have shown that glutamate is liberated by monoaminergic
acting at the N-terminal of the GABAB2 protein were re- neurones (Bezzi et al., 1998; Sulzer et al., 1998). An addi-
cently disclosed. Such ligands may, by analogy to the ac- tional and distinctive feature of the type III transporter is its
tions of BZPs at GABAA receptors, provide a route towards localization in soma and dendrites (raising the possibility of
novel classes of anxiolytic agent (Pin et al., 2001; Urwyler retrograde glutamatergic transmission), as well as in astro-
et al., 2001; Kerr et al., 2002). cytes, consistent with glial release of glutamate (Danbolt,
2001; Alger, 2002; Millan, 2002b).
Glutamate (together with other excitatory amino acids,
3. Excitatory amino acids such as aspartate) exerts its actions via several classes of
ionotropic and metabotropic receptor. Though ligands at glu-
3.1. Glutamatergic pathways tamate transporters have to date attracted little attention,
drugs acting at ionotropic and metabotropic receptors have
Glutamate, the major excitatory neurotransmitter in the been extensively examined in the VCT and other models of
mammalian CNS, fulfils a virtually universal role in in- potential anxiolytic activity.
formation transfer and synaptic plasticity. Prior to its exo-
cytotic release, it is loaded into vesicles by brain-specific, 3.2. NMDA receptors
Na+ -dependent “vesicular glutamate transporters”, of which
types I and II display complementary patterns of distribution 3.2.1. Cooperative operation of NMDA recognition and
in the brain: both have a predeliction for vesicles forming glycineB sites
asymmetrical synapses, a hallmark of glutamatergic termi- Of the three classes of ionotropic receptor which mediate
nals (Herzog et al., 2001; Varoqui et al., 2002). In light of glutamatergic transmission, cation-permeable, heteromeric
the omnipresence of glutamatergic pathways in corticolim- NMDA receptors are unique inasmuch as both glutamate
bic regions, it is hardly surprising that they fulfil—in inter- itself (or an alternative, excitatory amino acid) and a
action with GABAergic and monoaminergic networks—a “co-agonist”, glycine (or d-serine), are mandatory for their
plurality of roles in the emotional and cognitive response operation. The NMDA receptor complex is a tetramer (or
to fear. Correspondingly, psychological stress and exposure pentamer) assembled from two NR1 subunits (of which
to aversive stimuli is characterized by a pronounced in- eight splice variants exist) and two (or three) NR2 subunits
crease in extracellular levels of glutamate derived, at least (of which four members, A, B, C and D, are known) (Danysz
partially, from neuronal sources (Lowy et al., 1993, 1995; and Parsons, 1998; Ozawa et al., 1998; Dingledine et al.,
Moghaddam et al., 1994; Saulskaya and Marsden, 1995; 1999; Millan, 2002b). The precise subunit composition con-
Karreman and Moghaddam, 1996; Timmerman et al., fers specific properties upon NMDA receptors, including
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 109

kinetics, ligand recognition profiles, modulation and cerebral receptors. Studies effected in rats and in pigeons have gen-
distribution. NR1 and NR2 subunits are broadly-distributed erally revealed anxiolytic profiles of recognition site antag-
throughout the brain, including the cortex, amygdala and onists, though open channel blockers are less consistently
hippocampus, wherein NR2B and, to a much lesser degree, active, and primates have proven relatively irresponsive to
NR2C and NR2D subunits are also located (Danysz and both these classes of NMDA receptor antagonist (Bennett
Parsons, 1998; Ozawa et al., 1998; Dingledine et al., 1999; and Amrick, 1986; Dunn et al., 1989; Corbett and Dunn,
Millan et al., 2002). Incorporation of a more recently-dis- 1991, 1993; Koek and Colpaert, 1991; Willetts et al., 1991,
covered and developmentally-regulated NR3A subunit, lo- 1993; Wiley et al., 1992, 1998; Wiley and Balster, 1992; Xu
calized to adult hippocampus, cortex and PAG, suppresses and Commissaris, 1992). Microinjection studies have pro-
Ca2+ -permeability leading to a reduction in channel func- vided evidence that NMDA receptors located in the PAG
tionality (Al-Hallaq et al., 2002; Millan, 2002b). (Their as- and the hippocampus are involved in the anxiolytic actions
sociation with NR1 sites only generates excitatory channels of recognition site antagonists and open channel blockers
sensitive to glycine (Chatterton et al., 2002)). in the VCT, though other structures such as the nucleus ac-
Glutamate and glycine (d-serine) act at the NMDA recog- cumbens and amygdala are also implicated (Table 7 and
nition site and glycineB site, respectively, which are situated Fig. 3) (Bennett and Amrick, 1986; Falls et al., 1992; Plaznik
on different subunits—NR2 and NR1, respectively (Danysz et al., 1994a; Jessa et al., 1995, 1996b; Przegaliński et al.,
and Parsons, 1998; Ozawa et al., 1998; Dingledine et al., 1996; Sajdyk and Shekhar, 1997; Menard and Treit, 2000;
1999; Mothet et al., 2000; De Miranda et al., 2002; Millan, Carobrez et al., 2001; Martinez et al., 2002; Molchanov and
2002b). In addition to antagonists at NMDA recognition and Guimarães, 2002; Pilc et al., 2002a). Though few studies of
glycineB sites, agents acting within the cation-permeable long-term administration have been performed, anxiolytic
channel itself (“open channel blockers”) comprise poten- actions of recognition site antagonists and, probably, open
tial anxiolytic agents. Furthermore, degradation of trypto- channel blockers are maintained in the VCT and other con-
phan by the “kynurenine pathway” generates the NMDA flict procedures upon chronic administration (Wiley et al.,
recognition site agonist, quinolinic acid, and the antago- 1992; Xie et al., 1995a; Jessa et al., 1996a).
nist, kynurenic acid. Drugs targeting enzymes of this path- Anxiolytic actions of NMDA receptor antagonists in con-
way also represent potential therapeutic agents (Stone, 2001; flict tests are expressed independently of their analgesic
Millan, 2002b; Schwarcz and Pellicciari, 2002; Stone and properties and several studies have emphasized the speci-
Darlington, 2002). ficity of their actions in the VCT (Plaznik et al., 1994b,c;
In addition to their high density throughout corticol- Przegaliński et al., 1996; Molchanov and Guimarães, 2002;
imbic structures, interest in the role of NMDA/glycineB Pilc et al., 2002a). Investigations employing other conflict
receptors in the control of anxious states is encouraged by procedures underline the discrete influence of NMDA re-
numerous observations including: (1) the role of NMDA re- ceptor antagonists upon punished responses as compared to
ceptors in the amygdala—in association with metabotropic other behaviors (Plaznik et al., 1994b,c; Jessa et al., 1995,
receptors (Section 3.5.2)—in cognitive processes underly- 1996b; Przegaliński et al., 1996; Danysz and Parsons, 1998;
ing fear-conditioning, a paradigm highly pertinent to the Wiley et al., 1998; Pilc et al., 2002a). Nevertheless, the
VCT (Tang et al., 1999; Walker and Davis, 2000; Bauer therapeutic margin of recognition site antagonists and, in
et al., 2002); (2) the pronounced, modulatory influence particular, open channel blockers, for expression of their
of NMDA receptors upon the activity of monoaminergic anxiolytic versus undesirable actions (perturbation of motor
pathways (Takahata and Moghaddam, 1998; Morrow et al., performance, psychiatric side-effects, etc.) is unimpressive
2000; Babar et al., 2001; Del Arco and Mora, 2001; Adell (Plaznik et al., 1994c; Jessa et al., 1996b; Przegaliński et al.,
et al., 2002; Millan, 2002b); (3) the facilitatory influence 1996; Danysz and Parsons, 1998; Wiley et al., 1998; Krystal
of NMDA receptors upon limbic release of the anxiogenic et al., 1999; Millan, 2002b; Pilc et al., 2002a). Much interest
(Section 9.1.2) neuropeptide, CRF (Cratty and Birkle, has, thus, been directed towards the hypothesis that drugs
1999); (4) similarities in the behavioral effects of NMDA (partial agonists and antagonists) acting at the glycineB site
receptor antagonists and GABAergic agonists (Willetts may behave as anxiolytic agents with a superior therapeutic
et al., 1990) and (5) evidence for reciprocal, short and window.
long-term functional interactions amongst NMDA receptors
and BZPs in the modulation of anxious states (Sharma and 3.2.3. GlycineB site ligands
Kulkarni, 1993; Smith and Dudek, 1996; De Souza et al.,
1998; Przegaliński et al., 2000; Allison and Pratt, 2003). 3.2.3.1. Anxiolytic profiles. In fact, conflict procedures
and other paradigms have yielded a rather ambivalent
3.2.2. Anxiolytic actions of NMDA recognition site body of data concerning the potential anxiolytic properties
antagonists and open channel blockers of partial agonists, such as aminocyclopropyl carboxylic
Unlike certain other drug classes, conflict paradigms have acid (ACPC) and d-cycloserine, at glycineB sites. Overall,
been privileged in the characterization of potential anxiolytic however, it appears that they express modest anxiolytic ef-
properties of agents which interrupt transmission at NMDA fects in the virtual absence of a generalized disruption of
110 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Table 7
Influence of ligands acting at ionotropic glutamatergic receptors upon behavior in the Vogel Conflict Test
Drug Class Species Route (structure) Effect Active dose range Maximum Reference
(strain) change (%)
AP-7 NMDA (RS Ant) Rat (W) i.c.v. + 0.5 ␮g +94 Plaznik et al. (1994c)
Rat (W) i.c. (PAG) + 0.4 ␮g +60 Molchanov and Guimarães
(2002)
Rat (W) i.c. (hippocampus) + 0.5 ␮g +460 Jessa et al. (1995)
CGP37849 NMDA (RS Ant) Rat (W) i.p. + 1.25–5.0 mg/kg +171 Przegaliński et al. (1996)
Rat (W) i.p. + 2.5 mg/kg +220 Jessa et al. (1996a)
Rat (W) i.p. + 1.0 and 2.5 mg/kg +165 Plaznik et al. (1994c)
Rat (W) i.c. (hippocampus) IA 0.01–1.0 ␮g IA Przegaliński et al. (1996)
CGP39551 NMDA (RS Ant) Rat (W) i.p. + 5 and 20 mg/kg +100 Plaznik et al. (1994c)
Dizocilpine NMDA (OCB) Rat (W) i.p. + 0.0025 mg/kg +320 Jessa et al. (1996a)
Rat (W) i.p. + 0.005 and 0.01 mg/kg +450 Plaznik et al. (1994c)
Rat (W) i.c. (hippocampus) + 0.1 and 1.0 ␮g +110 Jessa et al. (1995)
Glycine Glycine B agonist Rat (W) i.p. IA >800 mg/kg IA Chojnacka-Wójcik et al.
(1996a)
ACPC Glycine B (PAGO) Rat (W) i.p. + 200 mg/kg +200 Przegaliński et al. (1996)
Rat (W) i.p. + 200 mg/kg +260 Chojnacka-Wójcik et al.
(1996a)
Rat (W) i.c. (hippocampus) + 3–30 ␮g +302 Przegaliński et al. (1996)
d-Cycloserine Glycine B (PAGO) Rat (W) i.p. + 200 and 300 mg/kg +337 Klodzińska and
Chojnacka-Wójcik (2000)
5,7-Dichloro- Glycine B Ant Rat (W) i.c.v. + 5 ␮g +104 Plaznik et al. (1994c)
kynurenate
L701,324 Glycine B Ant Rat (SD) i.p. IA 0.1–10 mg/kg IA Karcz-Kubiscka et al. (1997)
MRZ 21576 Glycine B Ant Rat (SD) i.p. IA 2.5–10 mg/kg IA Karcz-Kubiscka et al. (1997)

CNQX AMPA Ant Rat (W) i.p. IA 0.05–5 mg/kg IA Czlonkowska et al. (1997)
NBQX AMPA Ant Rat (W) i.p. IA 0.1–5 mg/kg IA Czlonkowska et al. (1997)
LY326,325 AMPA Ant Rat (SD) i.p. + 2.5–5.0 mg/kg +100 Kotlinska and Liljequist (1998b)
Piracetam AMPA positive Rat (W) p.o. + 500 mg/kg per day +95 Bhattacharyya et al. (1996)
modulator (14 days)
Abbreviations—AP-7: DL-2-amino-7-phosphonoheptanoic acid; ACPC: 1-aminocyclopropanecarboxylic acid; CNQX: 6-cyano-7-nitroquinoxaline-2,3-
dione; NBQX: 2,3-dihydroxy-6-nitro-7-sulfamoyl-benzo-quinoxaline; RS: recognition site; OCB: open channel blocker; PAGO: partial agonist; PAG:
periaqueductal gray and IA: no significant change (inactive). For other abbreviations, see Table 2.

behavior (Corbett and Dunn, 1991, 1993; Koek and Colpaert, 3.2.3.2. Significance of antagonist properties at glycineB
1991; Trullas et al., 1991; Dunn et al., 1992; Anthony sites. It is necessary to address the question of whether the
and Nevins, 1993; Danysz and Parsons, 1998; Dunn et al., anxiolytic actions of glycineB partial agonists reflect rein-
1998; Wiley et al., 1998). Correspondingly, selective anx- forcement or suppression of transmission at NMDA recep-
iolytic effects of glycineB ligands have been documented tors. Several arguments lend credence to the latter possibil-
for the VCT at doses which do not markedly modify motor ity.
function, nociception or other relevant parameters (Table 7) First, though there are circumstances under which spe-
(Chojnacka-Wójcik et al., 1996a; Przegaliński et al., 1996; cific populations of glycineB sites are not saturated, their
Klodzińska and Chojnacka-Wójcik, 2000). Of special note degree of occupation by endogenous ligands in vivo is
is the observation that anxiolytic actions of glycineB partial pronounced. It is likely to be particularly high under con-
agonists persist following their prolonged administration, ditions of stress owing to the enhanced availability of
in line with findings in other models (Skolnick et al., 1992; glycine and d-serine, both of which behave as full ago-
Przegaliński et al., 1999). Likewise in accordance with re- nists (Danysz and Parsons, 1998; Dingledine et al., 1999;
sults acquired with recognition site antagonists and open Millan, 2002b). The higher the concentration of these
channel blockers (Section 3.2.2), it has been shown that endogenous ligand(s) at glycineB sites, the greater the
glycineB sites in the hippocampus participate in the anxi- likelihood that exogenous partial agonists will act as an-
olytic actions of the partial agonist, ACPC, in the VCT, while tagonists in opposing their actions. Second, the relatively
observations with other paradigms suggest a prominent role high doses of d-cycloserine and ACPC required for anx-
for glycineB sites in the PAG (Table 7 and Fig. 3) (Matheus iolytic actions in the VCT and other models are com-
et al., 1994; Przegaliński et al., 1996; De Souza et al., 1998). mensurate with those exerting antagonist actions in other
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 111

functional paradigms (Millan, 2002b; Danysz and Parsons, 3.2.5. Modulatory sites on NR2B subunits: ifenprodil
1998). Third, glycine itself and NMDA (upon systemic or As mentioned above (Section 3.2.1), NMDA receptors
intracerebral administration) do not elicit anxiolytic proper- containing NR2B subunits are concentrated in corticolim-
ties in the VCT and other procedures (Table 7) (Anthony and bic structures of the forebrain. This raises the possibility
Nevins, 1993; Schmitt et al., 1995; Chojnacka-Wójcik et al., that, by analogy to glycineB ligands, an alternative route
1996a; De Souza et al., 1998). Moreover, glycine and/or towards anxiolytic agents with improved separation of
NMDA have been reported to interfere with the anxiolytic beneficial properties to side-effects might be obtained by
actions of ACPC and d-cycloserine (Chojnacka-Wójcik selectively targeting NR2B subunits (Danysz and Parsons,
et al., 1996a; Teixeira and Carobrez, 1999; Klodzińska 1998; Dingledine et al., 1999; Chizh et al., 2001; Millan,
and Chojnacka-Wójcik, 2000). Fourth, paralleling the ac- 2002b; Loftis and Janowsky, 2003). The neuroleptic,
tions of glycineB partial agonists, recognition site antag- haloperidol (Section 4.5) has high affinity for NR2B sites
onists and open channel blockers, in the VCT, anxiolytic and the phenylethanolamine, ifenprodil, has long been
effects have been documented upon i.c.v. administra- known to preferentially interact with this subunit. Though
tion of the poorly brain-penetrant glycineB antagonist, its actions at ␣1 -ARs, sigma1 receptors and other sites
5,7-dichlorokynurenate, and of the systemically-active, po- compromise its utility as a pharmacological tool, several
tent and selective antagonist, L701,324 (Table 7) (Kotlinska highly-selective agents for NR2B subunits have now been
and Liljequist, 1998a). Curiously, and possibly reflect- described (Williams, 1993, 2001; Chizh et al., 2001; Millan,
ing procedural variables, (Karcz-Kubicka et al., 1997) 2002b; Higgins et al., 2003; Loftis and Janowsky, 2003).
failed to demonstrate anxiolytic actions of L701,324 or a In comparison to subunit non-selective antagonists, these
further antagonist, MRZ 21576, in the VCT, suggesting chemically-diverse NR2B subunit-selective antagonists
the need for further studies of pure antagonists in this show an improved therapeutic window between doses ex-
paradigm (Table 7). Nevertheless, 5,7-dichlorokynurenate erting, for example, analgesic actions (reflecting blockade
and the chemically-distinct glycineB receptor antagonists, of NR2B sites in the dorsal horn as well as the forebrain)
(+)HA-966 (a pyrrolidinone derivative) and ACEA-1021 as compared to motor disruption and psychomimetic prop-
(a quinoxalinedione derivative), have proven active in a erties (Millan, 1999; Chizh et al., 2001; Narita et al., 2001;
variety of models of anxiolytic properties, including other Higgins et al., 2003; Loftis and Janowsky, 2003).
conflict procedures (Corbett and Dunn, 1991, 1993; Kehne Long-term administration of BZPs modulates the corti-
et al., 1991; Dunn et al., 1992; Anthony and Nevins, 1993; colimbic density of NR2B sites (Van Sickle et al., 2002).
Wiley et al., 1995; Baron et al., 1997; Carobrez et al., Further, it appears that they are involved in the acceleration
2001; Pilc et al., 2002a). Microinjection studies indicate of cortical 5-HT release by NMDA receptors (Fink et al.,
(in line with observations mentioned in Section 3.2.3.1) 1995), though they may not contribute to the induction of
that at least one locus of action is the PAG (Matheus mesolimbic DA release by stress (Serrano et al., 1989).
et al., 1994). A fifth and final argument supporting the Hippocampal populations of NR2B receptors have been
notion that a decrease in activity at glycineB sites is as- implicated in the control of cognitive function under fear
sociated with anxiolytic properties is provided by the conditioning Tang et al. (2001) and they contribute to the
finding that mice with genetically-disturbed function at anxious behavior which accompanies withdrawal from
glycineB receptors show a reduction in anxiety (Kew et al., chronic alcohol administration (Stork et al., 2002b). How-
2000). ever, ifenprodil was, in contrast to competitive NMDA
Despite these arguments, it still remains to be convinc- receptor antagonists, ineffective in a pigeon conflict proce-
ingly demonstrated that glycineB antagonists do, indeed, dure (Koek and Colpaert, 1991) and in a rat Geiler–Seifter
elicit robust, clinically-relevant anxiolytic actions with a suf- paradigm (Wiley et al., 1998). Further, it did not display a
ficiently wide margin relative to doses provoking undesir- coherent pattern of specific anxiolytic actions in a murine
able actions (Pilc et al., 2002b). open-field procedure (Fraser et al., 1996; Pilc et al., 2002b).
It will, thus, be of importance to characterize the influence
3.2.4. Interaction with BZPs of novel, highly-selective NR2B antagonists in the VCT
As mentioned above (Section 3.2.1), there is functional and other paradigms to establish whether such agents in-
evidence for short-term and long-term interactions between deed promise to be anxiolytic agents of improved tolerance
glutamatergic and GABAergic systems in the control of relative to drugs which indiscriminately interact with all
anxious states. In line with such observations, the BZP antag- isoforms of NR2B subunits.
onist, flumazenil, attenuated anxiolytic actions of the recog-
nition site antagonist, CGP37849, and of the glycineB partial 3.3. AMPA receptors
agonist, ACPC, in the VCT and other procedures (Kuribara
et al., 1990; Sharma and Kulkarni, 1993; Klodzińska and Ionotropic, heteromeric, cation-permeable AMPA recep-
Chojnacka-Wójcik, 2000; Przegaliński et al., 2000). The pre- tors are composed of four subunits (GluR1–4), of which
cise substrates and mechanisms underlying such interactions their “flop” as compared to their alternatively-spliced
remain to be ascertained. “flip” isoforms predominate in adult CNS (Sommer
112 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

et al., 1990; Bleakman and Lodge, 1998). The pre- potential (Turski et al., 1992; Kim et al., 1993; Benvenga
cise composition of AMPA receptor subunits determines et al., 1995; Tizzano, 2002).
their Ca2+ -permeability and other biophysical proper- Corticolimbic regions responsible for integrating the
ties. In addition to the recognition site for glutamate, anxiolytic actions of AMPA antagonists remain to be defini-
AMPA receptors bear multiple allosteric site(s) responsive tively identified, but the hippocampus, PAG, septum and
to chemically-diverse positive and negative modulators amygdala are likely candidates (Kim et al., 1993; Matheus
(Hollmann and Heinemann, 1994; Miu et al., 2001; Arai and Guimaraes, 1997; Sajdyk and Shekhar, 1997; Menard
et al., 2002; Quirk and Nisenbaum, 2002; Ruel et al., and Treit, 2000; Martinez et al., 2002). The precise neuronal
2002). These sites control the kinetics of desensitization substrates subserving the anxiolytic properties of AMPA
and (following removal of AMPA) deactivation. receptor antagonists also remain to be delineated. How-
AMPA receptors play a ubiquitous role in the CNS in me- ever, congruent with anatomical findings evoked above,
diating fast excitatory transmission. In line with a role in together with a suppression of the stress-induced activation
the control of anxious states, AMPA receptors have been vi- of raphe-derived serotonergic projections, VTA-derived
sualized in high concentrations in monoaminergic cell clus- dopaminergic neurones and LC-derived noradrenergic
ters, and in the PAG, amygdala, septum and hippocampus pathways, a (postsynaptic?) reinforcement of GABAergic
(Ozawa et al., 1998; Lees, 2000). Acute and prolonged stress transmission in limbic structures is probably of signifi-
(including social stress) have been shown to elicit both im- cance (Jedema and Moghaddam, 1996; Rasmussen et al.,
mediate and more sustained (delayed) alterations in AMPA 1996; Fedele et al., 1997; Tao et al., 1997; Adell et al.,
receptor expression in the hippocampus and cortex, though 2002; Wu et al., 2002).
no consensus has as yet been reached as concerns the precise In light of the above observations, it is curious that the
nature of these changes. Indeed, there appears to be a com- potential anxiolytic actions of negative modulators at al-
plex pattern of alterations in individual subunits and their losteric sites on AMPA receptors do not appear to have been
flip–flop isoforms, with an increase in the relative propor- described. On the other hand, it has been reported that the
tion of the former suggestive of enhanced permeability to pyrrolidinone derivative and positive modulator, piracetam,
calcium (Krugers et al., 1993; Bartanusz et al., 1995; Geiger possesses anxiolytic properties in the VCT and certain
et al., 1995; Schwendt and Jezová, 2000; Rosa et al., 2002). other paradigms (Table 7) (File et al., 1979; Bhattacharyya
If corroborated, this finding would be of importance since et al., 1996, 1993) while the chemically-related anirac-
excessive stimulation of AMPA receptors is implicated in etam revealed anxiolytic properties in mice (Nakamura and
the ultrastructural, molecular and other detrimental changes Kurasawa, 2001). Since there is no obvious mechanistic
associated with long-term exposure to stress (Magarinos and interpretation for these observations, they would merit sub-
McEwen, 1995; Rosa et al., 2002). It is possible that corti- stantiation (or refutation) with other classes (benzamide and
costeroids, which are likewise incriminated in these events, benzodiathiazine) of more selective, positive modulators at
contribute to the recruitment of AMPA receptors: further, AMPA receptors (Arai et al., 2002; Quirk and Nisenbaum,
they may exert their deleterious actions cooperatively with 2002).
those of glutamate at AMPA sites (Section 11) (Magarinos
and McEwen, 1995). 3.4. Kainate receptors
The above observations provide a framework for partic-
ipation of AMPA receptors in the emotional and cognitive Ionotropic, neuronal kainate receptors are, like their
changes elicited by fear and stress. Though remarkably few AMPA counterparts, composed of various combinations of
studies have addressed this question, activation of AMPA re- subunits (KA-1, KA-2 and GluR5–7). GluR5–7 subunits
ceptors is involved in the induction of dependence to BZPs, can assemble into functional homo- and heteromeric re-
as well as the rebound anxiety and other symptoms pro- ceptors, but KA-1 and KA-2 cannot and appear to confer
voked by their abrupt withdrawal (Steppuhn and Turski, specific pharmacological properties upon associations of
1993; Jackson et al., 2000; Izzo et al., 2001; Van Sickle GluR5–7 subunits (Bleakman, 1999; Dingledine et al., 1999;
and Tietz, 2002). In line with these findings, Kotlinska and Lerma et al., 2001; Swanson et al., 2002). Though gating
Liljequist (1998b) reported a significant increase in punished properties of kainate receptors differ from those of AMPA
response in the VCT with the highly-selective antagonist, receptors, structure-activity relationships for glutamatergic
LY326,325, at doses which failed to modify locomotor func- ligands have proven similar. This has complicated the dis-
tion or to elicit other non-specific actions (Table 7). Paradox- covery of selective agents permitting dissociation of their
ically, then, Czlonkowska et al. (1997) failed to demonstrate functional roles (Bleakman, 1999; Dingledine et al., 1999).
anxiolytic actions of the first generation antagonist, NBQX, Moreover, though novel drugs interacting selectively at
in a VCT, but the conditions of study may not have been kainate subunits, such as the GluR5 antagonist, LY382,884,
optimal for their detection (Table 7). Indeed, anxiolytic ac- have recently been disclosed, data concerning their actions
tions of these and chemically-distinct AMPA antagonists in in the VCT or other models of anxiolytic activity are not,
conflict (and other) procedures in mice and pigeons under- as yet, available (Conti et al., 2002). Such studies would be
pin the utility of the VCT for evaluation of their anxiolytic of considerable interest in light of: (1) the above-discussed
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 113

roles of AMPA and NMDA receptors in the control of 2002). Correspondingly, while the influence of group II
anxious states; (2) evidence for functionally-active, post- mGluR receptors upon 5-HT release in the hippocampus
and presynaptic (on glutamatergic fibers) kainate receptors and amygdala remains to be described, their induction of
in the hippocampus and amygdala and (3) a complex pat- 5-HT release in the FCX may reflect a disinhibitory action
tern of influence of kainate receptors upon corticolimbic effected via a reduction in tonic, GABAergic tone (Lee and
GABAergic and monoaminergic transmission (Bleakman, Croucher, 2003). The induction of 5-HT release by group
1999; Kathmann et al., 1999; Frerking and Nicoll, 2000; II/III mGluR agonists in the PAG may likewise reflect re-
Van Bockstaele, 2000; Lerma et al., 2001; Szabo and Blier, lief of GABAergic inhibition of serotonergic neurones, and
2001; Braga et al., 2003; Ghersi et al., 2003). this action may be related to their suppression of uncon-
ditioned, aversive (panic-like) states (Sections 1.5.3.2 and
3.5. Metabotropic receptors 4.3.4.4) (Graeff et al., 1993; Maione et al., 1998; Cartmell
and Schoepp, 2000). In contrast to group II mGluR recep-
3.5.1. Multiple classes of functionally-heterogeneous tors, no evidence has, as yet, been forwarded for a role of
metabotropic receptor group I mGluR receptors in the modulation of 5-HT re-
Metabotropic (mGlu) receptors are functional homod- lease (Maione et al., 1998; Lee and Croucher, 2003), though
imers (Kunishima et al., 2000) which can be divided into their facilitatory influence upon GABA and glutamate re-
three classes. Group I (mGluR1 and 5) receptors are posi- lease likely affects anxious states (De Novellis et al., 2003).
tively coupled via Gq to phospholipase C. Their activation Finally, the suppressive influence of group II mGluR re-
promotes the activity of colocalized NMDA receptors via ceptors upon hippocampal release of CCK (Section 9.1.1)
are a complex interplay of mechanisms ultimately involv- (Breukel et al., 1998; Hu et al., 1999) is consistent with anx-
ing (for GluR5 receptors) their protein kinase C-mediated iolytic actions in behavioral models (Section 3.5.2.1).
tyrosine phosphorylation. In contrast, group II (mGluR2
and mGluR3) and group III (mGluR4 and mGluR6–8) 3.5.2. Anxiolytic profiles of ligands at metabotropic
receptors are negatively coupled via Gi/o to adenylyl cy- mGluR receptors
clase (Conn and Pin, 1997; Bordi and Ugolini, 1999; Pin
et al., 1999, 2001; Schoepp et al., 1999; Bräuner-Osborne 3.5.2.1. Group II/III mGluR agonists. The preceding re-
et al., 2000; Schoepp, 2001; Benquet et al., 2002). Inas- marks provide a framework for observations that group II/III
much as inhibitory classes II and III receptors are localized mGluR agonists display, like NMDA and AMPA antago-
both pre- and postsynaptically with respect to glutamater- nists (Sections 3.2.2 and 3.3, respectively), anxiolytic prop-
gic pathways, agonists at these sites functionally counter erties in conflict and other models (Chojnacka-Wójcik et al.,
actions mediated by excitatory postsynaptic, ionotropic re- 1996a,b; Helton et al., 1998; Benvenga et al., 1999; Sporeen
ceptors both directly and indirectly—by suppressing the et al., 2000; Tatarczyńska et al., 2001b; Collado et al., 2002;
(stress-induced) release of glutamate. On the other hand, Tizzano et al., 2002) including a paradigm of panic-like be-
antagonists at excitatory group I receptors postsynaptic havior (Shekhar and Keim, 2002). In exerting its anxiolytic
and (possibly) presynaptic to glutamatergic pathways may actions, the group II (mGluR2/3) agonist, LY354,740, ap-
likewise oppose actions elicited by their ionotropic coun- pears to interrupt glutamatergic transmission in the amyg-
terparts (Karreman and Moghaddam, 1996; Stefani et al., dala via pre- and postsynaptic actions—and to act without
1998b; Cartmell and Schoepp, 2000; Thomas et al., 2001; modifying cognitive performance (Stanek et al., 2000;
Corti et al., 2002; Xi et al., 2002; De Novellis et al., 2003). Tizzano et al., 2002). LY354,740 is active in the rat VCT
All three classes of metabotropic receptor are enriched in both upon systemic administration and (in analogy to ag-
the hippocampus (Pin and Duvoisin, 1995; Blumcke et al., onists which fail to penetrate the blood-brain barrier) upon
1996; Shigemoto et al., 1997; Pin et al., 1999, 2001). Further introduction into the CA1 region of the dorsal hippocampus
group I mGluR5 receptors, as well as group II mGluR2 and 3 (Table 8 and Fig. 3) (Tatarczyńska et al., 2001b). Though the
receptors, are well-represented in the cortex, nucleus accum- amygdala has not to date been examined by use of the VCT,
bens and amygdala (Fotuhi et al., 1994; Bordi and Ugolini, this structure represents a substrate for anxiolytic properties
1999; Spooren et al., 2001; Schoepp and Marek, 2002). Of of group II agonists in a fear-potentiated startle paradigm
particular note, engagement of group II mGluR receptors (Walker et al., 2002). A group III mGluR agonist, “l-SOP”,
attenuates the anxiety-associated activation of LC-derived was likewise active in a rat VCT upon injection into the hip-
noradrenergic neurones and blunts the induction of corti- pocampus, but this result would benefit from corroboration
cal release of NA by stress (Vandergriff and Rasmussen, with more selective ligands inasmuch as a further serine ana-
1999; Schoepp, 2001; Schoepp and Marek, 2002). On the logue which behaves as an antagonist at group III mGluR
other hand, though conflicting data have been documented, sites, MSOP, likewise showed anxiolytic properties (Table 8
a majority of studies suggest that group II mGluR receptors and Fig. 3) (Chojnacka-Wójcik et al., 1997; Tatarczyńska
moderate hippocampal release of GABA, an action incon- et al., 2001b). It is currently controversial as to whether
sistent with the expression of anxiolytic properties (Cartmell GABAergic mechanisms participate in the expression of
and Schoepp, 2000; Ferris et al., 2001; Schoepp and Marek, anxiolytic actions of group II and III mGluR agonists (Ferris
114 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Table 8
Influence of ligands acting at metabotropic glutamatergic receptors upon behavior in the Vogel Conflict Test
Drug Class Species Route (structure) Effect Active dose range Maximum Reference
(strain) change (%)
(S)-4C3HPG mGluRI Ant Rat (W) i.c. (hippocampus) + 6.5 ␮g +225 Chojnacka-Wójcik et al. (1997)
mGluRII agonist
MPEP mGluRI Ant Rat (W) i.p. + 1 and 10 mg/kg +460 Tatarczynska et al. (2001a)
(S)-4CPG mGluRI Ant Rat (W) i.c. (hippocampus) + 20 ␮g +257 Tatarczynska et al. (2001b)
CPCCOEE mGluRI Ant Rat (W) i.c. (hippocampus) + 5 and 15 ␮g +264 Tatarczynska et al. (2001b)
LY354,740 mGluRII agonist Rat (W) i.p. + 0.5 and 1.0 mg/kg +200 Klodzińska et al. (1999)
Rat (W) i.c. (hippocampus) + 1 ␮g +230 Tatarczynska et al. (2001b)
l-CCG-1 mGluRII agonist Rat (W) i.c. (hippocampus) + 10 ␮g +242 Tatarczynska et al. (2001b)
MSOPPE mGluRII Ant Rat (W) i.c. (hippocampus) IA 2.5 and 6.5 ␮g IA Chojnacka-Wójcik et al. (1997)
l-SOP mGluRIII agonist Rat (W) i.c. (hippocampus) + 100 ␮g +164 Tatarczynska et al. (2001b)
MSOP mGluRIII Ant Rat (W) i.c. (hippocampus) + 6.5 ␮g +125 Chojnacka-Wójcik et al. (1997)
Abbreviations—(S)-4C3HPG: (S)-4-carboxy-3-hydroxyphenylglycine; MPEP: 2-methyl-6-(phenylethynyl)pyridine; (S)-4CPG: (S)-4-carboxyphenylglycine;
CPCCOEE: 7-(hydroxyimino)cyclopropa[b]chromen-1a-carboxylate ethyl ester; l-CCG-1: (2S,1 S,2 S)-2-(carboxycyclopropyl)glycine; l-SOP:
l-serine-O-phosphate; MSOP: (RS)-␣-methylserine-O-phosphate; and mGluRI-III: groups I, II and III metabotropic receptor, respectively. For other
abbreviations, see Table 2.

et al., 2001; Schoepp and Marek, 2002; Tizzano et al., tion of anxious states is not confined to short-term events:
2002). rather, they may also be involved in more sustained alter-
ations in emotionality and cognition of pertinence to clinical
3.5.2.2. Group I mGluR antagonists. As concerns anxiety disorders (Riedel et al., 1996, 2002; Schoepp et al.,
group I mGluR antagonists, parenteral administration of 1999). Though anxiolytic actions of mGluR5 antagonists in
2-methyl-6-(phenylethynyl)pyridine—a selective (non-com- the VCT and other models do not appear to reflect an in-
petitive) mGluR5 antagonist—or intrahippocampal appli- fluence upon mnesic function, these observations also serve
cation of other (lipophobic) antagonists, elicited anxiolytic to emphasize the fundamental notion that specific classes of
actions in the VCT as well as other conflict and non-conflict modulator (receptor) may be involved in diverse aspects of
procedures (Table 8 and Fig. 3) (Chojnacka-Wójcik et al., the response to stress and fear (Section 18.8).
1996b; Klodzińska et al., 1999, 2000; Spooren et al., 2000,
2001; Schulz et al., 2001; Tatarczyńska et al., 2001a,b; 3.6. Multiple glutamatergic mechanisms for anxiolysis
Brodkin et al., 2002; Pilc et al., 2002a,b; Schulz et al.,
2002). For these drugs, anxiolytic actions were seen both It appears, then, that several receptorial approaches for
upon acute and upon chronic administration (Pilc et al., the attenuation of excessive glutamatergic transmission
2002a,b) and were expressed at doses which did not perturb may—by analogy to multiple strategies for enhancement
cognitive function, motor performance or drinking behavior of GABAergic transmission—be associated with anxiolytic
(op. cit.). These results indicate that blockade of mGluR5 properties in the VCT and other conflict procedures. How-
receptors is associated with anxiolytic properties, mimick- ever, data for glutamatergic agents require consolidation
ing the above-discussed anxiolytic properties of NMDA with highly selective ligands (in particular, for subtypes of
receptor antagonists in the VCT (Section 3.2.2). Though metabotropic receptor), the precise neuronal mechanisms via
postsynaptic mGluR5 sites principally mediate these effects, which glutamatergic pathways control anxious states await
blockade of presynaptic populations excitatory to glutamate clarification, and the genuine therapeutic utility of glutama-
release may also be involved (Thomas et al., 2000, 2001). tergic drugs remains to be established (Kinsey et al., 2001).
This parallel between group I mGluR metabotropic and
NMDA receptor antagonists is notable in light of the co-joint
role of postsynaptic NMDA and mGluR5 receptors in medi- 4. Monoamines
ating neuronal excitation. Moreover, together with NMDA
receptors, mGluR5 receptors are required for the acquisi- 4.1. Noradrenaline
tion of fear-conditioning and long-term potentiation in the
amygdala and, perhaps, the hippocampus, structures impli- 4.1.1. Noradrenergic pathways
cated in glutamatergic mechanisms for the control of anxiety Ascending noradrenergic projections heavily innervate the
(Alagarsamy et al., 1999; Anwyl, 1999; Schulz et al., 2001, hippocampus, amygdala, PAG, cortex, hypothalamus and
2002; Fendt and Schmid, 2002; Rodrigues et al., 2002). essentially all corticolimbic regions involved in integrat-
This common role of metabotropic (mGluR5) and ionotropic ing the response to anxiety Lindvall and Björklund, 1984;
(NMDA) receptors in synaptic plasticity under conditions Aston-Jones et al., 1991; Valentino and Aston-Jones, 1995;
of stress suggests, moreover, that their role in the modula- Schatzberg and Schildkraut, 1995; Tanaka et al., 2000. A
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 115

substantial proportion of this input derives from the LC et al., 1992; Ferry et al., 1997; Boehm, 1999; Forray et al.,
(Lindvall and Björklund, 1984; Aston-Jones et al., 1991; 1999; Pralong et al., 2002).
Valentino and Aston-Jones, 1995). The marked and sus- Such neurochemical effects of ␣2 -AR agonists and par-
tained activation of noradrenergic pathways arising in this tial agonists likely participate in the expression of their
cell cluster by a confluence of anxiogenic and other stress- anxiolytic actions in procedures employing untrained be-
ful stimuli (Bremner et al., 1996a,b; McQuade et al., 1999; haviors (Handley and Mithani, 1984; Fontana et al., 1990;
Sands et al., 2000; Feenstra et al., 2001; Ishida et al., 2002; Lopez-Rubalcava and Fernandez-Guasti, 1994; Cole et al.,
McIntyre et al., 2002a) is accompanied by emotional, cog- 1995a,b; Sanchez, 2003), conflict procedures (Söderpalm,
nitive and autonomic manifestations of fear (Tanaka et al., 1989; Fontana et al., 1990) and the VCT (Table 9 and Fig. 6)
2000; Pardon et al., 2002; Schulz et al., 2002; Shekhar et al., (Gower and Tricklebank, 1988; MacDonald et al., 1988;
2002), and may be of particular significance to the induction Söderpalm, 1989; La Marca and Dunn, 1994; Söderpalm
of panic attacks (Uhde et al., 1984b; Gorman et al., 1989, et al., 1995a,b; Millan et al., 2000b,e). However, the anxi-
2000; Coupland et al., 1996; Sullivan et al., 1999; Shekhar olytic actions of ␣2 -AR agonists are generally exerted over a
et al., 2002). Accordingly, in response to fear and other stres- confined dose range. There are several possible explanations
sors, an acceleration of NA release has been quantified in for this observation. First, at high doses, ␣2 -AR agonists be-
several corticolimbic regions, including the FCX, amygdala gin to activate relatively-insensitive populations of ␣2 -ARs
and hippocampus (op. cit.). As emphasized throughout this postsynaptic to monoaminergic pathways, the engagement
article, then, the LC assimilates a vast number of conver- of which might, in theory, mediate anxiogenic properties.
gent and neurochemically-distinct inputs transducing the ef- However, there is currently little evidence to support this
fects of, and controlling the response to, anxiogenic stimuli hypothesis. Second, no ␣2 -AR agonist described to date
(Aston-Jones et al., 1991; Kawahara et al., 2000; Singewald shows absolute selectivity for ␣2 -ARs versus ␣1 -ARs and it
and Sharp, 2000; Lapiz et al., 2001). might be conjectured that the biphasic dose–response rela-
It is essential to consider the role of multiple classes of tionships of clonidine and other ␣2 -AR ligands in the VCT
␣1 -, ␣1 - and ␤-AR (Bylund et al., 1994; Hieble et al., 1995; reflect recruitment of anxiogenic, postsynaptic ␣1 -ARs at
Hieble, 2000; Piascik and Perez, 2001; Audinot et al., 2002) high doses (Table 9) (Söderpalm and Engel, 1988; Hieble
in the modulation of anxious states. ␣1A , ␣1B and ␣1D -AR et al., 1995; Millan et al., 2000e). Arguing against this
subtypes are positively coupled via Gq to phospholipase C, contention, however, is the general inactivity of ␣1 -AR lig-
the activation of which leads to an increase in intracellular ands in the VCT (Table 9) (Section 4.1.5). Third, a more
levels of Ca2+ : this excitatory influence is amplified by convincing and parsimonious explanation for the generally
their engagement of VDCCs. Contrariwise, the inhibitory bell-shaped dose–response curves of ␣2 -AR agonists in the
influence of ␣2A - (and its rodent homologue, ␣2D -), ␣2B - VCT and other models invokes a perturbation of motor
and ␣2C -ARs upon neuronal activity reflects their positive function at high doses. Indeed, a narrow therapeutic win-
coupling via Gi to adenylyl cylase, their suppression of dow between anxiolytic and motor-disruptive dose ranges
the activity of VDCCs and their opening of K+ -channels for ␣2 -AR agonists is hardly illogical since both these ac-
(Section 13). CNS-localized ␤1 - and ␤2 -ARs may be distin- tions reflect activation of ␣2 -ARs inhibitory to monoamin-
guished from ␣2 -ARs by their activation, via Gs, of adenylyl ergic pathways (Kable et al., 2000; Millan et al., 2000d,e;
cyclase. Millan, 2002a).
Evidently, the restricted range of active, non-sedative
4.1.2. α2 -AR agonists doses of ␣2 -AR (partial) agonists precludes their broad
␣2 -AR autoreceptors exert a tonic, inhibitory influence exploitation as anxiolytic agents. Nevertheless, there are cir-
upon the cerebral (and sympathetic) liberation of NA. Fur- cumstances under which the sedative—as well as haemosta-
ther, stimulation of ␣2 -AR heteroceptors brakes the release bilizing and analgesic—actions of ␣2 -AR agonists are
of 5-HT in corticolimbic structures: though it remains con- desirable: notably, the peri-operative environment. Corre-
troversial as to whether this control is exerted tonically or spondingly, clinical evidence has accumulated for mean-
phasically, the majority of evidence favors the latter pos- ingful anxiolytic actions of ␣2 -AR agonists upon their ad-
sibility (Hein et al., 1999; Bengtsson et al., 2000; Kable ministration prior to surgery (Ahmed and Takeshita, 1996;
et al., 2000; Millan et al., 2000b,c,d,e,f). These populations Thomson et al., 1998; Bitsios et al., 1998; Millan, 2002a).
of ␣2 -AR are considerably more sensitive than their postsy- In addition, the anxiolytic properties of ␣2 -AR agonists
naptic counterparts. As a consequence, ␣2 -AR agonists of contribute to their efficacy in the clinical management of
even modest efficacy diminish extracellular levels of NA and opioid withdrawal (Buccafusco, 1992; Devoto et al., 2002;
5-HT in corticolimbic structures—as well as DA release in Stine et al., 2002).
the frontal cortex and, under certain conditions, the nucleus
accumbens (Millan et al., 2000b,d,e; Tuinstra and Cools, 4.1.3. Significance of α2 -AR subtypes
2000; Adell et al., 2002). In addition, ␣2 -ARs have been
shown to suppress the release of glutamate in the cortex, 4.1.3.1. A key role for α2A -ARs. It is important to address
amygdala, hippocampus and other limbic regions (Kamisaki the question of the role of individual ␣2 -AR subtypes in
116 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Table 9
Influence of drugs interacting with ␣2 - and ␣1 -adrenoceptors upon behavior in the Vogel Conflict Test
Drug Class Species Route Effect Active dose range Maximum Reference
(strain) (structure) change (%)
S18616 ␣2 -AR agonist Rat (W) s.c. + 0.005 mg/kg +212 Millan et al. (2000b)
DMT ␣2 -AR agonist Rat (W) s.c. + 0.005 mg/kg +166 Millan et al. (2000b)
Medetomidine ␣2 -AR agonist Rat (W) s.c. + 0.001–0.005 mg/kg NI MacDonald et al. (1988)
␣2 -AR agonist Rat (SD) i.v. IA 0.002–0.005 mg/kg IA La Marca and Dunn (1994)
Guanabenz ␣2 -AR agonist Rat (SD) i.v. IA 0.03–0.1 mg/kg IA La Marca and Dunn (1994)
Guanfacine ␣2 -AR agonist Rat (SD) i.v. IA 0.1–0.5 mg/kg IA La Marca and Dunn (1994)
Clonidine ␣2 -AR PAGO Rat (SD) i.p. + 0.006 mg/kg +60 Söderpalm and Engel (1988)
Rat (SD) i.p. IA 0.01–0.1 mg/kg IA Gower and Tricklebank (1988)
Rat (WKY) i.p. + 0.025 mg/kg +40 Söderpalm and Engel (1989)
Rat (SHR) i.p. + 0.003 mg/kg +35 Söderpalm and Engel (1989)
Rat (W) s.c. IA 0.003–0.05 mg/kg IA Söderpalm and Engel (1989)
Rat (W) s.c. + 0.02 mg/kg +175 Millan et al. (2000b)
Rat (SD) i.v. + 0.02 mg/kg +210 La Marca and Dunn (1994)
Yohimbine ␣2 -AR Ant Rat (SD) i.p. IA >0.5 mg/kg IA Söderpalm and Engel (1990)
Rat (SD) i.p. + 4.0 mg/kg +90 Söderpalm et al. (1995a)
Rat (LH) i.p. + 2.5 mg/kg +85 Baldwin et al. (1989)
Rat (SD) s.c. + 0.25–2.0 mg/kg +870 Gower and Tricklebank (1988)
Rat (SD) i.v. IA 1.0 and 2.2 mg/kg +870 La Marca and Dunn (1994)
Idazoxan ␣2 -AR Ant Rat (SD) i.p. IA 0.5–2.0 mg/kg IA Söderpalm and Engel (1989)
Rat (SD) s.c. + 0.25–2.0 mg/kg +563 Gower and Tricklebank (1988)
Rat (SD) i.v. + 10.0 mg/kg +460 La Marca and Dunn (1994)
Piperoxan ␣2 -AR Ant Rat (SD) i.v. IA 1.0–5.0 mg/kg IA La Marca and Dunn (1994)
Rauwolscine ␣2 -AR Ant Rat (SD) i.v. + 2.2 mg/kg + 290 La Marca and Dunn (1994)
Atipamezole ␣2 -AR Ant Rat (W) s.c. IA >0.63 mg/kg IA Brocco (unpublished observations)
1-PP ␣2 -AR Ant Rat (SD) s.c. + 1.0–4.0 mg/kg +128 Gower and Tricklebank (1988)
WY26392 ␣2 -AR Ant Rat (SD) s.c. + 1.0 and 2.0 mg/kg +307 Gower and Tricklebank (1988)

ST 587 ␣1 -AR agonist Rat (SD) i.p. IA 0.5 and 2.0 mg/kg IA Söderpalm and Engel (1990)
Prazosin ␣1 -AR Ant Rat (W) i.p. IA 0.5 and 1.0 mg/kg IA Przegaliński et al. (1994b)
Rat (SD) i.p. IA >0.25 mg/kg IA Söderpalm and Engel (1990)
Rat (W) i.c. (hippocampus) IA 0.3–1.0 ␮g IA Przegaliński et al. (1994b)
Abbreviations—DMT: dexmedetomidine; 1-PP: 1-(2-pyrimidyl)piperazine; PAGO: partial agonist; WKY: Wistar Kyoto; SHR: spontaneously hypertensive;
IA: no significant change (inactive) and NI: not indicated. For other abbreviations, see Table 2.

the anxiolytic actions of ␣2 -AR agonists. Anatomical, phar- 2002). Of particular note, activation of ␣2A -ARs has
macological and genetic approaches all converge upon the been shown to blunt the cortical release of NA and DA
␣2A -AR subtype as of key significance in this respect. elicited under conditions of fear (Ihalainen and Tanila,
First, pre- and postsynaptic populations of ␣2A -AR are en- 2002). Third, the inhibitory influence of NA upon hip-
riched in many corticolimbic structures, including the amyg- pocampal liberation of glutamate is mediated by ␣2A -ARs
dala, PAG, cortex and hippocampus (Nicholas et al., 1996; (Boehm, 1999). Fourth, alterations in levels of (probably
Talley et al., 1996; Wang et al., 1996b; Milner et al., 1998). presynaptic) ␣2A -ARs in the hippocampus, amygdala and
Though little is known concerning neuronal substrates PAG have been detected upon acute and long-term expo-
transducing the anxiolytic effects of ␣2A -AR agonists, apart sure to stress (Nukina et al., 1987; Tejani-Butt et al., 1994;
from the LC, microinjection studies indicate that the amyg- Flügge, 1996; Flügge et al., 1997; Fulford and Marsden,
dala represents one major locus of action (Fendt et al., 1994; 1997; Fuchs and Flügge, 2002). Further, clinical studies
Glass et al., 2002). Second, both in rodents and in man, indicate that psychological stress modifies the functional
␣2A -AR autoreceptors exert a suppressive influence upon status of peripheral, platelet-localized ␣2A -ARs which serve
the cerebral release of NA, while ␣2A -heteroceptors mediate as a surrogate marker for their supraspinal counterparts
the inhibitory control exercised by noradrenergic pathways (Maes et al., 2002). Fifth, mice with a “null mutation” for
upon the corticolimbic release of 5-HT and DA (Palij and ␣2A -ARs reveal an anxious phenotype in behavioral models,
Stamford, 1996; Altman et al., 1999; Hein et al., 1999; a marked elevation in central turnover of NA and autonomic
Linner et al., 1999; Feuerstein et al., 2000; Kable et al., signs of anxiety (Schramm et al., 2001; Lähdesmäki et al.,
2000; Millan et al., 2000d,e; Hein, 2001; Hopwood and 2002). Finally, pharmacological studies employing the VCT
Stamford, 2001b; Scheibner et al., 2001; Bücheler et al., and other models have shown that selective blockade of
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 117

750
LICKS/3 MIN

500

250

0
VEH VEH 0.00125 0.0025 0.005 0.01 MG/KG, S.C.
DEXMEDETOMIDINE

NO SHOCK SHOCK

750
LICKS/3 MIN

500

250

0
VEH VEH 0.63 2.5 10.0 MG/KG, S.C.
REBOXETINE

NO SHOCK SHOCK

Fig. 6. Actions of the ␣2 -adrenoceptor agonist, dexmedetomidine, and of the noradrenaline reuptake inhibitor, reboxetine, in the Vogel Conflict Test.
Data are means ± S.E.M.s. Star indicates significance of differences to corresponding vehicle values. ( ) P < 0.05 to vehicle. Data are from Millan
et al. (2000e), and Brocco and Millan (unpublished observations).

␣2A -ARs, but not of ␣2B - or ␣2C -ARs, abrogates the anxi- other characteristics, are unlikely to express anxiolytic ac-
olytic properties of ␣2 -AR agonists (Millan et al., 2000e). tions in the absence of motor sedation and/or a perturbation
A predominant role of ␣2A -ARs in the suppression of of cardiovascular function (Millan et al., 2000b,e). Though
monoaminergic transmission may, then, underlie their me- a procognitive role for postsynaptic ␣2A -ARs in the frontal
diation of the anxiolytic properties of ␣2 -AR agonists. cortex has been advanced (Björklund et al., 1999, 2001;
However, this mechanism similarly accounts for their Friedman et al., 1999; Franowicz and Arnsten, 2002), this
motor-sedative actions (Kable et al., 2000; Millan et al., is of dubious therapeutic utility since: (1) it is primarily ex-
2000d,e; Hein, 2001; Franowicz and Arnsten, 2002). More- pressed under conditions where noradrenergic transmission
over, activation of central and peripheral populations of is reduced (in contrast to anxious states) and (2) ␣2A -ARs
␣2A -ARs inhibitory to sympathetic outflow underlies the agonists markedly suppress cholinergic and glutamatergic
hypotensive properties of ␣2A -AR agonists (Hein, 2001; Tan transmission (Section 4.1.2), actions incompatible with an
et al., 2002a). Indeed, ␣2A -ARs are localized on amygdala improvement of cognitive function (Tellez et al., 1997;
output neurones which project to the dorsal vagal complex, Forray et al., 1999; Linner et al., 1999; Li et al., 2001b;
and on catecholaminergic neurones in the nucleus tractus Shirazi-Southall et al., 2002; Gobert et al., 2003).
solitarious, a structure which likewise fulfils a crucial role
in integrating the autonomic response to stress (Kable et al., 4.1.3.2. α2C and α2B -ARs. A role for ␣2B -ARs in the
2000; Glass et al., 2001, 2002; Hein, 2001). control of anxious states appears unlikely inasmuch as they
In accordance with the above observations, ␣2A -AR sub- are sparse in the limbic system and do not appear to play
type selective agonists, irrespective of their efficacy and a role in the modulation of monoaminergic or GABAergic
118 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

transmission (Nicholas et al., 1996; Wang et al., 1996b; though the anxiogenic properties of idazoxan in man are
Hein, 2001). On the other hand, under certain condi- less pronounced than those of yohimbine (Table 9) (Krystal
tions, ␣2C -ARs may complement the inhibitory control et al., 1992; Hieble et al., 1995; Schmidt et al., 1999;
by ␣2A -ARs of corticolimbic and sympathetic monoamine Millan et al., 2000d,f). Second, as mentioned above (Section
release (Sallinen et al., 1997; Lee et al., 1998; Altman 4.1.2), any potential role of ␣2 -ARs postsynaptic to nora-
et al., 1999; Hein et al., 1999; Scheibner et al., 2001; drenergic networks in the control of anxious states remains
Bücheler et al., 2002; Ihalainen and Tanila, 2002). More- to be explored. Thus, it might be contended that anxiolytic
over, ␣2C -ARs postsynaptic to noradrenergic pathways properties of yohimbine and idazoxan reflect blockade of
are concentrated in the hippocampus, septum, amygdala “anxiogenic”, postsynaptic ␣2 -AR sites. Arguing against
and nucleus accumbens (Rosin et al., 1996; Dossin et al., this assumption, however, the highly-selective ␣2 -AR an-
2000). Nevertheless, though ␣2C -ARs may be involved in tagonists, fluparoxan, RX821,002 and atipamezole, do
the control of cognitive function, aggressive behavior and not show anxiolytic (or anxiogenic) actions in the VCT
the emotional and endocrine response to stress, any spe- (Table 9) (Brocco and Millan, unpublished observations).
cific role in the modulation of anxious states remains to be Third, in view of their modest selectivity at ␣2 -ARs versus
directly demonstrated (Sallinen et al., 1997, 1998, 1999; ␣1 -AR sites, yohimbine and idazoxan might exert anxiolytic
Björklund et al., 1999; Hein, 2001; Schramm et al., 2001). actions via blockade of postsynaptic ␣1 -ARs. However,
as outlined below (Section 4.1.5), evidence that activation
4.1.4. α2 -AR antagonists of postsynaptic ␣1 -ARs enhances anxious states is hardly
The ␣2 -AR antagonist, yohimbine, elicits anxiety overwhelming. Finally, inasmuch as ␣2 -AR antagonists
in man, most strikingly in (“trait”) anxious subjects, exert a modest facilitatory influence upon motor function
opioid-dependent patients and individuals susceptible to (Hieble, 2000; Millan et al., 2000e), it is conceivable that
panic attacks (Charney et al., 1983; Krystal et al., 1992; their “disinhibitory” properties contribute to the increase in
Bremner et al., 1996a,b; Stine et al., 2002). Robust anx- responding in the VCT.
iogenic actions of yohimbine have also been detected in Irrespective of the mechanisms underlying the anxiolytic
several (though not all) experimental models of untrained profile of yohimbine in the VCT, it must be conceded that
behavior (Handley and Mithani, 1984; Johnston and File, this procedure does not reveal its clinically-expressed, anx-
1988; Baldwin et al., 1989; Venault et al., 1993; Redfern iogenic properties. This discrepancy may—by analogy to
and Williams, 1995; De Boer and Koolhaas, 2003; see a further class of clinically-active anxiogenic compounds,
Cole et al., 1995a,b). Such actions involve activation of adenosine antagonists (Section 7.2)—reflect the fact that
the amygdala (Schroeder et al., 2003) and may reflect an these agents trigger “panic-like” states in man which are
induction of corticolimbic release of NA and DA (Millan poorly modeled by the VCT (Section 1.4).
et al., 2000f) together with a suppressive influence upon
the activity of GABAergic neurones (Maurin et al., 1985). 4.1.5. Role of α1 -ARs
However, the latter action awaits confirmation and several
authors have questioned the significance of ␣2 -AR receptor 4.1.5.1. Modulation of anxious states. Despite indications
blockade to the anxiogenic actions of yohimbine ((Handley for a small population of ␣1 -ARs in the LC, they are pre-
and Mithani, 1984; Johnston and File, 1988; Redfern and dominantly localized postsynaptically to noradrenergic neu-
Williams, 1995); see Cole et al., 1995a,b). Complicating rones in diverse corticolimbic territories (Section 4.1.5.2)
matters further, while certain investigators have documented (Nicholas, et al., 1996). Several reports of anxiolytic ac-
anxiogenic properties of yohimbine in conflict models, most tions of ␣1 -AR antagonists have appeared, though they are
studies, including those of the VCT, have reported anxi- generally ineffective in conflict models and have, with a
olytic actions (Table 9) (Gardner and Piper, 1982; Gower singular, unquantified exception (Gardner and Piper, 1982),
and Tricklebank, 1988; Baldwin et al., 1989; Söderpalm proven inactive in the VCT (Table 9) (Söderpalm and Engel,
and Engel, 1989, 1990; La Marca and Dunn, 1994; 1990; Przegaliński et al., 1994b; Sanchez, 2003; Brocco
Söderpalm et al., 1995a,b; see Cole et al., 1995a,b). A va- and Millan, unpublished observations). Likewise, selec-
riety of hypotheses have been formulated in an effort to tive ␣1 -AR agonists are essentially inert in this and other
explain these paradoxical anxiolytic actions of yohimbine. conflict procedures (Table 9) (Handley and Mithani, 1984;
First, the most compelling advocates a role of 5-HT1A Söderpalm and Engel, 1988, 1990; Salonen et al.,
autoreceptors, inasmuch as their stimulation by yohimbine 1992; La Marca and Dunn, 1994; Lopez-Rubalcava and
has been implicated in its reduction of cerebral 5-HT release Fernandez-Guasti, 1994; Söderpalm et al., 1995a,b). Never-
and several other of its functional actions (Söderpalm et al., theless, activation of ␣1 -ARs in the amygdala has been im-
1995a,b; Millan et al., 2000f; Shannon and Lutz, 2000). plicated in the induction of anxiety by stress (Cecchi et al.,
A similar combination of ␣2 -AR blockade and 5-HT1A 2002) and anxiogenic actions of centrally-administered
receptor agonism might account for the variable anxiolytic ␣1 -AR agonists in rodents have been attributed to recruit-
and anxiogenic effects of other ␣2 -AR antagonists, such as ment of CRF1 receptors (Berridge and Dunn, 1989; Yang
idazoxan, in the VCT and other experimental procedures— et al., 1990; Carrasco and Van de Kar, 2003). Further, it
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 119

has been suggested that the anxiogenic phenotype of mice they are, thus, strategically localized for engagement of
genetically lacking angiotensin (AT2 ) sites reflects activa- the hypothalamo-corticotrophic axis in response to fear
tion of ␣1 -ARs in the amygdala (Section 9.1.11) (Okuyama and other stressors (Day et al., 1999a). Reciprocally, the
et al., 1999b). In a similar vein, it has been posited (Section expression of ␣1B -ARs in these and other cerebral nuclei
4.1.2) that stimulation of ␣1 -ARs underlies the paradoxi- is subject to modulation by glucorticoids: most pertinently,
cal, anxiogenic actions of high doses of the ␣2 -AR partial they appear to obviate the reduction in the functional activ-
agonist, clonidine, in the VCT, though this supposition en- ity of ␣1B -ARs engendered by chronic stress (Sakaue and
joys little experimental support (Gardner and Piper, 1982; Hoffman, 1991; Day et al., 1999a; Stone et al., 2002). It has
Söderpalm and Engel, 1988, 1990; Przegaliński et al., been documented that mice lacking ␣1B -ARs display re-
1994b; Söderpalm et al., 1995a,b). In a opposite fashion, duced exploratory behavior in a novel environment, though
it has been proposed that engagement of ␣1 -ARs enhances any attribution of this observation to an increase in anxiety
the anxiolytic actions of BZPs in the VCT (Söderpalm and must be tentative in light of the generalized role of ␣1B -ARs
Engel, 1988, 1990; Söderpalm et al., 1995a,b). Indepen- in the control of spontaneous locomotor drive (Simon et al.,
dent support for this rather tenuous assertion of anxiolytic 1994; Stone et al., 1999, 2001; Knauber and Müller, 2000).
actions at ␣1 -ARs has likewise not been forthcoming. Finally, ␣1D -ARs predominate over other ␣1 -AR subtypes
Notwithstanding the limited and contradictory nature of in the hippocampus, they are highly expressed throughout
this evidence for a role of ␣1 -ARs in the modulation of anx- the cortex, and they are also present in several sub-territories
ious behavior, the balance of evidence favors an anxiogenic of the amygdala (Nicholas et al., 1996; Day et al., 1997).
action. This issue would merit further evaluation in view of Thus, ␣1B -ARs are currently the most promising candi-
the marked influence of ␣1 -ARs upon attentional and cog- dates for a role in the modulation (enhancement) of anxious
nitive processes under conditions of novelty (Hieble et al., behavior. The recent description of drugs discriminating in-
1995; Arnsten, 1997; Day et al., 1997; Aston-Jones et al., dividual ␣1 -AR subtypes should accelerate exploration of
1999; Stone and Quartermain, 1999; Nalepa et al., 2002; their functional roles in the control of anxious states (Hieble,
Tanoue et al., 2002). Moreover, protracted stress desensitises 2000).
cerebral populations of ␣1 -ARs (Stone et al., 1986, 2002;
Izumi et al., 1996). Conversely, long-term administration of 4.1.6. Role of β-ARs
antidepressant agents (which is accompanied by the devel- Several lines of evidence support a role of ␤-ARs in the
opment of anxiolytic properties) has generally, though not response to and control of anxious states. First, high con-
invariably, been found to circumvent this effect of stress and centrations of ␤1 - and ␤2 -ARs have been visualized in the
itself to amplify functional activity at central populations of cortex, amygdala, hippocampus, PAG and other limbic re-
␣1 -ARs (Section 4.4) (Vetulani et al., 1984; Vetulani and gions (Rainbow et al., 1984; Ordway et al., 1988, 1991;
Nalepa, 2000; Izumi et al., 1997; Maj et al., 2000). Schatzberg and Schildkraut, 1995; Nicholas et al., 1996).
Second, ␤-ARs have been shown to play a critical role in the
4.1.5.2. α1 -AR subtypes. Further insights into the signifi- modulation of emotion and cognition by effects in the amyg-
cance of ␣1 -ARs would likely be gained from a considera- dala, hippocampus and FCX, wherein they enhance NMDA
tion of individual ␣1A -, ␣1B - and ␣1D -AR subtypes which receptor-mediated transmission by both pre- and postsynap-
are differentially distributed in structures involved in the tic actions (Huang and Kandel, 1996; Huang et al., 1998;
control of anxious states (Pieribone et al., 1994; Hancock, Roozendaal et al., 1999; Zhang et al., 2001b; Crissman and
1996; Nicholas et al., 1996; Day et al., 1997; Domyancic O’Donnell, 2002; McGaugh et al., 2002; Pralong et al., 2002;
and Morilak, 1997). Wang et al., 2002a). Third, ␤1 - and ␤2 -ARs exert a pha-
␣1A -ARs are highly expressed throughout the cor- sic, facilitatory influence upon the release and synthesis of
tex, hypothalamus and amygdala, and they are also NA and, probably, 5-HT, in the FCX and subcortical struc-
well-represented in the hippocampus and PAG. While tures (Bouthillier et al., 1991; Murugaiah and O’Donnell,
␣1B -ARs share their predilection for the cortex, they oth- 1995; Gobert and Millan, 1999a; Tsuiki et al., 2000). Fourth,
erwise present a more confined distribution with a striking ␤-AR blockers are therapeutically utilized in the treatment of
density in the lateral nucleus (only) of the amygdala and, situation-specific/performance anxiety (Tyrer, 1988, 1992)
most remarkably, a high concentration in raphe nuclei. and can prevent the induction of panic attacks in human sub-
This observation suggests that activation of ␣1B -ARs is jects (Le Mellédo et al., 1998). Fifth, ␤-ARs participate in
responsible for the tonic, excitatory influence exerted by the recruitment of the hypothalamus–corticotropic axis by
NA upon the activity of ascending serotonergic pathways stress (Carrasco and Van de Kar, 2003). Finally, long-term
(Millan et al., 2000d; Hopwood and Stamford, 2001b; administration of several classes of antidepressant agent, in
Adell et al., 2002; Brown et al., 2002). ␣1B -ARs are like- parallel with the progressive development of their anxiolytic
wise facilitatory to VTA-derived dopaminergic pathways properties (Section 4.4), down-regulates corticolimbic (in-
(Auclair et al., 2002). An additional structure presenting cluding amygdala) populations of ␤1 -ARs (Ordway et al.,
a pronounced density of ␣1B -ARs is the hypothalamus 1988, 1991; Heal et al., 1989; Schatzberg and Schildkraut,
wherein they are situated on CRF-containing cell bodies: 1995; Newman-Tancredi et al., 1999).
120 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Table 10
Influence of drugs interacting with ␤-adrenoceptors in the Vogel Conflict Test
Drug Class Species Route (structure) Effect Active dose range Maximum Reference
(strain) change (%)
Betaxolol ␤1 -AR Ant Rat (W) i.p. IA 4.0–8.0 mg/kg IA Przegaliński et al. (1994a)
Rat (W) i.c. (hippocampus) IA 3.0 and 10.0 ␮g IA Przegaliński et al. (1995)
Metoprolol ␤1 -AR Ant Rat (SD) i.p. IA 0.5–2.0 mg/kg IA Söderpalm and Engel (1990)
Pindolol ␤1/2 -AR PAGO Rat (W) i.c. (hippocampus) + 0.1–3.0 ␮g +186 Przegaliński et al. (1995)a
Clenbuterol ␤2 -AR agonist Rat (SD) i.p. IA 0.25–1.0 mg/kg IA Söderpalm and Engel (1990)
ICI 118,551 ␤2 -AR Ant Rat (W) i.p. IA 4.0–8.0 mg/kg IA Przegaliński et al. (1994a,b)
Rat (W) i.c. (hippocampus) IA 3.0 and 10.0 ␮g IA Przegaliński et al. (1995)
Propranolol ␤1/2 -AR Ant Rat (SD) i.p. IA 1.0 and 2.0 mg/kg IA Söderpalm and Engel (1990)
Abbreviations—PAGO: partial agonist and IA: no significant change (inactive). For other abbreviations, see Table 2.
a 5-HT
1A receptor-mediated (Section 4.3.2.4).

Notwithstanding the above observations, anxiolytic prop- pared to 5-HT and NA, rather little attention has been de-
erties of ␤-AR antagonists have not, in general, been re- voted to the role of DA in the response to, and modula-
vealed in experimental models in rodents (Njung’e et al., tion of, anxious states (Thiel and Schwarting, 2001; Harari
1993; Sanchez et al., 1996; Cao and Rodgers, 1997b; Prut et al., 2002; Stein et al., 2002). This is curious in view
and Belzung, 2003; Sanchez, 2003), including conflict of the well-established role of stress-responsive mesocorti-
paradigms (Durel et al., 1986; Fontana and Commissaris, cal and mesolimbic dopaminergic pathways—both of which
1988; Fontana et al., 1989b; Söderpalm and Engel, 1990; originate in the VTA—in the control of mood (Swanson,
Terry and Salmon, 1991). The VCT is no exception inas- 1982; Le Moal and Simon, 1991; Finlay and Zigmond, 1997;
much as the systemic and intrahippocampal administration Millan et al., 2000d; Hasue and Shammah-Lagnado, 2002;
of ␤1 - and ␤2 -ARs antagonists fails to modify behavior Homberg et al., 2002). Indeed, by analogy to cognitive func-
in this procedure (Table 10) (Söderpalm and Engel, 1990; tion (Nieoullon, 2002), an “optimal” degree of dopaminergic
Przegaliński et al., 1994a, 1995)—though a study of their activity in the FCX may be requisite for an appropriate re-
direct injection into the amygdala would be of interest. sponse to stress and fear (Pezze et al., 2003). Anxious symp-
This irresponsiveness of the VCT and other experimen- toms are frequently co-morbid with (and may exacerbate)
tal paradigms to ␤-AR antagonists possibly reflects the drug abuse, affective and psychotic disorders, while anxi-
fact that: (1) they do not themselves modify monoamin- ety is a prominent, precocious and persistent symptom of
ergic transmission in rats—which is phasically enhanced Parkinson’s disease (Menza et al., 1993; Merikangas et al.,
by ␤1 /␤2 -AR agonists and (2) the anxiolytic properties of 1998; Maier and Falkai, 1999; Beekman et al., 2000; Shiba
␤-AR antagonists in man primarily reflect actions at pe- et al., 2000; Fenton, 2001; Marinus et al., 2002). Further, a
ripheral sites controlling autonomic and somatic function perturbation of dopaminergic input to the amygdala has been
(Durel et al., 1986; Tyrer, 1988, 1992; see Gobert and specifically implicated (by functional magnetic resonance
Millan, 1999a). imaging) in the abnormal emotional processing which char-
Though anxiolytic properties of the ␤-AR partial ago- acterizes parkinsonian patients (Benke et al., 1998; Braak
nist, pindolol, have been reported with the VCT and other and Braak, 2000; Tessitore et al., 2002).
models, its actions are probably mediated by stimula-
tion of 5-HT1A autoreceptors (Table 10) (Section 4.3.2.4) 4.2.1.2. Modulation of dopaminergic transmission by aver-
((Przegaliński et al., 1994a, 1995); Cao and Rodgers, 1997b; sive stimuli. In fact, the relationship of dopaminergic
Millan et al., 2002). Otherwise, information concerning the transmission to clinical anxiety disorders is complex and
potential influence of agonists at ␤1 - and ␤2 -ARs upon poorly understood. There is evidence that social phobia—
anxious states does not appear to be available. which is common in parkinsonian patients—is associated
with a sustained suppression both of dopaminergic trans-
4.2. Dopamine mission and of activity at dopaminergic receptors (Tiihonen
et al., 1997a; Grant et al., 1998; Schneier et al., 2000;
4.2.1. Dopaminergic pathways Kennedy et al., 2001; Stein et al., 2002). On the other
hand, non-parkinsonian subjects with unusually high lev-
4.2.1.1. Clinical disorders, dopamine and anxious states. els of “trait” anxiety and especially susceptible to panic
Central dopaminergic pathways have been most intensively attacks have revealed an enhancement in the activity of
studied within the perspective of their role in the pathogene- central dopaminergic pathways (Roy-Byrne et al., 1986;
sis of depression, drug abuse, schizophrenia and Parkinson’s Pitchot et al., 1992; Finlay and Zigmond, 1997; Mizuki
disease, disorders involving a marked (though contrasting) et al., 1997). In line with the latter findings, experimental
dysregulation of dopaminergic transmission. Indeed, as com- studies have shown that conditioned fear, anxiety and other
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 121

stressors elicit a (BZP-reversible) activation of VTA-derived In interpreting actions of dopaminergic ligands in mod-
dopaminergic pathways to the amygdala and adjacent bed els of potential anxiolytic properties, it should also be re-
nucleus of the stria terminals (Coco et al., 1992; Inglis and called that dopaminergic pathways fulfil an important role
Moghaddam, 1999; Greba et al., 2001; Kozicz, 2002; Suzuki in the formation, retention and extinction of fear-related as-
et al., 2002a,b), to the nucleus accumbens (McCullough sociations and memory (Morrow et al., 1999; Arnsten et al.,
and Salamone, 1992; Kalivas and Duffy, 1995; Pezze et al., 2000; Greba et al., 2001; Pezze et al., 2001; Nieoullon,
2001; Levita et al., 2002) and, most strikingly, to the FCX 2002).
(Ravard et al., 1989; Finlay and Zigmond, 1997; Wilkinson
et al., 1998; Broersen et al., 2000; Feenstra et al., 2000; 4.2.2. Multiple classes of dopamine receptor
Beaufour et al., 2001a; Dazzi et al., 2001; Del Arco et al., One difficulty confronted in the study of DA and anx-
2001). Very recently, providing a clinical correlate to these ious states is the particularly marked influence of genetic
studies, employing functional magnetic resonance imaging, background upon the actions of dopaminergic agonists in
it was reported that the psychostimulant and DA releaser, models of anxious (and other) behavior(s) (Gendreau et al.,
dextroamphetamine, enhances the response of the amyg- 1998; Kelly et al., 1998; Rodgers et al., 2002a,b). Fur-
dala to aversive stimuli in human subjects (Harari et al., thermore, dopaminergic receptors are localized both pre-
2002). This observation suggests that dopaminergic input and postsynaptically to dopaminergic projections (Section
to the amygdala is specifically involved in mediating the 4.2.3.1) (Levant, 1997; Vallone et al., 2000), while DA ex-
anxiogenic properties of dextroamphetamine and, possibly, erts its actions via multiple classes of dopamine receptor
other dopaminergic agents (Willick and Kokkinidis, 1995; poorly-differentiated by many widely-used agents (Shafer
Paine et al., 2002). A parallel electrophysiological inves- and Levant, 1998; Helmeste and Tang, 2000; Oak et al.,
tigation in the rat suggested that this action of DA in the 2000; Vallone et al., 2000; Millan et al., 2002).
amygdala reflects the simultaneous promotion of excitatory Closely-related D2 , D3 and D4 receptors share negative
input from the somato-sensory cortex, and suppression of coupling via Gi to adenylyl cyclase: D2 and D3 sites also
inhibitory input from the prefrontal cortex (Rosenkranz and couple negatively to GIRKs (Section 13.3), though such ac-
Grace, 2001, 2002). tions are less well established for D4 sites. On the other hand,
closely-related D1 and D5 receptors both stimulate adenylyl
4.2.1.3. Dopamine, reward and anxious states. Mesolim- cyclase via Gs (Sibley, 1999; Vallone et al., 2002).
bic dopaminergic projections are critically involved in mech-
anisms of motivation and reward, and a substantial body 4.2.3. Dopamine “D2 -like” receptors
of evidence suggests that their engagement contributes to
the preference for (non-aversive) novel stimuli (Hooks and 4.2.3.1. Cerebral distribution of D2 , D3 and D4 receptors:
Kalivas, 1995; Volkow et al., 2002; Wightman and Robinson, influence upon GABAergic transmission. Dopamine D2
2002). Correspondingly, and in particular in paradigms in- receptors are widely-distributed throughout the brain, dis-
volving the exploration of unfamiliar environments and other playing a high density in the FCX, nucleus accumbens,
measures of neophobia, the reinforcing effects of dopamin- amygdala and septum. They are present in far higher con-
ergic agents may interact with their influence upon anx- centrations than cerebral populations of D3 sites in rodents,
ious states per se (Homberg et al., 2002; Marinelli and though in primate (and human) brain, D3 receptors are bet-
Piazza, 2002). Indeed, it has been proposed that the positive ter represented (Levant, 1997; Sibley, 1999; Vallone et al.,
emotional effects associated with activation of mesolimbic 2000; Joyce, 2001). Further, across all species, D3 receptors
dopaminergic pathways can outweigh the negative impact of are principally localized in limbic regions, including the
stress (Marinelli and Piazza, 2002). This may explain why nucleus accumbens, hippocampus, olfactory tubercles, cere-
the imposition of anxiogenic and other modes of stressor bral cortex and thalamus (Jackson and Westlind-Danielsson,
augments the self-administration of dopaminergic drugs of 1994; Levant, 1997; Joyce, 2001). Dopamine D2 receptors
abuse (Piazza et al., 1990; Homberg et al., 2002)—in line predominate as autoreceptors on dopaminergic cell bodies
with the notion of self-medication as a coping behavior, an and terminals, though D3 receptors fulfil a complementary
issue evoked elsewhere in this review as concerns human role in the inhibitory control of cortical and subcortical DA
use (and abuse) of ethanol (Section 2.2.5). One further in- release, both as autoreceptors and, conceivably, via a post-
triguing illustration of the interrelationship between anxiety, synaptic feedback loop (Millan et al., 2000d; Zapata et al.,
reward and dopamine is provided by the phenomenon of 2001; Centonze et al., 2002; Joseph et al., 2002). Though
self-grooming in response to stress. This behavior, which in- likewise less prominent than their D2 receptor counterparts,
volves mesolimbic dopaminergic networks—and may serve the enrichment of D4 sites in the lateral septum, amygdala,
as a “dearousal” process contributing to the restoration of hypothalamus, hippocampus and, most strikingly, the FCX,
homeostasis following stress—predicts a high level of mo- is of particular pertinence in the present context, together
tivation for self-administration of the DA releaser, cocaine with their localization on cortical GABAergic interneurones
(Spruijt et al., 1992; Bandler et al., 2000; Homberg et al., in rodents and primates (Rossetti et al., 1989; Mrzlkak
2002). et al., 1996; Oak et al., 2000; Wedzony et al., 2000).
122 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Indeed, DA exerts a complex and receptor-dependent in- Oak et al., 2000; Vallone et al., 2000; Millan et al., 2002)
fluence upon GABAergic transmission reflecting both: (1) but the novel agent, S32504, which is devoid of activity at
presynaptic changes in the excitability of GABAergic neu- D4 receptors, reproduces their anxiolytic effects in the VCT
rones and of GABA release and (2) postsynaptic interactions (Table 11) (Dekeyne et al., 2001). Further, anxiolytic ac-
with GABAergic receptors. Dopamine D1 and D2 (D4 ) re- tions of S32504 are blocked by selective antagonists at D2
ceptors exert facilitatory and inhibitory effects, respectively, receptors, whereas D3 or D4 receptor antagonists are inef-
upon GABAergic inhibition of cortical pyramidal neurones. fective (Dekeyne et al., 2001; Brocco and Millan, unpub-
Moreover, activation of dopamine D4 receptors interferes lished observations). These data indicate that activation of
with the operation of topographically-adjacent postsynaptic D2 receptors underlies the anxiolytic actions of “D2 -like”
GABAA receptors by a complex mechanism involving re- agonists. Support for this contention has not, as yet, been
cruitment of protein kinase A (Grobin and Deutch, 1998; derived from D2 receptor knock-out mice, possibly since al-
Seamans et al., 2001; Wang et al., 2002b). However, the terations in motor function compound interpretation of their
interrelationship between GABAergic neurones and multi- phenotype Sibley, 1999; Waddington et al., 2001. Notwith-
ple classes of dopaminergic receptor remains poorly-defined standing several (preliminary) reports of anxiolytic actions
and this issue would be of interest to further explore in sub- of non-selective dopaminergic (D2 -like) agonists in man
cortical, limbic structures. (Ferrier et al., 1985; Mizuki et al., 1997), such findings re-
main to be corroborated by more extensive and rigorous in-
4.2.3.2. Dopamine D2 receptors. Varicose dopaminergic vestigations.
fibers originating in cell bodies of the VTA and hypothala- Dopamine D2 autoreceptors are substantially more re-
mus have been visualized in the DRN and LC (Kitahama sponsive than postsynaptic populations of D2 receptors,
et al., 2000), and certain studies have indicated a minor, manifesting greater sensitivity to low doses of agonists and
modulatory influence of D2 receptors upon the activity of responding even to ligands of modest efficacy (Levant, 1997;
corticolimbic serotonergic (facilitatory) and noradrenergic Millan et al., 2000d; Joyce, 2001). A role of highly-sensitive
(inhibitory) pathways. However, the physiological signifi- D2 autoreceptors in the anxiolytic properties of dopaminer-
cance of these roles remains unclear and neither agonists gic agonists would be consonant with the observation that
nor antagonists at dopamine D2 receptors elicit marked al- low doses are sufficient for expression of their actions in the
terations in the release of NA and 5-HT in corticolimbic VCT and other models (Hjorth et al., 1986; Borowski and
structures (Rossetti et al., 1989; Ferré et al., 1994; Mendlin Kokkinidis, 1996; Munro and Kokkinidis, 1997; Bartoszyk,
et al., 1999; Vanderschuren et al., 1999; Millan et al., 2000d; 1998; Rodgers et al., 2000). This interpretation is further
Adell et al., 2002; Szumlinski and Szechtman, 2002). underpinned by anxiolytic actions in the VCT of (−)PPP, a
Comparatively, few behavioral data have been forthcom- weak (partial) agonist known to preferentially stimulate pre-
ing in support of potential anxiolytic effects of D2 receptor as compared to postsynaptic populations of dopamine D2
agonists (Bartoszyk, 1998; Rodgers et al., 2000; Sanchez, receptor (Table 11) (Hjorth et al., 1987a). Further, local in-
2003). Nevertheless, the prototypical agonists, apomorphine jection of dopamine (D2 ) receptor agonists into the VTA is
and quinpirole, both elicited significant increases in pun- associated with anxiolytic properties (Gifkins et al., 2002).
ished responses in the VCT (Table 11) (Hjorth et al., 1986, A D2 autoreceptor-mediated reduction of stress-induced
1987a; Siemiatkowski et al., 2000b). Neither of these drugs DA release to the FCX (Section 4.2.1) may be of especial
distinguishes D2 from D3 and D4 receptors (Levant, 1997; importance to the actions of dopamine receptor agonists in

Table 11
Influence of dopaminergic agents upon behavior in the Vogel Conflict Test
Drug Class Species Route Effect Active dose Maximum Reference
(strain) (structure) range (mg/kg) change (%)
Apomorphine D1 /D2 agonist Rat (SD) s.c. + 0.006–0.025 +110 Hjorth et al. (1986)
(+)3-PPP D2 /D3 /D4 agonist Rat (SD) s.c. + 0.006–0.5 +45 Hjorth et al. (1987a,a)
(−)3-PPP D2 /D3 /D4 agonist Rat (SD) s.c. + 0.016–0.25 +50 Hjorth et al. (1987a,b)
Quinpirole D2 /D3 /D4 agonist Rat (W) i.p. + 1.0 +220 Siemiatkowski et al. (2000b)
S32504 D2 /D3 agonist Rat (W) s.c. + 5.0 +89 Brocco (unpublished observations)
Haloperidol D2 /D3 /D4 Ant Rat (W) s.c. IA 0.01–0.16 IA Millan et al. (1999a)
Rat (W) i.p. IA >0.2 IA Siemiatkowski et al. (2000b)
Sulpiride D2 /D3 Ant Rat (W) s.c. + 25.0 +160 Siemiatkowski et al. (2000b)
Nafadotride D3 > D2 Ant Rat (?) s.c. + 0.2–1.0 NI (Rogoz et al., 2000)
L745,870 D4 Ant Rat (W) s.c. IA 0.16–2.5 IA Brocco (unpublished observations)
S18126 D4 Ant Rat (W) s.c. IA 0.16–2.5 IA Brocco (unpublished observations)
SCH 23390 D1 Ant Rat (W) s.c. IA >0.01 IA Brocco (unpublished observations)
Abbreviations—(+)3-PPP: R(+)-3(3-hydroxyphenyl)-N-propylpiperidine; (−)3-PPP: preclamol; IA: no significant change (inactive) and NI: not indicated.
For other abbreviations, see Table 2.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 123

the VCT and other procedures (Ravard et al., 1990; Espejo, knock-out mice. This was tentatively attributed to a loss of
1997). However, the precise sites and mechanisms of action D4 receptor function in the FCX, despite the fact that D4
underlying anxiolytic properties of D2 receptor agonists receptors are inhibitory to GABAergic neurones in this re-
have, to date, eluded identification. gion (Section 4.2.1.2) (Rubinstein et al., 1997; Dulawa et al.,
1999; Sibley, 1999; Arnsten et al., 2000; Okuyama et al.,
4.2.3.3. Dopamine D3 receptors. Though isolated reports 2000; Falzone et al., 2002). While selective D4 receptor an-
of anxiogenic effects of “D2 receptor antagonists” in ex- tagonists do not display anxiolytic properties in the VCT
perimental studies (Bartoszyk, 1998; Timothy et al., 1999; and other procedures (Table 11) (Cao and Rodgers, 1997a;
Siemiatkowski et al., 2000b) would be consistent with Bartoszyk, 1998; Brocco and Millan, unpublished observa-
above-discussed findings indicative of an anxiolytic role tions), evaluation of the actions of selective D4 agonists
for D2 receptors, they are inactive in the VCT (Dekeyne would be of interest in such paradigms.
et al., 2001; Siemiatkowski et al., 2000b; Brocco and Mil-
lan, unpublished observations). Further, there have also 4.2.4. Dopamine “D1 -like” receptors
been reports (albeit conflicting) of anxiolytic actions of Dopamine D1 receptors are enriched in the FCX, nucleus
“D2 receptor antagonists” in conflict and other procedures accumbens, amygdala and hippocampus, wherein they dis-
(Pich and Samanin, 1986; Cools et al., 1993; Rodgers et al., play a complex pattern of interaction with GABAergic and
1994; Costall and Naylor, 1997; Bartoszyk, 1998; Cavazzuti glutamatergic neurones (Jackson and Westlind-Danielsson,
et al., 1999; De Boer and Koolhaas, 2003). Inasmuch as 1994; Ng et al., 2001). In line with this pattern of distribu-
drugs employed in these studies fail to discriminate D2 and tion, D1 receptors play an important role in the cognitive
D3 receptors—“D2 receptor antagonists” is something of a component (primarily extinction, rather than acquisition or
misnomer—their respective participation in such actions is consolidation) of fear-conditioning (Greba and Kokkinidis,
unclear. However, in the light of data discussed in the pre- 2000; El-Ghundi et al., 2001). Chronic stress is accompa-
ceding paragraphs, a specific role of D2 receptor blockade nied by a BZP-reversible decrease in D1 receptor density
in the expression of anxiolytic properties seems unlikely. in the nucleus accumbens (Giardino et al., 1998) and cer-
It might, thus, be postulated that D3 receptor blockade tain authors have speculated that the activation of D1 recep-
underlies potential anxiolytic actions of mixed D2 /D3 re- tors enhances, or elicits, anxious states (Simon et al., 1993;
ceptor antagonists. This contention is favoured by reports Siemiatkowski et al., 2000a; Cancela et al., 2001). However,
that knock-out mice deficient in D3 receptors generally dis- other studies contradict this assertion and there is little ev-
play an anxiolytic phenotype (Accili et al., 1996; Xu et al., idence for anxiolytic profiles of either dopamine D1 recep-
1997; Steiner et al., 1998; Karasinska et al., 2000; tor antagonists or D1 receptor-knock-out mice in either the
Blednov et al., 2001; Vallone et al., 2002). Further, anx- VCT or other procedures (Table 11) (Rodgers et al., 1994;
iolytic actions of preferential D3 antagonists have been Bartoszyk, 1998; El-Ghundi et al., 1999, 2001; Sibley, 1999;
reported in the VCT and certain other—though not all— Karasinska et al., 2000; Waddington et al., 2001; Blednov
procedures of anxiolytic activity (Gendreau et al., 1997; et al., 2002). Nevertheless, studies of D1 /D3 double mutant
Bartoszyk, 1998; Rodriguez-Arias et al., 1999; Rogoz et al., mice suggest that D1 receptors may interact with (possibly
2000; Dekeyne et al., 2001; Brocco and Millan, unpub- colocalized) D3 receptors in modulating the behavioral re-
lished observations). Such findings must be substantiated sponse to a novel, fear-inducing environment (Sibley, 1999;
by studies with highly-selective agents unambiguously dif- Karasinka et al., 2000; Joyce, 2001).
ferentiating D3 from D2 receptors. Nevertheless, supporting Dopamine D5 receptors are found predominantly in the
a role of D3 receptors in the facilitation of anxious states, hippocampus, but they are also localized in the FCX, hy-
stimulation of D3 receptors is associated with anxiogenic pothalamus, LC and DRN (Ariano et al., 1997; Ciliax et al.,
actions in several experimental models (Kahaya et al., 1996; 2000; Pralong et al., 2002). No drug is currently known to
Gendreau et al., 2000). differentiate D5 from D1 sites, but mice lacking D5 recep-
The—to date, tentative—possibility that D3 receptors tors have been described. Their phenotype reveals no signs
facilitate anxious behavior (Table 11) deserves additional of alterations in emotionality or in the response to stress
study inasmuch as it would provide an interesting parallel to suggesting that they do not play a major role in the con-
the opposite roles of (postsynaptic) D2 versus D3 receptors trol of anxious states (Sibley, 1999; Holmes et al., 2001;
in the enhancement and suppression of locomotor behavior, Waddington et al., 2001; Hollon et al., 2002).
respectively (Levant, 1997; Joyce, 2001). Interestingly, D5 receptors can form functional het-
erodimers with GABAA receptors. Assuming their occur-
4.2.3.4. Dopamine D4 receptors. Dopamine D4 receptor rence in the CNS, the potential significance of these con-
knock-out mice display alterations in their cognitive perfor- structs to anxious states would be of interest to explore (Liu
mance and in their motor response to novelty, complicating et al., 2001). Similarly, if—by analogy to D3 /D2 receptor
interpretation of their behavior in response to anxiogenic heterodimers—D1 /D2 and D1 /D3 heterodimers were found
stimuli. Nevertheless, a specific, BZP-preventable increase to exist, this would add a novel dimension to the roles
in unconditioned fear was recently detected in D4 receptor of multiple classes of dopamine receptor in the control of
124 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

emotion and other functions (Section 18.2) (Scarscelli et al., Beaufour et al., 2001b; Ishida et al., 2002; Miura et al.,
2001; George et al., 2002). 2002). Moreover, it has been demonstrated that responding
in the VCT is accompanied by a concurrent increase in 5-HT
4.3. Serotonin release in the hippocampus (Matsuo et al., 1996 though see
also Beaufour et al., 2001b). These observations underline
4.3.1. Serotonergic pathways the significance of serotonergic mechanisms to the control of
Though it is well known that the rate-limiting enzyme, performance in this procedure, though their precise role(s),
tryptophan hydroxylase, catalyses the synthesis of 5-HT, it and their involvement in the actions of BZPs, have been ex-
has long remained a mystery as to why its levels are so low tensively debated (Section 2.2.2.3) (Pratt, 1992).
in raphe nuclei as compared to the pineal gland—wherein As discussed below (Sections 4.3.2, 4.4 and 4.5), sero-
it undertakes the initial step in the generation of melatonin tonergic mechanisms participate in the influence of a broad
(Section 10). Even more surprisingly, mice deficient in this range of therapeutically-employed drugs upon emotionality
enzyme maintain essentially normal levels of 5-HT in the in general, and upon anxious states in particular. Moreover,
CNS despite a pronounced depletion of 5-HT in peripheral the 5-HT releaser, methylenedioxymethylamphetamine (ec-
tissues (Walther et al., 2003). This paradox was recently stasy), has been shown to modify anxiety states in a dose
resolved with the discovery of a second and brain-specific and test-dependent manner both in experimental studies and
form of tryptophan hydroxylase responsible for generation in human subjects (Parrott, 2000; Liechti and Vollenweider,
of 5-HT in cerebral tissue (Walther et al., 2003). 2001; Fone et al., 2002; Green and McGregor, 2002; Navarro
Serotonergic neurones originating in raphe nuclei provide and Maldonado, 2002). It’s escalating and illicit recreational
a massive input to corticolimbic structures involved in the abuse in many countries underscores the urgency of enhanc-
control of anxious states. The DRN primarily innervates the ing our understanding of the role of serotonergic mecha-
FCX, dorsal hippocampus and amygdala, while the median nisms in the control of anxious behaviour.
raphe nucleus (MRN) principally projects to the (dorsal and
ventral) hippocampus, septum, nucleus accumbens and hy- 4.3.2. 5-HT1A receptors
pothalamus (Azmitia and Segal, 1978; Gray, 1987; Deakin
and Graeff, 1991; Wilson et al., 1993; Eison and Eison, 4.3.2.1. Coupling and organization. The role of 5-HT1A
1994; Handley, 1995; Baumgarten and Grozdanovic, 1997; receptors in the modulation of anxious states has been par-
McQuade and Sharp, 1997; Vertes et al., 1999; Kirby et al., ticularly well studied (Chopin and Briley, 1987; Handley,
2003). These DRN and MRN-derived networks show con- 1995; Barnes and Sharp, 1999; Olivier et al., 1999, 2001;
trasting patterns of organization and modulation (Table 12) Gross et al., 2002).
(Azmitia and Segal, 1978; Kosofsky and Mollinver, 1987; Activation of Gi -coupled 5-HT1A receptors enhances
Deakin and Graeff, 1991; Wilson et al., 1993; Netto et al., K+ -currents (GIRKs) and inhibits the activity of adenylyl
2002; Kirby et al., 2003) and may fulfil differential roles cyclase: their engagement may also suppress the operation
in the control of anxious states. For example, serotonergic of VDCCs (Gerhardt and Heerikhuizen, 1997; Barnes and
pathways emanating from the DRN have been specifically Sharp, 1999; Raymond et al., 1999). 5-HT1A receptors,
implicated in the control of behavior in the VCT (Thiébot which are localized as inhibitory autoreceptors on the den-
et al., 1983; Pratt, 1992). A complex and non-uniform role drites of serotonergic cell bodies in raphe nuclei, are also
for 5-HT in the control of anxiety disorders has also been localized postsynaptically to serotonergic neurones in the
forwarded as a function of: (1) whether the anxious state is hippocampus, septum, amygdala, PAG, entorhinal cortex
provoked by conditioned or unconditioned fear and (2) con- and FCX: in the latter structure, they exert a “long-loop”,
trasting actions of 5-HT in specific cerebral regions, notably inhibitory feedback influence upon the electrical activity of
the amygdala as compared to the PAG (Deakin, 1988, 1991; serotonergic perikarya in the DRN (Celada et al., 2001).
Deakin and Graeff, 1991; Graeff et al., 1993, 1996a; Eison Interestingly, prolonged stress modifies cerebral levels of
and Eison, 1994; Handley, 1995; Green and McGregor, 5-HT1A sites and may preferentially densensitize pre- as
2002). Moreover, in assessing the role of 5-HT in the mod- compared to postsynaptic populations of 5-HT1A receptors
ulation of anxious states, it is advisable to consider the (McKittrick et al., 1995; Barton et al., 1999; Laaris et al.,
potentially confounding influence of serotonergic pathways 1999), a process contributing to the overall potentiation of
upon other behaviours, such as motor function, impulsiv- serotonergic transmission under these conditions.
ity and cognition (Evenden, 1999; Meneses, 1999; Dalley 5-HT1A receptor agonists elicit a pronounced increase in
et al., 2002; Green and McGregor, 2002; Harro, 2002). the activity of the hypothalamo-corticotropic axis, acting
Serotonergic pathways innervating structures such as the via multiple loci in the hypothalamus itself as well as in
FCX, amygdala, hypothalamus and hippocampus are acti- limbic structures. Further, activation of 5-HT1A receptors is
vated by anxiogenic stimuli, including psychosocial stress, involved in the induction of ACTH and corticosteroid secre-
conditioned fear and conflict procedures (Pratt, 1992; Wright tion in response to stress (Chaouloff, 1995; Fletcher et al.,
et al., 1992; Adell et al., 1997; Rueter et al., 1997; Amat 1996; Jorgensen et al., 1998, 2001; Carrasco and Van de Kar,
et al., 1998; Berton et al., 1998; Martinez et al., 1998; 2003). This action is of considerable pertinence inasmuch
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 125

Table 12
A comparison of the principle characteristics of the dorsal raphe nucleus (DRN) as compared to the median raphe nucleus (MRN)
MRN DRN

Major projection targets Olfactory bulb, ventral and dorsal hypothalamus, Frontal CX and pre-frontal CX, dorsal hippocampus,
parietal cortex, septum, nucleus accumbens amygdala, hypothalamus, nucleus accumbens
GABA inhibition (cell bodies) Mild Strong
␣1 -AR excitation (cell bodies) Strong Mild
5-HT1A inhibition (cell bodies) Mild Strong
5-HT1B inhibition (terminals) Strong Mild
Glutamatergic stimulation Mild Strong
(cell bodies)
Non-5-HT cells Many Few
Fibre type (projections) Varicose Fine
Focalized Diffuse, widespread
PCA, MDMA vulnerability Low High
SERT density Low High
Abbreviations—CX: cortex; GABA: ␥-amino-butyric acid; AR: adrenoceptor; 5-HT: serotonin; PCA: parachloroamphetamine; MDMA: methylenediethyl-
metamphetamine and SERT: 5-HT transporters. PCA and MDMA (ecstasy) are neurotoxins which acutely release, and chronically deplete, 5-HT.

as corticosteroids exert a reciprocal modulatory influence tissue-specific, “conditional” 5-HT1A receptor knock-out
upon the activity of pre- and postsynaptic populations of model. Thus, in mice in which functional 5-HT1A recep-
5-HT1A receptors and have been likewise implicated in the tors were preserved in the amygdala, hippocampus and
control of mood (Section 11.2). 5-HT1A receptor agonists, FCX, whereas raphe-localized 5-HT1A autoreceptors were
principally via central actions, also facilitate the release of disabled, the anxious phenotype was absent Gross et al.
NA from sympathetic neurones, an action which may secon- (2002)—though see Section 4.3.2.5.
darily enhance the activity of the hypothalamo-corticotropic 5-HT1A receptors exert an inhibitory influence upon
axis. Whether 5-HT1A receptors are involved in triggering limbic (e.g. amygdala) and cortical release of glutamate
the rapid increase in sympathetic outflow in response to Cheng et al., 1998; Lin et al., 2001; Wang et al., 2002a.
stress and fear remains, however, to be ascertained (McCall Correspondingly, a disinhibition of glutamatergic activity
and Clement, 1994; Chaouloff, 1995; Saphier and Welch, (Section 3) may contribute to anxious behavior in mice
1994; Korte et al., 1996). deficient in 5-HT1A receptors. Moreover, an ingenious hy-
pothesis for the anxiogenic phenotype of 5-HT1A receptor
4.3.2.2. Interactions with GABAergic, glutamatergic and knock-out mice invokes an interaction between postsynaptic
catecholaminergic neurones: the 5-HT1A receptor knock-out 5-HT1A receptors and glutamatergic transmission at pyra-
phenotype. Human volunteers possessing a low density midal cells in the frontal cortex and, possibly, subcortical
of central 5-HT1A receptors were found to be more anx- structures (Cai et al., 2002b). AMPA receptors excitatory
ious than control groups displaying higher levels (Tauscher to these neurones are under the facilitatory control of
et al., 2001; Condren et al., 2002). This observation is Ca2+ -calmodulin-dependent kinase II which: (1) enhances
reminiscent of findings in mice genetically lacking 5-HT1A their conductance by phosphorylation of GluR1 subunits
receptors which generally show an anxiogenic phenotype and (2) coordinates their anchoring at neuronal membranes
and an enhanced sensitivity to stress as regards both be- (Lee et al., 2000; Hayashi et al., 2000a,b). Serotonin1A
havioural and autonomic variables (Heisler et al., 1998b; receptors are inhibitory to Ca2+ -calmodulin-dependent
Sibille and Hen, 2001; Pattij et al., 2002a,b; Groenink et al., kinase II. Their genetic ablation is speculated, then, to
2003; Toth, 2003). Though the precise mechanistic bases potentiate the activity of presumptively anxiogenic (and
for this enhanced anxiety remain uncertain, it has been pos- colocalized) AMPA receptors (Section 3.3) (Cai et al.,
tulated to reflect a loss of activity at postsynaptic rather than 2002a,b). This mechanism would be consistent with the
presynaptic 5-HT1A sites inasmuch as extracellular levels anxiolytic phenotype of mice deficient in the ␣-isoform of
of 5-HT are little modified in animals lacking 5-HT1A re- Ca2+ -calmodulin-dependent kinase II itself (Chen et al.,
ceptor. This lack of alteration in 5-HT release may appear 1994).
surprising. However, it presumably reflects changes in the Alternatively, it has been contended that the anxious
control of serotonergic transmission compensatory to the phenotype of 5-HT1A receptor knock-out mice may involve
loss of 5-HT1A autoreceptors which exert, in any case, only an indirect disturbance of the function of amygdala and
a limited, tonic influence upon corticolimbic 5-HT release hippocampus-localized BZP/GABAA receptors (involving
(see Millan et al. (2000d)). Direct support for the contention a down-regulation of ␣2 -subunits), which are presumably
that a loss of postsynaptic 5-HT1A receptors underlies en- expressed by neurones bearing both 5-HT1A and GABAA
hanced anxiety may be derived from recent studies of a receptors. However, it is unclear as to how this change
126 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

is brought about and equivalent observations have not tion of 5-HT1A receptor antagonists into the MRN acts anx-
been obtained in all 5-HT1A receptor knock-out mice lines iolytically in the plus-maze was attributed to a (conjectural)
with divergent genetic backgrounds (Sibille et al., 2000; disinhibition of 5-HT release in the PAG (Canto-de-Souza
Belzung and Griebel, 2001; Olivier et al., 2001; Sibille et al., 2002). However, this remains to be formally demon-
and Hen, 2001; Gross et al., 2002; Pattij et al., 2002a,b; strated and neuronal substrates involved in the occasional
Toth, 2003). Irrespective of this uncertainty, other studies anxiolytic properties of 5-HT1A antagonists are still poorly
have provided evidence for a complex pattern of functional defined (Cao and Rodgers, 1997b; Griebel et al., 1999b,
interactions amongst 5-HT1A receptors and GABAergic 2000; Canto-de-Souza et al., 2002; Nunes-de-Souza et al.,
mechanisms (Peričić and Tvrdeić, 1993; Fernández-Guasti 2002; Blanchard et al., 2003; Bourin and Hascoët, 2003;
and López-Rubalcava, 1998; Sakaue et al., 2001; Pralong Prut and Belzung, 2003).
et al., 2002). Indeed, while colocalized GABAA/B and In contrast to antagonists at 5-HT1A receptors, systemic
5-HT1A receptors exert mutually inhibitory effects upon the administration of partial agonists and agonists is character-
activity of neurones (including raphe-localized serotonergic ized by reproducible anxiolytic actions in a broad range of
cell bodies), it has been shown that 5-HT1A receptors can procedures—though the plus-maze paradigm has yielded a
suppress the limbic release of GABA (Halasy et al., 1992; variable and equivocal pattern of data (Schreiber and De
Schmitz et al., 1995; Matsuyama et al., 1997; Koyama et al., Vry, 1993a,b; Charrier et al., 1994; De Vry, 1995; Molewijk
1999, 2002; Kishimoto et al., 2000a, 2001; Koyama et al., et al., 1995a; Millan et al., 1997; Graeff et al., 1998; Dekeyne
2002). As pointed out below, (Section 4.3.2.5), the latter et al., 2000b; Blanchard et al., 2003; Bourin and Hascoët,
action may participate in anxiogenic actions of 5-HT1A 2003; File and Seth, 2003; Prut and Belzung, 2003; Sanchez,
agonists at postsynaptic 5-HT1A receptors. 2003). Conflict models have generally revealed positive
Presynaptic (and, possibly, postsynaptic) populations of effects (Barrett, 1992; Hascoët et al., 1994; Stanhope
5-HT1A receptor exert an indirect and tonic facilitatory in- and Dourish, 1996; King et al., 1997). With one exception,
fluence upon mesolimbic and mesocortical dopaminergic 5-HT1A receptor antagonists have proven inactive in the
pathways, and upon corticolimbic noradrenergic projec- VCT. Conversely, a broad ensemble of chemically-diverse
tions (Millan et al., 2000d; Szabo and Blier, 2001). This agonists has displayed significant activity—though not
acceleration of the release of DA and NA by 5-HT1A re- necessarily with monophasic dose–response relationships
ceptor agonists may counteract the beneficial effects of (Table 13 and Fig. 7) (Shimizu et al., 1987; Wada and
their autoreceptor-mediated reduction in 5-HT release and Fukuda, 1991; Barrett, 1992; Stefański et al., 1992; Amano
account for the variable and/or biphasic actions of 5-HT1A et al., 1993; Meneses and Hong, 1993; Kennett et al., 1998;
receptor agonists in certain models of anxiolytic activity Griebel et al., 1999b, 2000; Vanover et al., 1999; Dekeyne
(Section 4.3.2.3). Furthermore, unrestrained corticolimbic et al., 2000b; Paluchowska et al., 2000; Vaidya et al., 2002).
release of NA may underlie the relative inefficacy of 5-HT1A Anxiolytic actions of 5-HT1A receptor ligands in the mouse
receptor agonists in the control of panic attacks (Section 4.1). VCT have not, as yet, been documented (Umezu, 1999).
On the other hand, it has been suggested that, under con- It is of note that the anxiolytic (anticonflict) actions of
ditions of stress, activation of 5-HT1A receptors attenu- 5-HT1A agonists such as buspirone and gepirone in the rat
ates the engagement of frontocortical dopaminergic path- VCT are more pronounced following their prolonged ad-
ways (Rasmusson et al., 1994). Moreover, the inhibitory ministration, an observation which parallels their compara-
influence of 5-HT1A receptors upon stress-induced CCK tively slow onset of efficacy in patients (Schefke et al., 1989;
release may be considered as favouring the expression of Yamashita et al., 1995; Silva and Brandao, 2000; Sramek
anxiolytic properties (Becker et al., 1999). et al., 2002). This delay may involve adaptive changes post-
Clearly, then, the role(s) of 5-HT1A receptors in the mod- synaptic to serotonergic neurones, such as the down-regulation
ulation of anxious states is complex, reflecting contrasting of anxiogenic 5-HT2C and 5-HT2A receptors (Section 4.3.4)
roles of pre versus postsynaptic sites and of individual pop- (Millan et al., 1997). However, in man, the retarded re-
ulations of 5-HT1A receptors in discrete supraspinal regions. sponse to buspirone may reflect prior treatment with BZPs
As outlined below, this assertion is borne out by pharmaco- to which extended exposure compromises expression of the
logic studies with the VCT and other experimental models. anxiolytic actions of 5-HT1A agonists (Section 18.3) (De
Martinis et al., 2000).
4.3.2.3. Anxiolytic properties of selective 5-HT1A recep-
tor agonists. Studies of 5-HT1A receptor knock-out mice 4.3.2.4. Role of 5-HT1A receptors in the anxiolytic ac-
(Section 4.3.2.2) notwithstanding, selective 5-HT1A recep- tions of diverse drug classes. The ␤-AR partial agonist,
tor antagonists do not display anxiogenic properties. More- pindolol (Section 4.1.6), is under intensive clinical inves-
over, their (rare) anxiolytic actions may involve blockade tigation for its ability to accelerate the therapeutic actions
of postsynaptic 5-HT1A receptors (probably on GABAergic of antidepressant agents in the treatment both of major de-
neurons in the ventral hippocampus) since they exert little pression and of anxiety disorders (Section 4.4) (McAskill
influence upon forebrain 5-HT release (Millan et al., 1997, et al., 1998; Hirschmann et al., 2000; Ziegenbein et al.,
2000d; Nunes-de-Souza et al., 2002). A report that injec- 2000; Artigas et al., 2001; Stein et al., 2001a,b). It has been
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 127

Table 13
Influence of 5-hydroxytryptophan (5-HTP) and of 5-HT1 receptor ligands upon behaviour in the Vogel Conflict Test
Drug Class Species Route (structure) Effect Active dose range Maximum Reference
(strain) change (%)
5-HTP 5-HT precursor Rat (SD) i.p. + 50 mg/kg +90 Hjorth et al. (1987)
Dihydroergosine 5-HT agonist Rat (W) i.p. + 10–50 mg/kg +380 Peričić and Tvrdeić (1993)
8-OH-DPAT 5-HT1A agonist Rat (SD) i.p. + 0.125–0.25 mg/kg +34 Engel et al. (1984)
Rat (W) i.p. + 0.025–0.05 mg/kg +67 Stefanski et al. (1992)
Rat (W) s.c. + 0.04 mg/kg +300 Dekeyne et al. (2000a)
Rat (LH) i.c. (DRN) + 20–500 ng +448 Higgins et al. (1988)
Rat (W) i.c. (hippocampus) IA >1 ␮g IA Belcheva et al. (1994)
Rat (W) i.c. (hippocampus) + 1–3 ␮g +275 Przegaliński et al. (1994b)
Rat (W) i.c. (hippocampus) + 0.5–1.0 ␮g +72 Stefanski et al. (1993a)
Rat (W) i.c. (N. accumbens) + 1 and 2.5 ␮g +67 Stefanski et al. (1993a)
Buspirone 5-HT1A PAGO Rat (W) i.p. + 5–10 mg/kg +105 Yamashita et al. (1995)
Rat (W) s.c. + 0.63 mg/kg +200 Dekeyne et al. (2000a)
Rat (SD) s.c. + 0.3–1.0 mg/kg +240 Gower and Tricklebank (1988)
Rat (W) p.o. + 10–20 mg/kg +192 Wada and Fukuda (1991)
Rat (LH) i.c. (DRN) + 0.2 ␮g +272 Higgins et al. (1988)
Rat (W) i.c. (hippocampus) + 1–3 ␮g +175 Przegaliński et al. (1994b)
Rat (W) i.c. (hippocampus) − 5 ␮g −65 Stefanski et al. (1993a)
Rat (W) i.c. (N. accumbens) − 5 ␮g −60 Stefanski et al. (1993a)
Ipsapirone 5-HT1A PAGO Rat (W) i.p. + 2.5–10 mg/kg +274 Przegaliński et al. (1992)
Rat (LH) i.c. (DRN) + 0.2 ␮g +198 Higgins et al. (1988)
Rat (W) i.c. (hippocampus) + 1 and 3 ␮g +175 Przegaliński et al. (1994a,b)
Gepirone 5-HT1A PAGO Rat (W) i.p. + 0.3 and 0.6 mg/kg +33 Stefanski et al. (1992)
Rat (W) i.p. + 2–10 mg/kg +220 Yamashita et al. (1995)
Rat (LH) i.c. (DRN) + 1 ␮g +300 Higgins et al. (1992)
Rat (W) i.c. (hippocampus) + 10 and 30 ␮g/side +175 Przegaliński et al. (1994b)
Tandospirone 5-HT1A PAGO Rat (SD) i.p. + 5.0 mg/kg +1000 Shimizu et al. (1987)
Rat (SD) s.c. + 1.0 mg/kg +420 Shimizu et al. (1992)
Rat (SD) i.c. (hippocampus) + 30 and 60 ␮g/side +442 Kataoka et al. (1991)
Indorenate 5-HT1A PAGO Rat (W) i.p. + 3–10 mg/kg +410 Meneses and Hong (1993)

S15535 5-HT1A PAGO Rat (W) s.c. + 2.5–40 mg/kg +300 Dekeyne et al. (2000a)
WAY100,635 5-HT1A Ant Rat (W) s.c. IA >0.63 mg/kg IA Dekeyne et al. (2000a)
Rat (W) s.c. IA 5–10 mg/kg IA Przegaliński et al. (1995)
Rat (W) s.c. + 0.3 and 1.0 mg/kg +325 Griebel et al. (2000)
Rat (SD) s.c. IA >0.3 mg/kg IA Kennett et al. (1998)
Rat (W) i.c. (hippocampus) IA 0.03–0.3 ␮g IA Przegaliński et al. (1995)
NAN190 5-HT1A Ant Rat (W) i.p. IA 0.5–1.0 mg/kg IA Przegaliński et al. (1994b)
Rat (W) i.c. (hippocampus) IA 0.1–1.0 ␮g IA Przegaliński et al. (1994b)
CGS12066B 5-HT1B agonist Rat (LH) i.c. (DRN) IA >2.5 ␮g IA Higgins et al. (1992)
SB224,289 5-HT1B Ant Rat (W) s.c. IA >10 mg/kg IA Brocco (unpublished observations)
Abbreviations—8-OH-DPAT: 8-hydroxydipropylaminotetralin; PAGO: partial agonist; N. accumbens: nucleus accumbens and IA: no significant change
(inactive). For other abbreviations, see Table 2.

proposed that its distinctive partial agonist properties at 5-HT1A autoreceptors by their direct blockade (Haddjeri
␤1 - and ␤2 -ARs facilitatory to corticolimbic dopaminergic et al., 1998; Artigas et al., 2001; Castro et al., 2003). One
and noradrenergic pathways contribute to its favourable difficulty with this hypothesis is that doses of pindolol em-
influence upon depressive—though not anxious—states ployed clinically are too low to attain marked occupation
(Gobert and Millan, 1999a). Nevertheless, “antagonist” ac- of 5-HT1A autoreceptors (Cremers et al., 2001; Martinez
tions of pindolol at 5-HT1A autoreceptors have generally et al., 2001; Rabiner et al., 2001). Further, several stud-
been advanced as the mechanism underlying its reinforce- ies have shown that pindolol can behave as an agonist
ment of the actions of antidepressant agents (Corradetti at 5-HT1A autoreceptors in rodents (Gobert and Millan,
et al., 1998; Artigas et al., 2001; Newman-Tancredi et al., 1999a; Lejeune and Millan, 2000; Sprouse et al., 2000;
2001). Pindolol is, thus, posited to rapidly mimic the grad- Artigas et al., 2001; Millan et al., 2001f). Though incon-
ual (long-term) antidepressant-induced desensitisation of sistent with long-term, adjunctive antidepressant properties,
128 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

750
LICKS/3 MIN
LICKS/3 MIN

500 500

250 250

0 0
VEH VEH 0.16 0.63 2.5 10.0 MG/KG, S.C. VEH VEH 0.04 0.16 0.63 2.5 MG/KG, S.C.
FLESINOXAN MDL 100,907
NO SHOCKS SHOCK NO SHOCK SHOCK

750
750
LICKS/3 MIN

LICKS/3 MIN

500
500

250
250

0 0
VEH VEH 0.63 2.5 10.0 40.0 MG/KG, I.P. VEH VEH 0.63 10.0 40.0 MG/KG, S.C.
SB 242,084 CITALOPRAM

NO SHOCK SHOCK NO SHOCK SHOCK

Fig. 7. Actions of the 5-HT1A receptor agonist, flesinoxan, the 5-HT2A antagonist, MDL100,907, the 5-HT2C antagonist, SB242,084, and the selective
5-HT reuptake inhibitor, citalopram, in the Vogel Conflict Test. Data are means ± S.E.M.s. Star indicates significance of differences to corresponding
vehicle values. ( ) P < 0.05. Data are from Gobert et al. (2000), Dekeyne et al. (2000a,b), and Brocco and Millan (unpublished observations).

the activation of presynaptic 5-HT1A receptors by pindolol VCT via activation of 5-HT1A autoreceptors (Millan et al.,
likely underlies its acute anxiolytic actions in the VCT and 1999a; Cussac et al., 2002a).
other procedures (Table 10) (Przegaliński et al., 1994a). In It has recently been recognized that many antiparkinson
this light, it is conceivable that the intrinsic partial agonist agents, such as apomorphine, possess significant affinity for
properties of several novel antidepressant agents at 5-HT1A 5-HT1A receptors (Seyfried et al., 1989; Newman-Tancredi
receptors, such as vilazadone (Bartoszyk et al., 1997; Treit et al., 1999, 2002a,b; Millan et al., 2002). They almost in-
et al., 2001; Page et al., 2002; Tordera et al., 2002), may variably behave as partial agonists exhibiting intrinsic ac-
permit rapid and sustained anxiolytic actions via a com- tivity sufficient to stimulate presynaptic, but not postsynap-
parable mechanism. Their effects in the VCT would merit tic, populations (Seyfried et al., 1989; Newman-Tancredi
evaluation. et al., 1999, 2002b; Millan et al., 2002). Interestingly, several
As argued elsewhere (Millan, 2000; Bantock et al., 2001), antiparkinsonian agents manifest anxiolytic properties in
stimulation of presynaptic 5-HT1A receptors imparts benefi- experimental models, including the VCT. While not chal-
cial properties to antipsychotic agents in terms of: (1) an aug- lenging the significance of D2 autoreceptor activation in the
mented therapeutic window of desirable versus secondary actions of those agents which preferentially act at D2 versus
actions: (2) improved control of cognitive-deficit symptoms 5-HT1A sites (Section 4.2.3.2), recruitment of 5-HT1A re-
and (3) auxiliary antidepressive and anxiolytic properties ceptors may contribute to the anxiolytic profiles of other an-
which improve patient compliance and, more generally, en- tiparkinsonian drugs—such as roxindole—which show more
courage patient recovery. As outlined in Section 4.5, sev- pronounced agonist properties at 5-HT1A than at D2 recep-
eral antipsychotic agents have been shown to interact with tors (Seyfried et al., 1989; Newman-Tancredi et al., 1999).
5-HT1A receptors—in addition to dopamine D2 receptors Anxious symptoms are a common and troubling feature of
and other monoaminergic sites. Of these, the benzopyrro- Parkinsons disease which may even anticipate motor dys-
lidine derivative, S16924, is of special note inasmuch as it function (Section 4.2.1), so it would be judicious to further
displays potent agonist properties at 5-HT1A receptors and characterize the role of 5-HT1A receptor activation in the
has been shown to express robust anxiolytic effects in the influence of antiparkinson drugs upon mood.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 129

Fig. 8. Schematic illustration of the sites of action of agents interacting with serotonergic receptors in the Vogel Conflict Test. The loci of action indicated
are based upon results published for the Vogel Conflict Test (see tables). Additional sites of action revealed by other procedures should be borne in
mind. Abbreviations—5-HT: serotonin; D/MRN: dorsal/median raphe nucleus and PAG: periaqueductal gray.

4.3.2.5. Pre- and postsynaptic sites of action of 5-HT1A the assertion that 5-HT1A autoreceptors play a major role
receptor agonists. Notwithstanding extensive efforts, it is in mediating the anxiolytic actions of 5-HT1A receptor ag-
currently impossible to specify the precise contribution of onists.
pre- as compared with postsynaptic receptors to the anxi- First, anxiogenic actions have been documented upon
olytic properties of 5-HT1A receptor agonists (Fig. 8). In- administration of 5-HT1A receptor agonists into the septum,
deed, it is arguably illusory to insist upon a unitary response hippocampus and amygdala (Hodges et al., 1987; Andrews
to this question in light of the plethora of factors which bear et al., 1994; Belcheva et al., 1994, 1997; Gonzalez et al.,
on the results of studies addressing this issue. For exam- 1996a; Micheau and Van Marrewijk, 1999; Nunes-de-Souza
ple: the choice of drug, its dose, duration and site of ap- et al., 2000, 2002), while recruitment of postsynaptic
plication; the species, strain and gender under investigation; 5-HT1A receptors has been implicated in the anxiogenic
the experimental model, its end-point and the magnitude of effects of CCK (Section 9.1.1) (Bickerdike et al., 1995).
stress provoked, and the degree of activation of serotoner- These anxiogenic effects may be provoked by activation
gic neurones (Blanchard et al., 1992; Crespi et al., 1992; of postsynaptic 5-HT1A receptors inhibitory to GABAergic
Schreiber and De Vry, 1993a,b; De Vry, 1995; Handley, neurones (Section 4.3.2.2) (Halasy et al., 1992; Matsuyama
1995; Menard and Treit, 1999; Olivier et al., 2001; Vaidya et al., 1997; Koyama et al., 1999; Kishimoto et al., 2000a,
et al., 2002). Moreover, as exemplified by (Jolas et al., 1995), 2001). Second, introduction of 5-HT1A receptor agonists
even microinjection studies may not provide definitive ev- into the DRN and (less prominently) the MRN is accompa-
idence in this regard owing to the rapid and extensive dif- nied by anxiolytic effects (Carli et al., 1989; Higgins et al.,
fusion through neuronal tissue of lipophylic ligands such as 1992; Schreiber and De Vry, 1993a; Andrews et al., 1994;
the 5-HT1A agonist, 8-OH-DPAT, which quickly attains the Maurel-Remy et al., 1996; Cervo et al., 2000). Third, par-
DRN even upon introduction into the dorsal hippocampus. enteral administration of selective 5-HT1A partial agonists
Supporting a role for 5-HT1A autoreceptors in the VCT: (1) possessing efficacy adequate for stimulation of pre but not
anxiolytic effects were reported upon direct introduction of postsynaptic populations may evoke anxiolytic actions more
5-HT1A receptor agonists into raphe nuclei (Eison et al., pronounced than those elicited by full agonists which like-
1986; Carli and Samanin, 1988); (2) agonists of modest ef- wise activate postsynaptic receptors (De Vry, 1995; Millan
ficacy which stimulate pre but not postsynaptic populations et al., 1997; Przegaliński et al., 1995; Dekeyne et al., 2000b).
of 5-HT1A receptor exert robust anxiolytic effects (Millan Fourth, anxiolytic effects of 5-HT1A partial agonists and
et al., 1997; Cervo et al., 2000; Dekeyne et al., 2000b; 5,7-dihydroxytryptamine lesions of serotonergic pathways
Bojarski et al., 2002) and (3) in certain studies, elimination in the VCT and other models are blunted by antagonists at
of serotonergic neurones exerted anxiolytic effects and inter- GABAA receptors (Söderpalm and Engel, 1991; Söderpalm
fered with the anxiolytic actions of 5-HT1A agonists (Engel et al., 1992, 1997; Fernández-Guasti and López-Rubalcava,
et al., 1984; Eison et al., 1986; Carli et al., 1989; Söderpalm 1998; Nazar et al., 1999a,b). These findings reflect either:
et al., 1992, 1997; Schreiber and De Vry, 1993a; Cervo and (1) indirect suppression of an inhibitory influence of postsy-
Samanin, 1995; Nazar et al., 1999a,b; Andrade and Graeff, naptic 5-HT1A receptors upon GABA release (see above);
2001; Green and McGregor, 2002; Netto et al., 2002). Data (2) release of a positive allosteric modulatory of GABAA
from other conflict models may also be advanced to bolster sites, such as neurosteroids (Söderpalm et al., 1997) and/or
130 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

(3) removal of the inhibitory influence of 5-HT2C receptors well-documented anxiogenic role of postsynaptic 5-HT2C
upon hippocampal populations of colocalized GABAA re- receptors (Section 4.3.4.4), activation of 5-HT1A autore-
ceptors (Section 4.3.4.4) (Huidobro-Toro et al., 1996). Fifth, ceptors, in moderating serotonergic transmission, will indi-
as considered in Section 4.4.2, the prolonged administration rectly relieve their engagement by reducing 5-HT release.
of SSRIs and other classes of antidepressant agent is ac- On the other hand, inasmuch as postsynaptic 5-HT1A recep-
companied by the progressive transformation of their acute, tors are inhibitory to neuronal activity, their activation will
anxiogenic properties into long-term anxiolytic actions. In interfere with the excitatory effects of co-expressed 5-HT2C
parallel, postsynaptic 5-HT1A receptors develop an increase receptors (Millan et al., 1992). Third, postsynaptic 5-HT1A
in sensitivity, a finding incompatible with a qualitatively receptors exert opposing functional influences upon
different influence of chronic versus acute treatment by an- GABAergic transmission. Thus, while 5-HT1A receptors
tidepressants upon anxious states (Blier and De Montigny, can directly inhibit the activity of GABAergic interneurones
1994; Newman et al., 1998; Haddjeri et al., 1998; Hjorth (Section 4.3.2.2), where colocalized with GABAA receptors
et al., 2000; Artigas et al., 2001; Hensler et al., 2002; Castro on individual neurones in the hippocampus and other struc-
et al., 2003). tures, 5-HT1A sites will synergistically act to suppress neu-
Contrariwise, in support of a role of postsynaptic sites, ronal excitability. This configuration is consistent with the
the anxiolytic properties of 5-HT1A receptor agonists in the concurrent expression of anxiolytic and anxiogenic actions
VCT were (in certain studies) unaffected by 5-HT depletion, by postsynaptic 5-HT1A receptors—perhaps subject to fine
and they are generally seen upon direct application into the tuning at the level of serotonergic terminals presynaptic to,
hippocampus—and nucleus accumbens (Commissaris et al., or in series with, GABAergic neurones. Fourth, no 5-HT1A
1981; Chojnacka-Wójcik and Przegaliński, 1991; Kataoka receptor agonist is absolutely selective, so it should not be
et al., 1991; Przegaliński et al., 1992, 1994b; Shimizu et al., forgotten that actions at sites other than 5-HT1A receptors
1992; Stefański et al., 1993a; Belcheva et al., 1994, 1997). A will modify (in a drug-specific fashion) actions in the VCT
predominant role of postsynaptic sites in the VCT is under- and other procedures.
scored by studies of other conflict procedures showing that Regardless of the role(s) of specific populations of
microinjection of 5-HT1A receptor agonists into the septum, 5-HT1A receptor, from a clinical perspective, the overall in-
hippocampus and/or amygdala is accompanied by anxiolytic fluence of drugs upon parenteral administration is a crucial
actions, while lesions of serotonergic neurones little modi- consideration. In this regard, in analogy to observations in
fies the effects of their systemic administration (Kostowski the VCT, partial agonists at 5-HT1A receptors have repro-
et al., 1989; Schreiber and De Vry, 1993a,b; Menard and ducibly shown anxiolytic activity in GAD patients—though,
Treit, 1998; Zangrossi et al., 1999; Groenink et al., 2000). as discussed elsewhere, any efficacy in the management of
Moreover, while bearing in mind the contrasting nature of other anxiety disorders remains to be proven (Pecknold,
the model (Section 1.3) (Graeff et al., 1993), direct introduc- 1997; Rickels et al., 1997; Laakman et al., 1998; Apter
tion of 5-HT1A agonists into the PAG eliminates the aver- and Allen, 1999; Broocks et al., 2000; Landen et al., 2001;
sive effects of its electrical stimulation: though they are in- Oshima et al., 2001). In healthy volunteers, furthermore,
active upon systemic administration, this may reflect their buspirone exerted anxiolytic actions against both condi-
acceleration of corticolimbic NA release (Section 4.3.2.2) tioned and, at higher doses, unconditioned aversive stimuli
(Noguiera and Graeff, 1995; Beckett and Marsden, 1997; (Hellewell et al., 1999; Bond et al., 2003).
Jenck et al., 1999; Millan et al., 2000d; Canto-de-Souza Thus, though the influence of 5-HT1A receptors upon anx-
et al., 2002; Graeff, 2002; Jacob et al., 2002). ious states is highly complex, their activation is generally as-
sociated with anxiolytic actions amenable to analysis by use
4.3.2.6. Multiple sites of action of 5-HT1A receptor lig- of the VCT. This procedure will prove of continued utility in
ands: clinical studies. From the above discussion of the further elucidation of their roles. Moreover, in analogy
innumerable—ostensibly contradictory—findings, the ques- to BZPs (Section 2.2.2), the anxiolytic properties of 5-HT1A
tion of the respective roles of pre versus postsynaptic partial agonists in the VCT provide an important parallel to
5-HT1A receptors may appear virtually insoluble. Indeed, clinical studies demonstrating their therapeutic efficacy in
it would seem unrealistic to categorically advocate an ex- the management of GAD.
clusive role for either in mediating the anxiolytic actions of
5-HT1A receptor agonists (Fig. 7). The following comments 4.3.3. 5-HT1B receptors
are of pertinence in this regard.
First, as a function of serotonergic tone, an intermedi- 4.3.3.1. Coupling, localization and interactions. 5-HT1B
ate degree of drug efficacy at 5-HT1A receptors sufficient receptors share the negative coupling of 5-HT1A receptors
for the maximal and submaximal activation of pre- and to adenylyl cyclase via engagement of Gi . Postsynaptic
postsynaptic sites, respectively, may result in optimal anxi- populations are found in the FCX, hippocampus, amygdala
olytic activity. Second, it is entirely plausible that both pre- and several other corticolimbic structures of rodents and,
and postsynaptic populations of 5-HT1A receptors medi- at particularly high levels, of man (Bruinvels et al., 1994;
ate anxiolytic properties. For example, in relation to the Bonaventure et al., 1998). As concerns their role in the
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 131

modulation of mood, most attention has, however, been fo- in a VCT (Gardner, 1985). Other 5-HT1B receptor agonists
cussed on inhibitory 5-HT1B autoreceptors localized on the have likewise failed to display anxiolytic effects either upon
terminals of serotonergic neurones (Bruinvels et al., 1994; systemic administration or upon direct introduction into the
Millan et al., 2000d; Hopwood and Stamford, 2001a; De DRN—wherein a minor population of 5-HT1B autorecep-
Groote et al., 2002). Complementary to dendritic 5-HT1A tors also occurs (Table 13) (Higgins et al., 1992; Millan
autoreceptors, these terminal-localized 5-HT1B autorecep- et al., 2000d; Brocco and Millan, unpublished observa-
tors play a key role in the phasic, inhibitory control of tions). Though anxiolytic actions of the poorly-selective
5-HT release (Barnes and Sharp, 1999; Millan et al., 1999b, serotonergic agonist, meta-chlorophenylpiperazine (mCPP)
2000d). Furthermore, these presynaptic populations have (Section 4.3.4.1), in the VCT were ascribed to activation
been specifically implicated in the control of anxiety under of 5-HT1B receptors, selective antagonists were not at that
conditions of stress (Clark et al., 2002). time available for study. Retrospectively, it is more likely
It should also be pointed out that postsynaptic popula- that other mechanisms were involved, probably engagement
tions of 5-HT1B receptor are inhibitory to glutamatergic and of 5-HT1A receptors or blockade of 5-HT3 sites (Sections
GABAergic neurones in the cortex and LC (Bobker and 4.3.2 and 4.3.5) (Chojnacka-Wójcik and Klodzińska,
Williams, 1989; Tanaka and North, 1993). 1992).
The above comments indicate that the engagement of
4.3.3.2. Control of anxious states. In view of the mutual postsynaptic 5-HT1B receptors may enhance anxious states.
inhibitory control exercised by 5-HT1B and 5-HT1A autore- In line with this interpretation, it has been reported that
ceptors upon 5-HT release—and arguing by analogy—it (presumably postsynaptic) 5-HT1B receptor blockade is as-
might be predicted that: (1) mice overexpressing 5-HT1B sociated with anxiolytic properties (Chopin et al., 1994).
receptors would present an anxiolytic phenotype; (2) mice However, such observations require confirmation with gen-
deficient in 5-HT1B receptors would mimic the anxiogenic uinely selective antagonists such as the highly-selective
profile of their 5-HT1A receptor knock-out counterparts agent, SB224,289, which was ineffective in the VCT
and (3) 5-HT1B agonists would elicit anxiolytic actions (Table 13). Further, in certain paradigms, evidence has
in parallel with a suppression of corticolimbic 5-HT re- been acquired for anxiolytic properties of 5-HT1B ago-
lease. However, overexpression of 5-HT1B autoreceptors nists (Bourin and Hascoët, 2003; Sanchez, 2003). Thus,
potentiated the anxious behaviour elicited by stress (Clark it would be premature to form any definitive conclusions
et al., 2002). Further, the rather unspectacular phenotype as regards the role of pre- and postsynaptic populations of
of mice lacking 5-HT1B receptors suggests a decrease 5-HT1B receptors in view of: (1) the relative poverty of
in anxiety (Ramboz et al., 1996; Malleret et al., 1999; data concerning actions of selective agonists and antago-
Zhuang et al., 1999; Belzung, 2001; Dirks et al., 2001; nists; (2) difficulties in pharmacologically distinguishing
Lesch, 2001). Alterations in coupling efficiency at pre- and the roles of pre versus postsynaptic populations of 5-HT1B
postsynaptic (amygdala-localized) populations of 5-HT1A receptors, an issue requiring direct exploration in mice
receptors in animals lacking 5-HT1B sites complicates in- sustaining regionally-circumscribed (pre or post synaptic)
terpretation of their phenotype (Knobelman et al., 2001; genetic elimination of 5-HT1B receptors; (3) lack of infor-
Ase et al., 2002), and a compensatory reduction in activ- mation concerning the neuronal substrates underlying the
ity at cerebral 5-HT2C receptors may contribute to their putative anxiogenic role of postsynaptic 5-HT1B sites and
low level of anxiety (Clifton et al., 2003). In any case, (4), the need for studies of the influence of 5-HT1B autore-
the lack of alteration in 5-HT release suggests (by anal- ceptors upon stress and fear-induced activation of seroton-
ogy to 5-HT1A receptor knock-out mice) (Section 4.3.2.2) ergic pathways. Additional analysis of the significance of
that the reduced anxiety cannot automatically be assigned 5-HT1B receptors would be of interest in the light of evi-
to a loss of inhibitory, presynaptic populations of 5-HT1B dence that adaptive changes (desensitization) in presynaptic
receptor. populations of 5-HT1B receptors upon long-administration
Indeed, a putative anxiogenic role of postsynaptic 5-HT1B of antidepressant agents develop with a time-course par-
receptors might explain the paradoxical finding that, at alleling that required for onset of their anxiolytic actions
doses which elicit reductions in 5-HT release comparable (Section 4.4.2) ((Blier and De Montigny, 1994; Anthony
to those induced by 5-HT1A agonists, 5-HT1B receptor ag- et al., 2000; Davidson and Stamford, 2000); Hjorth et al.,
onists do not, in general, elicit anxiolytic effects. Indeed, 2002).
several reports of anxiogenic actions of 5-HT1B agonists Regardless of the precise role of 5-HT1B receptors in the
in paradigms of exploratory behaviour were recently un- modulation of anxious states, the above comments underline
derpinned by a rigorous pharmacological demonstration arguments forwarded elsewhere (Sections 2.2.2.1, 4.3.2.4
that activation of 5-HT1B receptors enhances anxiety in and 18.1) that: (1) a reduction in 5-HT release is not invari-
the plus-maze procedure (Moret and Briley, 2000; Lin and ably accompanied by an anxiolytic profile; and (2) it would
Parsons, 2002). Data from conflict procedures are limited be naı̈ve to automatically attribute changes in anxious states
but, in line with these findings, it has been reported that the to actions at presynaptic as compared to postsynaptic popu-
5-HT1B agonist, RU24969, enhanced punished responding lations of 5-HT1B and 5-HT1A receptor.
132 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

4.3.4. Serotonin 5-HT2 receptor subtypes mented anxiolytic properties of 5-HT2A receptor antago-
nists (Stutzmann et al., 1991; Motta et al., 1992; Costall
4.3.4.1. The 5-HT2 receptor family: cellular coupling. and Naylor, 1995; Mora et al., 1997; Graeff et al., 1998).
The three closely-related members of the 5-HT2 recep- Anxiolytic actions of selective 5-HT2A antagonists have
tor family, 5-HT2A , 5-HT2B and 5-HT2C receptors, are also been seen—albeit at relatively high doses—in the
poorly differentiated by first-generation pharmacologi- VCT (Table 14 and Fig. 7) ((Griebel et al., 1997a); Brocco
cal agents such as the “5-HT2 ” receptor agonist, mCPP and Millan, unpublished observations); (4) at the postsy-
(which also interacts with 5-HT1A , 5-HT1B and 5-HT3 re- naptic level, stimulation of 5-HT2A receptors suppresses
ceptors), and the “5-HT2 ” receptor antagonist, ritanserin GABAergic transmission in the FCX (Feng et al., 2001)
(Barnes and Sharp, 1999). Only more recently, with via the protein kinase C-mediated phosphorylation of the
the discovery of highly-selective ligands distinguishing ␥2 subunit of colocalized GABAA receptors (Yan, 2002).
5-HT2A , 5-HT2B and 5-HT2C sites, and the generation of Whether similar interactions occur in other limbic struc-
5-HT2C receptor knock-out mice, has it become feasible tures awaits elucidation. (5) By analogy to the effects of
to address the issue of their individual roles (Barnes and long-term agonist administration, down-regulation (desen-
Sharp, 1999). Paralleling their similar patterns of ligand sitisation and internalisation) of cortical and sub-cortical
recognition, 5-HT2A , 5-HT2B and 5-HT2C receptors all populations of 5-HT2A receptors is provoked by chronic
stimulate the activity of phospholipase C (resulting in an treatment with antidepressant agents: this may contribute
increase in intracellular levels of Ca2+ ) via the recruitment to their delayed onset of anxiolytic (and other therapeutic)
of Gq (Gerhardt and Heerikhuizen, 1997; Barnes and Sharp, properties (Section 4.4) (Deakin, 1988, 1991; Graeff et al.,
1999; Cussac et al., 2002a,b). 1996a,b; Burnet et al., 1994, 1996; Bhatnagar et al., 2001;
Millan et al., 2001c,d; Bhattacharyya et al., 2002; Van
4.3.4.2. 5-HT2A receptors. Though a possible role of Oekelen et al., 2003).
5-HT2A receptors in the modulation of anxious states has Notwithstanding the above findings, 5-HT2A receptor an-
attracted comparatively little attention, they are found at tagonists have proven essentially inactive in several models
a high concentration in the entorhinal cortex, amygdala, of potential anxiolytic activity (Gleeson et al., 1989; Griebel
nucleus accumbens and hippocampus (Lopez-Giménez et al., 1997a; Setem et al., 1999; Blanchard et al., 2003;
et al., 1997, 1998; Cornea-Hébert et al., 1999; Xu and Bourin and Hascoët, 2003). Moreover, there are also reports
Pandey, 2000; Vaidya et al., 2001; Millan, 2002a). Further, of anxiogenic actions of 5-HT2A receptor antagonists, prin-
alterations in the functional status of 5-HT2A receptors cipally in models of untrained behaviours, but also including
have been seen upon acute imposition of stress, while a a conflict procedure in mice. One locus of action appears
pronounced elevation in the cortical density of 5-HT2A re- to be the amygdala (Olivier et al., 1998; Setem et al., 1999;
ceptors was evoked by protracted exposure to anxiogenic Maisonette et al., 2000). An explanation for the anxiogenic
stimuli (Chaouloff et al., 1994; Vaidya et al., 1999; Izumi effects of 5-HT2A receptor antagonists may be their relief
et al., 2002). Under resting and stressful conditions, 5-HT2A of a presynaptic, tonic, excitatory influence of 5-HT2A re-
receptor agonists exert a marked facilitatory influence upon ceptors upon GABAergic neurones in the amygdala and the
the activity of the hypothalamo-corticotropic axis and upon hippocampus (Cozzi and Nichols, 1996; Shen and Andrade,
sympathetic outflow (McCall and Clement, 1994; Welch 1998; Abi-Saab et al., 1999; Rainnie, 1999; Stutzmann
and Saphier, 1994; Chaouloff, 1995; McKittrick et al., and LeDoux, 1999; Zhou and Hablitz, 1999; Jakab and
1995; Nankai et al., 1995; Jorgenson et al., 1998; Rivet Goldman-Rakic, 2000; Xu and Pandey, 2000; Martin-Ruiz
et al., 2001; Hemrick-Lubecke and Evans, 2002; Carrasco et al., 2001). Thus, a loss of a spontaneous activity at 5-HT2A
and Van de Kar, 2003). receptors, via a reduction of GABA release, may lead to an
Several complementary lines of evidence suggest that the enhancement of anxious states. Further, in contrast to the
engagement of 5-HT2A receptors may contribute to anxious direct, excitatory influence of 5-HT2A receptors upon nora-
states: (1) activation of 5-HT2A receptors enhances the cor- drenergic perikarya in the LC (vide supra), they may exert an
ticolimbic efflux of glutamate, derived form both glial as indirect, inhibitory influence via the activation of GABAer-
well as neuronal sources Rocher et al., 1999; Meller et al., gic interneurones (Szabo and Blier, 2001). Despite the in-
2002; (2) 5-HT2A receptors are localized on cell bodies of activity of 5-HT2A receptor agonists in the VCT (Table 14),
the LC and the ventrotegmental area wherein they exert a similar, positive influence of 5-HT2A receptors upon
an excitatory influence upon corticolimbic noradrenergic GABAergic neurones in the PAG and the inferior colliculus
and dopaminergic pathways, respectively. 5-HT2A recep- may account for the generally BZP-like anxiolytic actions
tors also, via actions in the FCX and, possibly, the DRN of 5-HT2A agonists in models of anti-aversive (defensive)
itself, potentiate 5-HT output in cortical regions, though activity generated from these structures (Jenck et al., 1990,
not necessarily in other cerebral structures Gobert and 1998; Melo and Brandao, 1995; Nogueira and Graeff, 1995;
Millan, 1999b; Doherty and Pickel, 2000; Liu et al., 2000; Brandao et al., 2002; Castilho et al., 2002; Graeff, 2002;
Millan et al., 2000d; Martin-Ruiz et al., 2001; Pehek et al., Griffiths and Lovick, 2002; Jacob et al., 2002), as well as
2001; Nocjar et al., 2002. (3) Several studies have docu- their anxiolytic effects in several other procedures (Onaivi
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 133

Table 14
Influence of drugs interacting with 5-HT2 receptor subtypes upon behaviour in the Vogel Conflict Test
Drug Class Species Route (structure) Effect Active dose range Maximum Reference
(strain) change (%)
DOI 5-HT2A agonist Rat (W) s.c. IA >0.63 mg/kg IA Brocco (unpublished observations)
MDL100,907 5-HT2A Ant Rat (SD) i.p. + 1.0 mg/kg +181 Griebel et al. (1997a)
Rat (W) s.c. + 0.16–2.5 mg/kg +265 Brocco (unpublished observations)
Ketanserin 5-HT2A Ant Rat (W) s.c. IA >2.5 mg/kg IA Brocco (unpublished observations)
BW723C86 5-HT2B agonist Rat (SD) i.p. + 10 and 30 mg/kg +193 Kennett et al. (1998)
Rat (SD) i.c. (amygdala) IA 0.03–0.3 ␮g IA Duxon et al. (1997a)
SB204,741 5-HT2B Ant Rat (W) i.p. IA >40 mg/kg IA Brocco et al. (1998)
SB215,505 5-HT2B Ant Rat (SD) p.o. IA >3.0 mg/kg IA Kennett et al. (1998)
mCPP 5-HT2C agonist Rat (SD) i.p. IA 0.6–3.0 mg/kg IA Kennett et al. (1998)
Rat (W) i.p. + 0.12–0.5 mg/kg +367 Chojnacka-Wójcik and Klodzińska
(1992)
Rat (LH) i.c. (DRN) IA >12.5 ␮g IA Higgins et al. (1992)
Ro60-0175 5-HT2C agonist Rat (SD) s.c. IA 0.1–3.0 mg/kg IA Kennett et al. (2000)
Rat (W) s.c. IA >2.5 mg/kg IA Brocco (unpublished observations)
SB206,553 5-HT2C Ant Rat (SD) i.p. + 3–30 mg/kg +319 Griebel et al. (1997a)
Rat (W) i.p. + 10–40 mg/kg +610 Dekeyne et al. (2000a)
SB242,084 5-HT2C Ant Rat (W) i.p. + 0.16–2.5 mg/kg +50 Millan et al. (2001a,b,c,d,e,f)
Ritanserin 5-HT2C Ant Rat (W) i.p. + 2.5–5.0 mg/kg +57 Stefanski et al. (1992)
Rat (W) s.c. IA 0.63–40 mg/kg IA Brocco et al. (1990)
Abbreviations—DOI: 2,5-dimethoxy-4-iodoamphetamine; mCPP: 1-(3-chlorophenyl)piperazine and IA: no significant change (inactive). For other abbre-
viations, see Table 2.

et al., 1995; Hawkins et al., 2002; Sanchez, 2003). appear of some importance (Brunello et al., 1995; Sperling
These observations concur, further, with the ability of 5- and Demling, 1997; Meltzer, 1999; Millan et al., 2000a,c;
HT2A receptor antagonists to attenuate the suppression of Ninan et al., 2002).
fear-associated ultrasonic vocalizations in rats by SSRIs
(Section 4.4.2) (Schrieber et al., 1998; Sanchez and Mork, 4.3.4.3. 5-HT2B receptors. The low level of 5-HT2B sites
1999). in the CNS as compared to peripheral tissues has discour-
Thus, the opposing presynaptic (facilitatory) and post- aged interest in their potential role in the control of anx-
synaptic (inhibitory) influence of 5-HT2A receptors upon ious states. However, they have been detected in the amyg-
GABAergic transmission may contribute to their multiple, dala, lateral septum and hypothalamus (Duxon et al., 1997a;
contrasting roles in the modulation of anxious states. The in- Russel et al., 2002), and anxiolytic actions of the 5-HT2B
fluence of 5-HT2A receptor ligands clearly depends upon the receptor agonist, BW723C86, have been observed in several
type of anxious state under experimental or clinical study. experimental models (Kennett et al., 1996a; Duxon et al.,
Currently, the balance of evidence favors predominantly anx- 1997b). Further, anxiogenic properties of the 5-HT2B/2C
iolytic properties of 5-HT2A receptor antagonists, at least in agonist, Ro60,0175, at 5-HT2C receptors (Section 4.3.4.4)
models of conditioned fear, and there is a preliminary re- were revealed in the presence of selective 5-HT2B recep-
port of anxiolytic actions of a preferential 5-HT2A receptor tor antagonists to mask an anxiolytic component of activity
antagonist in man (Deakin, 1988; Deakin and Graeff, 1991; mediated by these sites (Lightowler et al., 2000). Microin-
Connell et al., 1995). Furthermore, as pointed out above, jection studies indicate that 5-HT2B sites in the (medial)
sustained activation of 5-HT2A receptors ultimately leads to amygdala transduce the anxiolytic actions of BW723C86
their functional disengagement by down-regulation. in the social interaction procedure (Hascoët et al., 2000).
In view of the extensive use of agents possessing 5-HT2A In the VCT, robust, dose-dependent and specific increases
antagonist properties in the management of anxiety dis- in responses were reported upon systemic administration of
orders (for example, the antidepressants, mirtazapine and BW723C86: significantly, in light of its limited selectivity,
trazodone), schizophrenia (for example, the antipsychotics, selective antagonists at 5-HT2B receptors abolished its anxi-
clozapine and risperidone) and depression (for example, the olytic properties (Table 14) (Kennett et al., 1998). However,
tricyclic agents, clorimipramine and mianserin) (Sections the population of 5-HT2B sites which mediates anxiolytic
4.4 and 4.5), further exploration of the significance of actions of BW723C86 in the VCT is not located in the amyg-
5-HT2A receptors in the control of anxious states would dala, and remains to be identified (Duxon et al., 1997b).
134 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Selective antagonists at 5-HT2B receptors are inactive in the interaction procedure (Gacsalyi et al., 1997; Griebel et al.,
VCT and several other procedures (Kennett et al., 1996a; 1997a; Mora et al., 1997; Graeff et al., 1998; Dekeyne et al.,
Brocco et al., 1998; Lightowler et al., 2000). 2000b; Millan et al., 2001a) and various conflict paradigms
(Chopin and Briley, 1987; Brocco et al., 1990, 1998; Kennett
4.3.4.4. 5-HT2C receptors. 5-HT2C receptors are enriched et al., 1995, 1996b, 1997a; Wood et al., 2001; Martin et al.,
in the FCX, hippocampus, amygdala, nucleus accumbens, 2002a). Similarly, these and other 5-HT2C receptor antago-
hypothalamus, PAG and septum, offering an anatomical nists enhance punished responses in the VCT (Table 14 and
foundation for their major role in the control of anxious Fig. 7) (Stefański et al., 1992; Gacsalyi et al., 1997; Griebel
states (Wright et al., 1995; Sharma et al., 1997; Clemett et al., 1997a; Dekeyne et al., 2000b; Millan et al., 2001a).
et al., 2000; Lopez-Giménez et al., 2001; Anderson et al., Such observations coincide with reports suggestive of di-
2002). Though 5-HT2C receptors (probably in the raphe minished anxiety in mice lacking 5-HT2C receptors (Heisler
nuclei) may exert a modulatory influence upon serotonergic et al., 1998a; Rocha et al., 2002). Further, anxiolytic actions
transmission under certain conditions, the functional sig- of ritanserin and a further “5-HT2 ” receptor antagonist, me-
nificance of this action remains unclear (Barnes and Sharp, tergoline, have been seen: (1) in patients with GAD and (2)
1999; Clemett et al., 2000; Gobert et al., 2000; Millan in healthy probands subjected to protocols of conditioned
et al., 2000d). Via activation of GABAergic interneurones, or anticipatory fear (Ceulemans et al., 1985; Graeff et al.,
5-HT2C receptors exert an indirect, inhibitory influence 1985; Hensman et al., 1991; Katz et al., 1993). Finally, a
upon the tonic and stress-induced activity of corticolimbic progressive reduction in the density of 5-HT2C receptors in
noradrenergic and mesocortical dopaminergic projections the frontal cortex, amygdala and hippocampus likely con-
(Gobert et al., 2000; Millan et al., 2000d; Di Giovanni et al., tributes to the delayed onset of anxiolytic actions of SS-
2001; Di Matteo et al., 2001; Pozzi et al., 2002). Conse- RIs and other classes of antidepressant agent (Section 4.4.2)
quently, the activity of these pathways is powerfully sup- (Deakin, 1988, 1991; Rocha et al., 1994; Bristow et al.,
pressed and enhanced by 5-HT2C agonists and antagonists, 2000; Dekeyne et al., 2000a; Van Oekelen et al., 2003).
respectively, actions which may well influence their control It is, thus, possible to marshall an impressive array of
of anxious states (see below). In limbic structures, activa- data in support of anxiolytic properties of 5-HT2C receptor
tion of 5-HT2C receptors, via a protein kinase C-dependent antagonists. Nevertheless, in a clinical model of uncon-
mechanism, interferes with the operation of colocalized ditioned fear (simulated public-speaking), ritanserin and
GABAA receptors, providing a mechanism for expression metergoline were found to potentiate anxiety, whereas
of their anxiogenic properties (Huidobro-Toro et al., 1996). anxiolytic effects were acquired with 5-HT releasers and
Historically, there has been a tendency to attribute anxio- a selective 5-HT2C receptor agonist (Graeff et al., 1985;
genic actions of the clinical, serotonergic probe and “5-HT2 ” Guimarães et al., 1997; Connell et al., 1998). Further, meter-
agonist, mCPP (Section 4.3.4.1), in animals and man to the goline abrogated the (rapid) relief by SSRIs by premenstrual
engagement of 5-HT2C receptors, whereas the (generally) dysphoric disorder (Roca et al., 2002), while 5-HT2C recep-
anxiolytic actions of the “5-HT2 ” receptor antagonist, ri- tor antagonists may aggravate panic attacks (Den Boer and
tanserin (see below), are considered to reflect their block- Westenberg, 1990; Graeff et al., 1993, 2001). These observa-
ade (Brocco et al., 1990; Kahn and Wetzler, 1991; Anderson tions can be assimilated into a framework whereby, in con-
et al., 2002). Subsequent studies of more selective ligands, trast to anxiogenic actions (for conditioned fear, GAD and
the focus of the following comments, largely support this phobias, etc.) mediated by 5-HT2C receptors in the amyg-
assumption. dala and/or hippocampus (Hodges et al., 1987), 5-HT2C
Anxiogenic actions of mCPP and preferential 5-HT2C re- receptors in the dorsomedial hypothalamus and the dorsal
ceptor agonists, such as Ro60,0175, have generally been re- PAG suppress—or fail to effect—non-conditioned, aversive
ported for tests based on spontaneous behavioural responses. stimuli (Deakin and Graeff, 1991; Graeff et al., 1996b,
In studies where anxiogenic effects were absent, this may 2001; Guimaraes et al., 1997; Zangrossi et al., 2001; Jacob
reflect their confounding actions at other classes of recep- et al., 2002; Blanchard et al., 2003). By analogy, 5-HT2C ag-
tor (notably, anxiolysis-mediating 5-HT2B receptors, Section onists rather than antagonists generally reduce the aversive
4.3.4.3), and their 5-HT2C receptor-mediated sedative and effects of PAG stimulation in rats (Jenck et al., 1989, 1990,
hypophagic properties (Kennett et al., 1989; Gibson et al., 1998; Mora et al., 1997; Graeff, 2002), a model thought
1994; Onaivi et al., 1995; Mora et al., 1997; Martin et al., to be related to panic attacks. Furthermore, in a procedure
1998; Setem et al., 1999; Kennett et al., 2000). Interestingly, of fear-induced ultrasonic vocalizations which has likewise
mCPP displayed mutual cross-generalization to the anxio- been related to panic disorders in man, this response is simi-
genic agent, pentylenetetrazol (which acts as an antagonist larly resistant to 5-HT2C receptor antagonists, yet attenuated
at the picrotoxin-sensitive site of GABAA receptors), in a by agonists (Molewijk et al., 1995b; Dekeyne et al., 2000b;
drug-discrimination procedure (Wallis and Lal, 1998). In ac- Millan et al., 2001a). The above-mentioned inhibitory and
cordance with anxiogenic actions of agonists, the 5-HT2C re- facilitatory influence of 5-HT2C agonists and antagonists,
ceptor antagonists, SB206,553, SB242,084 and SB243,213, respectively, upon corticolimbic noradrenergic transmission
display robust anxiolytic actions in models such as the social may be relevant to these findings, inasmuch as the triggering
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 135

of panic attacks has been imputed to a hyperactivation of ways, the specificity of these effects has been questioned and
noradrenergic pathways (Gorman et al., 1989, 2000; selective 5-HT3 receptor antagonists do not modify extra-
Zacharko et al., 1995; Lovick, 2000; Millan et al., 2000d). cellular level of DA, NA or 5-HT in cortical or sub-cortical
Nevertheless, not all observations are consistent with the structures (Barnes and Sharp, 1999; Millan et al., 2000d;
above scenario and both experimental and clinical evidence Morales and Wang, 2002). The detection of 5-HT3 receptors
that engagement of 5-HT2C receptors favors even uncondi- on a subset of GABAergic neurones in the amygdala, hip-
tioned fear and panic-like states is available (Charney et al., pocampus and FCX (Morales and Bloom, 1997; Jakab and
1987; Benjamin et al., 1999; Kennett et al., 2000; Jones Goldman-Rakic, 2000) provides an anatomical substrate for
et al., 2002a). There is, thus, a need for further study of an enhancement of GABA release in these structures (Ashby
the role(s) of 5-HT2C receptors in the control of panic-like, et al., 1991; Ropert and Guy, 1991; Diez-Ariza et al., 1998;
unconditioned, PAG-integrated, anxious states. Shen and Andrade, 1998; Stutzmann and LeDoux, 1999;
To summarize, despite the robust and clinically-relevant Zhou and Hablitz, 1999; Koyama et al., 2000, 2002). Such
anxiolytic actions of 5-HT2C receptor antagonists in the VCT an influence would be consistent with an anxiolytic role of
and several other models, it appears that 5-HT2C receptors 5-HT3 receptors, though no evidence to support this possibil-
do not fulfil a unitary role in the control of anxious states. ity has been forthcoming. On the other hand, the observation
Their clinical exploitation must be undertaken with this in that stimulation of 5-HT3 receptors augments cortical re-
mind. Definitive resolution of the therapeutic significance of lease of CCK (Paudice and Raiteri, 1991; Raiteri et al., 1993;
5-HT2C receptors to the control of anxiety will only come Morales and Bloom, 1997) is compatible with an anxiogenic
upon eagerly-awaited trials of genuinely-selective 5-HT2C role of 5-HT3 receptors “upstream” of CCK-containing neu-
receptor antagonists (and agonists) (Fitzgerald and Ennis, rones. Contrariwise, the reduction of CCK-induced panic
2002). attacks in human subjects by pre-treatment with 5-HT3 re-
ceptor antagonists (Dépôt et al., 1999) situates 5-HT3 sites
4.3.5. 5-HT3 receptors “downstream” of CCK receptors. The interrelationship of
5-HT3 receptors and CCK in the modulation of anxious
4.3.5.1. Coupling, localization and interactions. 5-HT3 states awaits, thus, further clarification.
receptors may be distinguished from all other classes of
5-HT receptor in that they comprise ligand-gated, pen- 4.3.5.2. Control of anxious states. By analogy to the
tameric ion channels structurally homologous to GABAA low-anxiety shown by mice genetically lacking 5-HT3A
receptors (Section 2.2) but functionally distinct in that they receptor subunits (Kelley et al., 2003), 5-HT3 receptor an-
are primarily permeable to cations. Activation of 5-HT3 tagonists generally exert positive effects in models based on
receptors has been shown to engage phospholipase C and exploratory behaviour, such as the plus-maze and the mouse
thereby elevate intracellular concentrations of Ca2+ , rein- light–dark (two-compartment) box though—bizarrely—
forcing their excitatory influence upon neuronal activity in the latter instance at virtually “homeopathic” doses
(Dubin et al., 1999; Millan, 2002a). Two subunits of 5-HT3 (Malgorzata et al., 1992; Bentley and Barnes, 1995; Costall
receptors have been cloned. In contrast to 5-HT3A sub- and Naylor, 1997; Greenshaw and Siverstone, 1997; Olivier
units, heterologous expression of 5-HT3B subunits alone et al., 2000; Bourin and Hascoët, 2003; Prut and Belzung,
does not yield functional receptors. Rather, they confer 2003). Their actions in paradigms of conditioned fear, in-
distinctive biophysical (current-gating) properties upon cluding the VCT, are less striking, usually being expressed
5-HT3A /5-HT3B heteromers as compared to pure 5-HT3A in a drug-dependent fashion, over a limited dose range and
homomers (Davis et al., 1999; Dubin et al., 1999; Hanna primarily when parameters are adjusted to elicit only a
et al., 2000; Brady et al., 2001). The suspicion that, by modest level of response suppression (Nevins and Anthony,
analogy to structurally-related GABAA (Section 2.2) and 1994; Greenshaw and Siverstone, 1997; Olivier et al.,
nicotinic (Section 6.2) receptors, 5-HT3 receptors may pos- 2000). Indeed, it remains unclear as to why certain 5-HT3
sess modulatory sites was recently afforded a concrete basis receptor antagonists but not others are effective in these
with the identification of an inhibitory, allosteric site re- paradigms (Table 15) (Jones et al., 1988; Piper et al., 1988;
sponsive to cannabinoids (Section 8.1). 5-HT3 receptors are Filip et al., 1992; Malgorzata et al., 1992; Stefański et al.,
expressed in the entorhinal cortex, hippocampus, septum, 1992, 1993b; Artaiz et al., 1995; Greenshaw and Siverstone,
amygdala and hypothalamus (Doucet et al., 2000; Miquel 1997; Olivier et al., 2000; Sanchez, 2003). Studies of con-
et al., 2002). Though mRNA encoding 5-HT3B subunits has flict and other procedures indicate that 5-HT3 sites in the
been detected in human brain, it has been suggested that hippocampus, nucleus accumbens and amygdala (though
CNS-localized (as compared to peripheral) neurones in the not for the VCT) transduce the anxiolytic properties of
rat synthesize exclusively 5-HT3A subunits—conceivably 5-HT3 receptor antagonists (Table 15 and Fig. 8) (Costall
in addition to other, as yet undescribed, subunits (Dubin et al., 1989; Higgins et al., 1991; Stefański et al., 1993b;
et al., 1999; Morales and Wang, 2002). Gargiulo et al., 1996). The modest and idiosyncratic effects
Though initial studies indicated a facilitatory influence of of various 5-HT3 receptor antagonists in the VCT and other
5-HT3 receptors upon dopaminergic and noradrenergic path- experimental models resemble their lack of reproducible
136 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Table 15
Influence of drugs interacting with 5-HT3 receptor ligands upon behaviour in the Vogel Conflict Test
Drug Class Species Route (structure) Effect Active dose range Maximum Reference
(strain) change (%)
Ondansetron 5-HT3 Ant Rat (W) i.p. + 0.1 and 0.3 mg/kg +465 Filip et al. (1992)
Rat (W) i.p. + 1.5 mg/kg +34 Stefanski et al. (1992)
Rat (W) i.p. + 0.05–1.6 +117 Jones (1988)
Rat (W) s.c. IA >0.63 mg/kg IA Brocco (unpublished
observations)
Rat (W) i.c. (hippocampus) + 1 and 2.5 ␮g +69 Stefanski et al. (1993b)
Rat (W) i.c. (N. accumbens) IA 0.01 and 15 ␮g IA Stefanski et al. (1993b)
Rat (LH) i.c. (amygdala) − 0.001–1.0 ␮g NI Higgins et al. (1991)
Granisetron 5-HT3 Ant Rat (W) i.p. + 0.1 mg/kg +92 Filip et al. (1992)
Rat (W) i.p. + 0.1 mg/kg +92 Aertaiz et al. (1995)
Tropisetron 5-HT3 Ant Rat (W) i.p. + 0.01 mg/kg +430 Filip et al. (1992)
Rat (W) i.p. + 0.25 and 0.5 mg/kg +430 Artaiz et al. (1995)
Rat (W) i.p. + 0.001–0.01 mg/kg +36 Stefanski et al. (1992)
Rat (W) i.c. (hippocampus) + 0.005 and 0.01 ␮g +125 Stefanski et al. (1993b)
Rat (W) i.c. (N. accumbens) + 0.01 and 0.25 ␮g +76 Stefanski et al. (1993b)
Rat (LH) i.c. (amygdala) − 0.001–0.1 ␮g NI Higgins et al. (1991)
MDL72,222 5-HT3 Ant Rat (W) i.p. IA 0.01–3.0 mg/kg IA Filip et al. (1992)
DAU6215 5-HT3 Ant Rat (W) s.c. IA 0.01–3.0 mg/kg IA Filip et al. (1992)
VA21B7 5-HT3 Ant Rat (W) i.p. + 1 and 2 mg/kg +85 Artaiz et al. (1995)
Zacopride 5-HT3 Ant Rat (W) i.p. + 0.1 and 1.0 mg/kg +215 Filip et al. (1992)
Abbreviations—N. accumbens: nucleus accumbens and IA: no significant change (inactive). For other abbreviations, see Table 2.

and convincing efficacy in clinical trials of GAD and panic However, their effects are not spectacular and 5-HT4 re-
attacks (see Greenshaw and Siverstone, 1997; Olivier et al., ceptor antagonists were found to be inactive in a conflict
2000). paradigm (Kennett et al., 1997b). Further, it has been re-
ported that 5-HT4 receptor antagonists interfere with the
4.3.6. Other classes of 5-HT receptor actions of BZPs and other drug classes in exploratory mod-
els of anxiolytic activity, suggesting that 5-HT4 receptors
4.3.6.1. 5-HT4 receptors. The presence of 5-HT4 recep- may participate in mechanisms inhibitory to anxious states.
tors (which engage adenylyl cyclase via Gs) in the hip- Correspondingly, it was recently reported that the genetic
pocampus and habenula in high levels, and in the septum and ablation of 5-HT4 receptors is associated with an anxio-
amygdala (of rodents) in more modest densities, is consis- genic phenotype (Cheng et al., 1994; Costall and Naylor,
tent with a role in the modulation of anxious states (Jakeman 1997; Gazzara et al., 2002). On the other hand, 5-HT4 re-
et al., 1994; Waeber et al., 1994; Bonaventure et al., 2000). ceptor antagonists did not modify the anxiolytic effects of
Though their influence upon GABAergic transmission in SSRIs in a model of fear-induced ultrasonic vocalization
subcortical structures does not appear to have been defined, (Schreiber et al., 1998).
5-HT4 receptors exert a complex, bi-modal influence upon In view of this contradictory pattern of data, further ex-
GABAA receptor-mediated currents in frontocortical pyra- ploration of the potential influence of 5-HT4 receptors upon
midal neurones: that is, GABAergic inhibition is accentuated anxious states is necessary before definitive conclusions can
in certain cells and relieved in others. This bi-directional ef- be made concerning their—currently equivocal—role.
fect reflects a protein kinase A-dependent phosphorylation
of ␤1 (reduction) and ␤3 (enhancement) subunits, respec- 4.3.6.2. 5-HT5 receptors. Of the two subtypes of 5-HT5
tively (Cai et al., 2002a; Yan, 2002). receptor, 5-HT5A and 5-HT5B , only the former is present in
5-HT4 receptors appear to exert a facilitatory influence human CNS (Grailhe et al., 2001). Activation of 5-HT5A re-
upon DRN-derived serotonergic neurones. However, the ceptors has been shown to inhibit adenylyl cyclase activity
contribution of direct and indirect mechanisms to this ac- and to promote K+ -currents, though the question of their
tion, and the conditions under which it is expressed, remain principle mode of intracellular coupling is incompletely re-
to be determined (Jakeman et al., 1994; Ge and Barnes, solved (Francken et al., 1998; Graı̈lhe et al., 2001; Millan,
1996; Lucas and Debonnel, 2002). Inhibition of seroton- 2002a). Though they have been largely neglected to date,
ergic transmission might be related to reports that 5-HT4 5-HT5A receptors are of special note in view of their high
receptor antagonists display anxiolytic properties in mod- concentration in the hippocampus, cortex, septum, habenula,
els based on spontaneous exploratory behaviour (Silvestre amygdala and the PAG, as well as the LC and raphe nuclei, of
et al., 1996; Kennett et al., 1997b; Menard and Treit, 1999). both rodent and human brain (Pasqualetti et al., 1998; Oliver
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 137

et al., 2000; Kinsey et al., 2001). Mice lacking 5-HT5A sites data concerning the actions of antisense probes directed
showed no clear alterations in behaviour in several models of against 5-HT7 receptors and of selective antagonists remain
anxiety (Grailhe et al., 1999), but the distinctive anatomical fragmentary and offer little support for such a possibility
organization of 5-HT5 sites in corticolimbic structures sup- (Clemett et al., 1998; Vanhoenacker et al., 2000). Thus, a
ports further exploration with the VCT and other paradigms more thorough evaluation of the potential significance of
of their potential implication in the modulation of anxious 5-HT7 receptors in the control of anxions states is required
states. before concrete conclusions can be reached.

4.3.6.3. 5-HT6 receptors. 5-HT6 receptors, which are 4.4. Antidepressant agents
positively coupled to adenylyl cyclase via Gs (Boess
et al., 1997), are abundant in the hippocampus, nucleus 4.4.1. Clinical actions of antidepressant agents in the
accumbens, amygdala, frontal cortex and entorhinal cortex treatment of anxious states
(Yoshioka et al., 1998; Hamon et al., 1999; Kinsey et al., Evaluation of the influence of antidepressant agents upon
2001; Millan, 2002a; Roberts et al., 2002b). There are sev- anxious states is of considerable interest inasmuch as: (1)
eral indications that their blockade may favor anxious states. genetic variation (polymorphisms) in the human 5-HT trans-
Thus, mice lacking 5-HT6 receptors present an anxiogenic porter (reuptake site) is associated with an altered response
phenotype (Tecott et al., 1998), while antisense probes neu- of the amygdala to fear (Lesch et al., 1996; Hariri et al.,
tralizing 5-HT6 receptors exerted anxiogenic properties in 2002a,b); (2) mice lacking 5-HT transporters show increased
rats (Hamon et al., 1999; Otano et al., 1999). Further, it anxiety (Wichems et al., 2000); (3) a substantial proportion
has been reported that selective 5-HT6 antagonists accel- of depressed patients display co-morbid anxious symptoms
erate hippocampal and frontocortical release of glutamate, (Nutt, 2000; Zimmerman et al., 2000) and (4) antidepressant
NA and/or DA, though contradictory data have been ob- agents are increasingly employed in the long-term treatment
tained (possibly reflecting differences amongst rat strains) of diverse anxious states—despite their tendency to exacer-
(Matsumoto et al., 1999; Dawson et al., 2000, 2001; Frantz bate symptoms upon commencing treatment (Sramek et al.,
et al., 2002; Gobert and Millan, unpublished observation). 2002).
On the other hand, there is some (albeit inconclusive) evi- In the latter respect, both SSRIs (Van Der Linden et al.,
dence for 5-HT6 receptor expression in raphe nuclei (Gérard 2000; Kasper and Resinger, 2001; Millan et al., 2001c,d;
et al., 1996; Kinsey et al., 2001). These findings may be Kent et al., 2002a,b; Liebowitz et al., 2002; Roca et al., 2002;
related to the ability of antisense probes against 5-HT6 Varia and Rauscher, 2002; Wagstaff et al., 2002) and mixed
receptors to attenuate the fear-induced release of 5-HT in 5-HT/NA reuptake inhibitors, such as venlafaxine and du-
FCX (Gérard et al., 1996; Yoshioka et al., 1998), though loxetine (Sasson et al., 1999; Allgulander et al., 2001; Millan
selective 5-HT6 receptor antagonists do not affect basal et al., 2001c,d; Goldstein et al., 2002; Montgomery et al.,
cerebral release of 5-HT (Dawson et al., 2000, 2001; Gobert 2002), have proven clinically effective against a broad range
and Millan, unpublished observations). Recently-described, of anxiety disorders. Indeed, together with BZPs and 5-HT1A
selective antagonists at 5-HT6 receptors (Slassi et al., 2002) agonists (both effective in the VCT), SSRIs represent the
could instructively be employed in the VCT and other pro- cornerstone of therapy for patients with GAD (Vetulani and
cedures for further elucidation of their currently-ambiguous Nalepa, 2000; Gorman, 2002; Sramek et al., 2002). More-
role in the control of anxious states. over, selective NA reuptake inhibitors also express clinical
anxiolytic properties upon sustained administration suggest-
4.3.6.4. 5-HT7 receptors. 5-HT7 sites potentiate the ac- ing that therapeutic efficacy can be achieved independently
tivity of adenylyl cyclase via recruitment of Gs (Hagan of serotonergic mechanisms (Sasson et al., 1999; Millan
et al., 2000). They are well-represented in the hippocam- et al., 2001c,d; Stahl et al., 2002; Versiani et al., 2002). The
pus, cortex, lateral septum, hypothalamus and amygdala clinical utility of such reuptake inhibitors is shared by agents
(Hagan et al., 2000; Kinsey et al., 2001; Neumaier et al., which behave as antagonists at 5-HT2C receptors including:
2001). It has been suggested that 5-HT7 receptors operate (1) mianserin—which also inhibits NA reuptake (Bjertnaes
as autoreceptors on serotonergic perikarya. Though these et al., 1982; Jenck et al., 1994); (2) mirtazapine, a potent
populations are of comparatively minor significance as 5-HT2C receptor (and ␣2 -AR) antagonist devoid of activ-
compared to dendritic 5-HT1A receptors (Section 4.3.2.1) ity at reuptake sites (Sperling and Demling, 1997; Feighner,
and terminal-localized 5-HT1B receptors (Section 4.3.3.1) 1999; Millan et al., 2000c,d) and (3) nefazodone—which
(Roberts et al., 2001, 2002a; Bonaventure et al., 2002), also inhibits 5-HT reuptake (Silva et al., 2001; Ninan et al.,
5-HT7 receptors may exert an indirect inhibitory influence 2002). Interestingly, by analogy to the 5-HT2 receptor antag-
upon DRN-localized serotonergic perikarya via the acti- onists, ritanserin and metergoline (Section 4.3.4), blockade
vation of local GABAergic interneurones (Roberts et al., of 5-HT2C sites was specifically implicated in the ability of
2002a). This configuration is coherent with a role of 5-HT7 nefazodone to blunt conditioned (but not unconditioned) fear
receptors in the modulation of anxiety via an interaction in normal human subjects (Nutt, 1996; Graeff et al., 2001;
with GABAergic and serotonergic mechanisms. However, Silva et al., 2001). Nefazodone has also revealed anxiolytic
138 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

activity in patients with social phobia or with co-morbid down-regulation/uncoupling of postsynaptic, corticolimbic
anxiety and depression and it reduces anxious symptoms populations of 5-HT2C and, possibly, 5-HT1A and 5-HT2A
in individuals undergoing marijuana withdrawal (Zajecka, receptors mediating anxiety (Sections 4.3.2 and 4.3.4)
1996; Van Ameringen et al., 1999; Haney et al., 2003). (Hollander et al., 1991; Kennett et al., 1994; Maj et al.,
1996; Skrebuhhova et al., 1999; Bristow et al., 2000; File
4.4.2. SSRIs and mixed 5-HT/NA reuptake inhibitors: et al., 2000; Berg et al., 2001; Kagaya et al., 2001; Millan
long-term, adaptive actions et al., 2001c,d; Hensler, 2002; Shen et al., 2002; Stout et al.,
In analogy to 5-HT2C receptor agonists (Section 4.3.4.4), 2002; Castro et al., 2003; Van Oekelen et al., 2003).
and reflecting indirect activation of 5-HT2C sites by an in- Other changes may also be implicated in the long-term
crease in extracellular levels of 5-HT, acute application of anxiolytic actions of drugs which suppress 5-HT (and
SSRIs and tricyclic agents moderates the aversive effects of 5-HT/NA) reuptake: a reduction in NPY, CCK, CRF, VP
PAG stimulation in rodents (Jenck et al., 1998; Schenberg and glucocorticoid-mediated mechanisms of anxiety, modu-
et al., 2002), a model of especial pertinence to unconditioned lation of GABA and neurosteroid production (Section 2.2.3),
fear and panic-like states (Section 1.5.3.2). A progressive and changes in the responsiveness of the hypothalamo-corti-
sensitisation of 5-HT2A/2C and 5-HT1A receptors in the cotropic axis (Barden et al., 1995; Montkowski et al., 1995;
PAG was, further, proposed to account for the long-term, Duncan et al., 1998; To et al., 1999; To and Bagdy, 1999;
beneficial actions of antidepressants drugs in unconditioned Makino et al., 2000, 2002a,b; Sanacora et al., 2002; Yau
(panic-like) anxiety states in man (Borsini et al., 2002; et al., 2002).
Graeff, 2002; Jacob et al., 2002; Castro et al., 2003). With one exception in mice (Hascoët et al., 2000), acute
This hypothesis remains, however, to be reconciled with administration of SSRIs or 5-HT/NA reuptake inhibitors
evidence discussed below that prolonged antidepressant has proven to be inffective or to increase anxiety in the VCT
exposure down-regulates levels of 5-HT2A/2C receptors in and other conflict procedures, mirroring above-mentioned
corticolimbic structures. clinical and experimental observations (Table 16 and Fig. 7)
In patients suffering from GAD or phobias, anxiety is (Mason et al., 1987; Fontana and Commissaris, 1988;
initially aggravated by treatment with antidepressant agents Fontana et al., 1989a; Kostowski et al., 1994; Beaufour et al.,
(Section 4.4.1). Correspondingly, acute application of SS- 1999; Borsini et al., 2002). In certain conflict paradigms, it
RIs and of 5-HT/NA reuptake inhibitors elicits anxiogenic has been shown that chronic treatment is accompanied by
effects in the rodent social interaction test. Though activa- the progressive induction of anxiolytic effects (Fontana and
tion of postsynaptic 5-HT1A sites has been incriminated, Commissaris, 1988; Fontana et al., 1989a; Beaufour et al.,
this is debatable (Schreiber et al., 1998; Duxon et al., 2000; 1999; Borsini et al., 2002) and, assuming its appropriate
File et al., 2000; Kõks et al., 2001a; Li et al., 2001b) and the manipulation to detect acute anxiogenic actions of SSRIs
indirect engagement of 5-HT2C receptors (by 5-HT) appears and 5-HT/NA reuptake inhibitors, such studies would be of
to be primarily involved (Bristow et al., 2000; Dekeyne interest to undertake with the VCT.
et al., 2000a; Bagdy et al., 2001; Kõks et al., 2001a;
Allikmets et al., 2002). Upon long-term administration of 4.4.3. Selective NA reuptake inhibitors
SSRIs, their anxiogenic effects wane and there may be a Selective inhibitors of NA reuptake, such as reboxetine,
gradual transition to anxiolytic actions (Bodnoff et al., 1989; mimic the anxiogenic effects of SSRIs upon acute admin-
Griebel et al., 1994; Beaufour et al., 1999; Skrebuhkova istration in the social interaction protocol ((Hascoët et al.,
et al., 1999; Bristow et al., 2000; Duxon et al., 1997a; Silva 1994); Dekeyne and Millan, unpublished observation).
and Brandao, 2000; Li et al., 2001c; Borsini et al., 2002; Though their long-term effects in this procedure remain
Jones et al., 2002b) paralleling the delayed onset of clinical to be evaluated, chronic (as compared to acute) admin-
anxiolytic, panicolytic (and antidepressant) properties in istration induced a gradual anxiolytic effect in a conflict
man. model (Fig. 6) (Beaufour et al., 1999; Lacerra et al., 1999).
This time-dependent institution of anxiolytic properties Interest in NA reuptake inhibitors it underpinned by their
may involve a progressive inactivation of the response of therapeutic efficacy in the control of GAD and other anxiety
raphe-derived serotonergic and/or LC-derived noradrenergic disorders (Section 4.4.1) and by the following observations:
neurones to stress: however, neurochemical studies have not (1) mice genetically deprived of NA transporters reveal
led to a consensus in this regard (Page and Abercrombie, sustained alterations in emotional behaviour (Haller et al.,
1997; Anthony et al., 2000; Szabo et al., 2000; Dazzi 2002a); (2) selective NA reuptake inhibitors are more effec-
et al., 2002; Page and Lucki, 2002). Further, owing to tive than SSRIs in moderating the behavioural symptoms of
down-regulation of inhibitory 5-HT1A and 5-HT1B autore- BZP withdrawal Lacerra et al. (1999) and (3) NA reuptake
ceptors (Sections 4.3.2 and 4.3.3), as well as ␣2 -AR hetero- inhibitors suppress the response of noradrenergic pathways
ceptors (Linner et al., 1999; Mateo et al., 2001), “resting” to stress (Valentino et al., 1990; Durand et al., 2000). It is
monoaminergic transmission is actually enhanced. It is possible that prolonged exposure to NA reuptake inhibitors
more likely, then, that the decisive events occur downstream induces adaptive processes simular to those engendered
of monoaminergic neurones, including a 5-HT-mediated by other antidepressant agents (Secton 4.4.2). However,
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 139

Table 16
Influence of antidepressant agents upon behaviour in the Vogel Conflict Test
Drug Class Species Route Effect Active dose Maximum Reference
(strain) (structure) range (mg/kg) change (%)
Citalopram SSRI Rat (W) s.c. IA 0.63–40.0 IA Brocco (unpublished observations)
Fluoxetine SSRI Rat (W) s.c. IA 0.63–40.0 IA Brocco (unpublished observations)
Maprotiline NARI Rat (SD) s.c. IA 3.0–30.0 IA Mason et al. (1987)
Reboxetine NARI Rat (W) i.p. IA 0.63–10.0 IA Brocco (unpublished observations)
Venlafaxine SNRI Rat (W) s.c. IA >10.0 IA Brocco (unpublished observations)
S33005 SNRI Rat (W) s.c. IA 0.16–10.0 IA Brocco (unpublished observations)
AS-8 SNRI Rat (W) i.p. IA 50–100 IA Kostowski et al. (1994)
Trazodone SSRI/5-HT2 Ant Rat (SD) s.c. + 0.1–10.0 +320 Mason et al. (1987)

Mianserin NARI/5-HT2A/2C Ant Rat (SD) s.c. + 3.0 +90 Mason et al. (1987)
Rat (SD) s.c. + 10.0 +290 Griebel et al. (1997a,b)
Rat (W) s.c. + 2.5 and 10.0 +90 Brocco (unpublished observations)
Mirtazapine 5-HT2A/2C Ant Rat (W) s.c. + 0.63–40.0 +230 Brocco (unpublished observations)
Abbreviations—SSRI: selective serotonin reuptake inhibitor; NARI: noradrenaline reuptake inhibitor; SNRI: serotonin and noradrenaline reuptake inhibitor
and IA: no significant change (inactive). For other abbreviations, see Table 2.

the neuronal mechanisms underlying the long-term, (and 5-HT2A ) receptors appears to more than compensate
therapeutic and experimental anxiolytic properties of NA for its inhibition of 5-HT reuptake (Table 16) (Mason et al.,
reuptake inhibitors remain obscure. 1987; Griebel et al., 1997a, Griebel et al., 1997).

4.4.4. Antidepressant agents possessing 5-HT2A/2C 4.4.5. Monoamine oxidase inhibitors


antagonist properties The mitochondrial enzyme, monoamine oxidase (MAO
In distinction to SSRIs and other agents which interact A), is principally located in catecholaminergic neurones and
with 5-HT transporters, the antidepressants, mianserin and in astrocytes. It displays high affinity for all monoamines
mirtazapine, are devoid of activity at these sites and share po- including 5-HT and catalyses their degradation by oxida-
tent antagonist properties at 5-HT2C receptors (Jenck et al., tive deamination. Mice genetically deprived of MAO A
1994; Sperling and Demling, 1997; Millan et al., 2000c,d). (or B) display elevated levels of 5-HT, NA and DA (or
In each case, they elicit robust and dose-dependent anxi- ␤-phenylethylamine, a trace amine) together with an en-
olytic actions in the VCT (Table 16) (Brocco and Millan, hanced reactivity to stress, though specific and consistent al-
unpublished observations). These observations suggest that terations in anxious behavior have not been reported (Cases
their clinical relief of anxious states might be more rapidly et al., 1995; Grimsby et al., 1997; Kim et al., 1997; Shih
expressed than in the case of SSRIs, though this remains et al., 1999). Inhibitors of monoamine oxidase A are of clini-
to be rigorously demonstrated by well-designed, compara- cal utility in the long-term administration of management of
tive therapeutic trials (Feighner, 1999; Van Hensbeek et al., anxious disorders, such as phobias and panic attacks (Buller,
2000; Quitkin et al., 2001; Borsini et al., 2002; Van Veen 1995; Feighner, 1999; Shih et al., 1999; Schneier, 2001;
et al., 2002). Mianserin and mirtazapine also show pro- Blanchard et al., 2001b; Borsini et al., 2002). Accordingly,
nounced anxiolytic actions in non-conflict paradigms sensi- chronic administration of MAO A inhibitors generally elic-
tive to 5-HT2C antagonists and, in such models, their actions its anxiolytic actions in conflict (and defensive behaviour)
are generally more pronounced upon long-term application paradigms, though no information is currently available for
(Chopin and Briley, 1987; Bodnoff et al., 1989; Benjamin the VCT (Chopin and Briley, 1987; Fontana et al., 1989a;
et al., 1992; Dazzi et al., 2001). The latter finding coincides Paslawski et al., 1996; Griebel et al., 1998a; Maki et al.,
with their ability (upon chronic treatment) to attenuate the re- 2000; Bonnet, 2002; Blanchard et al., 2003). Their pro-
sponse of FCX pools of NA to stress (Dazzi et al., 2002) and longed treatment also initiates adaptive events at pre- and
to down-regulate 5-HT2C and 5-HT2A sites, notably in the postsynaptic monoaminergic receptors (and in the response
amygdala (Zohar et al., 1988; Sanders-Bush, 1990; Rocha of monoaminergic pathways to stress) resembling those
et al., 1994; Anji et al., 2000; Van Oekelen et al., 2003). instigated by other classes of antidepressant agents which
This desensitization of 5-HT2A or 5-HT2C receptors, an ef- act at 5-HT2C receptors and/or 5-HT transporters (Sections
fect common to SSRIs (Sections 4.4.1 and 4.4.2), appears 4.4.2 and 4.4.4) (Twist et al., 1990; Miura et al., 1996;
to be of key significance to the long-term control of anxious Haddjeri et al., 1998; Bonnet, 2002). Such mechanisms are,
states and manifestly does not necessitate their direct acti- presumably, involved in the development of their anxiolytic
vation (Van Oekelen et al., 2003). Inasmuch as trazodone properties. Further, an increase in the activity of cortical
which is structurally related to nefazodone (Section 4.4.1), populations of GABAB receptors has been reported upon
exerts anxiolytic actions in the VCT, its blockade of 5-HT2C sustained exposure to MAO inhibitors (Sands et al., 2003).
140 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Catecholamines are also subject to catabolism by propensity to provoke extrapyramidal side-effects (Corbett
catechol-O-methyltransferase. Male mice genetically de- et al., 1993; Brunello et al., 1995; Millan et al., 1998, 2000a;
prived of this enzyme showed a pronounced augmenta- Millan, 2000). Clozapine reveals anxiolytic properties in
tion in DA levels and enhanced aggression, though they several experimental models, including conflict paradigms
displayed no propensity for alterations in anxious be- (Spealman et al., 1983; Bruhwyler et al., 1990; Wiley et al.,
haviour (Gogos et al., 1998). Contrariwise, notwithstand- 1993; Moore et al., 1994; Szewczak et al., 1995;
ing comparatively modest neurochemical changes, female Siemiatkowski et al., 2001; Manzaneque et al., 2002;
conspecifics manifested an anxiogenic phenotype. This Rademacher et al., 2002; Bourin and Hascoët, 2003), though
distinction pleads for interactions between hormones and it shows only borderline activity in the VCT (Table 17)
catechol-O-methyltransferase in the modulation of anxious (Millan et al., 1999a; Corbin et al., 2000). Amongst its many
behaviour, and anticipates a consideration in Section 12.3 of receptorial interactions, clozapine is a potent antagonist at
marked gender differences in the emotional response to fear. 5-HT2A and 5-HT2C receptors (Millan et al., 1998; Cussac
et al., 2000, 2002b). Their acute blockade and longer-term
4.5. Antipsychotic agents down-regulation likely contribute to its anxiolytic proper-
ties (Kuoppamaki et al., 1995; Burnet et al., 1996; Meltzer,
Inasmuch as anxiety is frequently a co-morbid symptom 1999; Van Oekelen et al., 2003), in analogy to a further,
of schizophrenia, it is important to evaluate the influence of structurally-distinct, multireceptorial antipsychotic, S18327,
antipsychotic agents upon anxious states (Penn et al., 1994; which does exert significant anxiolytic agents in the VCT
Emsley et al., 1999; Braunstein-Bercovitz, 2000; Fenton, (Millan et al., 2000a). 5-HT2A and 5-HT2C receptor antago-
2001). nism also likely accounts for the anxiolytic effects of other
Despite some early clinical reports (Raymond et al., 1957; antipsychotics with marked affinity for these sites, such as
Finnerty et al., 1976) that conventional neuroleptics, such as olanzapine, in the VCT and other conflict models (Table 17)
haloperidol, relieve anxious symptoms in “psychoneurotic (Engel et al., 1989; Moore et al., 1994; Benvenga and
patients”, they are essentially ineffective in alleviating anx- Leander, 1995; Millan et al., 2000a; Rademacher et al.,
ious symptoms of psychotic subjects: indeed, anxiety may 2002). Since sertindole likewise is a potent antagonist at
even be secondarily exacerbated by their extrapyramidal and 5-HT2A and 5-HT2C receptors, any potential activity in
dysphoric side-effects (Meltzer, 1995, 1999; Millan, 2000). the VCT may be compromised by its pronounced sedative
Correspondingly, with few exceptions (Pich and Samanin, properties (Table 17) (Sanchez et al., 1995).
1986), haloperidol has generally preven to be devoid of anx- Clozapine also possesses partial agonist actions at
iolytic properties in experimental models, including conflict 5-HT1A receptors, an action reinforced in several other (po-
procedures such as the VCT: indeed, it may even elicit anx- tential) antipsychotic agents (Bantock et al., 2001; Millan,
iogenic effects (Section 4.2.3.3) (Table 17) (Spealman et al., 2000; Cussac et al., 2002a). Notably, the mixed dopamine
1983; Mansbach et al., 1988; Wiley et al., 1993; Moore D2 /5-HT1A receptor partial agonist, PD158,778, and the
et al., 1994; Szewczak et al., 1995; Ninteman et al., 1996; dopamine D2 /5-HT2A /5-HT2C receptor antagonist/5-HT1A
Bartoszyk, 1998; Millan et al., 1999a; Corbin et al., 2000; receptor agonist, S16924, both of which manifest robust
Siemiatkowski et al., 2001). anxiolytic actions in the VCT (Table 17) (Ninteman et al.,
The “atypical” antipsychotic, clozapine, differs to 1996; Millan et al., 1999a; Cussac et al., 2002a). Indeed,
haloperidol (essentially a dopaminergic D2 /D3 /D4 receptor the broad anxiolytic profile of S16924 across numerous
antagonist) in recognizing multiple classes of monoaminer- procedures can largely be ascribed to its stimulation of
gic receptor, a distinctive profile which underlies its lesser 5-HT1A receptors—though blockade of 5-HT2C sites may

Table 17
Influence of antipsychotic agents upon behaviour in the Vogel Conflict Test
Drug Class Species Route Effect Active dose Maximum Reference
(strain) (structure) range (mg/kg) change (%)
Haloperidol Neuroleptic Rat (W) s.c. IA 0.01–0.16 IA Millan et al. (1999a,b)
Rat (SD) i.p. IA >0.25 IA Corbin et al. (2000)
Clozapine “Atypical” Rat (W) s.c. IA 0.63–5.0 IA Millan et al. (1999a,b)
Rat (SD) i.p. IA >10.0 IA Corbin et al. (2000)
Amperozide 5-HT2 Ant Rat (SD) s.c. + 0.4 +60 Engel et al. (1989)
Sertindole 5-HT2 /␣1 Ant Rat (?) s.c. IA >2.5 IA Sanchez et al. (1995)
S18327 5-HT2 /␣2 Ant Rat (W) s.c. + 2.5 +260 Millan et al. (2000a)
S16924 5-HT1A agonist/5-HT2 Ant Rat (W) s.c. + 0.04 and 0.16 +320 Millan et al. (1999a,b)
PD158,778 5-HT1A PAGO Rat (SD) i.p. + 1.0 and 2.0 +550 Ninteman et al. (1996)
All drugs possess antagonist properties at dopamine D2 receptors. PAGO: partial agonist and IA: no significant change (inactive). For other abbreviations,
see Table 2.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 141

also contribute (Millan et al., 1998, 1999a). An additional amygdala, PAG, septum, nucleus accumbens, cortex and hip-
antipsychotic which acts, albeit less markedly, by engage- pocampus: in addition, they are produced by monoaminer-
ment of 5-HT1A receptors is ziprasidone (Newman-Tancredi gic and GABAergic perikarya. Of particular note is the lo-
et al., 1998; Sprouse et al., 1999; Rollema et al., 2000; calization of inhibitory H3 autoreceptors on histaminergic
Cussac et al., 2002a). Though potential anxiolytic effects neurones (Leurs et al., 1998; Brown et al., 2001; Chazot
of ziprasidone in the VCT remain to be described, it allevi- et al., 2001; Drutel et al., 2001; Pillot et al., 2002; Haas and
ates anxious symptoms in psychotic patients (Schooler and Panula, 2003). Histaminergic pathways exert a direct facil-
Siu, 2000). itatory influence, via H1 and, less markedly, H2 receptors
Anxiolytic actions of antipsychotic agents are an impor- upon the activity of serotonergic and noradrenergic neurones
tant feature of their clinical profiles (Tollefson et al., 1998; originating in the DRN and LC, respectively—though it was
Schooler and Siu, 2000; Adityanjee and Schulz, 2002; recently suggested that, via recruitment of GABAergic in-
Barnett et al., 2002) which can be well-characterized em- terneurones, H1 receptors may indirectly inhibit serotoner-
ploying the VCT. The application of this procedure to gic transmission in mice (Brown et al., 2001, 2002; Son
novel antipsychotic drugs acting as 5-HT1A receptor ago- et al., 2001; Barbara et al., 2002; Millan, 2002a). In rats,
nists and/or as 5-HT2A/2C receptor antagonists—and which in contrast to H1 /H2 sites, histamine H3 receptors exert a
display anxiolytic actions in other models (Barnes et al., suppressive influence upon serotonergic and noradrenergic
1991; Szewczak et al., 1995; Ishida-Tokuda et al., 1996; pathways (Brown et al., 2001; Millan, 2002a).
Sakamoto et al., 1998)—would be of considerable interest.
Finally, by mechanisms which remain to be elucidated— 5.2. Control of anxious states
and in contrast to haloperidol—clozapine increases cortical
levels of allopregnanolone in rodents suggesting that neuros- The foregoing observations provide a neuroanatomical
teroids (Section 2.2.3) may be implicated in its distinctive substrate for several reports, employing models of sponta-
anxiolytic properties (Marx et al., 2003). neous behaviours, of anxiolytic effects of: (1) lesions of
histaminergic pathways; (2) genetic deletion of H1 recep-
tors in mice and (3) administration of H1 receptor antag-
5. Histamine onists (Imaizumi and Onodera, 1993; Frisch et al., 1998;
Privou et al., 1998; Yanai et al., 1998; Hasenöhrl et al., 1999;
5.1. Coupling, localization and interactions Malberg-Aiello et al., 2002; Sanchez, 2003). Contrariwise,
increases in endogenous levels of histamine and the admin-
Histaminergic projections derived from the tuberomam- istration of H1 receptor agonists are accompanied by an en-
millary hypothalamus heavily innervate corticolimbic struc- hancement of anxious behaviour (Imaizumi and Onodera,
tures, including the septum, hippocampus, PAG and amyg- 1993; Yuzurihara et al., 2000; Malberg-Aiello et al., 2002).
dala, as well as monoaminergic cell clusters. Interestingly, A clinical correlate of these findings is provided by con-
there is extensive colocalization of GABA with histamine in trolled, short- and long-term studies demonstrating efficacy
this network of neurones (Panula et al., 1989; Millan, 2002a; and safety of the preferential H1 receptor antagonist, hy-
Trottier et al., 2002; Haas and Panula, 2003). The activity of droxyzine, in the management of GAD. Its effects were ex-
histaminergic pathways subserving corticolimbic structures pressed in the absence of a withdrawal syndrome following
is enhanced in response to fear-evoking and other stressful its abrupt discontinuation. Moreover, this agent was partic-
stimuli (Ito et al., 1999; Brown et al., 2001; Cangioli et al., ularly efficacious against the psychic symptoms of anxious
2002; Westerink et al., 2002). states (Lader and Scotto, 1998; Llorca et al., 2002).
At least three classes of receptor for histamine have been Though the significance of H2 sites is less well-defined,
identified in the CNS. Histamine H1 receptors activate phos- they may fulfil an anxiolytic role opposite to that of H1
pholipase C via Gq and may also stimulate phospholipase receptors (Imaizumi and Onodera, 1993; Yuzurihara et al.,
A2 , whereas H2 receptors couple positively to adenylyl cy- 2000; Malberg-Aiello et al., 2002).
lase via Gs. In contrast to H1 and H2 sites, inhibitory H3 re- In the majority of studies, H3 receptor antagonists have
ceptors are coupled negatively via Gi to adenylyl cylase and revealed anxiogenic actions which may be explained by: (1)
they also suppress Ca2+ -currents (Hill et al., 1997; Hough, an increase in histamine release (which engages postsynap-
2001; Haas and Panula, 2003) (A fourth (H4 ) receptor, which tic H1 sites) and (2) blockade of inhibitory H3 sites on the
is virtually absent from CNS tissue (Hough, 2001; Gantner terminals of LC- and DRN-derived noradrenergic and sero-
et al., 2002), is not considered herein). tonergic projections, thereby accelerating corticolimbic re-
Histamine H1 and H2 receptors share high densities in lease of NA and 5-HT, respectively (Imaizumi and Onodera,
the hippocampus, amygdala, DRN and LC, and H1 sites 1993; Leurs et al., 1998; Pérez-Garcia et al., 1999;
are also prominent in the septum and cortex (Brown et al., Yuzurihara et al., 2000; Malberg-Aiello et al., 2002;
2001; Barbara et al., 2002; Haas and Panula, 2003). His- Westerink et al., 2002).
tamine H3 receptors display a more restricted distribution. In light of this intriguing pattern of data, further exami-
However, recent work has demonstrated H3 receptors in the nation of the influence of histaminergic mechanisms upon
142 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

anxious states could profitably be undertaken with the VCT distributed in corticolimbic structures, notably in the FCX,
and other models. hippocampus (wherein ␣7 sites are dominant) and septum.
Their co-existence as excitatory receptors on serotonergic
and noradrenergic cell bodies and, in the latter case, termi-
6. Acetylcholine nals deserves emphasis in the present context (Ribeiro et al.,
1993; Clark and Reuben, 1996; Wonnacott, 1997; Picciotto,
6.1. Cholinergic pathways and their modulation 1999; Bitner et al., 2000; Paterson and Nordberg, 2000;
Bitner and Nikkel, 2002; Pradhan et al., 2002; Seth et al.,
Cholinergic pathways, which fulfil a pivotal role in the 2002; Woo et al., 2002; Yasuda et al., 2002; Vincler and
control of arousal, motivation and cognitive-attentional func- Eisenach, 2003).
tion, ramify extensively in the hippocampus, amygdala, sep- In monoaminergic cell clusters and in limbic regions, ex-
tum, frontal cortex and several other limbic regions: therein, citatory ␣7 nicotinic receptors are both localized on, and
they reciprocally interact with monoaminergic, GABAergic postsynaptic to, GABAergic neurones (Yang et al., 1996;
and glutamatergic pathways and modulate the activity of the Frazier et al., 1998; Lu et al., 1998; Alkondon et al., 2000;
hypothalamo-corticotropic axis (Wonnacott, 1997; Bugajski Dani, 2001; Bitner and Nikkel, 2002; Kawai et al., 2002;
et al., 1998; Picciotto, 1999; Paterson and Nordberg, 2000; Mihailescu et al., 2002; Vincler and Eisenach, 2003). Fur-
Li et al., 2001b; Araki et al., 2002; Hajszan and Zabroszky, ther, there is considerable evidence for an excitatory in-
2002; Seth et al., 2002; Gobert et al., 2003). fluence of ␣7 and other classes of nicotinic receptor upon
The activity of frontocortical, hippocampal and lateral mesocortical and mesolimbic dopaminergic neurones. Their
septal cholinergic pathways is enhanced in response to anx- actions are mediated both directly and via multisynaptic
iogenic and stressful stimuli, and it has been proposed that circuits involving other transmitters (Pontieri et al., 1996;
an overactivity of cholinergic input to the FCX may con- Wonnacott, 1997; Schilström et al., 1998, 2003; Cohen et al.,
tribute to anxious states while septal cholinergic mecha- 2002; George et al., 2001). Interestingly, in contrast to the ef-
nisms participate in the pressor response to stress (Miczek fects of acute stimulation of nicotinic receptors, their chronic
and Lau, 1975; Gilad, 1987; Acquas et al., 1996; Berntson activation (presumably reflecting desensitisation) attenuates
et al., 1998; Ceccarelli et al., 1999; Hart et al., 1999; Sarter the induction of DA release in the FCX and nucleus accum-
and Bruno, 1999; Giovannini et al., 2001; Ichikawa et al., bens by conditioned fear (Schilström et al., 1998; George
2002; Kubo et al., 2003). Further, under conditions of stress, et al., 2000, 2001).
this anxiogenic role of ACh may involve the generation of
a distinct isoform of acetylcholinesterase (AChE-R, where 6.2.2. Influence upon anxious states
“R” stands for “read through”) which accumulates intracel- Zicotine elicits a bewilderingly complex pattern of anx-
lularly to interact with protein kinase C␤ (isoform II): this iogenic and anxiolytic effects reflecting: diverse sites of
kinase is concentrated in the amygdala, hippocampus and action in various corticolimbic structures (File et al., 1998;
cortex and is implicated in the induction of anxious states Kenny et al., 2000); interactions with glutamatergic (Perez
(Weeber et al., 2000; Birikh et al., 2003). De La Mora et al., 1991; Wonnacott, 1997), monoaminergic
However, cholinergic pathways do not fulfil a uniform (Section 6.2.1) and GABAergic (Section 6.2.1) pathways;
role in the control of anxious states (see below) and, in op- modulation of corticolimbic levels of various neurosteroids;
posite fashion, an overall enhancement of cholinergic trans- the differential involvement of ␣4 ␤2 versus ␣7 receptor sub-
mission in the hippocampus (elicited by local injection of types (Araki et al., 2002; Seth et al., 2002); procedural as-
acetylcholinesterase inhibitors) was accompanied by anx- pects, such as the time of testing, drug dose and duration of
iolytic properties (File et al., 1998; Degroot et al., 2001; administration; “basal” levels of anxiety; housing conditions
Degroot and Treit, 2002). and the experimental model employed (Vale and Green,
1986; Picciotto, 1999; Paterson and Nordberg, 2000; Cheeta
6.2. Nicotinic receptors et al., 2001a; Dani, 2001; George et al., 2001; Szyndler
et al., 2001; Araki et al., 2002; File et al., 2002 and Seth
6.2.1. Localization and interactions et al., 2002; Porcu et al., 2003). Both anxiogenic and anx-
Neuronal nicotinic receptors consist of ligand-gated, pen- iolytic properties of nicotine have likewise been detected in
tameric, cation-permeable channels assembled from ␣ and human subjects (Parrott, 1995; Netter et al., 1998; Picciotto,
␤-subunits, of which, respectively, nine and three isoforms 1999; Paterson and Nordberg, 2000). Not surprisingly,
exist in the CNS. The most prevalent central isoforms are recruitment of GABAergic neurones has been strongly im-
the ␣4 ␤2 heteromer and (less prominently) the ␣7 homo- plicated in the anxiolytic properties of nicotinic agonists,
mer, the latter of which is highly permeable to Ca2+ and whereas serotonergic mechanisms (in particular, involving
displays distinctively rapid kinetics (Picciotto, 1999; Lloyd postsynaptic 5-HT1A receptors, Section 4.3.2) are involved
and Williams, 2000; Paterson and Nordberg, 2000; Dani, in their anxiolytic and anxiogenic actions (op. cit.). Studies
2001; Araki et al., 2002; Lax et al., 2002; Mihailescu et al., of knock-out mice indicate a specific role for ␣7 recep-
2002). ␣4 , ␤2 , ␣7 and other nicotinic subunits are broadly tors in mediating the anxiolytic actions of nicotine, and of
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 143

their ␣4 ␤2 counterparts in transducing both anxiogenic and drenergic pathways may favor anxious (panic-like) states
anxiolytic effects, though their precise significance remains (Pudovkina et al., 2002). On the other hand, under certain
rather nebulous (Paylor et al., 1998; Picciotto, 1999; Ross conditions, muscarinic mechanisms counter anxiety via ac-
et al., 2000; Cheeta et al., 2001a; Labarca et al., 2001; tions in the hippocampus, possibly reflecting the modulation
Salas et al., 2002). Reconciliation of competing claims of of GABAergic transmission (Rodgers and Cole, 1995; File
anxiogenic versus anxiolytic roles for nicotinic receptors et al., 1998, 2000; Smythe et al., 1998; Wu et al., 2000). Fi-
will likely be aided not only by studies of mice genetically nally, muscarinic (M1 /M3 ) sites in the lateral septum have
deprived of individual nicotinic receptor subunits, but also been implicated in the pressor response to stress (Kubo et al.,
by use of pharmacological agents discriminating specific 2003).
cerebral isoforms. All of the above observations were derived from proce-
It is important to note that the above-described patterns of dures, such as the plus-maze, which employ exploratory be-
data were exclusively acquired in models employing natural haviours, and studies of the role of muscarinic mechanisms
behaviours, such as the social interaction and plus-maze tests in the control of behaviour in the VCT and other conflict
(Brioni et al., 1994; Ouagazzal et al., 1999; Picciotto, 1999; models remain to be undertaken. Moreover, though mice
Cheeta et al., 2000, 2001a; Kenny et al., 2000; Irvine et al., lacking specific muscarinic receptor subtypes have been gen-
2001). Indeed, with the exception of one (very dated) report erated, and M2 receptors demonstrated to facilitate corticos-
(Morrison, 1969), studies of the influence of nicotinic mech- terone secretion, there appear to be no reports specifically
anisms upon anxious states employing conflict paradigms focussing on their response to stress or anxious behaviour
are conspicuous by their absence. For an improved apprecia- (Gomeza et al., 1999; Hemrick-Lubecke et al., 2002).
tion of the broad pathophysiological and clinical pertinence
of nicotinic mechanisms in the control of anxious states it
would, then, appear imperative to undertake studies with 7. Adenosine
conflict procedures.
Though tricyclic antidepressants, fluoxetine and several 7.1. Generation, localization and coupling to multiple
other SSRIs behave as reversible, non-competitive inhibitors receptor subtypes
of nicotinic receptors, there is currently no evidence that
this mechanism participates in their control of anxious states Adenosine can, in principle, be produced throughout the
(Section 4.4) (Paterson and Nordberg, 2000). CNS by both neuronal and non-neuronal elements, and its
formation is interlinked with the generation of adenosine
6.3. Muscarinic receptors triphosphate and energy balance. As outlined elsewhere
(Latini and Pedata, 2001; Millan, 2002a), adenosine is
Of the five classes of muscarinic receptor known, three formed in neurones primarily by hydrolysis (catalysed by
(M1 , M3 and M5 ) are primarily coupled to phospholipase 5-nucleotidase) of 5-adenosine monophosphate. It is subse-
C via Gq underlying an excitatory influence upon neuronal quently released by a bi-directional, Na+ -dependent nucle-
activity, whereas two (M2 and M4 ) are inhibitory, reflect- oside carrier. In addition to this pool of secreted adenosine,
ing their Gi -mediated inhibition of adenylyl cyclase and it can be directly produced extracellularly from various
an enhancement of K+ -currents (McKinney, 1993; Wess, sources, of which its cleavage from adenosine monophos-
1996; Porter et al., 2002). Muscarinic M2 receptors are lo- phate via an ecto-5-nucleotidase is the most familiar mech-
calized presynaptically as inhibitory autoreceptors on the anism. Neuronal (and glial) transporters capture adenosine
terminals of corticolimbic cholinergic neurones in the cor- from the extracellular space and metabolise it (by adenosine
tex, hippocampus and other structures (Kitaichi et al., 1999; deaminase) into inosine which is eventually re-transformed
Ichikawa et al., 2000, 2002; Zhang et al., 2002). into adenosine triphosphate. Studies of the cerebral dis-
Little is known concerning the potential roles of multi- tribution of adenosine, adenosine deaminase and other
ple classes of muscarinic receptor in the response to, and markers have revealed maximal concentrations in the cor-
the control of, anxious states, despite their (M1 –M4 ) broad tex (frontal and parietal), lateral septum and hippocampus,
and differential occurrence in corticolimbic structures such as well as high levels in the amygdala, nucleus accumbens
as the FCX, hippocampus and amygdala, wherein M1 sites and PAG (Braas et al., 1986; Nagy et al., 1990). Further,
are especially prominent (Brann et al., 1988; Buckley et al., adenosine fulfils a co-transmitter role with NA—and pos-
1988; Levey et al., 1991; Gomeza et al., 1999; Piggott et al., sibly GABA—in the hippocampus and other supraspinal
2002). It has been proposed that postsynaptic muscarinic structures (Manzoni et al., 1994; Poelchen et al., 2001).
(M1 ) sites in the infralimbic region of the cortex partici- Adenosine exerts its actions via A1 , A2A , A2B and A3
pate in the induction of anxiety by cholinergic pathways receptors (Olah and Stiles, 1985, 1995; Impagnatiello et al.,
(Wall et al., 2001; Wall and Messier, 2002), and muscarinic 2000; Fredholm et al., 2001; Millan, 2002a). Although
mechanisms in the amygdala may facilitate (unconditioned) A2B receptor subtypes are found in the hippocampus at
fear (Power and McGaugh, 2002). Further, the excitatory modest levels, they are not specifically implicated in the
influence of muscarinic receptors upon LC-derived nora- control of anxious states, and the apparent presence of A3
144 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

sites in supraspinal structures awaits further verification result at variance with human experience (Table 18) (Beer
(Impagnatiello et al., 2000; Rivkees et al., 2000). Thus, the et al., 1972; Baldwin et al., 1989; Fredholm et al., 1999).
ensuing discussion focuses on A1 and A2A receptors. This puzzling incoherence between the VCT and the clini-
cal effects of A1 receptor blockade may reflect one or more
7.2. Adenosine A1 receptors of the following factors: (1) a lack of sensitivity of the
VCT to the relatively weak anxiogenic properties of caf-
Adenosine A1 sites are enriched in the hippocampus, sep- feine; (2) the fact that high doses of caffeine are needed to
tum and cortex (Rivkees et al., 1995; Swanson et al., 1995; elicit anxiety in man as compared to the relatively modest
Dixon et al., 1996; Ralevic and Burnstock, 1998; Ochiishi doses employed experimentally; (3) a role of A2A receptor
et al., 1999). They interact via Gi/o with adenylyl cyclase blockade (see below) in the anxiogenic actions of caffeine;
(inhibition), K+ -currents (potentiation) and VDCCs (reduc- (4) induction by caffeine in man of an (un-conditioned)
tion of Ca2+ -currents), actions accounting for the inhibitory anxious state qualitatively different to that reproduced by
influence of adenosine upon corticolimbic glutamatergic, the VCT in rodents and (5) actions of caffeine expressed
dopaminergic, serotonergic and other modes of transmission independently of adenosine receptors (Foukas et al., 2002).
(Impagnatiello et al., 2000; Johansson et al., 2001; Nikbakht Additional studies with more potent and selective A1 re-
and Stone, 2001; Okada et al., 2001; Abrams et al., 2002; ceptor antagonists would be desirable to resolve this issue
Masino et al., 2002; Solinas et al., 2002). Mice lacking A1 since the lack of anxiogenic properties of caffeine in the
receptors display enhanced anxiety (Johansson et al., 2001; VCT represents an apparently major discrepancy between
Giménez-Llort et al., 2002) and the anxiogenic actions of this model and results in human subjects.
adenosine antagonists, such as caffeine, in animals and man
have generally been attributed to blockade of A1 sites (Uhde 7.3. Adenosine A2A receptors
et al., 1984a; Loke et al., 1985; File et al., 1988; Baldwin
and File, 1989; McCloskey et al., 1990; Nickell and Uhde, Counterbalancing the neuronal actions of inhibitory A1
1994; Jain et al., 1995; Fredholm et al., 1999). Indeed, sites, excitatory A2A receptors are positively coupled via Gs
studies employing models based on exploratory behaviour to adenylyl cyclase and (via protein kinase A) to VDCCs
indicate that selective stimulation and antagonism of A1 re- (Ralevic and Burnstock, 1998; Impagnatiello et al., 2000).
ceptors is associated with anxiolytic and anxiogenic actions, Adenosine A2A receptors are concentrated in the basal gan-
respectively (Baldwin and File, 1989; Jain et al., 1995; Florio glia and the nucleus accumbens. Though they have also
et al., 1998). been detected in the LC, DRN, hippocampus, cortex and
However, conflict paradigms in rodents and primates amygdala (Olah and Stiles, 1995; Dixon et al., 1996; Rosin
have not invariably yielded evidence in favour of a specific et al., 1998; Impagnatiello et al., 2000; Fredholm et al.,
anxiolytic role of A1 sites (Coffin and Spealman, 1985; 2001; Phillis, 2001; DeMet and Chicz-DeMet, 2002), and
Commissaris et al., 1990; Haraguchi and Kuribara, 1991; appear to share the localization of A1 sites on serotoniner-
Thiébot et al., 1991). Moreover, caffeine was reported to gic, glutamatergic and GABAergic neurones (Impagnatiello
either act anxiolytically, or to be ineffective, in the VCT, a et al., 2000; Nikbakht and Stone, 2001), the quantitative

Table 18
Influence of adenosine receptor ligands, of melatonin and testosterone, of NO synthase inhibitors, of calcium channel blockers, and of sigma ligands
upon behaviour in the Vogel Conflict Test
Drug Class Species Route Effect Active dose Maximum Reference
(strain) (structure) range (mg/kg) change (%)
Theophylline A1/2 Ant Rat (?) i.p. + 75 and 100 + 400 Beer et al. (1972)
Caffeine A1/2 Ant Rat (?) i.p. + 25–100 + 800 Beer et al. (1972)
Rat (LH) i.p. IA 40.0 IA Baldwin et al. (1989)
Dibutyryl cyclic AMP A1/2 Ant Rat (?) i.p. + 100–400 +450 Beer et al. (1972)
Melatonin Melatonin agonist Rat (W) s.c. IA 0.63–80.0 IA Millan et al. (2002)
Testosterone Androgen Rat (W) s.c. + 5 + 102 Bing et al. (1998)
l-NAME NO (synth. inhib.) Rat (?) i.v. + 54.8 and 100 NI Dunn et al. (1995)
l-NOARG NO (synth. inhib.) Rat (?) i.v. + 30 NI Dunn et al. (1995)
Flunarizine Ca2+ channel Ant Rat (W) i.p. + 10 and 20 +70 Matsumoto et al. (1994)
Verapamil Ca2+ channel Ant Rat (W) i.p. + 20 +95 Matsumoto et al. (1994)
Nicardipine Ca2+ channel Ant Rat (W) i.p. + 20 +90 Matsumoto et al. (1994)
PD144418 Sigma1 “Ago” Rat (W) i.p. IA 5 and 15 IA Akunno et al. (1997)
Cyclazocine Sigma1 “Ago” Mouse (ICR) s.c. IA 0.1–3 IA Umezi (1999)
Siramesine Sigma2 “Ago” Rat (W) i.p. + 10 and 22 +65 Sanchez et al. (1997)
Abbreviations—l-NAME: N(G)-nitro-l-arginine methyl ester; l-NOARG: N(G)-nitro-l-arginine; A: adenosine; NO (synth. inhib.): nitric oxide synthase
inhibitor; IA: no significant change (inactive) and NI: not indicated. For other abbreviations, see Table 2.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 145

distribution, precise identity and functional status of Currently, it appears that cannabinoids exert their cere-
extra-striatal A2A sites remains controversial. Irrespective of bral actions solely via CB1 receptors, though the poten-
this uncertainty, there is evidence that A2A receptors exert a tial existence of additional CB sites remains under discus-
facilitatory influence upon GABA release in the septum and sion, not least in view of the preservation of several func-
hippocampus, actions which may be related to the observa- tional actions of “selective” CB1 receptor antagonists in
tion that mice genetically lacking A2A receptors show an mice lacking these sites (Pertwee, 1997, 2000; DiMarzo
exaggerated response to anxiogenic stimuli. However, A2A et al., 2000; Breivogel et al., 2001; Haller et al., 2002b;
receptors are also excitatory to 5-HT and glutamate release Millan, 2002a). Furthermore, actions of anandamide at sites
in the hippocampus (Popoli et al., 1995; Impagnatiello et al., other than CB1 receptors, such as vanilloid receptors, VD-
2000; Okada et al., 2001) and no consistent evidence for CCs (T-type) and muscarinic receptors, should be borne in
anxiolytic effects of A2A receptor stimulation has, to date, mind (Chemin et al., 2001; Millan, 2002a) and it was re-
been obtained (Baldwin and File, 1989; Griebel et al., 1991; cently reported that anandamide non-competitively reduces
Jain et al., 1995; Ledent et al., 1997). Unfortunately, infor- the operation (cation-conductance) of 5-HT3 receptors by
mation from the VCT concerning the potential significance actions at a novel, allosteric site refractory to selective CB1
of A2A sites in the control of anxious states is unavailable. receptor antagonists (Section 4.3.5) (Barann et al., 2002;
In light of the above comments, there is a need for a Molderings-Godlewski et al., 2002). Though CB1 receptors
thorough and interdisciplinary reappraisal of the role(s) of can engage various intracellular mediators, their inhibition
adenosine in the modulation of anxious states employing of adenylyl cyclase via Gi/o and their suppression and po-
ligands highly-selective for specific adenosine receptor sub- tentiation of Ca2+ -currents and K+ -currents, respectively,
types, as well as agents modulating its synthesis and degra- provide a substrate for their generally inhibitory influence
dation (Impagnatiello et al., 2000; Millan, 2002a). upon neuronal excitability (Pertwee, 1997).

7.4. Purinoceptors 8.2. Influence upon anxious states

As mentioned above (Section 7.1), one source of neu- Though comparatively little information is available, the
ronal pools of adenosine is adenosine triphosphate which following observations support a role of CB1 receptors in
itself acts via ionotropic P2X and metabotropic P2Y recep- the modulation of anxious states.
tors (Ralevic and Burnstock, 1998). Though their influence First, cannabinoids are produced throughout the brain and
upon anxious states remains to be defined, in light of their CB1 receptors are particularly well-represented in the cor-
occurrence in corticolimbic regions and their modulatory in- tex (entorhinal and cingulate), hippocampus, lateral septum,
fluence upon monoaminergic pathways—for example, P2X nucleus accumbens, amygdala and PAG (Tsou et al., 1998;
sites are excitatory to noradrenergic perikarya—this issue Ameri, 1999; Katona et al., 1999; Davis et al., 2002).
would be of interest to explore (Ralevic and Burnstock, Second, cannabinoids modulate the release of several
1998; Sansum et al., 1998; Williams and Jarvis, 2000). transmitters implicated in the control of anxious states.
They suppress the outflow of glutamate in the hippocampus
and PAG—possibly reflecting a retrograde action at presy-
8. Cannabinoids naptic terminals—though glutamate release in the FCX is
(indirectly) enhanced by cannabinoids (Shen et al., 1996;
8.1. Generation and coupling to CB1 receptors Robe et al., 2001; Stella and Piomelli, 2001; Tomasini
et al., 2002). Cannabinoids are inhibitory to corticolimbic
Cannabinoids comprise a family of lipids, including release of NA and DA (Schlicker et al., 1997; Schlicker and
anandamide and 2-arachidonylglycerol, which are phasi- Kathman, 2001), 5-HT (Nakazi et al., 2000; Schlicker and
cally generated from phospholipids in an activity-dependent Kathman, 2001; Hermann et al., 2002) and the anxiogenic
fashion (Piomelli et al., 2000; Giuffrida et al., 2001; neuropeptides, CCK and CRF (Ameri, 1999; Rodriguez
Stella and Piomelli, 2001; Millan, 2002a). Following their de Fonseca et al., 1997; Schlicker and Kathman, 2001).
(non-vesicular) release, they are rapidly taken up into neu- On the other hand, by presynaptic mechanisms, they in-
rones and glia and catabolized by the microsomal enzyme, terfere with GABAergic transmission in the amygdala,
fatty acid amide hydrolase (amidohydrolase) and by the ser- hippocampus, FCX and other regions (Katona et al., 1999,
ine hydrolase, monogluceride lipase (Piomelli et al., 2000; 2001; Hajos et al., 2000a; Hoffman and Lupica, 2000;
Giuffrida et al., 2001; Stella and Piomelli, 2001; Bracey Ferraro et al., 2001; Schlicker and Kathman, 2001; Wilson
et al., 2002; Dinh et al., 2002; Millan, 2002a). Interestingly, et al., 2001; Marsicano et al., 2002; Millan, 2002a; Pistis
cannabinoids exert actions both anterogradely (conven- et al., 2002). This interruption of GABAergic activity
tional transmission) and, in analogy to NO (Section 14) may underlie their indirect disinhibition of cortical glu-
and neurotrophins (Section 15), retrogradely at presynaptic tamatergic and dopaminergic transmission pathways in
terminals (Ohno-Shosaku et al., 2001; Tao and Poo, 2001; the FCX and nucleus accumbens (Navarro et al., 1993;
Alger, 2002; Millan, 2002a). French, 1997; Ameri, 1999; Schlicker and Kathman, 2001;
146 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Cohen et al., 2002; Pistis et al., 2002; Tomasini et al., in sulphated and non-sulphated forms of various length—
2002). exerts its actions via excitatory CCK1 (CCKA ) and CCK2
Third, probably reflecting this complex pattern of influ- (CCKB ) receptors, both of which activate phospholipase C
ence upon neurotransmitters known to divergently modu- via Gq (Woodruff et al., 1991; Noble and Roques, 1999).
late anxious states, cannabinoids elicit dose-dependent and Together with CCK itself, CCK2 receptors are widespread
environment-dependent anxiolytic and anxiogenic effects in throughout the limbic system (including the amygdala and
rodent models incorporating exploratory behaviour (Onaivi hippocampus) and cortex. While the distribution of CCK1
et al., 1990; Rodriguez de Fonseca et al., 1996, 1997; sites is more restricted, their presence in the hypothalamus,
Navarro et al., 1997; Martin et al., 2002b). In line with DRN, hippocampus and septum is of note in the present con-
this broad spectrum of actions in experimental models, as text (Rehfeld and Nielsen, 1995; Van Megen et al., 1996;
a function of personality and the precise circumstances of Noble and Roques, 1999; Adell et al., 2002). These neu-
testing, cannabinoids have been found to either relieve or roanatomical observations, the responsiveness of cortical
to induce anxiety in man (Ashton, 2001; Porter and Felder, and amygdala pools of CCK2 receptors and CCK to anxiety
2001; Robson, 2001). and other stressful stimuli in rats (Harro et al., 1990; Del Bel
Fourth, notwithstanding this diverse pattern of experi- and Guimaraes, 1997; Becker et al., 1999, 2001), and func-
mental and clinical observations with cannabinoid agonists, tional interactions amongst CCK-containing, monoaminer-
CB1 receptor antagonists generally provoke an increase in gic and GABAergic neurones (Boden et al., 1991; Woodruff,
anxiety in rodents (Navarro et al., 1997; Arevalo et al., 2001; 1992; Bickerdike et al., 1995; Ferraro et al., 1999; Noble
Maccarone et al., 2002; Martin et al., 2002b). Correspond- and Roques, 1999; Hamilton et al., 2001; Kõks et al., 2001b;
ingly, mice genetically lacking CB1 receptors reveal an anx- Tanganelli et al., 2001; Adell et al., 2002; Siniscalchi et al.,
iogenic profile (Haller et al., 2002b): however, in this study, 2003), collectively support a role of CCK in the modula-
anxiolytic properties of CB1 receptor antagonists were, sur- tion of anxious states (Ravard and Dourish, 1990; Bradwejn,
prisingly, observed. Moreover, their persistence in the CB1 1993; Wettstein et al., 1994; Van Megen et al., 1996; Bourin
knock-out population was interpreted as support for the no- et al., 1998; Daugé and Léna, 1998; Kõks et al., 2000;
tion that an “anxiogenic” CB receptor exists (Haller et al., Bradwejn and Koszycki, 2001).
2002b). This hypothesis awaits verification and an alter-
native, more conservative, interpretation would be that the 9.1.1.2. Role of CCK2 sites in the control of anxious states.
antagonists in question recognize other non-cannabinoid CCK2 sites in several structures, including the nucleus
mechanisms mediating anxiety. tractus solitarius, amygdala, nucleus accumbens, FCX,
In addition to mice deficient in CB1 receptors, popula- hippocampus and DRN, have been implicated in the anx-
tions lacking fatty acid amide hydrolase have been generated iogenic properties of CCK (Frankland et al., 1997; Noble
(Navarro et al., 1997; Cravatt et al., 2001; Maccarone et al., and Roques, 1999; Becker et al., 1999, 2001; Kõks et al.,
2002; Martin et al., 2002b). Moreover, chemical inhibitors of 2000; Adell et al., 2002; Wunderlich et al., 2002). Ac-
this cannabinoid-degrading enzyme, modulators of cannabi- cordingly, the anxiogenic (panicogenic) actions of CCK4
noid transport, and selective agonists and antagonists at CB1 and the synthetic pentapeptide analogue, pentagastrin, in
receptors are all available for exploitation in the VCT and animals and man are abolished by selective CCK2 recep-
other models (Pertwee, 1997, 2000; Piomelli et al., 2000; tor antagonists (Singh et al., 1991; Bradwejn, 1993; Lines
Giuffrida et al., 2001; Millan, 2002a). There is increasing et al., 1995; Bourin et al., 1998; Kõks et al., 2000; Geraci
interest in the therapeutic use of cannabinoidergic agents for et al., 2002). Such experimental studies, and complementary
the management of pain, schizophrenia and drug abuse, dis- work demonstrating intrinsic anxiolytic actions of CCK2
orders which display substantial comorbidity with anxiety antagonists, have principally been successful upon employ-
states (Pertwee, 1997, 2000; Porter and Felder, 2001; Millan, ing models of exploratory behaviour, notably the elevated
2002a). A clarification of the poorly-understood role of plus-maze (Costall et al., 1991; Singh et al., 1991; Frankland
cannabinoidergic mechanisms in the modulation of anxious et al., 1997; Tsutsumi et al., 1999; Wunderlich et al., 2002;
states would, consequently, appear to be of some urgency. Blanchard et al., 2003). Contrariwise, anxiolytic actions of
CCK2 (and CCK1 ) antagonists in conflict paradigms are
weak, inconsistent and rarely dose-dependent (Hendrie and
9. Neuropeptides Dourish, 1990; Powell and Barrett, 1991; Singh et al., 1991;
Charrier et al., 1995; Dawson et al., 1995; Van Megen et al.,
9.1. Neuropeptides/receptors previously examined 1996; Griebel et al., 1997a,b). In line with these ambivalent
by use of the VCT findings, in a study of the VCT, selective antagonists at
either CCK1 or CCK2 receptors were ineffective (Table 19)
9.1.1. Cholecystokinin (Griebel et al., 1997b). Moreover, mice lacking CCK2 (or
CCK1 ) receptors show at most modest alterations in anxious
9.1.1.1. Localization and coupling to CCK1 and CCK2 sites. behaviour, and may even reveal an anxiogenic phenotype
Like the closely-related peptide, gastrin, CCK—which exists (Kobayashi et al., 1996; Nomoto et al., 1999; Chambers
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 147

Table 19
Influence of neuropeptide receptor ligands upon behaviour in the Vogel Conflict Test
Drug Class Species Route Effect Active dose Maximum Reference
(strain) (structure) range change (%)
Lorglumide CCK1 Ant Rat (SD) i.p. IA 0.3–1.0 mg/kg IA Griebel et al. (1997b)
PD135,158 CCK2 Ant Rat (SD) i.p. IA 0.01–1.0 mg/kg IA Griebel et al. (1997b)
LY288,513 CCK2 Ant Rat (SD) i.p. IA 0.3–10.0 mg/kg IA Griebel et al. (1997b)
CRF CRF1(2) agonist Rat (W) i.c.v. − 10 ␮g −40 Britton et al. (1985)
CRF9–41 CRF Ant Rat (?) i.c. (lateral + 0.25–0.5 ␮g NI Thomas et al. (2002)
septum)
CP154,526 CRF1 Ant Rat (W) i.p. + 80 mg/kg +400 Millan et al. (2001a)
Rat (W) i.p. IA 0.62–20 mg/kg IA Griebel et al. (1998a,b)
DMP695 CRF1 Ant Rat (W) i.p. + 40 mg/kg +300 Millan et al. (2001a)
SR125,543A CRF1 Ant Rat (SD) i.p. + 20 and 30 mg/kg +188 Griebel et al. (2002b)
SR149,415 VP1B Ant Rat (W) i.p. + 3 and 10 mg/kg +148 Serradiel-Gal et al. (2002)
GR205,171 NK1 Ant Rat (W) i.p. + 40 mg/kg +131 Brocco (unpublished
observations)
Neuropeptide Y Y1 agonist Rat (SD) i.c.v. + 0.2–5.0 nmol +320 Heilig et al. (1989)
Neuropeptide Y 13-36 Y2 agonist Rat (SD) i.c.v. IA 0.4 and 2.0 nmol IA Heilig et al. (1989)
Glucagon-like peptide Agonist Rat (SD) i.c.v. − 10 ␮g −74 Moller et al. (2002)
Galanin Agonist Rat (SD) i.c.v. + 3 nmol +95 Bing et al. (1993)
Rat (SD) i.c. (amygdala) − 0.2 and 0.6 nmol −50 Moller et al. (1999)
Morphine ␮ Opioid agonist Rat (W) ip IA 5–10 mg/kg IA Ågmo et al. (1995)
U50,488H ␬-Opioid agonist Mouse (ddY) sc + 1.0 mg/kg +117 Tsuda et al. (1996)
Naloxone Opioid Ant Rat (W) ip IA 2.5 and 20 mg/kg IA Ågmo et al. (1995)
Mouse (ddY) sc − 3.0 mg/kg −36 Tsuda et al. (1996)
OrphaninFQ ORL agonist Rat (SD) i.c.v. + 0.3–1.0 nmol NI Lu et al. (2002)
ACTH1–24 MC4 agonist Rat (SD) i.c.v. − 0.3–10 ␮g −80 Corda et al. (1990)
␣-MSH MC4 agonist Rat (SD) i.c.v. − 1.0–5.0 mg −84 Corda et al. (1990)
SNAP-7941 MCH1 Ant Rat (W) i.p. + 40 mg/kg +60 Brocco (unpublished
observations)
Angiotensin II AT1 /AT2 Ago Rat (W) i.c.v. − 0.1–0.5 ␮g −80 Georgiev et al. (1990)
Abbreviations—CCK: cholecystokinin; CRF: corticotropin releasing factor; VP: vasopressin; NK: neurokinin; ORL: orphan opioid receptor; MCH1 :
melanin concentrating hormone; AT: angiotensin; ACTH: adrenocorticotropic hormone; MC: melanocortin; IA: no significant change (inactive) and NI:
not indicated. For other abbreviations, see Table 2.

and Fletcher, 2000; Daugé et al., 2001; Kõks et al., 2001b; neuropeptide, CRF, is generated in numerous cerebral struc-
Miyasaka et al., 2002; Noble and Roques, 2002). tures, including the cortex, amygdala, hippocampus and,
These findings question the anxiolytic potential of CCK2 most prominently, the hypothalamus (Gray and Magnuson,
(and CCK1 ) receptor antagonists. Correspondingly, in clin- 1992; Koob et al., 1994; De Souza, 1995; Steckler and Hols-
ical studies, despite their ability to block the anxiogenic boer, 1999; Eckart et al., 2002). CRF exerts its actions via
(panicogenic) actions of CCK4 , CCK2 receptor antagonists CRF1 and (with less pronounced affinity) CRF2 receptors
have not proven consistently efficacious as anxiolytic agents which share positive coupling (via the activation of Gs) to
(Kramer et al., 1995; Sramek et al., 1995, 2002; Van Megen adenylyl cyclase: recruitment of phospholipase C via Gq has
et al., 1996; Goddard et al., 1999; Pande et al., 1999b; Cham- also been demonstrated in neuronal tissue (De Souza, 1995;
bers and Fletcher, 2000; Blanchard et al., 2001b; Radu et al., Steckler and Holsboer, 1999; Dautzenberg and Hauger,
2003). Moreover, it is unlikely that appreciable amounts of 2002; Eckart et al., 2002; Blank et al., 2003a; Hauger et al.,
CCK4 or pentagastrin penetrate into CNS tissue upon i.v. 2003). Amongst several isoforms, CRF2␣ receptors pre-
application: this suggests that the susceptibility of their ef- dominate in supraspinal structures, while a CRF2␥ isoform
fects to blockade by CCK2 antagonists reflects actions at was recently isolated from human amygdala (Steckler and
gastrointestinal or vagal, rather than cerebral, loci. The lack Holsboer, 1999; Dautzenberg and Hauger, 2002; Hauger
of activity of CCK2 receptor antagonists in the VCT in ro- et al., 2003). In mammalian brain, an additional CRF-related
dents is, then, mirrored by their limited efficacy in man. peptide, urocortin, displays comparable affinity at CRF1 as
compared to CRF2 sites. Two further peptides, urocortin II
9.1.2. Corticotropin releasing factor (also known as stresscopin-related peptide) and urocortin
III (or stresscopin)—both 38 amino acids in length—were
9.1.2.1. CRF and related peptides: localization and cou- recently shown to behave as selective agonists at CRF2 ver-
pling to CRF1 and CRF2 receptors. The 41 amino acid sus CRF1 sites (Hsu and Hsueh, 2001; Lewis et al., 2001;
148 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Reyes et al., 2001; Dautzenberg and Hauger, 2002; Eckart bic as compared to hypothalamo-hypophyseal populations
et al., 2002; Hauger et al., 2003). The pharmacology of CRF of CRF1 receptors likewise revealed an anxiolytic pheno-
is further complicated by the existence of a CRF-binding type: this finding underscores the importance of cerebral
protein which recognises (and neutralizes) CRF and uro- populations of CRF1 receptors in the “direct” control of
cortin, though not urocortin II and III (Behan et al., 1995; mood as compared to “indirect” effects of populations
Seasholtz et al., 2001; Eckart et al., 2002). modulating activity of the hypothalamo-corticotropic axis
CRF1 receptors mediate the facilitatory influence of CRF and, ultimately, glucocorticoid secretion (Section 11) (Reul
upon ACTH secretion from the pituitary in response to and Holsboer, 2002). As judged by behavioural and auto-
stressors and other stimuli (De Souza, 1995; Arborelius nomic criteria, mice overexpressing CRF centrally reveal
et al., 2000; McElroy et al., 2002; Ohata et al., 2002; increased anxiety together with a dysregulation of the
Carrasco and Van de Kar, 2003) with CRF2 sites fulfilling hypothalamus–corticotropic axis: mice lacking CRF-binding
a more complex, modulatory and time-dependent role (Ma protein also show enhanced anxiety (Stenzel-Poore et al.,
et al., 1997; Coste et al., 2000, 2001; Kishimoto et al., 1994; Heinrichs et al., 1997b; Holsboer, 1999; Karolyi et al.,
2000b; Bale et al., 2002; Dautzenberg and Hauger, 2002; 1999; Coste et al., 2001; Clément al., 2002; Dirks et al.,
Eckart et al., 2002; Gold and Chrousos, 2002; Reul and 2002; Groenink et al., 2002, 2003; Van Gaalen et al., 2002).
Holsboer, 2002). Via actions at CRF1 (and CRF2 ) receptors Paradoxically, mice deprived of CRF reveal few changes in
in the brain, CRF fulfills a more general role in regulating behaviour indicative of reduced anxiety or insensitivity to
the endocrine, cognitive, emotional and autonomic response stress (Dunn and Swiergiel, 1999; Weninger et al., 1999;
to stress, which is characterized by robust alterations in Muglia et al., 2001 see Groenink et al., 2003). However, in
levels of CRF in the amygdala and several other limbic interpreting these findings, it should be noted that urocortin
regions (Albeck et al., 1997; Hsu et al., 1998; Merali et al., may compensate for decreases (or increases) in the activity
1998; Makino et al., 1999, 2002a; Steckler and Holsboer, of CRF.
1999; Bakshi and Halin, 2000; Gilligan et al., 2000; Gold Particularly, under conditions of high stress, anxiogenic
and Chrousos, 2002; Roozendaal et al., 2002). properties of CRF itself, and anxiolytic actions of peptider-
gic and non-peptidergic CRF1 antagonists, are expressed in
9.1.2.2. CRF1 receptors. CRF1 receptors are particularly several rodent and primate paradigms (De Fonseca et al.,
concentrated in the paraventricular hypothalamus, cortex, 1996; Heinrichs et al., 1997b, 2002; Griebel et al., 1998b;
PAG, hippocampus and amygdala (Sanchez et al., 1991a; Habib et al., 2000; Okuyama et al., 1999a; Keck et al.,
Chalmers et al., 1995; Chen et al., 2000b). Further, they 2001; Millan et al., 2001a; Gehlert et al., 2002; McElroy
mediate the excitatory influence of CRF upon LC-derived et al., 2002; Zorilla et al., 2002; Blanchard et al., 2003;
noradrenergic pathways, an action implicated in their acti- Blank et al., 2003b). Structures implicated in these ac-
vation by psychological and other stressors (Valentino et al., tions include the amygdala and the interconnected bed
1992; Arborelius et al., 2000; Kawahara et al., 2000; Mil- nucleus of the stria terminalis (Liebsch et al., 1995; Davis
lan et al., 2001a; Sauvage and Steckler, 2001; Valentino and Shi, 1999; Sajdyk et al., 1999a; Sajdyk and Gehlert,
and Van Bockstaele, 2001; Griebel et al., 2002b; Lejeune 2000; Bakshi et al., 2002; Keim et al., 2002), the PAG
and Millan, 2003). More tentative evidence for a facili- (Martins et al., 1997) and the lateral septum (Takahashi,
tatory influence of CRF1 receptors upon (a subpopulation 2001; Bakshi et al., 2002; Thomas et al., 2002). Consistent
of) DRN-localized serotonergic neurones has also been pre- with these observations, centrally-administered antisense
sented: however, there are conflicting data and the precise probes directed against CRF1 receptors display anxiolytic
nature of this interaction awaits clarification (Ruggiero et al., properties expressed, at least partially, via actions in the
1999; Isogawa et al., 2000; Millan et al., 2001a; Hammack amygdala (Liebsch et al., 1995, 1999; Heinrichs et al.,
et al., 2002, 2003; Linthorst et al., 2002a,b; Commons et al., 1997a). A preliminary (open-label) report that blockade of
2003; Roche et al., 2003; Temel et al., 2003; Thomas et al., CRF1 receptors elicits anxiolytic actions in depressive pa-
2003). Similarly, while it has been conjectured that CRF1 tients awaits extension to controlled studies in subjects with
receptors participate in the engagement of serotonergic path- “pure” anxious disorders (Zobel et al., 2000).
ways by stress, this hypothesis remains to be corroborated Relatively little information is available concerning con-
(Arborelius et al., 2000; Hammack et al., 2002; Linthorst flict procedures. Despite the absence of changes in VCT per-
et al., 2002a,b; Penalva et al., 2002). formance in mice overexpressing CRF, i.c.v. administration
The above observations may be related to alterations in of CRF is anxiogenic in the rat VCT (Table 19) (Britton et al.,
emotionality in mice lacking CRF1 receptors which man- 1985, 2000; Liebsch et al., 1995; Van Gaalen et al., 2002).
ifest an unambiguous anxiolytic phenotype in an array of Further, the anxiogenic effects of social defeat in the VCT
experimental procedures (Smith et al., 1998a; Timpl et al., were abolished in mice lacking CRF1 receptors (Liebsch
1998; Contarino et al., 1999a,b; Contarino and Gold, 2002). et al., 1995; Steckler and Holsboer, 1999). More recently,
Notably, a reduction in circulating levels of glucocorticoids the selective CRF1 antagonists, CP154,256 and DMP695,
could be dissociated from these changes. Moreover, a line were documented to elicit specific and dose-dependent ac-
of “conditioned” knock-out mice deprived of corticolim- tions in the VCT of a magnitude comparable to the BZP,
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 149

750 750
LICKS/ 3 MIN
LICKS/3 MIN

500 500

250 250

0 0
VEH VEH 5.0 40.0 80.0 MG/KG, I.P. VEH VEH 0.63 10.0 40.0 MG/KG, I.P.
CP 154,626 GR 205,171
NO SHOCK SHOCK NO SHOCK NO SHOCK

750
LICKS/3 MIN

500

250

0
VEH VEH 10.0 40.0 MG/KG, I.P.
SNAP-7941

NO SHOCK SHOCK

Fig. 9. Actions of the CRF1 receptor antagonist, CP154,526, of the NK1 receptor antagonist, GR205,171, and of the MCH1 receptor antagonist,
SNAP-7941, in the Vogel Conflict Test. Data are means ± S.E.M.s. Star indicates significance of differences to corresponding vehicle values. ( )
P < 0.05. Data are from Millan et al. (2001a), and Brocco and Millan (unpublished observations).

chlordiazepoxide (Table 19 and Fig. 9) (Griebel et al., 1998b; sponse to serotonergic neurones to stress, though this action
Millan et al., 2001a). In contrast to this benzodiazepine, remains to be confirmed (Chalmers et al., 1995; Sanchez
CRF1 receptor antagonists were active at doses devoid of et al., 1999; Van Pett et al., 2000; Lin et al., 2001; Higelin
sedative properties. Griebel et al. (2002b) reported simi- et al., 2001; Smagin et al., 2001; Hammack et al., 2002,
lar anxiolytic actions of two further CRF1 receptor antag- 2003; Linthorst et al., 2002).
onists, SSR125543A and antalarmin, in the VCT. Thus, The phenotype of CRF2 receptor-deficient mice has
chemically-diverse and selective antagonists at CRF1 recep- proven variable with reports of both reduced and—
tors reproducibly express anxiolytic actions in the VCT. In predominantly—heightened anxiety, together with a
view of the equivocal actions of CRF1 antagonists in certain hyper-responsiveness to stress. The predisposition to en-
rodent paradigms (see Millan et al., 2001a; Griebel et al., hanced anxiety may, at least partially, reflect a compen-
2002b), and the stressful nature of the VCT, it would ap- satory up-regulation of the actions of CRF (or urocortin) at
pear particularly well-suited to the future characterization of anxiogenic CRF1 sites. However, these observations likely
CRF1 antagonists as potential anxiolytic agents. unveil a more complex role of CRF2 receptors whereby they
Very recently, employing a microinjection approach, it mediate both anxiogenic and (possibly delayed) anxiolytic
was shown that the lateral septum contributes to the anxi- actions (Heinrichs et al., 1997a; Bale et al., 2000, 2002;
olytic properties of CRF1 receptor antagonists in the VCT Coste et al., 2000; Kishimoto et al., 2000b; Takahashi, 2001;
(Thomas et al., 2002). Takahishi et al., 2001a; Radulovic et al., 2002; Todorovic
et al., 2002; Valdez et al., 2002a,b). Possibly related to an
9.1.2.3. CRF2 receptors. CRF2 receptors are localized in anxiolytic role for CRF2 sites, mice lacking urocortin reveal
the lateral septum, amygdala, ventromedial hypothalamus an anxiogenic phenotype (Vetter et al., 2002), while i.c.v.
and hippocampus (primates), while modest levels are found application of urocortin III suppresses anxious behaviours
in the DRN: therein they have been implicated in the re- in rats (Valdez et al., 2002b). Contrariwise, though antisense
150 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

probes provoking modest reductions (∼20%) in the central nate in parvocellular neurones of the hypothalamic paraven-
density of CRF2 receptors levels did not markedly affect tricular nucleus, the amygdala and the PAG), V1A receptors
anxiety in rats, a more pronounced (∼70%) loss of CRF2 are enriched in the cortex as well as the septum, hip-
receptors in the lateral septum—a site at which CRF itself pocampus and amygdala (Burbach et al., 1993, 1995, 1998;
behaves anxiogenically—was associated with an anxiolytic Hallbeck et al., 1999). V1B receptors have been identified
profile (Radulovic et al., 1999; Kishimoto et al., 2000b; Ho in the amygdala, cingulate cortex, frontal cortex, nucleus
et al., 2001; Takahashi, 2001). In line with these variable accumbens and hippocampus, as well as in the DRN and the
findings, blockade of CRF2 sites by injection of the pep- LC (Lolait et al., 1992, 1995; Burbach et al., 1995; Vaccari
tidergic antagonist, antisauvagine-30, into the septum (and, et al., 1998; Brinton et al., 2000; Hernando et al., 2001).
possibly, the amygdala) was accompanied by anxiolytic Functional evidence for a role of vasopressin in the mod-
actions in several studies, whereas others observed a lack ulation of anxious states is provided by observations of an
of effect (Radulovic et al., 1999; Ho et al., 2001; Taka- induction of central vasopressin synthesis and release in the
hashi, 2001; Takahashi et al., 2001a; Pelleymounter et al., amygdala, septum and elsewhere upon exposure to stress
2002). Regrettably, in none of the above experiments was (Chen and Herbert, 1995; Albeck et al., 1997; Wotjak et al.,
the significance of CRF2 receptors evaluated employing 1998; Aguilera and Radadan-Diehl, 2000; Ebner et al., 2002;
punishment-based conflict procedures. Makino et al., 2002a). Further, Brattleboro rats, which genet-
ically lack vasopressin, show low levels of anxiety (Williams
9.1.2.4. Multiple roles of CRF. To summarize, a convinc- et al., 1985; Herman et al., 1986), while anxiolytic prop-
ing body of evidence supports a critical role of CRF1 re- erties of mixed V1A /V1B receptors antagonists have been
ceptors in the response to stress, and it has been clearly documented (Liebsch et al., 1996). Antisense knockdown
demonstrated that selective blockade of CRF1 receptors di- of V1A sites in the septum suppressed anxious behaviour
minishes anxious states. However, there is a complex pat- in rats (Landgraf et al., 1995) and V1A receptors have been
tern of—not necessarily discordant—data (Contarino et al., implicated in the control of social behaviour in rodents and
1999b; Bale et al., 2002; Dautzenberg and Hauger, 2002; other species (Hammock and Young, 2002). Further, though
Holsboer, 1999; Valdez et al., 2002a) suggesting that: (1) the phenotype of mice lacking V1B sites was interpreted
stimulation of CRF2 receptors amplifies anxiogenic actions as anxiogenic, the novel, selective V1B receptor antagonist,
transduced by CRF1 sites and/or (2) possibly via actions in SR149,415, elicited a selective and significant increase in
other regions and over different time-scales (during recov- punished responses in the VCT in parallel with anxiolytic ac-
ery from stress), activation of CRF2 receptors opposes the tions in other paradigms of untrained behaviours (Table 19)
anxiogenic properties elicited via CRF1 sites. The applica- (Lolait et al., 2000; Hernando et al., 2001; Griebel et al.,
tion of the VCT and other conflict models would likely pro- 2002a; Serradeil-Le Gal et al., 2002).
vide important insights into these multiple role(s) of CRF in Thus, while there is a reasonable body of evidence in-
the response to stress and the modulation of anxious states. dicating that central, stress-responsive VP-containing path-
Currently, it would be premature to attempt concrete predic- ways are dedicated to the control of mood, the precise roles
tions regarding the influence of (systemic administration of) of V1A and V1B sites in the control of anxious states remain
CRF2 receptor antagonists upon anxiety disorders in man. poorly-understood.

9.1.3. Vasopressin and oxytocin 9.1.3.2. Oxytocin and oxytocin receptors. By analogy to
vasopressin, oxytocin is synthetized by magnocellular neu-
9.1.3.1. Vasopressin, V1A and V1B receptors. The actions rones situated in the supraoptic and paraventricular nuclei
of vasopressin are mediated by V1A , V1B and V2 receptors, of the hypothalamus which project to the median eminence
of which both the V1A subtype (which stimulates phos- and the posterior lobe of the pituitary. On the other hand,
pholipase C via Gq) and the V1B subtype (which likewise parvocellular neurones of the paraventricular nucleus deliver
engages phospholipase C via Gq) are found in the CNS OT-containing afferents to a broad range of CNS structures
(Lolait et al., 1992, 1995; Hallbeck et al., 1999). Hypophy- including the cortex, hippocampus, LC, DRN, septum and
seal populations of V1B receptors transduce the facilitatory amygdala (Gimpl and Fahrenholz, 2001; Millan, 2002a).
influence of vasopressin (mediated in synergy with CRF) Oxytocin-containing neurones manifest pronounced respon-
upon the outflow of ACTH—including both basal release sivity to emotional stress, and the presence of eponymous
and its induction by anxiogenic stimuli. Conversely, hy- oxytocin receptors (which are coupled via Gq to phospholi-
pothalamic populations of V1A receptors indirectly con- pase C) throughout corticolimbic structures—including the
tribute to activation of the corticotropic axis (Antoni, 1993; hypothalamus and amygdala—offers a substrate for a po-
Burbach et al., 1995; Aguilera and Radadan-Diehl, 2000; tential role in the modulation of anxious behaviour (Bale
Birnbaumer, 2000; Morimoto et al., 2000; Keck et al., et al., 1995; Burbach et al., 1995; Nishioka et al., 1998;
2002; Serradeil-Le Gal et al., 2002; Carrasco and Van de Uvnäs-Moberg, 1998; Vaccari et al., 1998; Wotjak et al.,
Kar, 2003). Tracking the broad, supraspinal distribution of 1998; Gimpl and Fahrenholz, 2001; Millan, 2002a; Winslow
vasopressin-containing pathways (which principally origi- and Insel, 2002).
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 151

Surprisingly, little functional information concerning OT There is still much confusion concerning the—probably
and anxious behaviour is currently available. Nevertheless, multiple—roles of NK1 receptors in the modulation of anx-
anxiolytic properties of OT have been reported upon cen- ious states. Thus, SP fragments have been reported to exert
tral administration to rodents and the amygdala plays a key anxiolytic-like effects in primates (Barros et al., 2002),
role in its control of anxious states (Stoehr et al., 1992; while SP exerts anxiolytic actions upon administration into
Uvnäs-Moberg et al., 1994; McCarthy et al., 1996; Windle the ventral pallidum/nucleus basalis magnocellularis of ro-
et al., 1997; Bale et al., 2001). Further, mice with genetically dents (Hasenöhrl et al., 1998; Nikolaus et al., 2000, see also
disabled oxytocin receptors were found to display an anx- Zernig et al., 1993). Mice strains possessing low levels of SP
iogenic profile, though conflicting findings have also been manifested a high level of anxiety in the VCT—possibly due
reported (DeVries et al., 1997; Mantella and Amico, 2002; to a reduction in the activity of hippocampal populations of
Winslow and Insel, 2002). GABAergic neurones (Sloviter et al., 2001; Sudakov et al.,
The above observations indicate that, in addition to the 2001). Further, the functional inactivation of NK1 receptors
familiar role of oxytocin in the modulation of social, ag- (indirectly) reinforces the activity of corticolimbic nora-
gressive and maternal behaviour, this neuropeptide may con- drenergic, dopaminergic and—principally upon long-term
tribute to the control of anxious states (Witt, 1995; DeVries administration—serotonergic pathways (Froger et al., 2001;
et al., 1997; Popik and Van Ree, 1998; Uvnäs-Moberg, 1998; Millan et al., 2001e; Santarelli et al., 2001; Adell et al., 2002;
Argiola, 1999; Winslow and Insel, 2002). Conley et al., 2002; Léger et al., 2002; Lejeune et al., 2002;
Intriguingly, then, it appears that OT and VP may fulfil Liu et al., 2002; Maubach et al., 2002; Commons et al.,
opposite anxiolytic and anxiogenic roles, respectively. 2003). These actions highlight the attractiveness of NK1
receptors as innovative targets for the improvement of de-
9.1.4. Substance P and other tachykinins pressive states but, arguably, are inconsistent with the acute
expression of anxiolytic properties. However, as mentioned
9.1.4.1. Multiple classes of neurokinin receptor. The below, at doses which do not themselves affect noradrener-
mammalian tachykinins, substance P (SP, 11 amino acids gic pathways, NK1 antagonists can (probably directly) pre-
in length), neurokinin A and neurokinin B, are thought vent their activation by CRF and stress (Hahn and Bannon,
to exert their actions principally via excitatory neurokinin 1999; Bert et al., 2002; Steinberg et al., 2002). More-
(NK)1 , NK2 and NK3 receptors, respectively, each of over, by analogy to other classes of antidepressant agent
which stimulates phospholipase C via recruitment of (Section 4.4), it may be speculated that the sustained en-
Gq (Maggi, 1995; Quartara and Maggi, 1997; Mantyh, gagement of corticolimbic monoaminergic projections by
2002). chronic administration of NK1 receptor antagonists initiates
postsynaptic adaptive changes ultimately reflected in the
9.1.4.2. SP and NK1 receptors. Moderate to high con- long-term alleviation of anxious states.
centrations of SP and NK1 receptors have been detected Moreover, an extensive body of evidence has accrued in-
in limbic regions of the brain, including the FCX, amyg- dicating that activation of NK1 receptors triggers or aggra-
dala, hypothalamus, LC, PAG and hippocampus: in the lat- vates anxious states.
ter structure, a sub-population of NK1 receptors is localized First, upon introduction into sites in the PAG, amygdala
on GABAergic neurones. In addition to broadly-distributed and lateral septum, SP elicited defensive, anxiogenic and
clusters of local neurones, several SP-containing projections aversive actions (Ravard et al., 1994; Aguiar and Brandao,
have been identified, including pathways from the septum 1996; Mongeau, 1998; Gavioli et al., 1999, 2002a; Rupniak
and mammillary bodies to the hippocampus, from the amyg- and Kramer, 1999; Smith et al., 1999; Baretta et al., 2001;
dala to the hypothalamus and from the parabrachial nucleus De Araújo et al., 2001; Rupniak et al., 2001).
to the amygdala itself (Pernow, 1983; Mantyh et al., 1984; Second, local application of SP into the LC (directly)
Kiyama et al., 1993; Otsuka and Yoshioka, 1993; Nakaya enhanced the activity of noradrenergic neurones, while (as
et al., 1994; Saria, 1999; Ribeiro-da-Silva and Hökfelt, 2000; mentioned above) acute, systemic administration of NK1
Sloviter et al., 2001). Interestingly, SP is colocalized with receptor antagonists abrogated their activation by CRF and
raphe-derived serotonergic pathways in man, though not nec- stress (Hahn and Bannon, 1999; Chen et al., 2000a; Bert
essarily in other species (Sergeyev et al., 1999). These obser- et al., 2002; Steinberg et al., 2002). Further, in analogy to
vations, the responsiveness of corticolimbic SP-containing BZPs, NK1 receptor antagonists blocked the induction of
neurones to stress (Maggi, 1995; De Felipe et al., 1998; frontocortical DA release by stress (Barton et al., 1999).
Smith et al., 1999; Steinberg et al., 2002), their interaction Third, in accordance with these findings, in a diversity of
with monoaminergic pathways (vide infra), and the facilita- models (Rupniak and Kramer, 1999; File, 2000; Rupniak
tory influence of NK1 receptors upon corticolimbic release et al., 2000, 2001; Vassout et al., 2000; Ballard et al., 2001;
of GABA and glutamate (Sloviter et al., 2001; Stacey et al., Baretta et al., 2001; Cheeta et al., 2001b; Santarelli et al.,
2002a,b) underpin interest in a role of SP in the modula- 2001, 2002; Steinberg et al., 2002; Varty et al., 2002a,b),
tion of anxious states (Rupniak and Kramer, 1999; Mantyh, specific anxiolytic actions of non-peptidergic NK1 receptor
2002). antagonists have been reported in several species. Though
152 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

results concerning effects of NK1 receptor antagonists in display anxiolytic properties in the VCT (Walsh et al., 1995;
conflict procedures have, to date, been conspicuous by their Teixeira et al., 1996; Griebel et al., 2001a,c; Ribeiro and
absence, anxiolytic actions have been documented in a De Lima, 2002). This inactivity is somewhat perplexing in
paradigm of conditioned fear (Ballard et al., 2001). Further, view of the interrelated observations that: (1) NK2 recep-
we recently found that the potent and selective antago- tor antagonists interfere with the activation of LC-derived
nist, GR205,171, elicits robust and stereospecific anxiolytic noradrenergic neurones by CRF (Steinberg et al., 2001)
properties in the VCT (Table 19 and Fig. 9). Inasmuch as and that (2) anxiolytic actions of CRF1 receptor antago-
clinically-available 5-HT1A receptor agonists are poorly nists in the VCT involve their suppressive influence upon
effective in subjects pre-treated with BZPs (Section 18.3) stress-induced release of NA (Section 9.1.2). An alterna-
(De Martinis et al., 2000), it would be of importance to tive, non-noradrenergic explanation for anxiolytic actions
confirm observations that NK1 antagonists suppress the anx- of NK2 receptor antagonists in models of exploratory be-
iety which accompanies their abrupt withdrawal in rodents haviour invokes their modulatory influence upon the activity
(Ribeiro and De Lima, 2002). of DRN-derived serotonergic pathways (Walsh et al., 1995).
Fourth, behavioural and neurochemical analyses of mice NK3 receptors are concentrated in several limbic regions,
lacking NK1 sites revealed a diminution of anxious be- including the PAG, amygdala and hypothalamus, and they
haviour and stress-related responses (Rupniak and Kramer, are also found in the cerebral cortex (Ding et al., 1996; Yip
1999; Rupniak et al., 2000, 2001; Vassout et al., 2000; and Chahl, 1997, 2001; Mileusnic et al., 1999). They exert a
Ballard et al., 2001; Santarelli et al., 2001, 2002). More re- facilitatory influence upon corticolimbic noradrenergic and
cently, selective deletion of the “tacl” gene which encodes dopaminergic—but not serotonergic—neurones (Seabrook
SP (and NK A) yielded an anxiolytic profile in a broad range et al., 1995; Jung et al., 1996; Marco et al., 1998; Panocka
of experimental procedures, including a conflict paradigm et al., 2001; Bert et al., 2002; Léger et al., 2002. These find-
(Bilkei-Gorzo et al., 2002). ings are consistent with a role in the modulation of anx-
Fifth, reduced anxiety was reported upon ablation of NK1 ious states, and of mood in general. However, information
receptor-bearing neurones in the amygdala by local injection remains fragmentary (Ribeiro et al., 1999; Panocka et al.,
of a SP-saporin, cytotoxic conjugate which enters cells via 2001) and data from the VCT and other conflict procedures
NK1 receptor internalisation (Mantyh, 2002). are not, as yet, available.
Finally, there is evidence that prolonged exposure to an- There is, thus, a need for additional study of the roles of
tidepressant agents depletes central levels of SP, for exam- multiple classes of NK receptor in the control of anxious
ple, in the amygdala, observations which have been related behaviour by use of the VCT and other procedures. This
to their treatment of affective disorders (Jones et al., 1985; would be of particular importance in view of the ongoing,
Shirayama et al., 1996; Burnet et al., 2001). The possibil- therapeutic evaluation of selective antagonists at NK1 , NK2
ity that a modulation of SP/NK1 receptor-mediated trans- and NK3 receptors for the treatment of various categories of
mission by antidepressants is involved in the long-term de- psychiatric and neurologic disorder Raffa (1998).
velopment of their anxiolytic properties would, thus, be of
interest to evaluate. 9.1.5. Neuropeptide Y
Anxiolytic actions of a NK1 receptor antagonist (aprepi-
tant) were reported in a clinical study undertaken in sub- 9.1.5.1. Localization and coupling to multiple receptor sub-
jects suffering from major depression (Kramer et al., 1998; types. Together with peptide YY and pancreatic polypep-
Rupniak and Kramer, 1999; Krishnan, 2002). Though in- tide PP, likewise 36 amino acids in length, NPY exerts its
formation from patients presenting with pure GAD or other actions via multiple classes of receptor, of which Y1 , Y2 ,
anxious states does not appear to have been divulgued, this Y4 and Y5 receptors occur in rat and human CNS and are
highly-publicized therapeutic trial has sparked a resurgence of pertinence to the modulation of anxious states—a further
of interest in the clinical utility of NK1 antagonists. In- Y6 receptor possibly exists in murine CNS (Heilig et al.,
deed, these encouraging observations will hopefully crys- 1994; Wettstein et al., 1995; Dumont et al., 1996a,b; Michel
tallize into definitive proof that NK1 receptor antagonists et al., 1998; Zimanyi et al., 1998; Bischoff and Michel, 1999;
genuinely represent a novel approach to the management of Redrobe et al., 2002b; Thorsell and Heilig, 2002). Y1 , Y2 ,
anxious (and depressive) disorders. Y4 and Y5 receptors all exert an inhibitory influence upon
adenylyl cyclase via recruitment of Gi . The high levels of
9.1.4.3. NK2 and NK3 receptors. Though NK2 receptors NPY (principally in interneurones) in corticolimbic tissues,
are sparse in the CNS and have proven difficult to visualize, including the nucleus accumbens, PAG, septum and amyg-
compelling evidence for their existence in the septum, hip- dala, underscore interest in its role in the control of anxious
pocampus and cerebral cortex of rodents and man was re- states. Further, in the amygdala, NPY is partially colocalized
cently reported (Steinberg et al., 1998; Bensaid et al., 2001; with GABA (McDonald and Pearson, 1989; Aoki and Pickel,
Saffroy et al., 2001, 2003; Candenas et al., 2002). Despite 1990; Caberlotto et al., 2000; Kask et al., 2002) or with NA
their positive effects in models based on exploratory and (Wettstein et al., 1995; Zimanyi and Poindexter, 2000). Y1
defensive behaviours, NK2 receptor antagonists failed to receptors (the most abundant subtype in the CNS) are highly
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 153

expressed in the hippocampus, PAG, frontal cortex, hypotha- 1999, 2000, 2002). Consistent with these findings, Heilig
lamus and amygdala while Y2 receptors, which share their et al. (1989) demonstrated that NPY is highly active in el-
occurrence in the hippocampus, amygdala and hypothala- evating punished responses in the VCT (Table 19). Though
mus, are also found in the lateral septum, nucleus accum- the (polyvalent) role of NPY in the control of nociception
bens, raphe nuclei and the LC. In contrast to Y1 sites, Y2 (Millan, 1999; Kask et al., 2002) suggests caution in the in-
receptors may operate as autoreceptors on NPY-containing terpretation of these data, it appears that the anxiolytic ac-
neurones (Dumont et al., 1996a,b, 2000; Naveilhan et al., tions of NPY are specifically expressed in the absence of
1998; Parker and Herzog, 1999; Caberlotto et al., 2000; alterations in nociceptive thresholds and water appetite. In-
Migita et al., 2001; Kask et al., 2002; Redrobe et al., 2002b). terestingly, in light of the robust anxiolytic actions of CRF1
receptor antagonists in the VCT (Section 9.1.2.2), it has been
9.1.5.2. Modulation. Modulation of the cerebral synthe- proposed that NPY opposes the anxiogenic actions of CRF
sis of NPY by a variety of acute and chronic, fear-inducing in at least two cerebral regions, the septum and the amygdala
stimuli, including an acceleration of its production in the (Heilig et al., 1994; Kask et al., 1997, 2001a, 2002; Britton
amygdala in parallel with behavioural habituation to stress, et al., 2000). While engagement of GABAergic mechanisms
has prompted the hypothesis that NPY “buffers” the ef- has been implicated in the sedative and sleep-promoting
fects of long-term exposure to anxiogenic stimuli (Krysiak properties of NPY, the relationship of GABAA /BZP recep-
et al., 2000; Kask et al., 2002; Thorsell and Heilig, 2002). tors to anxiolytic actions mediated by Y1 receptors remains
This assertion is supported by the relative insensitivity to to be clarified (Antonijevic et al., 2000; Naveilhan et al.,
fear-inducing situations and stress of transgenic mice over- 2001).
expressing (hippocampal) NPY: moreover, they display an In the above-evoked studies of the central administration
anxiolytic profile in the VCT (Inui et al., 1998; Thorsell of NPY and related sequences, it was concluded that Y2
et al., 1999, 2000). Contrariwise, mice lacking NPY are receptors play an insignificant role as compared to their
prone to exaggerated anxiety upon exposure to fear-inducing Y1 counterparts in transducing the anxiolytic actions of
situations (Palmiter et al., 1998; Bannon et al., 2000). Inter- NPY (Table 19) (Heilig et al., 1989, 1992, 1993, 1994;
estingly, levels of NPY in the cerebrospinal fluid of patients Wahlestedt et al., 1993; Broqua et al., 1995; Britton et al.,
with major depression were inversely proportional to the in- 1997; Kask et al., 2002). Indeed, activation of Y2 recep-
tensity of concomitant anxious symptom: this suggests that tors in the amygdala may actually be accompanied by
subjects with a low rate of NPY release had a greater procliv- anxiogenic effects (Nakajima et al., 1998; Sajdyk et al.,
ity towards anxious behaviour (Widerlöv et al., 1989). Other 2002a,b; Thorsell and Heilig, 2002), and an anxiolytic
alterations in NPY levels in patients with affective and anx- phenotype has been documented for mice lacking Y2 re-
ious disorders were likewise interpreted as congruent with ceptors (Redrobe et al., 2002a, 2003). However, Y2 sites
a role in the control of emotivity (Kask et al., 2002). Com- have been reported to mediate anxiolytic actions via loci in
plementing these observations, NPY exerts an inhibitory in- (or near) the LC. This observation is consonant with the
fluence upon ACTH secretion in man Antonijevic et al., predominant role of Y2 as compared to Y1 receptors in the
2000. NPY-mediated inhibition of noradrenergic pathways, acti-
vation of which may aggravate anxious states (Section 4.1.1)
9.1.5.3. Key roles of Y1 and Y2 sites in the control of anx- (Finta et al., 1992; Illes et al., 1993; Kask et al., 1998a, 2000,
ious states. In accordance with the above comments, cen- 2002).
tral administration of NPY (and fragments thereof) results
in robust anxiolytic actions in a diversity of paradigms in- 9.1.5.4. Other NPY receptor subtypes. Most investigations
cluding both natural and conditioned behaviours and con- have focussed on the appetite-modulating role of Y5 recep-
flict tests. Microinjection studies indicate that these effects tors, but their discovery in regions controlling emotionality,
are exerted in at least three structures: the amygdala (Heilig such as the amygdala, hippocampus and septum, indicates
et al., 1993; Heilig, 1995; Sajdyk et al., 1999b), the dorsolat- that they may also contribute to the control of anxious states
eral septum (Kask et al., 2001a,b) and the PAG (Kask et al., (Bischoff and Michel, 1999; Nichol et al., 1999; Parker and
1998b), though the possible implication of additional sites in Herzog, 1999; Zimanyi and Poindexter, 2000; Thorsell and
the hypothalamus, hippocampus and elsewhere should not Heilig, 2002). Further, Y5 receptors have been identified
be discounted (Kask et al., 2002). Characterization of anti- on limbic populations of GABAergic and CRF-containing
sense probes, and of NPY fragments possessing differential neurones (Bischoff and Michel, 1999; Grove et al., 2000).
potency at Y1 versus Y2 receptors, revealed that activation A study with the Y5 receptor antagonist, CGP71683A, re-
of the former was responsible for these anxiolytic actions: vealed only a modest influence upon anxious states un-
indeed, a decrease in activity at Y1 sites in the PAG is asso- der resting conditions, but an anxiogenic effect under con-
ciated with anxiogenic effects suggestive of a tonic role of ditions of stress, coinciding with a recent report that Y5
Y1 receptors in the control of anxious states (Heilig et al., receptors transduce anxiolytic actions in the (basolateral)
1989, 1993, 1994; Wahlestedt et al., 1993; Broqua et al., amygdala (Broqua et al., 1995; Kask et al., 2001c, 2002;
1995; Heilig, 1995; Britton et al., 1997; Kask et al., 1998b, Sajdyk et al., 2002a). Thus, additional work with further
154 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

selective antagonists and agonists—and with models such Under conditions of aversive stimulation, it has been pos-
as the VCT—would be instructive in defining their potential tulated that the harnessing of local or LC-derived sources of
role. galanin negates the anxiogenic effects of NA in the amyg-
Y4 receptors have been detected in the amygdala, the hip- dala (Crawley et al., 2002; Khoshbouei et al., 2002a,b). In
pocampus and the mammillary bodies, raising the possibil- this light, it is something of a conundrum that direct intro-
ity of a role in the modulation of emotionality and anxious duction of galanin into the amygdala resulted in a specific
states (Zimanyi et al., 1998; Parker and Herzog, 1999). How- suppression of punished responses in the VCT (Table 19)
ever, there is currently no direct support for this contention (Bing et al., 1993; Möller et al., 1999). Moreover, though the
inasmuch as mice lacking pancreatic polypeptide (which dis- i.c.v. administration of galanin reduced anxiety in the VCT,
plays marked affinity for Y4 receptors) did not manifest sig- the dose–response curve was bell-shaped and higher doses
nificant alterations in their behaviour in models of potential increased anxiety. Mice overexpressing GAL showed lit-
anxiolytic activity (Asakawa et al., 1999). tle change in anxious behaviour, though anxiogenic actions
There is, thus, a compelling body of evidence that NPY of the ␣2 -antagonist, yohimbine, were abolished (Crawley
fulfils a crucial role in the response to, and modulation of, et al., 2002; Holmes et al., 2002a). Several reasons may un-
anxious states. Continued use of the VCT, together with derlie these apparently discordant data.
subtype-selective ligands, should provide further insights First, galanin may exert contrasting anxiogenic and anxi-
into the significance of NPY in the etiology and, optimisti- olytic actions via different receptor subtypes, of which three
cally, management of anxiety disorders. are currently recognized. Inhibitory “GAL1 ” and “GAL3 ”
receptors are negatively coupled via Gi to adenylyl cylase,
9.1.6. Glucagon-like peptide-1 whereas excitatory “GAL2 ” receptors are positively coupled
While the major processing product of pre-pro-glucagon to phospholipase C via Gq (Branchek et al., 2000; Waters and
in the pancreas is glucagon, the primary splicing products Krause, 2000). In the amygdala, PAG and LC, GAL1 sites
in the CNS are glucagon-like peptide-1 (7-36 amide) and predominate, but GAL2 and GAL3 sites are also present.
glicentin peptide (Lopez et al., 1983; Flint et al., 1998). Further, GAL1 , GAL2 and GAL3 receptors are all local-
Glucagon-like peptide-1 (which activates adenylyl cylase) ized in the hypothalamus, PAG, hippocampus and cortex
plays a key role in the control of appetite, likely reflect- (O’Donnell et al., 1999; Waters and Krause, 2000). Though
ing actions in the nucleus solitarius and the hypothalamus little is known of the functional roles of these GAL recep-
(Lopez et al., 1983; Flint et al., 1998). However, brainstem tor subtypes, mice genetically deficient in GAL1 receptors
neurones producing glucagon-like peptide-1 (localized near exhibited an anxiogenic profile, suggesting that GAL1 sites
the nucleus of the solitary tract) also innervate the DRN, may counter anxious states (Holmes et al., 2002b). Second,
LC and many limbic regions, including the PAG, hippocam- the influence of galanin upon anxious states may depend
pus, septum and amygdala: these structures are also en- upon its site of action. For example, via actions at both the
riched in receptors for this peptide (Turton et al., 1996; Jin cell body and terminal level, a “negative feedback” autore-
et al., 1988; Merchenthaler et al., 1999). Direct injection of ceptor action of galanin upon the activity of LC-derived no-
glucagon-like peptide-1 into the amygdala specifically sup- radrenergic neurones (Hökfelt et al., 1998; Perez et al., 2001;
pressed punished responding in the VCT, indicative of an Millan, 2002a; Yoshitake et al., 2003a) may act anxiolyt-
anxiogenic role in this structure (Table 19) (Möller et al., ically. On the other hand, anxiogenic and/or anxiolytic
2002). Though virtually nothing else is known of its role of actions may be exerted postsynaptically to noradrenergic
the control of emotionality, underpinning these findings in projections in limbic regions (vide supra). Third, in addition
the VCT, it was recently shown that glucagon-like peptide-1 to its co-storage with NA in noradrenergic neurones, GAL
also exerts aversive properties by actions in the amygdala is colocalized with 5-HT (Hökfelt et al., 1998; Perez et al.,
(Kinzig et al., 2002). 2001). Indeed, galanin exerts a modulatory influence upon
serotonergic transmission and displays a complex pattern of
9.1.7. Galanin (generally inhibitory) interactions with 5-HT1A autorecep-
The 29 amino acid, galanin, is broadly distributed tors in raphe nuclei, as well as with postsynaptic 5-HT1A
throughout the corticolimbic system, including the hip- sites in the amygdala and hippocampus (Hökfelt et al., 1998;
pocampus, the nucleus accumbens, the amygdala, the con- Misane et al., 1998; Diaz-Cabiale et al., 2000; Kehr et al.,
tiguous bed nucleus of the stria terminalis and the PAG 2002a,a,b; Yoshitake et al., 2003b). Finally, galanin exerts
(Skofitsch and Jacobowitz, 1985; Bartfai et al., 1992, 1993; an excitatory influence upon the mesolimbic release of
Merchenthaler et al., 1993; Waters and Krause, 2000; Perez dopamine (Ericsson and Ahlenius, 1999). Galanin may, thus,
et al., 2001). In these structures, galanin is both synthesized act via several substrates in the modulation of anxious states.
locally and released from ascending noradrenergic path- Much work remains to be performed with the VCT and
ways wherein it co-exists with NA: both of these pools of other models to determine the precise influence of GAL upon
galanin are recruited in response to stress (Holmes et al., anxious states and, in particular, to delineate the respective
1995, 2002a,b; Hokfelt et al., 1998; Crawley et al., 2002; roles and potential clinical pertinence of GAL1 , GAL2 and
Khoshbouei et al., 2002a,b). GAL3 receptor subtypes.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 155

9.1.8. Neurotensin and GABAergic networks, including an inhibitory influence


Neurotensin (13 amino acids in length) and the related upon GABAergic neurones in the hippocampus, amygdala
sequence, neuromedin N (six amino acids in length), are and PAG (Mansour et al., 1988; Cohen et al., 1992; Herz
processed from a common precursor peptide. They exert et al., 1993; Sugita and North, 1993; Kang et al., 2000;
their actions via “NT1 ” and “NT2 ” receptors—the func- Kieffer and Gavériaux-Ruff, 2002; Kozicz, 2002; Liberzon
tional status of a third neurotensin binding entity (identical et al., 2002; Millan, 2002a; Tao and Auerbach, 2002a,b).
to the protein, sortilin) is questionable since it is not a Indicative of contrasting roles of multiple opioid receptor
G-protein-coupled receptor (Vincent et al., 1999; McMahon subtypes, whereas mice lacking ␬-opioid receptors do not
et al., 2002). NT1 and NT2 receptors are both positively reveal an altered phenotype, populations deprived of either
coupled via Gq to phospholipase C, and differentially inter- enkephalins or ␦-opioid receptors reveal enhanced anxiety.
act with several other intracellular cascades (Rostène and These findings suggest an anxiolytic role for enkephalins
Alexander, 1997; Vincent et al., 1999; McMahon et al., at ␦-receptors. Correspondingly, pharmacological evidence
2002). While NT1 receptors are particularly abundant in the was forwarded that ␦-opioid receptors participate in express-
lateral septum, NT2 receptors are concentrated in the cortex, ing the anxiolytic properties of cannabinoids (Köning et al.,
hippocampus and hypothalamus (Sarret et al., 1998; Walker 1996; Filliol et al., 2000; Ragnauth et al., 2001; Berrandero
et al., 1998). Stress is accompanied by alterations in central and Maldonado, 2002).
(hypothalamic) levels of neurotensin. Under these condi- It has been reported that high doses of the opioid receptor
tions, via recruitment of the anxiogenic peptide, CRF, neu- antagonist, naloxone, interfere with the anxiolytic actions of
rotensin exerts a facilitatory influence upon ACTH secretion BZPs in the VCT and several other procedures (Millan and
(Rowe et al., 1995; Helmreich et al., 1999; Seta et al., 2001). Duka, 1981; Ågmo et al., 1995; Tsuda et al., 1996). Inas-
The above comments, together with the facilitatory influ- much as naloxone preferentially blocks ␮- and ␦-opioid as
ence of NT upon serotonergic pathways (Shipley et al., 1987; compared to ␬-opioid receptors, the need for a high dose in-
Li et al., 2001a), support a role of neurotensin in the control dicated that blockade of ␬-opioid rather than ␮- or ␦-opioid
of emotivity. Though the selective NT1 antagonist, SR48692, receptors was involved, an assertion supported by anxiolytic
was ineffective in several rodent paradigms, including the actions of ␬-opioid receptor agonists in the mouse VCT
VCT (Griebel et al., 2001a), analysis of its actions in a mouse (Table 19) (Tsuda et al., 1996) and a plus-maze procedure
defensive behaviour procedure suggested that engagement in rats (Privette and Terrian, 1995). This contention is, how-
of NT1 receptors may contribute to the adaptive response ever, inconsistent with studies of mice deficient ␬-opioid
to unavoidable stress (Griebel et al., 2001a). Further, any receptors which do not show an anxious phenotype (vide
role of the more widespread NT2 receptors in the modula- supra). Further, supporting evidence for an anxiolytic role of
tion of anxious states remains to be explored. While selec- ␬-opioid receptors is unavailable and their activation is gen-
tive antagonists for NT2 sites are not apparently available, erally anxiogenic and aversive to rodents and higher species
centrally-active, neurotensin analogues (agonists) have re- (Millan, 1990; Bals-Kubik et al., 1993; Nobre et al., 2000;
cently been described and their actions in the VCT and re- Sante et al., 2000; Marin et al., 2003). Thus, a contribu-
lated protocols would be of interest to evaluate (Boules et al., tion of ␦-opioid receptors to the anxiolytic effects of BZPs
2001; McMahon et al., 2002). would offer an alternative interpretation. Though such a role
of ␦- or ␬-opioid receptors in mediating the anxiolytic ac-
9.1.9. Multiple opioid peptides and melanocortins tions of BZPs appears intuitively improbable—and clearly
requires corroboration—it would be of interest to examine
9.1.9.1. Endorphins, µ-, δ- and κ-receptors. Multiple opi- further inasmuch as naloxone was reported to attenuate the
oid peptides exert their actions via ␮-, ␦- and ␬-opioid re- anxiolytic actions of BZPs in human subjects (Millan and
ceptors, for which ␤-endorphin/endomorphins, enkephalins Duka, 1981; Duka et al., 1982).
and dynorphin-related peptides, respectively, comprise their Surprisingly, mice genetically deprived of ␮-receptors
endogenous ligands. ␮-, ␦- and ␬-opioid receptors all share showed an anxiolytic phenotype suggesting that, in
an inhibitory influence upon neuronal activity attributable distinction to ␦-receptors, ␮-opioid receptor activation
to their suppression and enhancement of Ca2+ - and K+ - may enhance anxiety (Filliol et al., 2000; Kieffer and
currents, respectively, and to their stimulation of adenylyl Gavériaux-Ruff, 2002). One possible explanation may be
cyclase (Kachaturian et al., 1985; Herz et al., 1993; Mansour relief of the tonic, inhibitory influence of ␮-opioid recep-
et al., 1995; Satoh and Minami, 1995; Millan, 2002a). Like tors upon limbic populations of GABAergic interneurones
stress-responsive pools of opioid peptides, multiple opi- (Cohen et al., 1992; Sugita and North, 1993; Millan, 2002a).
oid receptors are highly (though differentially) expressed Further, in contrast to studies mentioned above (Millan and
throughout corticolimbic structures (Kachaturian et al., Duka, 1981; Ågmo et al., 1995; Tsuda et al., 1996), em-
1985; Millan, 1986, 1990, 2002a; Herz et al., 1993; Mansour ploying non-conflict paradigms, other investigations have
et al., 1995; Martin-Schild et al., 1999; Kozicz et al., 2002). found that modest doses (sufficient to block ␮-receptors) of
They fulfil contrasting roles in the control of mood, the opioid antagonists, naloxone and naltrexone, potentiate
and display reciprocal interactions with monoaminergic anxiolytic properties of BZPs (Belzung and Agmo, 1997;
156 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Belzung and Dubreuil, 1998; Frussa-Filho et al., 1999; Slowe et al., 2001; Millan, 2002a). In the amygdala, or-
Silva and Frussa-Filho, 2002). These observations are con- phaninFQ suppresses the release of both glutamate and
ceivably related to anecdotal reports of nervousness and GABA (Meis and Pape, 2001). Further, in addition to corti-
dysphoria upon a first administration of morphine to naive, colimbic structures, expression of ORL1 receptors in the LC
human subjects. However, naloxone and naltrexone en- and raphe nuclei provides a substrate for the pronounced
hance or fail to affect anxious states in man (Grevert and inhibitory influence of OFQ upon the activity of noradrener-
Goldstein, 1978; Millan and Duka, 1981), pharmacological gic and serotonergic pathways, respectively (Connor et al.,
blockade of ␮-opioid receptors alone is not associated with 1999; Calo et al., 2000; Sbrenna et al., 2000; Schlicker and
anxiolytic properties in rodents or primates, and mice lack- Morari, 2000; Millan, 2002a; Rominger et al., 2002; Marti
ing ␤-endorphin showed no changes in anxious behaviour et al., 2003).
(Millan and Duka, 1981; Nobre et al., 2000; Sante et al., Upon central administration—to circumvent problems of
2000; Kieffer and Gavériaux-Ruff, 2002). Further, ␮-opioid limited CNS penetration—both orphaninFQ and a synthetic
analgesics (such as morphine) are reliably active in a con- ORL1 receptor agonist, Ro64-6198, elicited a coherent pat-
ditioned emotional response procedure in rats, as well as tern of anxiolytic actions in various experimental models,
certain other models of anxiolytic activity, while their calm- including conflict procedures, at doses which did not perturb
ing and mood-improving properties are quintessential to the (though see Higgins et al., 2001) motor or cognitive func-
unique quality of long-term pain-relief which they afford tion (Jenck et al., 1997; Mogil and Pasternak, 2001; Carey
(Millan and Duka, 1981; Millan, 1990, 1999, 2002a; Herz and Varty, 2002). (Further, orphaninFQ elicits hyperalge-
et al., 1993; Filliol et al., 2000; Berrandero and Maldonado, sia rather than antinociception via actions in supraspinal
2002; De Boer and Koolhaas, 2003; Sanchez, 2003). Thus, structures (Millan, 2002a)). Very recently, in line with these
the possibly bimodal (anxiolytic and anxiogenic) role of findings, i.c.v. administration of orphaninFQ was reported
␮-opioid receptors in the control of anxious states remains to display anxiolytic properties in the VCT (Table 19) (Lu
puzzling and deserves further investigation. et al., 2002). Amongst other cerebral regions, ORL1 recep-
The profound analgesic properties of opioid agonists, tors in the amygdala and the PAG are involved in transduc-
most strikingly at ␮-opioid receptors (Millan, 1999, 2002a), ing the anxiolytic actions of orphaninFQ (Jenck et al., 1997,
might be expected to compromise evaluation of their actions 2000; Griebel et al., 1999b; Kyuhou and Gemba, 1999;
in models incorporating noxious stimuli. In fact, ␮-opioid Mogil and Pasternak, 2001; Millan, 2002a; Sajdyk et al.,
agonists show negligible activity in the VCT and only weak 2002c). Underscoring the tonic role of OFQ in the control
and variable actions in the Geller–Seifter conflict model of of anxious states, ORL1 receptor antagonists exhibit anx-
anxiolytic activity, underpinning the argument that antinoci- iogenic properties. Further, though ORL1 knock-out mice
ceptive actions are not, in principle, sufficient for disinhi- show few changes, animals lacking the OFQ gene reveal
bition of punished responses (Table 19) (Vogel et al., 1971; an increased vulnerability to stress, impaired adaptation to
Millan and Duka, 1981; Pollard and Howard, 1989; Heilig anxiogenic situations and elevated circulating levels of glu-
et al., 1992; Millan and Brocco, 2003). Indeed, though the cocorticoids (Köster et al., 1999; Reinscheid et al., 2000;
danger of “false positives” should systematically be exam- Reinscheid and Civelli, 2002; Sajdyk et al., 2002c).
ined for all drugs possessing pronounced analgesic effects, Confounding assessment of the pathophysiological role(s)
irrespective of pharmacological class, there appears to be a of orphaninFQ, the pre-pro-orphaninFQ precursor from
general dissociation between antinociceptive properties and which it is spliced gives rise to several, further bioactive
anxiolytic actions in punishment-based conflict procedures sequences, notably nocistatin (Calo et al., 2000; Mogil and
such as the VCT and other behavioural models (Millan and Pasternak, 2001; Millan, 2002a). The latter sequence can
Brocco, 2003). interfere (via actions at a separate site) with the effects of
orphaninFQ and was recently reported to exert anxiogenic
9.1.9.2. OrphaninFQ and ORL1 receptors. Recently, con- properties in mice (De Lima et al., 2002; Gavioli et al.,
siderable attention has been devoted to the role of a novel 2002b). Further, it was recently suggested that orphaninFQ
member of the opioid family, orphaninFQ (OFQ, also re- and nocistatin interact with GABAergic systems in the con-
ferred to as nociceptin), in the modulation of anxious states trol of anxious states (De Lima et al., 2002). Whether, by
(Bertorelli et al., 2000; Mogil and Pasternak, 2001). Its analogy to ORL1 receptor agonists, nocistatin antagonists
(naloxone-insensitive) receptor, “ORL1 ”, shares the cou- possess anxiolytic properties remains to be established. The
pling profile of its ␮-, ␦- and ␬-opioid receptor counterparts interrelationship between OFQ and nocistatin in the modu-
to diverse intracellular signals and similarly exerts an in- lation of anxious behaviour would clearly be conducive to
hibitory influence upon neuronal activity (Section 9.1.9.1) further study employing the VCT.
(Hawes et al., 2000a). Like ORL1 receptors, orphaninFQ is
widely distributed in virtually all centres involved in the eti- 9.1.9.3. Melanocortins and related peptides. Both in the
ology and modulation of anxious states: the hippocampus, pituitary and in perikarya localized in the arcuate hypotha-
the amygdala, the hypothalamus, the nucleus accumbens, the lamus and nucleus tractus solitarious, posttranslational
PAG, the cortex and the lateral septum (Neal et al., 1999a,b; cleavage of the pro-opiomelanocortin precursor to generate
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 157

␤-endorphin (Section 9.1.9.1) yields several other bioactive MC4 receptor antagonist reversed the anxiogenic effects of
neuropeptides. These include the melanocortins: ACTH, pre-exposure to stress in the light–dark box procedure in
␣-melanocyte-stimulating hormone (MSH)—the N-terminal mice and a plus-maze in rats (Chaki et al., 2003a).
13 amino acids fragment of ACTH-␤-MSH, and ␥-MSH Interestingly, an endogenous antagonist (inverse ago-
(Bagnol et al., 1999; Adan and Gispen, 2000; MacNeil et al., nist) of MC4 and MC3 receptors has been discovered
2002; Millan, 2002b). Melanocortins exert their actions via (agouti-related peptide) which, like melanocortins, is syn-
five classes of receptor, all of which—in contrast to opi- thetised in the arcuate nucleus and shares the localization
oid receptors—couple positively via Gs to adenylyl cyclase of MC4 and MC3 receptors in the bed nucleus of the stria
(Wikberg, 1999; Adan and Gispen, 2000). ␣-MSH has high terminalis and the lateral septum. Its physiological and
affinity for each subtype of melanocortin receptor except potential therapeutic significance to the control of anxious
MC2 , while ACTH is the only melanocortin to show high states awaits clarification (Haskell-Luevano et al., 1999;
affinity for all subtypes including MC2 receptors which are Haskell-Luevano and Monck, 2001).
concentrated in the adrenal gland (Adan and Gispen, 2000). Neurones containing melanocortins also produce a 132
Like MC2 sites, MC1 receptors are poorly represented in amino acid peptide termed cocaine- and amphetamine-
the CNS, though a small population has been located in the regulated transcript (CART) which is also synthesized in
PAG (Adan and Gispen, 2000; Millan, 2002b). Modest lev- several limbic structures, including the nucleus accum-
els of MC5 receptors have been detected supraspinally but bens, the amygdala and the hippocampus (Koylu et al.,
the principle sites of relevance to the modulation of anxious 1998; Kuhar and DallVechia, 1999). The synthesis of
states are MC3 and MC4 receptors. The former, which may CART is elevated by stress, and CART markedly activates
act as autoreceptors on melanocortin-containing neurones, CRF-containing neurones in the hypothalamus and accel-
are highly expressed in various hypothalamic nuclei and erates the secretion of ACTH (Vrang et al., 1999; Stanley
they are also found in the septem, nucleus accumbens, hip- et al., 2001; Chaki et al., 2003b). Moreover, its bioactive
pocampus, amygdala, raphe and ventrotegmental area. On fragment, CART55-102 , elicited anxiety upon i.c.v. admin-
the other hand, MC4 sites are enriched in several hypotha- istration to mice and enhanced the electrical activity of
lamic nuclei, the amygdala, the lateral septum, the bed the LC in rats, a structure in which it is co-localized with
nucleus of the stria terminalis and the PAG, while they are adrenergic neurones (Chaki et al., 2003b).
also detectable in the LC, DRN and, despite scarce inner- These mutual anxiogenic properties of CART and
vation by melanocortin-containing fibers, the hippocampus melanocortins are all the more interesting inasmuch as they
and cortex (Mountjoy et al., 1994; Lindblom et al., 1998; likewise share the ability to inhibit appetite and to elicit hy-
Roselli-Rehfuss et al., 1993; Bagnol et al., 1999; MacNeil peralgesia (Adan and Gispen, 2000; MacNeil et al., 2002;
et al., 2002; Kishi et al., 2003). In distinction to opioids, Millan, 2002b).
melanocortins may exert an excitatory influence upon Clearly, then, exploration of the interrelationships of
adrenergic neurones in the LC (René et al., 1998) and they melanocortins, CART, agouti-related peptide and ␤-endor-
have been shown to activate CRF-containing neurones in phin in the modulation of anxious states is likely to provide
the parvocellular division of the paraventricular nucleus of fertile territory for research over the coming years.
the hypothalamus (MacNeil et al., 2002).
By analogy to ␤-endorphin (Section 9.1.9.1), hypophy- 9.1.10. Melanin Concentrating Hormone
seal and supraspinal pools of melanocortins are responsive Melanin Concentrating Hormone (MCH) is a cyclic
to stress (Millan, 1986, 2002a; Adan and Gispen, 2000). neuropeptide of 19 amino acids which is intimately in-
Further, via actions in the PAG and other cerebral struc- volved in the control of energy balance: it is synthetized
tures, both ACTH and ␣-MSH elicit grooming behaviour in the lateral hypothalamus—and zona incerta (Nahon,
in rodents—a characteristic stress-related response (Section 1994; Shimada et al., 1998). MCH-containing projections
4.2.1)—while antibodies directed against ACTH suppress from these regions extensively innervate corticolimbic
the grooming response to novelty Dunn et al. (1979), Adan regions of the brain, including the FCX, hippocampus,
et al. (1999), Adan and Gispen (2000). More direct evidence septum, PAG and VTA (Bittencourt et al., 1992; Casatti
for an anxiogenic role of ACTH and ␣-MSH is provided by et al., 2002; Dallvechia-Adams et al., 2002). These cen-
their suppression of social interaction in rodents, their in- tral MCH-containing pathways, which are inhibitory to the
duction of anxiety upon infusion into the ventromedial nu- hypothalamo-corticotropic axis, are subject to modulation
cleus and, most pertinently, their robust and dose-dependent both by CRF and by stress, concordant with a role in the
reduction of responses in the VCT upon i.c.v. administra- control of emotionality (Nahon, 1994; Bluet-Pajot et al.,
tion (Table 19) (File and Clarke, 1980; Corda et al., 1990; 1995; Griffond and Baker, 2002).
Gonzalez et al., 1996). Though a role of MC3 receptors MCH exerts its actions via MCH1 receptors (originally
should not be discounted, it is likely that MC4 sites are im- designated the orphan receptor, SLC-1/GPR24) and MCH2
plicated in these actions. Thus, in “preliminary studies”, a (SLT/S643b) receptors, both of which couple via Gq to phos-
synthetic MC4 agonist was similarly found to display anxio- pholipase C and intracellular pools of Ca2+ . Further, MCH1
genic effects in the VCT while a selective, non-peptidergic receptors are inhibitory to adenylyl cyclase, depress the
158 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

operation of VDCCs and activate K+ -channels via Gi/o , of angiotensin II are peripheral, the entire multi-enzyme
though their major mode(s) of transduction remain(s) to renin–angiotensin system is found in the brain. Indeed,
be specified (Hawes et al., 2000b; Audinot et al., 2001a,b; though the percentage of angiotensinogen synthesized by
Rodriguez et al., 2001; Griffond and Baker, 2002; Gao and neurones as compared to astrocytes may be modest, it
Van Der Pol, 2002). The distribution of central MCH1 recep- has been detected in the amygdala, hippocampus, septum,
tors parallels that of MCH itself, with notably high concen- hypothalamus and cortex (Bunnemann et al., 1993; Gard,
trations in the cortex (entorhinal, frontal, cingulate and tem- 2002; Morimoto and Sigmund, 2002). Moreover, AT1 recep-
poral), the hippocampus, lateral septum, nucleus accumbens, tors, which stimulate phospholipase C and inhibit adenylyl
amygdala, hypothalamus and the mammillary nuclei (Her- cyclase via recruitment of Gq and Gi , respectively, are
vieu et al., 2000; Saito et al., 2001; Borowski et al., 2002). localized in the FCX, hippocampus, hypothalamus, PAG,
The occurrence of MCH1 sites in the LC and DRN implies amygdala and LC (Wright and Harding, 1995; Lenkei et al.,
a role for MCH in the modulation of ascending noradren- 1997; Allen et al., 1998; De Gasparo et al., 2000). AT2 sites
ergic and serotonergic transmission, a possibility requiring (for which coupling mechanisms are poorly defined, though
direct neurochemical and electrophysiological verification. they are inhibitory to neuronal activity) have been detected
Mice genetically devoid of MCH or MCH1 receptors in the entorhinal cortex, amygdala, PAG, raphe nuclei and
display, predictably, alterations in appetite and energy LC (Wright and Harding, 1995; Lenkei et al., 1997; De
metabolism, but potential changes in their emotionality, re- Gasparo et al., 2000).
sponse to stress and anxious behaviour do not appear to have Employing models of exploratory behaviour, data have
been examined (Shimada et al., 1998; Chen et al., 2002a; been acquired indicating that the selective AT1 receptor an-
Marsh et al., 2002). Further, there are conflicting reports con- tagonist, losartan, possesses anxiolytic properties (Barnes
cerning the influence of central administration of MCH upon et al., 1991; Okuyama et al., 1999b; Gard, 2002). Such
anxious behaviour in rats. These disparities conceivably re- findings are congruent with reports of anxiolytic actions
flect the expression of anxiolytic as compared to anxiogenic of angiotensin converting enzyme inhibitors, and with anx-
actions in different CNS regions, such as the amygdala, the iomimetic effects of angiotensin II itself in several models,
hippocampus and the hypothalamus, though the respective including the VCT, though caution must be exercised in in-
roles of these structures remain to be clarified (Gonzalez terpreting the latter action in view of the role of angiotensin
et al., 1996b; Monzon and De Barioglio, 1999; Monzon in the control of water appetite (Costall et al., 1990; Georgiev
et al., 2001). Selective, non-peptidergic antagonists at MCH1 et al., 1990; Kaiser et al., 1992; Tsuda et al., 1992; Costall
sites have recently been disclosed and shown to possess and Naylor, 1997; Okuyama et al., 1999a; Gard, 2002). An-
anxiolytic properties upon parenteral administration in a so- giotensin II also reduced defensive burying behaviour in
cial interaction procedure (Borowski et al., 2002; Takekawa mice (Tsuda et al., 1992). At variance with these observa-
et al., 2002). In view of their modest efficacy in the VCT tions, upon acute administration, in a time-dependent fash-
(Table 19 and Fig. 9) (Brocco and Millan, unpublished ob- ion, angiotensin II elicited a delayed “rebound” reduction
servations), MCH1 sites appear a promising novel target for in anxiety, and not all studies with AT1 antagonists have
the management of anxious and other psychiatric disorders. observed anxiolytic effects (Gard, 2002). The role of an-
MCH2 receptors share the localization of their MCH1 giotensin II in the control of anxious states may, thus, be
counterparts in the cortex (entorhinal and temporal), amyg- complex, reflecting multiple and contrasting roles of AT1
dala and hippocampus of primates, including man. How- and AT2 receptors (Shepherd et al., 1996; Okuyama et al.,
ever, functional MCH2 receptors appear to be absent from 1999b; Gard, 2002). Indeed, mice lacking AT2 receptors
rats and mice—which possess inactive, truncated forms— showed heightened anxiety-like behaviour, possibly medi-
invalidating rodents in the exploration of their functional ated (indirectly) via actions of NA at ␣1 -adrenoceptors in
significance (Griffond and Baker, 2002; Tan et al., 2002b). the amygdala (Section 4.1.5) (Ichiki et al., 1995; Walther
Thus, any potential role of MCH2 sites in the modulation of et al., 1999; Okuyama et al., 1999b).
anxious states will be difficult to discern. Finally, AT4 receptors—for which angiotensin IV (derived
from angiotensin II) may be the major endogenous ligand—
9.1.11. Angiotensin/AT receptors are located in the amygdala, PAG and hippocampus, sug-
Angiotensin II (an octomer) is derived from angiotensin gesting that they may also be involved in the modulation of
I by an action of angiotensin converting enzyme, while an- anxious states (Wright et al., 1995; Allen et al., 1998; Chai
giotensin I itself is cleaved from angiotensinogen by renin et al., 2000; Gard, 2002; Turner and Hooper, 2002).
(Morimoto and Sigmund, 2002; Turner and Hooper, 2002).
The familiar role of angiotensin II in the control of cardio- 9.2. Peptides/receptors as yet to be evaluated in the VCT
vascular function, water homeostasis and sodium reabsorp-
tion reflects actions at AT1␣ , AT1␤ and AT2 receptors—each 9.2.1. Vasoactive intestinal peptide and pituitary adenlyl
of which is differentially implicated in its functional effects cyclase activating peptide
(Allen et al., 1998; Morimoto and Sigmund, 2002; Turner Vasoactive intestinal peptide and pituitary adenyl cy-
and Hooper, 2002). Though the primary loci of action clase activating peptide (PACAP)—which belong to a
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 159

family of peptides including glucagon (Section 9.1.6)—are respect, recently-discovered non-peptidergic antagonists
structurally-related peptides of 28 and 38 amino acids in at sst2 receptors should prove instructive tools (Crider,
length, respectively. Both share high affinity for so-called 2002). Despite anatomical studies indicating their broad
“VPAC1 ” and “VPAC2 ” receptors. These sites, which are localization in diverse corticolimbic regions, the functional
positively coupled to adenylyl cyclase via Gs, are localized significance of other sst receptor types in the modulation
in the hippocampus, amygdala and cortex, congruent with of anxious states remains obscure (Csaba and Dournaud,
an, as yet unexplored, role in the modulation of anxiety 2001).
(Vertongen et al., 1997; Dumont et al., 2000; Vaudry et al.,
2000; Laburthe and Couvineau, 2002). However, PACAP 9.2.3. Natriuretic peptides
also potently activates excitatory “PAC1 ” receptors which Atrial natriuretic peptide (28 residues in length), together
stimulate phospholipase C via Gq. Like PACAP, PAC1 with brain natriuretic peptide (26 amino acids) and C-type
receptors are concentrated in the hippocampus, septum, natriuretic peptide (34 amino acids), play important roles
amygdala, hypothalamus, hippocampus and cortex, while in the central and peripheral regulation of endocrine and
they have also been found in the LC and DRN, from which autonomic function (Gutkowska and Nemer, 1989; Imura
noradrenergic and serotonergic pathways, respectively, arise et al., 1992).
(Hashimoto et al., 1996; Piggins et al., 1996; Laburthe and The high levels of atrial natriuretic peptide and “ANP-A”
Couvineau, 2002). receptors in the entorhinal cortex, FCX, hippocampus,
These anatomical studies provide a framework for a amygdala, septum, hypothalamus, as well as the DRN, sug-
potential influence of PACAP upon anxious states, pos- gest a role in the control of emotionality (Langub et al.,
sibly involving a facilitatory influence upon the activity 1995). This supposition is supported by increases in its se-
of monoaminergic projections. In line with this notion, cretion in patients undergoing panic attacks (Kellner et al.,
knock-out mice in which PAC1 receptors were deleted (and, 1997). Furthermore, in human subjects, atrial natriuretic
possibly, mice lacking PACAP itself) display a reduction peptide concurrently inhibited the anxiogenic response to
in anxious behaviour: this phenotype apparently reflects a CCK4 and suppressed its induction of ACTH secretion
loss of tonic activity at PAC1 sites in the LC, amygdala and (Kellner et al., 2002). In line with these findings, both the
septum (Sauvage et al., 2000; Hashimoto et al., 2001; Otto i.c.v. and intra-amygdala administration of atrial natriuretic
et al., 2001). Regrettably, however, pharmacological studies peptide and brain natriuretic peptide elicited anxiolytic
have not, as yet, been undertaken concerning the role of effects in rats, while isartin, a putative, endogenous in-
these peptides in the control of anxious states. hibitor of ANP-A receptors exerted anxiogenic properties
(Bidzseranova et al., 1992; Glover et al., 1995; Bhatta-
9.2.2. Somatostatin charyya et al., 1996; Biro et al., 1996; Ströhle et al., 1997).
Somatostatin 14 and the N-terminal extended somatastin Supporting a role in the control of behaviour under condi-
28, together with various cortistatins, interact with at least tions of anxiety, both atrial and brain natriuretic peptides
five classes of receptor (Dournaud et al., 1998, 2000; facilitate fear-motivated learning in rats (Bidzseranova
Hervieu and Emson, 1998; Csaba and Dournaud, 2001). et al., 1992; Biro et al., 1996; Ströhle et al., 1997). Though
Like somatostatin, sst2 receptors (which inhibit adenylyl GABA does not appear to participate in the modulation of
cyclase via Gi ) are predominantly localized in corticolim- anxiety by atrial natriuretic peptide, monoaminergic mech-
bic structures, including the PAG, amygdala, cortex and anisms may be involved (Vatta et al., 1994; Bhattacharyya
hippocampus. Further, they are also located in the LC, et al., 1996; Biro et al., 1996; Griebel, 1999a). In contrast
coinciding with the modulatory influence of somatostatin to these data, atrial natriuretic peptide failed to modify pun-
via sst2 —and other—receptor subtypes upon noradrener- ished responses in a Geller–Conflict model (Heilig et al.,
gic neurones (Epelbaum et al., 1994; Chessell et al., 1996; 1992). Additional studies of the potential actions of atrial
Dournaud et al., 1996, 1998, 2000; Selmer et al., 2000; natriuretic peptide would, thus, be informative to resolve its
Csaba and Dournaud, 2001; Pittaluga et al., 2001). Inter- role in the modulation of anxious states.
estingly, numerous somatostatin-containing neurones in the C-type natriuretic peptide, which is concentrated in limbic
amygdala, though devoid of CCK and vasoactive intestinal regions such as the hippocampus (Langub et al., 1995), is
polypeptide, contain GABA and NPY, transmitters inti- a preferential ligand at “ANP-B” sites (Koller et al., 1991).
mately involved in the modulation of anxious states and Though the functional role of C-type natriuretic peptide is
behavioural performance in the VCT (Sections 2.2.2 and less well-defined than that of atrial natriuretic peptide, there
9.1.5, respectively) (McDonald and Mascagni, 2002). Mice is evidence that it promotes the CRF-induced secretion of
with a null mutation for sst2 receptors displayed a pro- ACTH in man and that it concomitantly elicits anxiety-like
nounced enhancement in emotional reactivity, in anxious symptoms, actions opposite to those of ANP (Kellner et al.,
behaviour and in their response to stress (Viollet et al., 1997, 2002; Montkowski et al., 1998). Interestingly, then,
2000). These observations encourage the use of the VCT its central administration to rodents proved to be anxio-
in the examination of a potential suppressive influence of genic, an action which possibly reflects recruitment of CRF
somatostatin via sst2 sites upon anxious behaviour. In this (Jahn et al., 1997, 2001; Montkowski et al., 1998). Inasmuch
160 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

as CRF controls behaviour in the VCT, this procedure would show no alterations in anxious behaviour, populations with a
evidently be appropriate for further characterization of the null mutation for BB2 receptors display an increase in social
influence of C-type natriuretic peptide upon anxious be- behaviour: this was suggested to reflect an enhancement in
haviour. activity at GABAA receptors though it remains unclear as to
how this is brought about (Wada et al., 1997; Merali et al.,
9.2.4. Bombesin-related peptides 1998; Yamada et al., 2000b, 2001, 2002b). Though only
The bombesin-like family of peptides incorporates the peptidergic antagonists for BB2 sites are currently available
mammalian neuromedin B and gastrin-releasing peptide. (Jensen and Coy, 1991), the foregoing observations incite
They are 32 and 27 amino acids in length, respectively, but it further exploration of their role in the modulation of anxious
is probably their 10 amino acid N-terminal sequences which states (Yamada et al., 2002a).
constitute their bioactive fragments (Battey and Wada, 1991; BB3 receptors are most prevalent in the hypothalamus,
Kroog et al., 1995; Ohki-Hamazaki, 2000). Mammalian neu- but they are also found in the amygdala (Ohki-Hamazaki
romedin B and gastrin-releasing peptide are preferential ag- et al., 1997; Yamada et al., 2000a). In addition to metabolic
onists of “BB1 ” and “BB2 ” receptors, respectively, though changes, mice in which BB3 receptors have been genetically
they also display modest affinity for a third “BB3 ” site for ablated show alterations in non-aggressive social behaviour,
which an endogenous ligand remains to be isolated (Benya and they present an anxiolytic profile in both the plus-maze
et al., 1995; Kroog et al., 1995; Ohki-Hamazaki, 2000). and light-box procedures (Yamada et al., 2000a). Thus, once
BB1 , BB2 and BB3 receptors all engage phospholipase C via discovered, it would be of interest to examine potential anx-
Gq/11 , though they may also recruit (independently of phos- iolytic properties of selective antagonists at these sites.
pholipase C) other intracellular signals, such as adenylyl Collectively, then, the above data suggest that all three
cyclase and mitogen-activated protein kinase (Benya et al., subtypes of BB receptor are (perhaps differentially) involved
1995; Kroog et al., 1995; Ohki-Hamazaki, 2000). in the control of emotionality and anxious states.
Neuromedin B and BB1 receptors have been visualized in
the lateral septum, frontal cortex, PAG, hippocampus, amyg-
dala, LC and raphe nuclei (Wada et al., 1990, 1991; Ohki- 10. Melatonin
Hamazaki, 2000). The latter observation has attracted par-
ticular attention inasmuch as neuromedin B and bombesin 10.1. Coupling and localization
exert an excitatory influence upon raphe-localized seroton-
ergic neurones via the activation of BB1 receptors (Pinnock Melatonin, which is synthesized in the pineal gland,
and Woodruff, 1991; Pinnock et al., 1994), Further, in mice plays an important chronobiotic role in the control of diur-
lacking BB1 receptors, the facilitatory influence of stress nal activity, of seasonal rhythms and of sleep, all of which
upon serotonergic neurones is abrogated (Yamano et al., are perturbed in anxious states (Borjigin et al., 1999). Its
2002). Though these mice did not exhibit any modification secretion is enhanced by activation of stress-sensitive, sym-
of behaviour in the plus-maze, they showed alterations in pathetic noradrenergic input (Seggie et al., 1985; Joshi et al.,
emotionality in other procedures, heightened susceptibility 1986; Golombek et al., 1996). Melatonin acts via “MLT1 ”
to stress, and a down-regulation of 5-HT1A receptors both (Mel1A ) and “MLT2 ” (Mel1B ) receptors. Both are coupled
pre- and postsynaptically to serotonergic pathways (Yamada via Gi/o to inhibition of adenylyl cyclase. MLT1 receptors
et al., 2002a,b,c; Yamano et al., 2002). Further, stress induces also regulate the activity of phopholipase A via Gq, while ac-
the release of bombesin-related peptides in the amygdala and tivation of MLT2 receptors diminishes intracellular levels of
hypothalamus (Kent et al., 1998; Merali et al., 1998), while cyclic guanidine monophosphate (Reppert et al., 1996; Von
neuromedin B interacts with the anxiogenic peptide, CRF, Gall et al., 2002; Witt-Enderby et al., 2003). The cerebral
in the modulation of appetite and satiety (Ohki-Hamazaki, pattern of distribution of melatonin receptors is strikingly
2000). Thus, inasmuch as the first non-peptidergic antag- species-specific. Though MLT1 sites have been visualized
onists at BB1 receptors have been described (Ryan et al., in the cortex, the amygdala, septum and other limbic struc-
1999), it would be instructive to evaluate their actions in the tures of various species, such observations have not to date,
VCT and other procedures of potential anxiolytic properties. been made in mice, rats (or man). Nevertheless, MLT2 sites
Though the distribution of gastrin-releasing peptide and have been identified in the hippocampus and FCX of rodents
BB2 receptors in the CNS is distinct to that of neuromedin B and primates, consistent with a role in the control of anx-
and BB1 receptors, like the latter, they are expressed in the ious states (Reppert et al., 1995; Mazzucchelli et al., 1996;
hippocampus, the hypothalamus, the amygdala, the PAG and Vanecek, 1998; Wan et al., 1999; Witt-Enderby et al., 2003).
the LC (Wada et al., 1990, 1991; Battey and Wada, 1991).
Stress promotes the release of gastrin-releasing peptide in 10.2. Control of anxious states
the amygdala. Therein, following release from glutamatergic
terminals, it excites GABAergic interneurones and plays a Though melatonin was not effective in a rat VCT over
role in the modulation of fear-induced memory (Shumyatsky a broad dose range (Table 18) (Brocco et al., unpublished
et al., 2002). Though mice lacking gastrin-releasing peptide observations), it exerted anxiolytic actions (preventable
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 161

by a specific MLT1/2 receptor antagonist) in other mod- and McGaugh, 1998; Chen and Qiu, 2001; Gass et al.,
els, primarily those based on exploratory behaviour 2001; Haller, 2001; Korte, 2001; Makino et al., 2002b).
(Guardiola-Lemaı̂tre et al., 1992; Golombek et al., 1993, The familiar genomic actions of glucocorticoids reflect
1996; Kopp et al., 1999a, 2000; Nava and Carta, 2001). The their docking to a cytosolic receptor which is then translo-
blockade by flumazenil (an antagonist at the BZP binding cated into the nucleus: therein, in conjunction with other
site on GABAA ) receptors (Section 2.2.2), of the anxiolytic transcription proteins, glucocorticoids bind to DNA and
actions of melatonin gels with other behavioural evidence modify gene expression (Joëls and Vreugdenhil, 1998; Chen
suggesting that an interaction with GABAergic mechanisms and Qiu, 2001; Kellendonk et al., 2002). The transcription
contributes to its influence upon behaviour, (Pierrefiche factors, type I mineralocorticoid receptors (MR) and type II
et al., 1993; Golombek et al., 1996; Kopp et al., 1999a,b, glucocorticoid (GR) receptors, which mediate these actions
2000; Wan et al., 1999; Wang et al., 2003). However, this of glucocorticoids, are well-represented in the CNS, wherein
has been disputed and the precise mechanisms underlying they appear to fulfil contrasting functional roles. The latter
anxiogenic effects of melatonin remain puzzling inasmuch are broadly distributed in the amygdala, hippocampus and
as MLT2 receptors exert an inhibitory control upon GABAA other corticolimbic structures (including sites involved in
receptor-mediated transmission in the hippocampus (Wan the feedback control of the hypothalamo-corticotropic axis),
et al., 1999). Further, it should be pointed out that melatonin and they are also localized in the LC and DRN. In contrast,
is not involved in the anxiolytic effects of BZPs which actu- the distribution of MR receptors is more confined in that
ally exert a suppressive influence upon melatonin secretion they are highly expressed in the hippocampus, amygdala
(Winters et al., 1991; Djeridane and Touitou, 2001). and septum (De Kloet et al., 1991; Morimoto et al., 1996;
Though clinical studies of melatonin as a pre-medication De Kloet et al., 1998; Joëls and Vreugdenhil, 1998; Gass
prior to surgery are consistent with reduced anxiety, its et al., 2001; Kellendonk et al., 2002). MR receptors dis-
sedative and hypnotic properties are clearly relevant to its play high affinity for glucocorticoids and may be virtually
influence upon mood under such conditions (Naguib and saturated under “resting” conditions, whereas low-affinity
Samarkandi, 2000). Indeed, while prescription of melatonin GR receptors are phasically occupied upon imposition of
improves the compliance of patients discontinuing BZP stressful stimuli (op. cit.).
treatment for sleep disorders, there is no evidence that this It has been proposed that chronic exposure to stress
action reflects an influence upon anxious states and there results, via mechanisms implicating glucocorticoids, in
are currently no compelling arguments to support the ther- deleterious morphological and functional changes in the
apeutic utility of melatonin in the management of anxious hippocampus, amygdala and cortex, including an impair-
disorders (Garfinkel et al., 1999; Cardinali et al., 2002). ment of synaptic transmission (notably, of “long-term
Thus, negative findings with melatonin in the VCT are not potentiation”), a (probably reversible) dendritic atrophy,
surprising and any potential utility of melatonin alone in the suspension of oligocyte proliferation and a reduction in
control of anxious states may primarily reflect its chronobi- neurogenesis, effects contributing to the emotional and
otic role in normalizing perturbed circadian (and seasonal) cognitive deficits of persistent anxious and depressive
rhythms rather than anxiolytic properties per se. Indeed, in states. Changes in the hippocampus involve a variety of
certain paradigms, anxiolytic effects of melatonin have only transmitters and reflect effects mediated both locally and
been seen upon night-time administration (Golombek et al., “upstream” in other structures such as the amygdala with
1993). On the other hand, it is conceivable that the associa- which it is interlinked (Akirav and Richter-Levin, 1999;
tion of agonist properties at melatonin receptors with other Gould and Tanapat, 1999; Sherwood-Brown et al., 1999;
components of activity (either in a single molecule or as an Kim et al., 2001; Alfarez et al., 2002; Kim and Diamond,
adjunctive agent) may be beneficial in the treatment of clin- 2002; Makino et al., 2002a,b; Venero et al., 2002).
ical anxiety disorders. On an opposite time-scale, several organs, including the
CNS, can respond rapidly (within seconds or minutes) to
glucocorticoids indicative of non-genomic actions elicited
11. Glucocorticoids independently of gene transcription (Borski, 2000; Sapolsky,
2000a,b; Sapolsky et al., 2000; Chen and Qiu, 2001; Makara
11.1. Genomic and non-genomic actions in the CNS and Haller, 2001; Kellendonk et al., 2002). These effects
may require engagement of a membrane-bound receptor
Following their rapid secretion by the adrenal cortex into structurally similar to GR sites. An inhibitory influence of
the systemic circulation in response to stress, glucocorticoids glucocorticoids upon neuronal excitability likely reflects
mobilize energy reserves by attenuating cellular glucose up- a diminution in the activity of adenylyl cyclase and their
take and inducing hepatic gluconeogenesis. More recently, opening of K+ -channels, though, based on studies effected
numerous studies have yielded convincing evidence that, in the hippocampus and cortex, modulation of Ca2+ -influx
under conditions of stress and fear, glucocorticoids exert a (via VDCCs) and of intracellular Ca2+ -concentrations can
pronounced influence upon cerebral function and plastic- underlie excitatory actions of glucocorticoids. Thus, the
ity via both long-term and more short-term actions (Cahill effects of glucocorticoids are complex, and both genomic
162 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

and non-genomic mechanisms trigger tissue-dependent glucocorticoids (mediated via GR receptors) upon nucleus
alterations in cellular excitability (Sze and Iqbal, 1994; accumbens release of DA, in engaging reward mechanisms,
Ffrench-Mullen, 1995; Qiu et al., 1998; Borski, 2000; Karst counteracts aversive states and reduces vulnerability to stress
et al., 2000; Chen and Qiu, 2001). (Section 4.2.1) (Marinelli and Piazza, 2002).
It is also important to consider the temporal framework
11.2. Interactions with neurotransmitters inasmuch as the roles and mechanisms of action of gluco-
corticoids depend upon: (1) the timing of their administra-
Though the relative contribution of rapid versus long-term tion (release) relative to the stressor (before, during or af-
and genomic versus non-genomic processes to the central ter) and (2) the duration of their secretion in response to
actions of glucocorticoids remains to be clarified, there is stress. For example, it has been conjectured that a failure of
a vast literature (Haller et al., 1998; Meijer and De Kloet, the counter-regulatory glucocorticoid feedback mechanism
1998; Raber, 1998; Sapolsky, 2000a,b; Sapolsky et al., 2000; to contain excessive and protracted CRF and vasopressin re-
Korte, 2001; Makara and Haller, 2001; Kellendonk et al., lease under conditions of chronic stress contributes to the
2002) dedicated to the interaction of central pools of gluco- development of pathological anxious states (Meijer and De
corticoids with corticolimbic noradrenergic (including ␣1 Kloet, 1998; Kovacs et al., 2000; Makino et al., 2002a,b).
and ␣2 -ARs) (Stone et al., 1986, 2002; Stone and Quarter- This argument, as discussed in Sections 9.1.2 and 9.1.3, is
main, 1999; Makino et al., 2002b), dopaminergic (Piazza coherent with the major roles of CRF (and vasopressin) in
et al., 1996; Rougé-Pont et al., 1998; Sillaber et al., 1998), the induction of anxiety.
serotonergic (including 5-HT1A , 5-HT2A and 5-HT2C recep- On the other hand, upon transient exposure to stress, the
tors and 5-HT transporters) (Chaouloff, 1995; Maines et al., enhanced occupation of central GR receptors by glucocor-
1999; Cyr et al., 2001; Czyrak et al., 2002), glutamater- ticoids may reinforce the activity of raphe-derived seroton-
gic (including NMDA receptors) (Lowy et al., 1993; ergic projections in relieving their inhibition by 5-HT1A au-
Moghaddam et al., 1994) and GABAergic transmission toreceptors (Meijer and De Kloet, 1998; Man et al., 2002).
(Haller et al., 1998; Falkenstein et al., 2000; Haller, 2001). Indeed, many investigations have examined the intricate in-
This complex pattern of actions provides a diversity of terrelationship between glucocorticoids and both corticolim-
substrates for the modulation of anxious states. bic and raphe-localized populations of 5-HT1A receptor—
and other classes of 5-HT receptor. The majority of data
11.3. Control of anxious states suggest an inhibitory influence of glucocorticoids upon the
functional status of 5-HT1A receptors both pre- and post-
Studies of glucocorticoid depletors (Plihal et al., 1996; synaptic to serotonergic pathways (Chaouloff, 1995; Lopez
Murphy, 1997), of ligands selective for MR and/or GR re- et al., 1998; Stutzmann et al., 1998; Karten et al., 1999;
ceptors (Korte et al., 1995; Plihal et al., 1996; Murphy, 1997; Sibug et al., 2000; Wissink et al., 2000; Gur et al., 2001;
Korte, 2001), of transgenic mice expressing an antisense se- Czyrak et al., 2002; Farisse et al., 2000; Man et al., 2002;
quence directed against GR sites (Montkowski et al., 1995; Mueller and Beck, 2000; Müller et al., 2002). Though the
Steckler and Holsboer, 1999; Müller et al., 2002), and of precise nature of these actions remain both confusing and
the behaviour of mice in which the gene encoding central controversial, presumably reflecting such a reduction in ac-
GR receptors has been impaired, deleted or over-expressed tivity at 5-HT1A sites, glucocorticoids interfere with anxi-
(Rochford et al., 1997; Tronche et al., 1999; Gass et al., 2001; olytic properties of the 5-HT1A receptor partial agonist, bus-
Oitzl et al., 2001; Clément et al., 2002; Müller et al., 2002), pirone (Haller et al., 1998, 2001).
indicate that short- or long-term interference with central ac- Glucocorticoids probably also intervene in the interfer-
tions of glucocorticoids generally alleviates anxious states ence by stress with the anxiolytic actions of BZPs (Haller
and improves mood, whereas a reinforcement in their ac- and Halasz, 2000). Though anxiolytic properties of CRF1
tions results in their aggravation. These actions are largely receptor antagonists are, in contrast, more pronounced un-
mediated independently of adrenal corticosterone secretion. der conditions of stress (Section 9.1.2), this observation
However, not all observations conform to this schema, al- may be explained by an induction of CRF synthesis and
terations in cognitive function complicate interpretation of release rather than actions of glucorticocoids (Steckler and
data, most studies have employed models of spontaneous, Holsboer, 1999). Interestingly, an augmentation of CRF gene
exploratory behaviour rather than conflict paradigms, and expression in the amygdala and adjacent bed nucleus of
contrasting functional roles of MR and GR receptors should the stria terminalis by chronic administration of glucocor-
be borne in mind (Korte et al., 1995; Montkowski et al., ticoids may participate in their long-term exacerbation of
1995; Roozendaal et al., 1996; Rochford et al., 1997; Ströhle anxious states (Schulkin et al., 1998; Shepard et al., 2000;
et al., 1998; Tronche et al., 1999; Calvo and Volosin, 2000; Gass et al., 2001; Makino et al., 1999, 2002a). Similarly,
Gass et al., 2001; Korte, 2001; Clément et al., 2002; Gold in mice overexpressing GR receptors, the amplification of
and Chrousos, 2002; Kellendonk et al., 2002; Makino et al., CRF gene expression in the amygdala may contribute to
2002a,b). For example, it has been argued—on the basis of their anxious phenotype (Wei et al., 2001; Müller et al.,
pharmacological studies—that the facilitatory influence of 2002).
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 163

In view of the responsiveness of the VCT to 5-HT1A re- et al., 2001; Wise et al., 2001a,b; Vasudevan et al.,
ceptor ligands, BZPs and CRF1 antagonists (Sections 2.2.2, 2002).
4.3.2 and 9.1.2), its stressful nature, and the involvement However, estrogen also elicits rapid (over seconds)
of glucocorticoids in the emotional and cognitive response changes in neuronal activity reflecting non-genomic actions.
to conditioned fear (Korte, 2001), it would be well-adapted Though still poorly understood, they involve: alterations
to further exploration of the modulatory influence of gluco- in the activity of G-protein receptor-coupled transduc-
cortocoids upon anxious states and the actions of anxiolytic tion cascades, diverse protein kinases, adenylyl cyclase
agents. and nitric oxide; modulation of current flux through ion
channels, and interactions with neurotrophins (Falkenstein
11.4. The complex role of glucocorticoids et al., 2000; Belcher and Zsarnovszky, 2001; Driggers and
Segars, 2002; Littleton-Kearney et al., 2002; Segars and
Despite the pivotal importance of glucocorticoids in in- Driggers, 2002). These actions—which can evoke further
tegrating the peripheral and central, acute-and long-term downstream changes in cellular gene expression—may,
response to stress—including its emotional dimension— at least partially, be mediated by membrane-localized re-
remarkably little is known concerning their precise influence ceptors closely-related to intracellular ER-␣ and ER-␤
upon anxious states. Similarly, comparatively little informa- sites, though other mechanisms probably also intervene
tion is available concerning the relevance of glucorticocoids (Kelly and Wagner, 1999; Razandi et al., 1999; Nadal
to the management of anxious disorders by BZPs (Section et al., 2000; Belcher and Zsarnovszky, 2001; Wise et al.,
2.2.2), antidepressants (Section 4.4.2) and other classes of 2001a,b). As an additional complication, it should be men-
psychotropic agent. In this context, it should be recalled tioned that ER receptors are susceptible to activation by
that anxiety itself is accompanied by an enhancement in the phosphorylation, providing a potential mechanism for their
activity of the hypothalamo-corticotropic axis, underlining estrogen-independent recruitment by a variety of intracel-
the reciprocal interrelationship between glucocorticoid se- lular signals (Cenni and Picard, 1999; Litttleton-Kearney
cretion and anxious states (Chaouloff, 1995; Herman, 1999; et al., 2002; Segars and Driggers, 2002).
Sibug et al., 2000; Makino et al., 2002a,b). In light of the
crucial pathophysiological role of glucocorticoids in the 12.2. Control of anxious states
central and peripheral response to stress, there is a need for
additional study of their role in the modulation of anxious The following observations lend credence to the view that
states. the central effects of estrogen are involved in the induction
and modulation of anxious states.
First, both rapid and longer-term actions of estrogen are
12. Estrogen expressed in the hippocampus, septum, amygdala, PAG,
frontal cortex and other corticolimbic regions known to be
12.1. Genomic and non-genomic actions in the CNS rich in (differentially-distributed) ER-␣ and ER-␤ receptors.
For example, both are found in the hypothalamus, amyg-
Over recent years, it has become apparent that the gonadal dala and LC, while ER-␣ sites predominate in the PAG
steroid hormone, 17-␤-estradiol (or estrogen), exerts a com- and ER-␤ sites are prevalent in the cortex, hippocampus
plex, pleiotropic influence upon the development and func- and DRN (Kuiper et al., 1997; Shughrue et al., 1997; Van
tion of many tissue types other than those of the reproductive Amelsvoort et al., 2001).
axis (Falkenstein et al., 2000; Driggers and Segars, 2002; Second, central actions of estrogens can profoundly
Littleton-Kearney et al., 2002; Carrasco and Van de Kar, modify mood, including depressive episodes (Halbreich,
2003). The CNS is no exception in this regard inasmuch as 1997; Van Amelsvoort et al., 2001; Brinton, 2002; Littleton-
estrogen exerts pronounced effects upon neurones and, un- Kearney et al., 2002). Suggestive of an influence upon anx-
der certain conditions, astrocytes and microglia (Belcher and ious states, acute administration of estrogen displayed an
Zsarnovszky, 2001; Koenig, 2001; Van Amelsvoort et al., anxiolytic profile in a plus-maze procedure in rats (Rubinow
2001; Wise et al., 2001a,b; Liang et al., 2002). The genomic et al., 1998; Figueiredo et al., 2002) and facilitated the
actions of estrogen are expressed via intracellular “ER-␣‘’ anxiolytic properties of oxytocin in mice (McCarthy et al.,
and “ER-␤‘’ receptors: following a conformational change, 1996). Further, mice genetically lacking ER-␣ receptors
receptor-estrogen dimers bind to a palindromic DNA se- show alterations in emotionality (Ogawa et al., 1997).
quence which, together with several co-factors, leads to al- Third, mechanistic studies have demonstrated that es-
terations in the gene transcription of “estrogen-responsive” trogens impact upon several transmitter pathways known
elements (McDonnell, 1999; Driggers and Segars, 2002; to fulfil a crucial role in the control of anxious states.
Littleton-Kearney et al., 2002; Vasudevan et al., 2002). Thus, (1) experimental and clinical investigations indi-
These genomic actions, which are exerted over a time-scale cate a pronounced influence of estrogen upon serotonergic
of hours or more, are responsible for the neurotrophic transmission including, upon prolonged administration, al-
and neuroprotective actions of estrogen (Van Amelsvoort terations in 5-HT1A receptor binding in the limbic system
164 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

of rats and an elevation in the density of cortical 5-HT2A a population in which the role of estrogen in the control of
receptors: the latter finding has, notably, been repro- mood and cognition is under intensive scrutiny. Indeed, syn-
duced in postmenopausal women (Sumner and Fink, 1995; thetic “selective estrogen receptor modulators”, engineered
Osterlund and Hurd, 1998; Rubinow et al., 1998; Moses to optimise procognitive and neuroprotective properties and
et al., 2000; Roca et al., 2002); (2) estrogen potentiates to minimize adverse side-effects, are under serious consider-
non-NMDA receptor-mediated glutamatergic transmission ation as a treatment option for postmenopausal women pos-
in the hippocampus via the recruitment of protein kinase sessing low levels of ovarian steroids (Koenig, 2001; Van
A (Gu and Moss, 1996). Moreover, following long-term Amelsvoort et al., 2001; Wise et al., 2001a,b; Brinton, 2002;
treatment, estrogen increases spine density and intensifies Littleton-Kearney et al., 2002).
NMDA receptor-mediated signalling in hippocampal pyra- As regards males, it should be pointed out that testos-
midal neurones (Pozzo-Miller et al., 1999). On the other terone exerts a marked influence upon mood both: (1) fol-
hand, estrogen elicits a (delayed) acceleration of gluta- lowing its aromatisation to estradiol in the brain, via actions
mate clearance via an up-regulation of astrocytic glutamate at ER receptors and (2) following its metabolism to neu-
transporters in the human cortex (Liang et al., 2002). (3) rosteroids, via actions at GABAA receptors (Section 2.2.3).
Estrogen attenuates the potency of agonists at GABAB re- Intriguingly, testosterone has been reported to increase pun-
ceptors (Kelly and Wagner, 1999). Contrariwise, there is ished responses in the VCT (Table 18). Though it was ten-
evidence that estrogen reinforces transmission at GABAA tatively suggested that this action should be attributed to its
receptors, possibly by modifying their subunit composition: pro-impulsive properties rather than to a reduction in anxi-
this action translates into a diminution in their sensitivity to ety, testosterone displays anxiolytic actions in other models
the anxiogenic properties of GABAA receptor antagonists, (Bitran et al., 1993; Su et al., 1993; Bing et al., 1998; Mc-
though any alteration in the anxiolytic actions of allosteric Intyre et al., 2002b; Yang et al., 2002). It will interesting to
modulators awaits description (Herbison and Fénelon, 1995; determine the extent to which actions at ERs as compared
Jung et al., 2002). (4) Short-term exposure to estrogen both to GABAA receptors are implicated in its influence upon
suppresses the synthesis of ␣2A -ARs and interferes with anxious states.
their inhibitory influence upon NA release in rodent cor- Though long disregarded, inter-sex disparities in the re-
tex (Karkanias and Etgen, 1993; Karkanias et al., 1997). sponse to fear and stress and the actions of anxiolytic agents
Further, estrogen exerts a more persistent, sex-dependent are attracting increasing attention (Rubinow et al., 1998;
facilitatory (female) and inhibitory (male) influence upon Bale et al., 2002; Canli et al., 2002; Jung et al., 2002; McIn-
LC-derived noradrenergic projections (Thanky et al., 2002). tyre et al., 2002a,b; Monleon et al., 2002; Stroud et al., 2002;
Fifth, estrogen exerts a pronounced but complex pattern of Thanky et al., 2002; see Palanza, 2001). Notably, there is
influence upon the hypothalamo-pituitary–corticotropic axis substantial evidence for sexual dimorphism in the behaviour
as a function of the condition of study (resting state or stress), of males as compared to females (which are seldom em-
gender, species and parameter under evaluation (Carrasco ployed in pharmacological studies) in tests of anxious be-
and Van de Kar, 2003). Notably, in postmenopausal women, haviour (Gray, 1971; Archer, 1975; Johnston and File, 1991;
estrogen replacement therapy moderates the corticotropic Gogos et al., 1998; Palanza, 2001). Sex differences have
response to emotional stress (Lindheim et al., 1992). been documented for conflict tests in general and for the
In light of the above, it is not unreasonable to propose VCT in particular. However, the contribution of enhanced
that the dual genomic and non-genomic, central actions of emotionality to the lower level of punished responses emit-
estrogen are expressed in, as yet little-evaluated, alterations ted by female as compared to male rats in the VCT remains
in anxious states. However, in view of the diversity of these uncertain since this difference may, at least partially, reflect
interactions and the paucity of behavioural data, it would their greater sensitivity to noxious and other modes of aver-
be perilous to attempt concrete predictions concerning the sive stimuli, in addition to enhanced anxiety per se (Johnston
overall influence of ER ligands upon anxious states in man. and File, 1991; Peričić and Pivac, 1995; Palanza, 2001).
Indeed, the interpretation of test-dependent differences be-
12.3. Clinical pertinence: gender and anxiety tween genders is fraught with difficulty inasmuch as certain
variables which systematically differ between males and fe-
Impetus to studies of the functional significance of estro- males, such as appetite, endocrine status and motor func-
gens in the control of anxious states is provided by human tion, impact upon performance in models of anxious be-
exposure (voluntary and involuntary) to many environmental haviour. Thus, though the significance of inter-sex differ-
estrogen mimics and antagonists. They are generally clas- ences in the response to and control of anxious states is
sified as: (1) xenoestrogens (including pollutants and pes- widely acknowledged, there still remain some uncertainty
ticides) and (2) phytoestrogens (Belcher and Zsarnovszky, concerning their precise nature and significance—an issue
2001; Naftolin and Guadalupe Stanbury, 2002). The lat- of not inconsiderable interest to the researcher and lay public
ter, which are derived from plants, bacteria and funghi, are alike!
present in the diet and have been advocated as alternatives In analogy to the central actions of glucorticoids, then,
to estrogen replacement therapy in postmenopausal women, estrogen has been strongly implicated in the control of mood,
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 165

and numerous biochemical studies have demonstrated its tion and/or extinction is still debated (Bauer et al., 2002;
interaction with corticolimbic neurotransmitters involved in Cain et al., 2002); (4) L-type VDCCs trigger a sustained
the modulation of anxious states. Nevertheless, there is a neuronal influx of Ca2+ upon depolarisation which, via a
need for considerable further study of its influence upon diversity of soluble messengers and transcription factors,
behaviour under conditions of stress and fear in order to initiates long-term processes of synaptic plasticity in the
more accurately define its relevance to the pathogenesis and hippocampus and amygdala: such alterations in calcium
management of anxiety disorders. signalling are implicated both in the control of emotion
and in the pathology of mood disorders (Helmeste and
Tang, 1998; Yamawaki et al., 1998; Dolmetsch et al., 2001)
13. Ion channels and 5), Most pertinently within the present context, L-type
VDDC antagonists, such as nicardipine, display specific
13.1. Calcium channels anxiolytic actions in the VCT (Table 18) (Matsumoto et al.,
1994).
13.1.1. Operation and localization
Calcium influx into central neurones is mediated, in re- 13.1.3. Actions of GABApentin and pregabalin
sponse to depolarisation, by VDCCs: that is, voltage-gated, Notwithstanding the importance of GABAA receptors
heteromeric (4 and 5 subunits) Ca2+ -channels composed (Section 2.2.4), direct actions at VDCCs may partici-
of ␣1 -subunits, membrane-anchored (largely extracellular) pate in the anxiolytic properties of non-volatile anes-
␣2␦ -subunits, intracellular ␤-subunits and transmembrane thetics such as propofol (Kitayama et al., 2002). Interest
␥-subunits. The ␣1 -subunit, which constitutes the cur- in VDCCs as targets for potential anxiolytic agents is
rent sensor and conduction pore, bears most regulatory also encouraged by recent observations with pregabalin
sites. Ten forms of ␣1 -subunit are differentially incorpo- (S-isobutylGABA) and GABApentin (neurontin), both of
rated into, and determine the characteristics of, several which are structurally-related to GABA. As mentioned in
types of Ca2+ -channel, L, N, P, Q, R and T (Miljanich Section 2.3.1, it was recently suggested that GABApentin
and Ramachandran, 1995; Hofmman et al., 1999; Ertel behaves as a positive modulator of GABAB receptors, ac-
et al., 2000). Three (rodents) or four (in man) subtypes of tivation of which suppresses VDCC-induced transmitter
␣2␦ -subunit, four subtypes of ␤-subunit and several subtypes release (Wu and Saggau, 1997; Taylor et al., 1998; Ng et al.,
of ␥-subunit also exist. Though their exact roles remain 2001). Thus, an action of GABApentin at GABAB sites
unclear, co-expression of the ␣2␦ moiety with other VDCC might contribute to its anxiolytic profile. However, this con-
subunits promotes current flow and yields additional bind- tention has been strongly challenged (Stringer and Lorenzo,
ing sites (Klugbauer et al., 1999; Ertel et al., 2000; Hobom 1999; Lanneau et al., 2001; Bowery et al., 2002). Indeed, a
et al., 2000; Marais et al., 2001). ␣2␦ -subunits are compelling body of evidence has accumulated showing that
well-represented in corticolimbic structures. While all sub- pregabalin and GABApentin potently interact with ␣2 /␦1
types are present in the cortex, ␣2 /␦1 - and ␣2 /␦3 -subtypes and ␣2 /␦2 (but not ␣2 /␦3 ) subunits of VDCCs (Gee et al.,
are highly expressed in the hippocampus, and ␣2 /␦2 -subunits 1996; Stefani et al., 1998a; Bryans and Wustrow, 1999;
are concentrated in the septum (Hill et al., 1993; Klugbauer Gong et al., 2001; Marais et al., 2001), and actions at these
et al., 1999; Hobom et al., 2000; Marais et al., 2001; Fink sites appear to mediate their anxiolytic actions in conflict
et al., 2002). and other paradigms (Singh et al., 1996; Field et al., 2001).
Specifically, one component of the anxiolytic actions of
13.1.2. Control of anxious states GABApentin and pregabalin may reflect interference with
Though the composition of specific populations of VDCC P/Q type Ca2+ -channels controlling transmitter outflow.
involved in the control of anxious states cannot, as yet, Thus, they suppress NA release both directly and indirectly
be stated, their potential importance is underpinned by the via a reduction of glutamate release: the latter action leads
following observations: (1) in line with the above-described to a moderation of the excitatory tone of AMPA receptors
localization of VDCC subunits, L-type VDCCs are virtu- at noradrenergic terminals. Further, a decrease in the release
ally ubiquitous in the CNS and radiolabelled antagonists of glutamate onto corticolimbic NMDA receptors may also
intensely label binding sites in the hippocampus, amygdala be involved in the anxiolytic properties of GABApentin
and cortex (Gould et al., 1985; Hobom et al., 2000); (2) (Dooley et al., 2000a,b, 2002; Fink et al., 2002; Van Hooft
VDCC antagonists modulate monoaminergic, glutamatergic et al., 2002).
and GABAergic transmission (Harvey et al., 1996; Dooley Inasmuch as pregabalin and GABApentin express marked
et al., 2000a,b; Katagiri et al., 2001; Chen et al., 2002b; anxiolytic properties in man (Beydoun et al., 1995; Pande
Fink et al., 2002; Kitayama et al., 2002); (3) high-threshold et al., 1999a, 2000; Kasper et al., 2002; Rickels et al., 2002;
L-type VDCCs fulfil a role complementary to that of Smith et al., 2002b), their actions in the rat VCT would be
NMDA receptors in the mediation of fear-conditioning of interest to evaluate. Indeed, very recently, a preliminary
in the amygdala, a process with parallels to the VCT, report of anxiolytic properties of pregabalin in a mice VCT
though their precise implication in mechanisms of acquisi- appeared (Lotarski et al., 2002).
166 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

13.2. Sodium channels demonstrated that mice genetically deprived of the ␤-subunit
of Kv1.1 channels display an anxiolytic phenotype raising
Of the many types of voltage-gated Na+ -channels now the possibility that—an, as yet undefined population of—
recognized, several tetrodotoxin-sensitive, heteromeric (one these sites may favour anxious states (Brennan et al., 2002).
␣-, one ␤1 - and one ␤2 -subunit) classes are concentrated in Of the seven families of inwardly-rectifying K+ -channels
the cerebral cortex and hippocampus (Catterall, 2000; Clare currently accepted, the Kir3 class of G-protein activated
et al., 2000; Denac et al., 2000; Novakovic et al., 2001; K+ -channels (GIRKs) is activated by ␤␥-subunits released
Lindia and Abbadie, 2003). In the latter structure, via en- from pertussis toxin-sensitive Gi/o -proteins coupled to
gagement of VDCCs, Na+ -influx accelerates the release of several classes of GPCR, including GABAB receptors,
NA (Gerevich et al., 2001). To date, studies of the functional opioid and OFQ receptors, 5-HT1A , D2 and D3 receptors,
significance of central Na+ -channels have focussed on pain M2 -receptors, ␤2 - and ␣2 -ARs and A1 receptors: they may
and anesthesia—reflecting actions at thalamo-cortical affer- also be directly engaged by ethanol (North, 1989; Dascal,
ents and peripheral nerves (Schwarz and Puil, 1998; Millan, 1997; Lewohl et al., 1999; Reimann and Ashcroft, 1999;
1999; Waxman et al., 1999; Clare et al., 2000; Laughlin Mark and Herlitze, 2000; Lavine et al., 2002). More re-
et al., 2002)—and on neuroprotection and epilepsy (Taylor cently, attention has also been directed to the inhibition
and Narasimhan, 1997; Ragsdale and Avoli, 1998; Clare of GIRKs by GPCRs coupled to (pertussis-insensitive)
et al., 2000; Anger et al., 2001). More recently, attention Gq-proteins: they appear to act in concert with phospholi-
has been directed towards the influence of lamotrigine—an pase C by facilitating the diffusion away from the channel
anticonvulsant which acts as an antagonist of voltage-gated of the positive modulator, phosphatidylinositol phosphate
Na+ -channels (as well as VDCCs)—upon the limbic re- (Lei et al., 2001). Of the four subunits to date cloned, neu-
lease of glutamate and GABA in relation to its influence ronal GIRK tetramers are composed of GIRK1, GIRK2 and,
upon mood and affective disorders (Xie et al., 1995b; Cun- possibly, GIRK3 subunits—generally two of each, each of
ningham and Jones, 2000; Dursun and Deakin, 2001; Braga which bear binding sites for ␤␥. The broad distribution
et al., 2002; Spadoni et al., 2002). A further multi-target drug of GIRKs in the FCX, amygdala, bed nucleus of the stria
which interferes with glutamate release—and also shows terminalis, hippocampus, LC and ventrotegmental area, for
modest antagonist properties at NMDA receptors—is rilu- example, underpins interest in their role in the modulation
zole, currently approved for the treatment of amyotrophic of emotional states (Kobayashi et al., 1995; Karschin et al.,
lateral sclerosis. It has been reported to abrogate anxiogenic 1996; Reimann and Ashcroft, 1999; Lavine et al., 2002;
properties of inverse agonists at BZP receptors in rodents, Sadja et al., 2002). Indeed, mice deficient in GIRKs reveal
and to suppress activation of the hypothalamo-corticotrophic a low level of anxiety (Blednov et al., 2001, 2002).
axis by stress in human subjects (Stutzmann et al., 1989; Though little else is known concerning a potential role
Kniest et al., 2001). It would be instructive to evaluate a of K+ -channels in the control of anxious states, it has been
possible influence of lamotrigine, riluzole and related drugs speculated that the blockade of GIRK2s by fluoxetine, to-
acting via central voltage-gated Na+ -channels upon anxious gether with its inhibition of voltage-gated (Kv1.1 and Kv3.1)
behaviour in the VCT. channels, contributes to its influence upon anxious and other
psychiatric disorders (Choi et al., 2001; Kobayashi et al.,
13.3. Potassium-channels 2003).

The electrical activity of neurones is regulated by a strik-


ing diversity of K+ -channels sub-classified into several fam- 14. Nitric oxide
ilies as a function of their regulation by calcium, adenosine
triphosphate, membrane current, GPCRs and other signals The short-lived, highly-reactive free radical, NO, is
(Aronsen, 1992; Kobayashi et al., 1995; Dascal, 1997; Doyle generated from the cleavage of l-arginine by various
et al., 1998; Kaczorowski and Garcia, 1999; Reimann and forms of NO synthase (I, II and III) in both neuronal and
Ashcroft, 1999). The role of the majority of these multi- non-neuronal elements throughout the CNS: it modulates
meric K+ -channels in the induction and control of anxious neurotransmission by anterograde and retrograde actions
states remains essentially unexplored, largely reflecting the (Zhang and Snyder, 1995; Millan, 1999; Tao and Poo,
limited availability of agents for their selective manipula- 2001; Wiesinger, 2001; Alger, 2002). Constitutively-active,
tion. This paucity of experimental tools applies in particular cytosolic, Ca2+ -dependent neuronal NO synthase (or NO
to voltage-gated and GPCR-gated K+ -channels for which, synthase I), is present in several corticolimbic structures,
nevertheless, the following observations justify mention. including the amygdala and hippocampus (Egberongbe
Various classes of voltage-dependent, heteromeric et al., 1994; Braissant et al., 1999). It shows a broad pat-
K+ -channels are composed of four (␣ and ␤) subunits of tern of interaction with monoaminergic pathways and also
which the latter are responsible for their functional regu- modulates GABAergic transmission (Szabo, 1996; Segieth
lation and rapid inactivation (Aronsen, 1992; Kaczorowski et al., 2001; Adell et al., 2002; Millan, 2002a). Thus, NO
and Garcia, 1999). Recently, a preliminary communication appears to interfere with the activity of GABAA receptors,
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 167

and acute application of neuronal NO synthase inhibitors Evidently, various pools of NO are differentially impli-
potentiates the actions of allosteric modulators at GABAA cated in the response to and control of anxious states, and
sites (Robello et al., 1996). Further, their long-term admin- this mediator is unlikely to fulfil a uniform role. A prerequi-
istration elicits a widespread increase in the cerebral density site for an improved understanding of its significance would
of GABAA recognition and BZP binding sites, together be the more precise delineation of the functional roles of all
with a complex pattern of alterations in the gene expression three isoforms of NO-synthase, and a more thorough func-
of various GABAA receptor subunits (Kim and Oh, 2002). tional evaluation of the interaction of NO with glutamater-
On the other hand, inhibitors of NO synthase enhance the gic, GABAergic and monoaminergic pathways in discrete
activity of GABA-transaminase, thereby accelerating the corticolimbic structures.
catabolic transformation of GABA (Jayakumar et al., 1999).
Though many G-protein-coupled receptors can modulate
neuronal synthesis of NO (Christopoulos and El-Fakahany, 15. Neurotrophins
1999), most interest has focussed on the activation of NO
by NMDA receptors via an increase in intracellular levels of 15.1. Actions of neurotrophins at “Trk” receptors:
Ca2+ , and NO may itself modulate glutamatergic transmis- cerebral localization of BDNF
sion (Millan, 1999; Wiesinger, 2001). Further by analogy to
glutamatergic transmission and NMDA receptors, acute and Neurotrophins act via members of the tropomyosin-related
chronic exposure to fear and other stressors enhances the kinase (Trk) family of tyrosine kinases. Nerve growth fac-
activity of neuronal NO synthase in the PAG, amygdala and tor (NGF) activates TrkA receptors; brain-derived growth
other cerebral tissues (De Oliveira et al., 2000; Echeverry factor (BNDF), neurotrophin4/5 and, to a lesser degree,
et al., 2002; Masood et al., 2003. neurotrophin3 , are ligands of TrkB sites, while neurotrophin3
One might expect from the above comments that inhibi- engages TrkC sites (Millan, 1999; Kaplan and Miller, 2000;
tion of neuronal NO synthase would mimic functional ac- Patapoutian and Reichardt, 2001).
tions of positive GABAA receptor modulators (Section 2.2) The localization of TrkB receptors and BDNF in the DRN,
and of NMDA receptor antagonists (Section 3.2.3.1). This LC, hypothalamus, amygdala, cortex, PAG, septum and,
has proven the case in models of, for example, antinocicep- most strikingly, the hippocampus suggests a role in the con-
tive properties (Harkin et al., 1999; Millan, 1999). Further, trol of mood via both long-term (“neurotrophic”) and more
inhibitors of NO synthase have been reported to display rapid (transmitter-like) actions (Altar et al., 1996; Yan et al.,
anxiolytic properties, while l-arginine blunts the anxiolytic 1997; Black, 1999; Drake et al., 1999; Akbarian et al., 2002;
properties of BZPs (Faria et al., 1997; Volke et al., 1997, Reiriz et al., 2002; Tolwany et al., 2002). As concerns the lat-
1998; Dunn et al., 1998). Notably, the anxiolytic actions ter, BDNF can recruit non-selective cation channels via stim-
of NO synthase inhibitors in the VCT were blocked by ulation of phospholipase C. However, its excitatory (depo-
l-arginine demonstrating the specificity of these observa- larising) influence upon neuronal activity primarily reflects
tions (Table 18) (Dunn et al., 1995). the rapid activation of a specific class of voltage-dependent
However, serious reservations concerning the generality Na+ -channels (Li et al., 1999; Blum et al., 2002). By anal-
of anxiolytic properties of neuronal NO synthase inhibitors ogy to NO (Section 14) and cannabinoids (Section 8), BDNF
are raised by other studies indicative of anxiolytic and anx- can act both via conventional anterograde and also via ret-
iogenic actions of NO and NO synthase inhibitors, respec- rograde modes of signalling (Millan, 1999; Tao and Poo,
tively (Vale et al., 1998; Monzon et al., 2001; Li et al., 2003; 2001; Alger, 2002).
Masood et al., 2003)). In addition, the latter agents inter-
fered with anxiolytic actions of BZPs in mice (Quock and 15.2. Role of BDNF in the control of anxious states
Nguyen, 1992).
Long-term treatment with BZPs induced NO synthase II The gene expression of BDNF is positively regulated by
in immunocompetent, cerebral populations of microglia and protein kinase A, following recruitment of adenylyl cyclase.
astroglia. Whether this action is related to their anxiolytic It is also engendered by an increase in intracellular concen-
properties to be ascertained (Suzuki et al., 2002a,b). Like- trations of Ca2+ provoked either by activation of phospho-
wise, inasmuch as chronic (but not acute) exposure to SSRIs lipase C and/or the engagement of VDCCs. Thus, the gen-
similarly augmented the gene expression of NO synthase II, eration of BDNF may, at least in theory, be enhanced by
it would be interesting to establish the relationship of this ac- numerous mechanisms discussed in this review, of which
tion to their long-term alleviation of anxious states (Suzuki glutamatergic mechanisms appear to be of particular impor-
et al., 2002a,b)). tance (Hartmann et al., 2001; Mackowiak et al., 2002).
While the above-described behavioural actions reflect Acute and chronic exposure to stress or corticosteroids
roles of either neuronal NO synthase or NO synthase II, ge- elicits dendritic retraction and remodelling, provokes neu-
netic deletion of endothelial NO (synthase III) led, via an un- ronal atrophy and (perhaps causally) suppresses BDNF ex-
clear mechanism, to an anxiogenic phenotype (Frisch et al., pression in the hippocampus, amygdala and other cerebral
2000), indicative of an anxiolytic role for this source of NO. areas (Smith et al., 1995b; Altar, 1999; Duman et al., 1999;
168 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Nibuya et al., 1999; Vaidya et al., 1999; Sousa et al., 2000; status of BDNF-expressing neurones in the hippocampus
Fuchs and Flügge, 2002; Vyas et al., 2002). These patholog- and other corticolimbic regions.
ical changes can be countered by long-term administration
of SSRIs and other classes of antidepressant agent which 15.3. BDNF/TrkB receptors as a therapeutic target
amplify the gene expression of BDNF in the hippocampus
and amygdala: the accompanying alterations in synaptic ar- Despite the general significance of BDNF in the con-
chitecture and intracellular signalling may be involved in trol of mood, and its (potential) role in the anxiolytic and
their delayed onset of therapeutic efficacy (Siuciak et al., antidepressant effects of long-term administration of psy-
1997; Altar, 1999; Duman et al., 1999; Zetterström et al., chotropic agents, there remain several caveats concerning
1999; Mössner et al., 2000; Conti et al., 2002; Nestler et al., its validity as a “target” for improved management of anx-
2002; Tolwany et al., 2002; Saarelainen et al., 2003). Indeed, ious (and depressive) states. First, even if highly-selective,
though such events have widely been interpreted within the brain-penetrant, bioactive mimics (or blockers) of BDNF
perspective of depressive states and their treatment, it is pos- can be identified, they will likely elicit marked actions at
sible that they are also related to their long-term manage- non-cerebral TrkB sites localized in the spinal cord and the
ment of anxiety disorders. peripheral nervous system wherein BDNF fulfils important
The following observations further support the argument roles both developmentally and in the modulation of sensory
that acute and chronic actions of BDNF in limbic structures and neuroendocrine function in adults (Lewin and Barde,
may be of relevance to anxious as well as depressive states 1996; Millan, 1999; Pollack and Harper, 2002). Second, it is
(Drake et al., 1999; Duman et al., 1999; Duman et al., 2001; still controversial whether BDNF is merely a marker of drug
Shirayama et al., 2002). First, in the hippocampus, BDNF actions, or whether it genuinely initiates processes underly-
is partially colocalized with GABAergic neurones. In this ing the delayed onset of therapeutic activity in the treatment
structure, in addition to its developmental role in regulat- of affective disorders (Duman et al., 2001; Popoli et al.,
ing the maturation of GABAergic neurones and the forma- 2002; Nestler et al., 2002). Third, assuming that neuroge-
tion of GABAergic synapses, BDNF rapidly down-regulates nesis and other structural changes involving BDNF/TrkB
postsynaptic GABAA receptors in the adult hippocampus, receptors do participate in the long-term actions of SSRIs
thereby attenuating GABAergic transmission (Brünig et al., and other agents, and that they are causative of clinical
2001; Henneberger et al., 2002). Second, the facilitatory improvement (Duman et al., 1999; Nestler et al., 2002), it
influence of BDNF upon the activity of DRN-derived sero- would be desirable to devise clinical protocols permitting
tonergic and—though less-well-characterized—LC-derived the demonstration that such events occur in the course of
noradrenergic neurons (Siuciak et al., 1995, 1996, 1998; therapeutic trials. Fourth, it is difficult to predict with any
Celada et al., 1996; Madhav et al., 2001) has been (causally) certainty the long-term consequences of sustained and direct
related to its positive actions in models of potential an- manipulation (facilitation or disruption) of neurotrophin sig-
tidepressant activity. However, potential alterations in be- nalling: for example, whether the presumptive modifications
haviour in models of anxious states may also be envisaged of synaptic architecture and transmission—desirable and/or
and, by analogy to SSRIs, acute induction of corticolimbic undesirable—would be reversible.
release of 5-HT and NA by BDNF may well be accompa- For these and other reasons, it may be more prudent to
nied by a transient accentuation of anxiety. Third, the ac- therapeutically intervene upstream of BDNF/TrkB recep-
celeration by BDNF of hippocampal glutamate release may tors employing drugs which permit the precise targeting
similarly contribute to its influence upon mood—though the of discrete, therapeutically-pertinent, corticolimbic pools of
participation of glutamatergic mechanisms to the changes in BDNF and TrkB receptors implicated in the etiology and,
synaptic strength evoked by BDNF remains under discus- potentially, management of mood disorders. Clearly, such
sion (Gooney and Lynch, 2001; Messaoudi et al., 2002). agents are already available, as represented by SSRIs and
Fourth, heterozygotic BDNF (±) mice show alterations other classes of antidepressant agents (Duman et al., 2001;
in aggressive behaviour, in performance in tests of antide- Popoli et al., 2002; Nestler et al., 2002), though it is likely
pressant activity, in hippocampal excitability and in sero- that innovative and more efficacious approaches for modu-
tonergic transmission (Croll et al., 1999; Lyons et al., 1999; lation of the activity of BDNF will be developed.
MacQueen et al., 2001), though potential alterations in anx- Within this perspective, then, considerable further re-
ious behaviour have not, as yet, been reported. search is needed to decipher the role of BDNF in the patho-
The use of genetically-transformed mice either deficient genesis and treatment of anxiety and other mood disorders
in (or overexpressing) BDNF, together with studies of prior to its potential exploitation as a therapeutic target.
BDNF itself and small molecules recognizing TrkB recep-
tors (Obara and Nakahata, 2002; Pollack and Harper, 2002), 15.4. Other neurotrophins
should permit the future designation of the role of BDNF in
the control of anxious states. It would also be informative to In this regard, the possible significance of NGF/TrkA re-
determine the influence of long-term treatment with BZPs ceptors and NT3/TrkC receptors would also justify evalua-
and other classes of anxiolytic agent upon the functional tion since, though data are far less extensive, they are also
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 169

implicated in the cerebral and endocrine response to psy- tifies evaluation employing the VCT and other experimen-
chosocial stress, and in the modulation of monoaminergic tal strategies. However, cytokine signalling in the CNS and
transmission both in animals and in man (Alleva and Aloe, other organs is crucial to the control of immune function
1989; Smith et al., 1995a,b; McLay et al., 1997; Schaaf et al., both under “resting” conditions and in the response to stress
1997; Winkler et al., 2000; Akbarian et al., 2001; Alleva and and illness. Thus, notwithstanding accumulating evidence
Santucci, 2001; Aloe et al., 2002; Haddad, 2002). for a modulatory influence of cytokines upon neuronal com-
munication and emotion, great care would need to be taken
in their hypothetical exploitation as therapeutic targets for
16. Cytokines the treatment of anxiety—or other mood—disorders.

A diverse group of proteins termed cytokines, of which


tissue necrosis factor␣ and Interleukin1␤ are representative 17. Sigma1 and sigma2 sites
members, play key roles in communication between immune
cells (McLay et al., 1997; Müller and Ackenheil, 1998; Had- 17.1. Sigma1 sites
dad, 2002). More recently, cytokines—derived both from
peripheral and from intrinsic, cerebral sources—have been “Sigma receptors” were originally welcomed into the fold
implicated in the modulation of neuronal activity in regions of opioid receptors but were subsequently banished on the
such as the amygdala, the hippocampus, the hypothalamus basis of their insensitivity to the “universal” opioid receptor
and the cortex (Besedovsky and Del Rey, 1996; Elenkov antagonist, naloxone (Section 9.1.9) (Quirion et al., 1992;
et al., 2000). In addition to their interaction with GABAer- Herz et al., 1993). Subsequently, they were confounded with
gic neurones, cytokines potentiate the activity of monoamin- the “phencyclidine receptor” which transpired to be a bind-
ergic and CRF-containing pathways in these corticolimbic ing site located in the ion channel coupled to NMDA recep-
structures (Hopkins and Rothwell, 1995; Merali et al., 1997; tors (Section 3.2.1) (Dingledine et al., 1999; Quirion et al.,
Kamikawa et al., 1998; Cho et al., 1999; Day et al., 1999b; 1992). Indeed, the odyssey of “sigma receptor” classification
Millan, 1999; Bonaccorso et al., 2000; Nickola et al., 2000; only concluded with the discovery of genuinely-discriminant
Kronfol and Remick, 2000; Schaefer et al., 2002). These ob- ligands (Quirion et al., 1992) and, ultimately, the cloning
servations, together with the pronounced response of central of sigma1 sites. This 223 amino acid protein, which pos-
and peripheral pools of cytokines to stress and their facil- sesses a single transmembrane domain, bears conspicuously
itatory influence upon the hypothalamo-corticotrophic axis little resemblance to other mammalian proteins yet discon-
(Kung, 1999; Kronfol and Remick, 2000; Schaefer et al., certing similarities to a yeast enzyme (C8–C7 isomerase) of
2002; Leonard and Song, 2002), have prompted interest in ergosterol synthesis, the fungal counterpart of cholesterol—
the role of cytokines in the pathogenesis of affective and though sigma1 sites do not exhibit equivalent enzymatic ac-
anxious states (Bluthé et al., 1992; Connor and Leonard, tivity (Hanner et al., 1996; Moebius et al., 1997; Prasad et al.,
1998; Müller and Ackenheil, 1998; Bonaccorso et al., 2000; 1998; Seth et al., 1998). Further, sigma1 sites bear a bind-
Kronfol and Remick, 2000; Schaefer et al., 2002). Indeed, ing domain for steroids (Section 2.2.3) which shows high
alterations in circulating levels of interleukin1␤ have been affinity for progesterone (an antagonist) and pregnanelone
demonstrated in patients with anxiety disorders (Brambilla (an agonist), though other neurosteroids such as allopreg-
et al., 1994; Zorrilla et al., 1994; Leonard and Song, 2002) nanalone are less potent ligands (Maurice et al., 2001).
and both experimental and clinical investigations have high- The functional status of sigma1 sites and the notion of
lighted the marked influence of antidepressant agents upon “agonists” and “antagonists” is still ambivalent inasmuch
the production of cytokines (Kenis and Maes, 2002). they have no apparent endogenous ligand—with the possi-
As concerns anxious states, overexpression of cytokines ble exception of steroids (Alonso et al., 2000; Maurice et al.,
in autoimmune mice is associated with behaviours indicative 2001). Further, they possess an endoplasmic reticulum re-
of enhanced anxiety, and mice overexpressing interleukin6 tention signal and are enriched on mitochondria and other
or tissue necrosis factor␣ likewise reveal an anxiogenic intracellular membranes (Hanner et al., 1996; Dussossoy
phenotype: underpinning these findings, anxiogenic actions et al., 1999; Maurice et al., 1999). Mechanisms of intracel-
of cytokine administration have been reported in rodents lular transduction also remain poorly-defined. The balance
(Bluet-Pajot et al., 1992; Bluthé et al., 1992; Sakic et al., of evidence suggests, in line with their structure, that they do
1994; Schrott and Crnic, 1996; Connor and Leonard, 1998; not directly couple to G-proteins though they may, upon ac-
Flore et al., 1998; Bonaccorso et al., 2000). On the other tivation and recruitment to plasma membranes, interact with
hand, certain cytokines may be inhibitory to processes in- phospholipase C and VDCCs (Morin-Surun et al., 1999;
ducing anxious states, inasmuch as deletion of the gene Tokuyama et al., 1999; Hayashi et al., 2000a,b; Hong and
encoding Interferon␥ led to heightened anxiety (Flore et al., Werling, 2000; Romero et al., 2000). Further, they modulate
1998; Kustova et al., 1998). intracellular concentrations of calcium by a novel mecha-
In light of the above comments, the heterogeneous role nism. Thus, activation of sigma1 sites complexed to the cy-
of specific cytokines in the control of anxious states jus- toskeletal adaptor protein, ankyrin, on the endoplasmic retic-
170 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

ulum leads to their dissociation from inositol triphosphate behavioural models (Umezu, 1999). To date, there appear
receptors: this results in an enhancement in the binding of in- to be little or no data regarding the potential influence of
ositol phosphates to the latter and an efflux of calcium from genuinely-selective sigma1 antagonists upon anxious states,
the endoplasmic reticulum (Hayashi et al., 2000a; Hayashi an issue which would be of interest to examine by use of
and Su, 2001; Maurice et al., 2001). Very recently, evidence the VCT and other procedures.
that the modulation of calcium signalling (intra and extra- The tricyclic compound, opipramol, which shows high
cellular) by sigma1 ligands contributes to their behavioural affinity for sigma1 sites, demonstrated anxiolytic efficacy in
(antidepressant) actions was acquired (Urani et al., 2002). a controlled trial of GAD in man (Rao et al., 1990; Ferris
Sigma1 sites may also inhibit potassium currents (Wilke et al., 1990; Möller et al., 2001). However, while opipramol
et al., 1999) which would further enhance cellular excitabil- has no affinity for monoamine transporters, its antagonist
ity. In rodents, primates and man, sigma1 sites are present actions at 5-HT2 or histamine1 receptors are likely of greater
in high concentrations in neurones in the hippocampus, ol- significance to its anxiolytic properties than its actions at
factory bulb and hypothalamus and a moderate density has sigma1 sites in view of the above-evoked pattern of negative
been described in the cortex, lateral septum, amygdala and data from experimental models.
PAG (Largent et al., 1986; Mash and Zabetian, 1992; Alonso
et al., 2000; Kitaichi et al., 2000; Maurice et al., 2001). 17.2. Sigma2 sites
Central sigma1 receptors modulate corticosterone secre-
tion, though the precise nature of their influence remains Certain ligands which bind to sigma1 sites, such as DTG,
unclear (Matheson et al., 1991). They also display func- (+)PPP and haloperidol, also recognize sigma2 sites which
tional interactions with NMDA receptors (Whitemore and can be distinguished by their relatively low affinity for the
Woodward, 1997; Kitaichi et al., 2000). Notably, sigma1 (+)- as compared to (−)-isomers of SKF10,047, penta-
agonists potentiate the facilitatory influence of NMDA re- zocine and cyclazocine, whereas sigma1 sites show the op-
ceptors upon hippocampal NA release (Monnet et al., 1996 posite pattern of preference: (+) > (−) (Quirion et al., 1992;
though see Gonzalez-Alvear and Werling, 1995) and may, Hellewell et al., 1994). Unfortunately, sigma2 sites have not
administered alone, enhance extracellular levels of NA in as yet been sequenced and their identity remains nebulous.
the cortex (Cooke et al., 1998) though this remains to be Further, despite the (to date, unconfirmed) suggestion that
confirmed. Further, sigma1 site agonists reproducibly accel- they also modulate release of intracellular calcium stores
erate the activity of ventrotegmental dopaminergic perikarya (Vilner and Bowen, 2000), their mode of coupling remains
innervating the limbic system and cortex (Ceci et al., 1988; obscure. More encouragingly, in line with earlier studies
Zhang et al., 1993; Matsumo et al., 1995; Volonté et al., of non-selective radioligands, the novel, radiolabelled and
1995). The influence of sigma1 ligands upon serotonergic selective sigma2 ligand, siramesine (Lu28-719), bound to
transmission is less pronounced, though certain agonists sites in the cortex, hippocampus, locus coeruleus and amyg-
elicit a long-term enhancement in the activity of serotonergic dala of rat brain, though quantitative differences amongst
neurones (Bermack and Debonnel, 2001; Takahashi et al., discrete cerebral structures were modest (Bouchard and
2001b). Quirion, 1997; Heading, 2001; Soby et al., 2002). Sirame-
The above data have been interpreted as consistent with sine showed exceptionally potent anxiolytic activity (rem-
potential antidepressant actions of sigma1 ligands. Corre- iniscent of miniscule doses of 5-HT3 antagonists (Section
spondingly, the ability of the “agonist”, (+)SKF10,047, to 4.3.5)) in a rat light–dark box (0.00018 ng/kg) and was
suppress conditioned freezing in rats—an action blocked also potently active in a model of active social interaction
by the selective antagonist, NE 100—was attributed to (0.1 ng/kg) though only substantially (million-fold) higher
antidepressant rather than anxiolytic properties (Kamei doses were active in the VCT (10 and 22 mg/kg) (Sanchez
et al., 1996) and may be related to the facilitatory influence et al., 1997). Inasmuch as siramesine was well tolerated in
of sigma1 sites upon mesolimbic dopaminergic pathways a Phase I trial (Heading, 2001), clinical data pertaining to
(Kamei et al., 1994, 1997). This action was accompanied the anxiolytic potential of sigma2 ligands may be imminent.
by alterations in the binding density of central sigma1 sites
indicating their responsiveness to stress (Nabeshima et al., 17.3. Perspectives for sigma ligands as anxiolytic
1988). Anxiogenic effects were, in fact, reported for sigma1 agents
agonists in a preliminary communication concerning the rat
VCT (Lai et al., 1989) but Akunne et al., 1997 reported In view of the unclear nature of transduction mechanisms
that the selective sigma1 agonist, PD144418, was inactive for sigma2 sites, their resistance to cloning and the limited
in this model. In a mice edition of this procedure, Umezu, quantity of experimental data, their physiological and ther-
1999 likewise failed to detect significant anxiolytic proper- apeutic significance to the control of anxious states remains
ties of a mixed sigma1/2 agonist, racemic cyclazocine (vide moot. For sigma1 sites, despite their distribution in cor-
infra). Finally, McMillan et al., 1991 reported that sigma1 ticolimbic structures, interaction with monoaminergic and
agonists were ineffective in a Geller–Seifter conflict proce- glutamatergic pathways and better-defined mechanisms of
dure in rats, underscoring findings from the VCT and other coupling, the lack of selectivity of certain ligands—and,
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 171

probably as a consequence, surprising differences between plementary techniques including: local drug and antisense
the actions of individual ligands in functional models— administration; selective inactivation (lesions) of discrete
complicates assessment of their pathophysiological perti- neuronal pathways, and monitoring of neurochemical and
nence. Though acute studies of agonists indicate no influ- electrophysiological events in specific brain regions. More-
ence upon anxious states, by analogy to other classes of an- over, “genetic” techniques permitting the manipulation of
tidepressant agent, examination of their actions (and those function in circumscribed cerebral structures—for example,
of antagonists) upon chronic administration would certainly deletion of post as compared to presynaptic populations of
be justified. receptors—offer novel and powerful strategies to address
this issue.
In the above light, as exemplified by studies of seroton-
18. General discussion: open questions, conceptual ergic, cholinergic, NPY and CRF-containing neurones, it
issues and future research directions should be recalled that certain transmitters—and even indi-
vidual classes of receptor—can mediate contrasting actions
18.1. Multiple substrates of action as a function of, for example, their localization in different
cerebral structures (Figs. 10 and 11); their occurrence at pre-
In concluding this survey of neuronal mechanisms in- as compared to postsynaptic sites; the duration of their ac-
volved in the modulation of anxious states, certain themes tivation, and the type of anxiety under investigation. On the
likely to inform future research merit brief recapitulation. other hand, certain mediators may exert a unitary influence
To reiterate comments made throughout this article, upon anxious states in acting synergistically at multiple loci.
the actions of numerous modulators implicated in the re- The most obvious illustration is provided by GABAA recep-
sponse to stress and in the control of anxious states have tors, of which numerous supraspinal populations transduce
not, as yet, been evaluated by use of the VCT—or other the powerful anxiolytic properties of BZP (Fig. 10). Indeed,
punishment-based conflict procedures. Even for the major- this broad-based spectrum of actions may account for the
ity of those modulators which have been shown to exert distinctively robust and rapid anxiolytic profile of BZPs—
anxiolytic and/or anxiogenic properties in these (and other) in analogy to the unexcelled pain-relieving properties of
models, comparatively little information is available con- ␮-opioid analgesics which simultaneously act at multiple
cerning the precise neuronal substrates underlying their levels of the neuroaxis (Millan, 2002a). Irrespective of neu-
actions (Figs. 10 and 11). In the resolution of such ques- ronal mechanisms underlying the control of anxious states,
tions, it will be of importance to adopt a variety of com- however, it must be emphasized that the overall influence of

Fig. 10. Schematic representation of the sites of actions of various neurotransmitters and their receptors in the modulation of anxious states. The effects
indicated are a (non-exhaustive) synthesis of observations across all procedures. For the actions of certain transmitters and receptors, contradictory
data have been documented—even for the same procedure—and the action indicated represents an overall assessment for each structure. Anxiolytic
actions are indicated by a “+” sign and anxiogenic actions are indicated by a “−” sign: “±” indicates that a broad and balanced mixture of anxiolytic
and anxiogenic actions has been reported. A “0” indicates no effect though, to avoid over-encumbering the figure, only certain negative observations
are indicated. Abbreviations—GABA: ␥-amino-butyric acid; NMDA: N-methyl-d-aspartate; GLYB : glycineB ; MTB: metabotropic glutamatergic; 5-HT:
serotonin; ACh/n: acetylcholine/nicotinic and ACh/m: acetylcholine/muscarinic. For further details, see text.
172 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Fig. 11. Schematic representation of the sites of actions of various neuropeptides and their receptors in the modulation of anxious states. The effects
indicated are a (non-exhaustive) synthesis of observations across all procedures. For the actions of certain neuropeptides and their receptors, contradictory
data have been documented—even for the same procedure—and the action indicated represents an overall assessment for each structure. Anxiolytic
actions are indicated by a “+” sign and anxiogenic actions are indicated by a “−” sign: “±” indicates that a broad and balanced mixture of anxiolytic
and anxiogenic actions has been reported. A “0” indicates no effect though, to avoid over-encumbering the figure, only certain negative observations are
indicated. Note that, in comparison to the hippocampus, the amygdala is a “hot-bed” of neuropeptide activity whereas, for “conventional” neurotransmitters
depicted in Fig. 10, both structures have been implicated to a comparable degree. Abbreviations are as follows; CCK: cholecystokinin; SP/NK1 : substance
P/neurokinin1 ; CRF: corticotropin releasing factor; VP: vasopressin; OT: ocytocin; NPY: neuropeptide Y; PAC1 : receptor for Pituitatry Adenyl Cyclase
Activating Peptide; GAL: galanin; OFQ: orphaninFQ; GLP1 : glucagon-like peptide-1 and ANP: atrial natruetic peptide. For further details, see text.

drugs upon their systemic (generally oral) administration is bic monoaminergic mechanisms in the control of anxious
of decisive importance in assessing their potential for thera- states, as emphasized in Section 2.2.2.3, the assumption
peutic utilization. Hence, the continued importance—in ad- that the suppression/enhancement of 5-HT, NA or DA re-
dition to mechanistic analyses—of examining the actions of lease is synonymous with anxiolytic/anxiogenic properties,
all drug classes upon parenteral application. respectively, is clearly fallacious. Accordingly, anxiolytic
In elucidating the functional significance of specific me- and anxiogenic properties can be exerted independently of
diators, information concerning their localization is critical. changes in the release of these monoamines, and various
Not only their regional distribution in cerebral structures, classes of anxiolytic agent exert contrasting patterns of in-
but also their precise patterns of neuronal connectivity fluence upon extracellular levels of 5-HT, NA and DA—as
in relation to other modulators involved in the control of exemplified in Table 20, the data of which were generated
anxious states. As accentuated in the present review, knowl- in parallel under identical conditions.
edge of the relationships between specific neurotransmitters
and: (1) GABAergic interneurones; (2) serotonergic, no- 18.2. Cellular mechanisms of action
radrenergic and—as increasingly accepted—dopaminergic
projections and (3) key input and output channels of corti- In the preceding paragraphs, the importance of defining
colimbic structures modulating anxious states, can provide the actions of specific modulators at the neuronal level was
crucial insights into their potential influence upon the re- evoked. In this regard, though it is imperative to study cel-
sponse to stress and fear. Pursuing this line of reasoning, lular coupling in “simple” transfection systems, it will be
delineation of the short- and long-term influence of specific increasingly necessary to integrate the following concepts in
receptor types upon neuronal activity via their coupling to interpreting—and predicting—the influence of drugs upon
diverse intracellular signals (soluble second messengers and anxious states.
ion channels) can clarify their potential functional roles. First, neurotransmitters and other modulators manifest
For example, agents which reinforce GABAergic trans- extensive co-storage and co-release, and operate in a coor-
mission in the amygdala may reasonably be anticipated to dinated rather than an independent fashion (Hökfelt et al.,
diminish anxiety, whereas the excitation of CRF-containing 2000; Merighi, 2002; Millan, 2002a). Observations made
neurones would likely exert an opposite anxiogenic effect. in heterologously-expressed, pure populations of cloned
On the other hand, despite the significance of corticolim- receptors should not automatically be extrapolated to cere-
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 173

Table 20
Influence of acute administration of diverse classes of anxiolytic agent upon extracellular levels of monoamines in the frontal cortex of freely-moving
rats as compared to their actions in the Vogel Conflict Test
Class Drug Dose Route Change (%) (vs. control) = 0% Change (%) (vs. basal) = 0%
(mg/kg)
VCT 5-HT NA DA

BZP Ago Diazepam 10.0 i.p. +186∗ −29 ± 4∗ −21 ± 4∗ −34 ± 3∗


Triazolo-BZP Ago Triazolam 2.5 i.p. +348∗ −48 ± 8∗ +5 ± 8 −16 ± 3
5-HT1A Ago Flesinoxan 10.0 s.c. +266∗ −50 ± 6∗ +104 ± 19∗ +60 ± 8∗
5-HT2A Ant MDL100,907 0.16 s.c. +239∗ −1 ± 8 +18 ± 8 0 ± 5
5-HT2C Ant SB242,084 10.0 i.p. +102∗ +6 ± 8 +90 ± 30∗ +73 ± 10∗
SSRI Citalopram 10.0 s.c. −3 +280 ± 32∗ +75 ± 7∗ +30 ± 15
SNRI Venlafaxine 10.0 s.c. +19 +251 ± 19∗ +360 ± 33∗ +190 ± 29∗
NARI Reboxetine 10.0 i.p. +43 0 ± 7 +330 ± 31∗ +220 ± 36∗
5-HT2C /␣2 AR Ant Mirtazapine 10.0 s.c. +156∗ +10 ± 5 +330 ± 47∗ +160 ± 27∗
␣2 -AR Ago Dexmedetomidine 0.005 s.c. +168∗ −45 ± 6∗ −79 ± 10∗ −35 ± 7∗
CRF1 Ant CP154,526 40.0 i.p. +135∗ 0 ± 8 +10 ± 5 +2 ± 5
NK1 Ant GR205,171 40.0 i.p. +131∗ 0 ± 0 +77 ± 9∗ +63 ± 12∗
Abbreviations—BZP: benzodiazepine; SSRI: selective serotonin reuptake inhibitor; SNRI: serotonin and noradrenaline reuptake inhibitor; NARI: nora-
drenaline reuptake inhibitor; AR: adrenergic receptor; CRF: corticotropin releasing factor and NK: neurokinin. All data were generated in the author’s
laboratory under identical conditions (Gobert and Millan, 1999a; Gobert et al., 2000; Dekeyne et al., 2000a,b; Millan et al., 2000b,c,d,e, 2001a,b,c,d;
Lejeune and Milan, 2002; Gobert and Millan, unpublished observations). The doses of drugs examined correspond to those at which they exert significant
anxiolytic actions in the VCT, with the exception of the reuptake inhibitors which are inactive in this paradigm. For these antidepressant agents, doses
are equivalent to those at which they exert anxiogenic actions in the social interaction procedure and antidepressant actions in other procedures (Dekeyne
et al., 2000a; Millan et al., 2001c,d). Note the diverse pattern of influence of drugs upon the levels of various monoamines. Star indicates significant
differences vs. control or basal values: ∗ P < 0.05.

bral neurones inasmuch as receptors may be coupled in a in the modulation of anxious states (Bonhaus et al., 1998;
region- and neuron-specific manner to intracellular trans- Cordeaux et al., 2000; Kukkonen et al., 2001). Assuming
duction mechanisms ((Jin et al., 2001; Porter et al., 2002; contrasting functional correlates of alternative cellular path-
Blank et al., 2003a,b)). Moreover, at specific neurones (and ways of signal transduction, selective engagement of one
even synapses), multiple transmitters/receptors converge may permit improved dissociation of beneficial properties
upon common and interacting transduction mechanisms—a from untoward side-effects (Kenakin, 1995, 2002; Cussac
phenomenon sometimes referred to as “cross-talk”. Colo- et al., 2002b). This notion of “divergent” coupling mecha-
calized receptors may not only interact at the level of nisms cascading from a single receptor is, in a sense, a mir-
intracellular messengers, but even physically associate to ror image of the “convergence” of multiple receptors onto
form heterodimers with functional properties differing from a common tranduction signal, as outlined in the preceding
those of the respective homomers. For example, functional paragraph.
heterodimers of ␣2C -AR/M2 , 5-HT1A /5-HT1B , D2 /D3 , Third, receptors may be “constitutively-active”: that
␮-opioid/ORL1 , mGluR/A1 and adenosine A1 /mGluR1 re- is, they may show spontaneous, ligand-independent cou-
ceptors have all been shown to assemble, at least in cultured pling to G-proteins. Though the physiological pertinence
all lines. Though their suspected existence in neurones of this phenomenon at neuronal populations of receptor
remains to be unequivocally proven, one example of a func- (which are generally occupied by endogenous agonists) is
tional CNS-localized heterodimer is provided by GABAB still disputed, so-called “inverse agonists” exert actions at
receptors, which are composed of structurally-distinct constitutively-active receptors opposite to those of agonists,
GABAB1 and GABAB2 proteins (Section 2.3.1) (Maggio and the effects of both are blocked by “neutral” antago-
et al., 1993; Bouvier, 2001; Devi, 2001; Marshall, 2001; nists (Millan et al., 1999b; Morisset et al., 2000; Chavigne
Scarselli et al., 2001; George et al., 2002; Pan et al., 2002). et al., 2001). Heterologously-expressed 5-HT1A , 5-HT1B ,
Second, via “agonist-directed trafficking”, individual 5-HT2A and 5-HT2C receptors, ␣2A -ARs, histamine H3 re-
drugs may differentially favour the activation of one of sev- ceptors, CB1 receptor and M2 receptors provide examples
eral alternative intracellular cascades coupled to the same of constitutively-active receptors potentially relevant to anx-
receptor. For example, certain ligands differentially recruit ious states (Barker et al., 1994; Morisset et al., 2000; Murrin
the G-protein subtype Gi as compared to Gq via actions et al., 2000; De Ligt et al., 2000; Menzaghi et al., 2002;
at 5-HT1A or 5-HT2C receptors (Watson et al., 2000; Berg Newman-Tancredi et al., 2002a; Seifert and Wenzel-Seifert,
et al., 2001; Cussac et al., 2002b,c; Stout et al., 2002; 2002). Assuming the existence of constitutive activity at
Newman-Tancredi et al., 2002a; Blank et al., 2003a). Traf- receptor types contributing to the generation of anxious
ficking has also been reported at, for example, ␣2A -ARs, states, inverse agonists could, in theory, prove to be more
CB1 receptors and A1 receptors, which likewise play a role efficacious anxiolytic agents than neutral antagonists.
174 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

As for other fields of neurobiological research, these and Table 21


related concepts, despite their still uncertain pathophysio- Influence of acute administration of diverse classes of anxiolytic agent in
the Vogel Conflict Test as compared to other procedures
logical and therapeutic significance, will certainly enrich re-
search into the causes and cures of anxiety disorders. Class Drug VCT SINT PM USV

BZP Ago Diazepam + + + +


18.3. Long-term actions of anxiolytic agents Triazolo-BZP Ago Triazolam + + + +
5-HT1A Ago Flesinoxan + + IA +
Relatively little information is available concerning the 5-HT2A Ant MDL100,907 + IA IA IA
5-HT2C Ant SB242,084 + + IA IA
long-term actions of drugs in the VCT and other pro- SSRI Citalopram IA − − IA
cedures. Obviously, the potential development of toler- SNRI Venlafaxine IA − − IA
ance/desensitization to the anxiolytic actions of an agonist NARI Reboxetine IA − − IA
upon its recurrent administration is an issue of fundamental 5-HT2C /␣2 -AR Ant Mirtazapine + + IA +
importance. Moreover, upon suspension of chronic drug ␣2 -AR Ago Dexmedetomidine + IA IA +
CRF1 Ant CP154,526 + + IA IA
administration, the possible induction of a withdrawal syn- NK1 Ant GR205,171 + IA IA IA
drome (including rebound anxiety) necessitates evaluation.
A related issue requiring experimental investigation is the Abbreviations—SINT: social interaction test; PM: elevated plus-maze test;
USV: ultrasonic vocalization test; BZP: benzodiazepine; SSRI: selective
potential consequences of long-term pre-treatment with one serotonin reuptake inhibitor; SNRI: serotonin and noradrenaline reuptake
drug upon the subsequent effects of a second agent. This inhibitor; NARI: noradrenaline reuptake inhibitor; AR: adrenergic recep-
mirrors the clinical scenario whereby the discontinuation tor; CRF: corticotropin releasing factor; NK: neurokinin; MCH: melanin
of one drug (generally due to poor efficacy or unaccept- concentrating hormone; +, anxiolytic; −, anxiogenic; IA: inactive. All
able tolerance) is followed by the introduction of a second data were generated in the author’s laboratory under identical conditions
(Dekeyne et al., 2000a,b; Millan et al., 1997, 1998, 2000e, 2001a,b;
drug (often) of a different mechanistic class. For example, Brocco and Millan, unpublished observations). The tests are briefly out-
therapeutic trials revealed that buspirone aggravates the lined in Table 1. Note the diverse pattern of anxiolytic profiles.
withdrawal syndrome to BZPs, and that it is virtually in-
effective in patients recently treated with these agents (De
Martinis et al., 2000). For the appropriate design of clinical eral models of untrained behaviours are also well-adapted
trials, and for the switching of drugs in clinical practice, for the detection of anxiogenic properties: notably the
such information is essential. plus-maze and the social interaction procedures (Table 1
and Section 1.3). The ability of such experimental protocols
18.4. Detection of anxiogenic and anxiolytic properties to detect heightened anxiety is also a precondition for their
exploratory use in the characterization of models of “trait”
In studies where the actions of drugs are not, a priori, anxiety, such as novel populations of genetically-engineered
known, it is important that the parameters of the VCT and mice. Indeed, such an approach allows for the demonstra-
other paradigms be adjusted to render them responsive to tion that potential therapeutic agents can “normalize” aber-
both anxiolytic and anxiogenic agents (e.g. Sepinwall and rant reactivity to stressful/fear-inducing stimuli or exces-
Cook, 1980; Liljequist and Engel, 1984a,b; Meneses and sive “spontaneous” anxiety, rather than merely decreasing
Hong, 1993; Olivier et al., 2000). Otherwise, valuable in- anxiety below a “normal” baseline.
formation may be lost. For example, if under appropriate
conditions, anxiogenic actions of SSRIs were expressed in 18.5. Comparing actions in the VCT to other models
the VCT, it would provide an instructive model for study
of their delayed onset of anxiolytic efficacy (Section 4.4). As mentioned towards the beginning of this article
However, the VCT has not, regrettably, been extensively (Section 1.4) and summarized in Table 1, the VCT is a
employed for the detection of anxiogenic properties whereas paradigm of transient, “state” anxiety which incorporates
a related paradigm of “safety signal withdrawal” has proven a punishment component and which usually requires the
especially instructive in this regard. This model measures performance of a conditioned behaviour. Despite its broad
the influence of drugs upon food-rewarded responses upon utility and compelling predictive, face and construct valid-
the unexpected suppression of a “safety signal” (cue light) ity, like all other models currently available, it cannot be
indicating that the subsequent session will not be punished. considered alone as sufficient for the full characterization of
Thus, identical conditions can be used to characterize both any of the mechanisms discussed herein. Further, by analogy
anxiolytic and anxiogenic drugs and this intriguing proce- to other conflict models of response suppression, despite
dure has revealed robust anxiogenic actions of, for example, its empirical utility for drug characterization, it should not
adenosine antagonists, GABAA antagonists and inverse ag- be regarded as a “model” of clinical anxiety per se. Use of
onists at BZP receptors (Sections 2.2.2.1 and 7) (Thiébot the VCT must, then, be embedded in an experimental strat-
et al., 1991; Charrier et al., 1994). Upon appropriate adjust- egy incorporating complementary models reflecting other
ment of experimental conditions (essentially to moderate aspects of anxious states. Within this perspective, Table 21
the intensity of anxiety experienced during the test), sev- illustrates the strikingly diverse profiles of functional ac-
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 175

tivity displayed by various classes of agent in the VCT as pected to exert anxiolytic properties. An instructive anomaly
compared to other models of potential anxiolytic activity is provided by work with CB1 receptor-deficient mice in
performed in parallel under identical conditions. which the actions of selective antagonists persisted, sug-
gesting the existence of additional, central “CB1 ” receptors
18.6. Complementary pharmacological and genetic controlling anxious states (Section 8).
strategies These comments serve to illustrate the synergistic and
indispensable nature of genetic and pharmacological ap-
Genetic (knock-out and transgenic) strategies have been proaches to the characterization of anxious states and their
enthusiastically embraced by many groups, and they proven management.
extremely powerful tools in the identification and charac-
terization of mechanisms for the control of anxious states. 18.7. Neuronal networks and anxious states
However, the extraordinary proliferation (see Section 1.1)
of “anxiogenic” (and, less frequently, “anxiolytic”) pheno- As argued in Section 1.5.3, it would be naı̈ve to conceive
types upon disruption of diverse genes raises the interrelated of the amygdala or of any other corticolimbic region as act-
questions of (1) whether the changes seen are genuinely in- ing in isolation in the induction, modulation and experience
dicative of a key role in the control of anxious states, rather of anxious states in view of the intermeshing of function
than a modulatory role of perhaps subsidiary importance— across diverse structures and modulators. Nonetheless, as in
at least from a clinical point of view (Section 18.10); (2) other spheres of biological research, the field of anxiety is
whether changes in anxious behaviour observed upon modi- dominated by a reductionist drive to disaggregate systems
fication of gene function are truly indicative of physiological into their individual components. This is explained by the
and pathological significance when operational—we mea- relative intractability of “systems” as units of study—despite
sure behaviour in their absence not presence; (3) whether the utility of computer models of network organization, func-
alterations in behaviour in experimental models of stress, tional imaging of brain function in man, genetic elimination
fear and anxiety, etc. are primary events specifically reflect- of specific transmitters and other useful approaches. Thus, an
ing emotionality rather than an epiphenomenon of other ac- inevitable and initial step in understanding the operation of
tions relevant to performance under these conditions, such systems committed to the control of anxious states remains
as cognitive-attentional function, motor behaviour, or im- the “top down” approach which focuses successively on cir-
pulsivity (Section 18.9). cumscribed cerebral structures, their sub-territories, specific
These questions should not, of course, be construed as classes of neurone, discrete transmitters, individual classes
uniquely applicable to such genetic strategies—they are of receptor, their subunits or isoforms, coupling to defined
no less pertinent to studies of, for example, discrete brain intracellular signals, etc. (Of these elements, receptors re-
lesions or chronic administration of antagonists. Neverthe- main, arguably, the lynch-pin of research in view of their
less, after a decade or so of exploitation, we are currently accessibility as targets for therapeutic agents). However, the
witnessing a “demystification” of genetic strategies with conceptual re-assembly of these components into functional
the realization that these approaches, like all others, have networks poses major analytical difficulties. Moving from
their advantages and disadvantages and present certain the bottom up, in a non-linear fashion, each higher stratum
inherent interpretational difficulties—genetic background, will yield “emergent” properties not immediately apparent
strain differences and developmental compensation, etc. from study of its constituents and their interactions—and
(Section 1.1). Indeed, while there is clearly no such thing as more than the sum total of their properties. This uncertainty
a definitive pharmacological study, even the most parochial is compounded by the enormous degree of redundancy of
champion of genetic strategies would now accept that no mechanisms for the control of anxious states (Section 18.9).
single study of knock-out or transgenic mice, irrespective For example, from this holistic perspective, comprehensive
of its strengths, is sufficient to unequivocally define the role knowledge (even if possible) of the individual roles of all
of any mechanism for the control of anxious states. 5-HT or NPY receptor subtypes cannot reveal the overall
Nevertheless, more sophisticated approaches encompass- significance of these transmitters in the response to and con-
ing (for example) conditioned (“inducible”), region-specific trol of anxious states. Likewise, a complete description of
gene deletion and point mutations, together with more thor- the actions and interactions of each neurochemically and
ough controls of the specificity and generality of changes ob- anatomically-defined class of neuron in the amygdala will
served, continue to strengthen the use of genetic approaches not suffice to understand the global role of the amygdala per
to the evaluation of mechanisms underlying the etiology se in the modulation of anxious states. The major challenges
and management of anxious states. Further, the characteriza- remain then, of: (1) understanding the global function and
tion in genetically-modified mice of pharmacological agents operating properties of networks incorporating diverse brain
“selective” for the gene under study offers an elegant and regions and modulators and (2) making realistic predictions
mutual control for these complementary pharmacological about the clinical consequence of manipulating individual
and genetic strategies. For example, in mice lacking NK1 units of broad-based systems for the control of anxious be-
receptors, highly selective NK1 antagonists would not be ex- haviours.
176 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

18.8. Anxious states and cognitive function 18.9. Redundancy in the control of anxious states

In Section 2.2.2.6, within the framework of developing Recent research has implicated an extravagant repertoire
novel ligands for subunits of GABAA receptors, it was of modulators in the response to stress and fear, and in the
pointed out that there is a reciprocal and—unfortunately— modulation of anxious states. This panoply of mechanisms
negative interrelationship between anxious states and sleep: will inexorably be further increased by ongoing studies of
accordingly, by analogy to other neuropsychiatric indica- novel genes derived from the human genome. Indeed, in
tions, it is crucial that drugs exert a positive (or, at least, no line with other psychiatric diseases, GAD, social phobias
deleterious) influence upon sleep. and other anxiety disorders are: (1) polygenic (determined
As evoked throughout this article, there is also an intimate, by multiple genes), (2) reflect a complex interplay of hered-
complex and reciprocal pattern of interactions between itary, developmental and environmental factors and (3)
anxious states and cognitive function which are integrated involve the aberrant functioning of a diversity of cerebral
and modulated in common structures, such as the amyg- mechanisms (Heim and Nemeroff, 1999; Scherrer et al.,
dala, hippocampus and cortex (Section 1.5) (Clément 2000; Gratacos et al., 2001; Hettema et al., 2001; Stein
and Chapoutier, 1998; LeDoux, 2000; Sarter et al., 2001; et al., 2001a,b; Clément et al., 2002; Freimer et al., 2002;
Carter et al., 2003; McGaugh et al., 2002; McKell-Carter Glatt and Freimer, 2002). Correspondingly, there exists an
et al., 2003; Ohl et al., 2003). Indeed, it might seriously be impressively large palette of potential targets for the im-
questioned whether any mechanism which profoundly influ- proved treatment of anxiety disorders. This raises several
ences one of these functions will not, at least indirectly, im- points.
pact upon the other. In this light, it is interesting that certain First, like all crucial, existential functions integrated cen-
classes of anxiolytic agent exert a detrimental influence upon trally (Section 1.5.1), such as pain, appetite and motor be-
cognitive function: most strikingly, conventional BZPs elicit haviour, anxious states are subject to a plethora of concurrent
a marked anterograde amnesia epitomized by impaired in- positive and negative influences. Indeed, it is arguable that
formation acquisition and recall (Lister, 1985; Clément and the most effective means to guarantee the operation and sta-
Chapoutier, 1998; Mitler, 2000). NMDA receptor antago- bility of critical, adaptive functions is by a hierarchy of mul-
nists provide a further example of anxiolytic agents which tiple, opposing controls. (This axiom is probably universal
compromise cognitive function (Danysz and Parsons, 1998; as a “design principle” extending from cells—transduction
Tang et al., 2001). However, there are certain exceptions in- mechanisms and cycles of cell growth—though organs,
dicating that the phenonema can be dissociated: CRF2 and whole organisms and societies to ecosystems, etc.). How-
NK1 receptor antagonists, as well as 5-HT1A receptor partial ever, several fundamental questions still remain: (1) which
agonists, for example, exert anxiolytic properties without mechanisms control the equilibrium between intrinsic anxio-
compromising mnesic function (Maggi, 1995; Meneses, genic and anxiolytic mechanisms, and how is the “set-point”
1999; Steckler and Holsboer, 1999; Blank et al., 2003b), determined for a specific level of reactivity to stress and
while prototypical amnesic agents such as muscarinic an- fear; (2) why do these tightly-regulated homeostatic mecha-
tagonists clearly do not present a consistent and robust nisms sometimes fail, leading to the development of anxiety
anxiolytic profile (Section 6.3). disorders; (3) what is the precise contribution of hereditary,
More generally, as pointed out in Section 2.2.2.6, it developmental (aversive events during childhood or ado-
must not be naively assumed that an “enhancement” of lescence) and environmental (transitory and/or repetitive
cognitive-attentional function will necessarily be favourable exposure to banal and/or extreme stressors) factors to their
for an anxiolytic agent. From an empirical, patient-based pathogenesis and (4) from a phylogenetic perspective, how
perspective, many may actually wish (at least temporarily) to and when did such an astonishing profusion of controls
“forget”. In addition, a disruption of processes of extinction evolve?
is a cardinal feature of certain classes of anxiety disorder Second, with respect to the latter point (4) many media-
(Section 1.5.1). From a conceptual point of view, moreover, tors responsive to fear and stress influence other functions,
it has been suggested that excessive attention and vigilance such as cognition, motor behaviour, nociception and en-
may be maladaptive in amplifying the anxiogenic effects of docrine secretion, in addition in emotion. In certain cases,
fear-inducing stimuli, a process which may also secondarily modulation of mood may constitute a subsidiary rather than
compromise the performance of cognitive tasks (Clément a primary function. In this light, it is possible that the evolu-
and Chapoutier, 1998; LeDoux, 2000; Sarter et al., 2001). tion of multiple mechanisms for the control of anxious states
The roles of specific mechanisms discussed herein in the illustrates the principle whereby a modulator which plays
control of mnesic processes in relation to their modulation one functional and adaptive role, say in mnesic processes
of anxious states is clearly a critical issue requiring, in most related to conditioning, is then “co-opted” for a further role,
cases, considerable further study. In this light, as for sleep such as control of mood, under similar conditions. Though
patterns, it is imperative that for all drugs intended for the virtually unexplored, this phenomenon of “exaptation”, at
clinical control of anxious states, their influence upon cog- the level of intracellular signals, receptors, neurotransmit-
nitive function be carefully evaluated. ters, neurones and cerebral circuits, is arguably fundamen-
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 177

tal to the evolution of multiple and redundant controls of of action may emerge to be particularly effective therapeu-
anxiety and other critical CNS functions. tic agents in the control of anxiety disorders. For example,
Third, one positive aspect of the ongoing (genome-driven) drugs which reinforce endogenous mechanisms countering
effort to identify novel mechanisms for the modulation anxious states while concurrently inhibiting those responsi-
of anxious state is that we can now realistically aspire to ble for their induction.
possess all the pieces of the metaphorical jigsaw puzzle
underlying anxious disorders and their treatment. However, 18.10. Clinical validation of hypotheses
rather than a static, invariant and two-dimensional image,
the brain is a dynamic, flexible and multidimensional entity In view of the intricate and dynamic pattern of recipro-
(temporally as well as spatially) and all of these pieces (mod- cal interactions amongst mechanisms controlling anxious
ulators) reciprocally interact and possess variable “forms” states, the clinical consequences of their manipulation
(functional properties). This point underpins the argument are difficult to predict. Thus, despite the utility of the
made above (Section 18.7) that studies of individual con- VCT and other experimental models, considerable caution
stituents of networks of controlling anxious states—though should be exercised in extrapolating hypotheses to man.
crucial—are unlikely to provide a full understanding of This difficulty is compounded by the heterogeneous and
mechanisms underlying their pathogenesis and treatment. co-morbid nature of anxiety disorders in patients, and by
Further, apart from the “rational” element drug of discovery the more complex nature of cognitive and emotional fac-
(genome →gene →protein →“lead”), the creative “reality” tors in man as compared to other species. Very few of the
of target validation (elucidation of function(s)—likely mechanisms discussed herein have been therapeutically as-
multiple, and both beneficial and deleterious), transposi- sessed and considerable feedback from the clinic will be
tion of “leads” into real drugs and their thorough clinical necessary to confirm, refine or refute hypotheses, and to
validation still remains the essence of innovative and suc- focus research strategies aimed at the improved manage-
cessful drug development for anxiety and other disorders. ment of anxious states. Indeed, bearing in mind the sheer
In this respect, the imaginative exploitation of the VCT abundance of mechanisms which control anxious states,
and other experimental models, and both pharmacological as implied above, the challenge will be to identify by
and genetic strategies, will continue to play a vital role in well-designed experimental and therapeutic studies those
ongoing efforts to ameliorate the management of anxiety targets of which the manipulation (either alone or together
disorders. with others) will lead to efficacious and well-tolerated
Fourth, certain mechanisms discussed herein may make agents for the rapid and sustained management of anxiety
only subliminal, minor contributions to the overall induc- disorders.
tion or suppression of clinical anxious states. It is essential, Inasmuch as many mechanisms evoked herein await clin-
thus, to identify those mechanisms which fulfil crucial and ical feedback—while experimental data are often scarce—
clinically-relevant roles in the generation of anxious states it would appear premature to emphatically assert which
and, correspondingly, represent genuine targets for their ther- would be currently the most likely to offer optimal man-
apeutic relief. agement of clinical anxiety in terms of: enhanced efficacy,
Fifth, a related corollary of the striking redundancy of rapid onset and sustained duration of action in the absence
endogenous mechanisms either favouring or opposing anx- of dependence, preservation of cognitive function and sleep
ious states is that drugs interacting with a single component patterns, and acceptable tolerance. In any event, it is un-
may be insufficient to achieve robust and broad therapeutic likely that any single mechanism will offer a panacea for
efficacy. BZPs and SSRIs, the most widely-used drugs for the improved treatment of all anxiolytic states. This tenet
clinical treatment of anxiety disorders, may appear to con- applies in particular to drugs with highly-selective mecha-
tradict this assertion in view of their selectivity for GABAA nisms of action—the current tendency in exploiting innova-
receptors and 5-HT transporters respectively. However, re- tive, genome-derived targets.
flecting the operation of GABAergic inhibition throughout Unambiguous clinical proof of anxiolytic properties for
the CNS, BZPs effectively act at diverse corticolimbic loci specific classes of agent may prove more difficult to ac-
(Sections 2.2.2.2 and 18.1). Further, in view of the contrast- quire than might be imagined since differences between
ing subunit composition of various populations of GABAA active drugs and placebo in the treatment of anxious (and
receptor, one might even conceive BZPs as “multi-target” depressive) states are frequently unclear. Such ambivalent
agents. As for SSRIs—and other classes of antidepressant outcomes do not necessarily undermine the validity of the
agent—they indirectly recruit a diversity of receptorial and drug and the hypothesis under evaluation but, as discussed
adaptive mechanisms (Section 4.4), including 14 classes of elsewhere (Amsterdam and Brunswick, 2002; Greist et al.,
postsynaptic 5-HT receptor. Thus, in analogy to the use of 2002; Katz et al., 2002; Khan et al., 2002; Klein et al.,
multireceptorial agents such as clozapine in the treatment 2002; Mayberg et al., 2002), raise awkward questions about
of a further multifactorial disease, schizophrenia (Brunello the conditions under which therapeutic trials are currently
et al., 1995; Meltzer, 1995, 1999; Millan, 2000; Chakos being undertaken in psychiatry, and about our ability to
et al., 2001), agents possessing dual or multiple mechanisms definitively validate or refute concepts in man.
178 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

19. Concluding comments Adell, A., Casanovas, J.M., Artigas, F., 1997. Comparative study in the
rat of the actions of different types of stress on the release of 5-HT
in raphe nuclei and forebrain areas. Neuropharmacology 36, 735–741.
In parallel with other experimental approaches, the VCT Adell, A., Celada, P., Abellan, M.T., Artigas, F., 2002. Origin and
has proven highly instructive in exploring the multitude of functional role of the extracellular serotonin in the midbrain raphe
mechanisms implicated in the control of anxious states. As nuclei. Brain Res. Rev. 39, 154–180.
emphasized herein, as part of a broader programme of re- Adityanjee, M.D., Schulz, S.C., 2002. Clinical use of quetiapine in
disease states other than schizophrenia. J. Clin. Psychiatry 63, 32–38.
search, there remains considerable scope for its continued
Adkins, C.E., Pillai, G.V., Kerby, J., Bonnert, T.P., Haldon, C., McKernan,
utilization in the identification and characterization of novel R.M., Gonzalez, J.E., Oades, K., Simpson, P.B., 2001. ␣4 ␤3 ␦
classes of anxiolytic agent. Enormous progress has been GABA(A) receptors characterized by fluorescence resonance energy
achieved in recent years as concerns our knowledge of en- transfered measurements of membrane potential. J. Biol. Chem. 276,
dogenous mechanisms for the modulation (enhancement and 38934–38939.
Aggleton, J.P., 1993. The contribution of the amygdala to normal and
suppression) of anxious states. The major challenge remains abnormal emotional states. Trends Neurosci. 16, 328–333.
of translating these gains in our understanding of the plural- Ågmo, A., Pruneda, R., Guzman, M., Gutiérrez. 1991. GABAergic drugs
ity of neuronal substrates underlying anxiety disorders into and conflict behavior in the rat: lack of similarities with the actions
tangible benefits for patients. of benzodiazepines. Naunyn Schmiedebergs Arch. Pharmacol. 344,
314–322.
Ågmo, A., Galvan, A., Heredia, A., Morales, M., 1995. Naloxone
blocks the antianxiety but not the motor effects of benzodiazepines
Acknowledgements and pentobarbital: experimental studies and literature review.
Psychopharmacology 120, 186–194.
Aguiar, M.S., Brandao, M.L., 1996. Effects of microinjections of the
The author would like to thank M. Soubeyran, M. Brocco,
neuropeptide substance P in the dorsal periaqueductal gray on the
A. Gobert, A. Dekeyne, B. Di Cara, J.-M. Rivet and S. behaviour of rats in the plus-maze test. Physiol. Behav. 60, 1183–1186.
Dapremont for invaluable assistance in the preparation of Aguilera, G., Radadan-Diehl, C., 2000. Vasopressinergic regulation
this manuscript. G. Dawson, B. Olivier and T. Sharp are of the hypothalamic–pituitary–adrenal axis: implications for stress
thanked for helpful comments on the manuscript. adaptation. Regul. Pept. 96, 23–29.
Ahmed, I., Takeshita, J., 1996. Clonidine: a critical review of its role in
the treatment of psychiatric disorders. CNS Drugs 6, 53–70.
Akbarian, S., Bates, B., Liu, R.-J., Skirboll, S.L., Pejchal, T., Coppola,
References V., Sun, L.D., Fan, G., Wilson, M.A., Tessarollo, L., Kosofsky, B.E.,
Taylor, J.R., Bothwell, M., Nestler, E.J., Aghajanian, G.K., Jaenisch,
Abadie, P., Boulenger, J., Benali, K., Brre, L., Zarifian, E., Baron, J., R., 2001. Neurotrophin3 modulates noradrenergic neuron function and
1999. Relationships between trait and state anxiety and the central opiate withdrawal. Mol. Psychiatry 6, 593–604.
benzodiazepine receptor: a PET study. Eur. J. Neurosci. 11, 1470–1478. Akbarian, S., Rios, M., Liu, R.-J., Gold, S.J., Fong, H.-F., Zeiler, S.,
Abellan, M.T., Jolas, T., Aghajanian, G.K., Artigas, F., 2000. Dual Coppola, V., Tessarollo, L., Jones, K.R., Nestler, E.J., Aghajanian,
control of dorsal raphe serotonergic neurons by GABAB receptors. G.K., Jaenisch, R., 2002. Brain-derived neurotrophic factor is essential
Electrophysiological and microdialysis studies. Synapse 36, 21–34. for opiate-induced plasticity of noradrenergic neurons. J. Neurosci.
Abi-Saab, W., Bubser, M., Roth, R.H., Deutch, A.Y., 1999. 5-HT2 22, 4153–4162.
Akirav, L., Richter-Levin, G., 1999. Biphasic modulation of hippocampal
receptor regulation of extracellular GABA levels in the prefrontal
plasticity by behavioral stress and basolateral amygdala stimulation
cortex. Neuropsychopharmacology 21, 92–96.
in the rat. J. Neurosci. 19, 10530–10535.
Abrams, J.K., Johnson, P.L., Shekhar, A., Lowry, C.A., 2002. Different
Akunne, H.C., Whetzel, S.Z., Wiley, J.N., Corbin, A.E., Ninteman,
anxiogenic drugs activate a common, topographically distinct
F.W., Tecle, H., Pei, Y., Pugsley, T.A., Heffner, T.G., 1997. The
subpopulation of serotonergic neurones in the rat dorsal raphe nucleus.
pharmacology of the novel and selective sigma ligand, PD 144418.
Soc. Neurosci. Abstr. 75, 1.
Neuropharmacology 36, 51–62.
Accili, D., Fishburn, C.S., Drago, J., Steiner, H., Lachowitcz, J.E.,
Akwa, Y., Purdy, R.H., Koob, G.F., Britton, K.T., 1999. The amygdala
Park, B.H., Gauda, E.B., Lee, E.J., Cool, M.H., Sibley, D.R., Gerfen,
mediates the anxiolytic-like effect of the neurosteroid allopregnanolone
C.R., Westphal, H., Fuchs, S., 1996. A targeted mutation of the D3
in rat. Behav. Brain Res. 106, 119–125.
dopamine receptor gene is associated with hyperactivity in mice. Proc. Alagarsamy, S., Marino, M.J., Rouse, S.T., Gereau, R.W., Heinemann,
Natl. Acad. Sci. U.S.A. 93, 1945–1949. S.F., Conn, P.J., 1999. Activation of NMDA receptors reverses
Acquas, E., Wilson, C., Fibiger, H.C., 1996. Conditioned and desensitization of mGluR5 in native and recombinant systems. Nat.
unconditioned stimuli increase frontal cortical and hippocampal Neurosci. 2, 234–240.
acetylcholine release: effects of novelty, habituation and fear. J. Albeck, D.S., McKittrick, C.R., Blanchard, D.C., Blanchard, R.J.,
Neurosci. 16, 3089–3096. Nikulina, J., McEwen, B.S., Sakai, R.R., 1997. Chronic social stress
Adan, R.A.H., Gispen, W.H., 2000. Melanocortins and the brain: from alters levels of corticotropin-releasing factor and arginine vasopressin
effects via receptors to drug targets. Eur. J. Pharmacol. 405, 13–24. mRNA in rat brain. J. Neurosci. 17, 4895–4903.
Adan, R.A.H., Szklarczyk, A.W., Oosterom, J., Brakkee, J.H., Alfarez, D.N., Wiegert, A., Joëls, M., Krugers, H.J., 2002. Corticosterone
Nijenhuis, W.A., Schaaper, W.M., Meloen, R.H., Gispen, W.H., 1999. and stress reduce synaptic potentiation in mouse hippocampal slices
Characterization of melanocortin receptor ligands on cloned brain with mild stimulation. Neuroscience 115, 1119–1126.
melanocortin receptors and on grooming behavior in the rat. Eur. J. Alger, B.E., 2002. Retrograde signaling in the regulation of synaptic
Pharmacol. 378, 249–258. transmission: focus on endocannabinoids. Prog. Neurobiol. 68, 247–
Addolorato, G., Caputo, F., Capristo, E., Domenicali, M., Bernardi, M., 286.
Janiri, L., Agabio, R., Colombo, G., Gessa, G.L., Gasbarrini, G., 2002. Al-Hallaq, R.A., Jarabek, B.R., Fu, Z., Vicini, S., Wolfe, B.B., Yasuda,
Baclofen efficacy in reducing alcohol craving and intake: a preliminary, R.P., 2002. Association of NR3A with the N-methyl-d-aspartate
double-blind, randomized controlled study. Alcohol 37, 504–508. receptor NR1 and NR2 subunits. Mol. Pharmacol. 62, 1119–1127.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 179

Alkondon, M., Braga, M.F.M., Pereira, E.F.R., Maelicke, A., Andrews, N., File, S.E., 1993. Handling history of rats modifies
Alburquerque, E.X., 2000. ␣7 -Nicotinic acetylcholine receptors and behavioural effects of drugs in the elevated plus-maze test of anxiety.
modulation of GABAergic synaptic transmission in the hippocampus. Eur. J. Pharmacol. 235, 109–112.
Eur. J. Pharmacol. 393, 59–67. Andrews, N., Hogg, S., Gonzalez, L.E., File, S.E., 1994. 5-HT1A
Allen, A.M., Moeller, I., Jenkins, T.A., Zhuo, J., Aldred, G.P., Chai, receptors in the median raphe nucleus and dorsal hippocampus may
S.Y., Mendelshon, F.A.O., 1998. Angiotensin receptors in the nervous mediate anxiolytic and anxiogenic behaviours respectively. Eur. J.
system. Brain Res. Bull. 47, 17–28. Pharmacol. 264, 259–264.
Alleva, E., Aloe, L., 1989. Physiological roles of NGF in adult rodents: Anger, T., Madge, D.J., Mulla, M., Riddall, D., 2001. Medicinal chemistry
a biobehavioral perspective. Int. J. Comp. Psychol. 2, 147–163. of neuronal voltage-gated sodium channel blockers. J. Med. Chem.
Alleva, E., Santucci, D., 2001. Psychosocial vs. “physical” stress 44, 115–137.
situations in rodents and humans: role of neurotrophins. Physiol. Anji, A., Kumari, M., Hanley, S.N.R., Bryan, G.L., Hensler, J.G., 2000.
Behav. 73, 313–320. Regulation of 5-HT2A receptor mRNA levels and binding sites in
Allgaier, C., Scheibler, P., Müller, D., Feuerstein, T.J., Illes, P., 1999. rat frontal cortex by the agonist DOI and the antagonist mianserin.
NMDA receptor characterization and subunit expression in rat cultured Neuropharmacology 39, 1996–2005.
mesencephalic neurones. Br. J. Pharmacol. 126, 121–130. Anthony, E.W., Nevins, M.E., 1993. Anxiolytic-like effects of N-methyl-
Allgulander, C., Hackett, D., Salinas, E., 2001. Venlafaxine extended d-aspartate-associated glycine receptor ligands in the rat potentiated
release (ER) in the treatment of generalised anxiety disorder. Br. J. startle test. Eur. J. Pharmacol. 250, 317–324.
Psychiatry 179, 15–22. Anthony, J.P., Sexton, T.J., Neumaier, J.F., 2000. Antidepressant induced
Allikmets, L., Pruus, K., Matto, V., 2002. Do the serotonin receptors regulation of 5-HT1B mRNA in rat dorsal raphe nucleus reverses
mediate anxiogenic-like effect of antidepressants in rats? Eur. rapidly following drug discontinuation. J. Neurosci. Res. 61, 82–87.
Neuropsychopharmacol. 12, S190–S191. Antoni, F.A., 1993. Vasopressinergic control of the pituitary
Allison, C., Pratt, J.A., 2003. Neuroadaptive processes in GABAergic adrenocorticotropin secretion comes of age. Front. Neuroendocrinol.
and glutamatergic systems in benzodiazepine dependence. Pharmacol. 14, 76–122.
Ther. 98, 171–195. Antonijevic, I.A., Murck, H., Bohlhalter, S., Friboes, R.M., Holsboer, F.,
Aloe, L., Alleva, E., Fiore, M., 2002. Stress and nerve growth factor Steiger, A., 2000. Neuropeptide Y promotes sleep and inhibits ACTH
findings in animal models and humans. Pharmacol. Biochem. Behav. and cortisol release in young men. Neuropharmacology 39, 1474–1481.
73, 159–166. Anwyl, R., 1999. Metabotropic glutamate receptors: electrophysiological
Alonso, G., Phan, V.L., Guillemain, I., Saunier, M., Legrand, A., Anoal, properties and role in plasticity. Brain Res. Rev. 29, 83–120.
Aoki, C., Pickel, V.M., 1990. Neuropeptide Y in cortex and
M., Maurice, T., 2000. Immunocytochemical localization of the
striatum: ultrastructural distribution and coexistence with classical
sigma1 receptor in the adult rat central nervous system. Neuroscience
neurotransmitters and neuropeptides. Ann. N. Y. Acad. Sci. 611,
97, 155–170.
186–205.
Altar, C.A., 1999. Neurotrophins and depression. Trends Pharmacol. Sci.
Apter, J.T., Allen, L.A., 1999. Buspirone: future directions. J. Clin.
20, 59–61.
Psychopharmacol. 19, 86–93.
Altar, C.A., Siuciak, J.A., Tran, T.M., Tepper, J.M., 1996. Local infusion
Arai, A.C., Xia, Y.-F., Rogers, G., Lynch, G., Kessler, M., 2002.
of BDNF modifies the pattern of dorsal raphe serotonergic neurons.
Benzamide-type AMPA receptor modulators form two subfamilies with
Brain Res. 712, 293–298.
distinct modes of action. J. Pharmacol. Exp. Ther. 303, 1075–1085.
Altman, J., Everitt, B.J., Glautier, S., Markou, A., Nutt, D., Oretti, R.,
Araki, H., Suemaru, K., Gomita, Y., 2002. Neuronal nicotinic receptor
Phillips, G.D., Robbins, T.W., 1996. The biological, social and clinical
and psychiatric disorders: functional and behavioral effects of nicotine.
bases of drug addiction: commentary and debate. Pschopharmacology
Jpn. J. Pharmacol. 88, 133–138.
125, 285–345.
Arborelius, L., Skelton, K., Thrivikraman, K., Plotsky, P., Schulz, D.,
Altman, J.D., Trendeleburg, A.U., MacMillan, L., Bernstein, D., Limbird,
Owens, M., 2000. Chronic administration of the selective corticotropin
L., Starke, K., Kobilka, B.K., Hein, L., 1999. Abnormal regulation of
releasing factor1 receptor antagonist CP-154,526: behavioral, endocrine
the sympathetic nervous system in ␣2A -adrenergic receptor knock-out
and neurochemical effects in the rat. J. Pharmacol. Exp. Ther. 294,
mice. Mol. Pharmacol. 56, 154–161.
588–597.
Amano, M., Goto, A., Sakai, A., Achiha, M., Takahashi, N., Hara, C., Archer, J., 1975. Rodent sex differences in emotional and related
Ogawa, N., 1993. Comparison of the anticonflict effect of buspirone behavior. Behav. Biol. 14, 451–479.
and its major metabolite 1-(2-pyrimidinyl)-piperazine (1-PP) in rats. Arevalo, C., De Miguel, R., Hernandez-Tristan, R., 2001.
Jpn. J. Pharmacol. 61, 311–317. Cannabinoid effects on anxiety-related behaviours and hypothalamic
Amara, S.G., Fontana, A.C.K., 2002. Excitatory amino acid transporters: neurotransmitters. Pharmacol. Biochem. Behav. 70, 123–131.
keeping up with glutamate. Neurochem. Int. 41, 313–318. Argiola, A., 1999. Neuropeptides and sexual behavior. Neurosci.
Amat, J., Matus-Amat, P., Watkins, L.R., Maier, S.F., 1998. Escapable Biobehav. Rev. 23, 1127–1142.
and inescapable stress differentially alter extracellular levels of 5-HT Ariano, M.A., Wang, J., Noblett, K.L., Larson, E.R., Sibley, D.R., 1997.
in the basolateral amygdala of the rat. Brain Res. 812, 113–120. Cellular distribution of the rat D2B receptor in central nervous system
Ameri, A., 1999. The effects of cannabinoids on the brain. Prog. using anti-receptor antisera. Brain Res. 746, 141–150.
Neurobiol. 58, 315–348. Arnsten, A.F.T., 1997. Catecholamine regulation of the prefrontal cortex.
Amin, J., 1999. A single hydrophobic residue confers barbiturate J. Psychopharmacol. 11, 151–162.
sensitivity to ␥-aminobutyric acid type C receptor. Mol. Pharmacol. Arnsten, A.F.T., Murphy, B., Merchant, K., 2000. The selective dopamine
55, 411–423. D4 receptor antagonist, PNU-101387G, prevents stress-induced
Amsterdam, J.D., Brunswick, D.J., 2002. Site variability in treatment cognitive deficits in monkeys. Neuropharmacology 23, 405–410.
outcome in antidepressant trials. Prog. Neuropsychopharmacol. Biol. Aronsen, J.K., 1992. K+ channels in nervous tissue. Biochem. Pharmacol.
Psychiatry 26, 989–993. 43, 11–14.
Anderson, I.M., Clark, L., Elliott, R., Kulkarni, B., Williams, S.R., Deakin, Artaiz, I., Romero, G., Zazpe, A., Monge, A., Calderó, J.M., Roca,
J.F.W., 2002. 5-HT2C receptor activation by m-chlorophenylpiperazine J., Lasheras, B., Del Rı́o, J., 1995. The pharmacology of VA21B7:
detected in humans with fMRI. NeuroReport 13, 1547–1551. an atypical 5-HT3 receptor antagonist with anxiolytic properties in
Andrade, T.G.C., Graeff, F.G., 2001. Effect of electrolytic and neurotoxic animal models. J. Psychopharmacol. 117, 137–148.
lesions of the median raphe nucleus on anxiety and stress. Pharmacol. Artigas, F., Celeda, P., Laruelle, M., Adell, A., 2001. How does pindolol
Biochem. Behav. 70, 1–14. improve antidepressant action? Trends Pharmacol. Sci. 22, 224–228.
180 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Asakawa, A., Inui, A., Ueno, N., Fujimiya, M., Fujino, M.A., Kasuga, Baldwin, H.A., File, S.E., 1989. Caffeine-induced anxiogenesis: the role
M., 1999. Mouse pancreatic polypeptide modulates food intake, while of adenosine, benzodiazepine and noradrenergic receptors. Pharmacol.
not influencing anxiety in mice. Peptides 20, 1445–1448. Biochem. Behav. 32, 181–186.
Ase, A.R., Sénécal, J., Reader, T.A., Hen, R., Descarries, L., 2002. Baldwin, H.A., Johnston, A.L., File, S.E., 1989. Antagonistic effects of
Decreased G-protein coupling of serotonin 5-HT1A receptors in the caffeine and yohimbine in animal tests of anxiety. Eur. J. Pharmacol.
brain of 5-HT1B knock-out mouse. Neuropharmacology 42, 941–949. 159, 211–215.
Ashby, C.R., Minabe, Y., Edwards, E., Wang, R.Y., 1991. 5-HT3 -like Bale, T.L., Pedersen, C.A., Dorsa, D.M., 1995. CNS oxytocin receptor
receptors in the rat medial prefrontal cortex: an electrophysiological mRNA expression and regulation by gonadal steroids. Adv. Exp. Med.
study. Brain Res. 550, 181–191. Biol. 395, 269–280.
Ashton, C.H., 1994. The effect of drugs on sleep. In: Cooper, R. (Ed.), Bale, T.L., Contarino, A., Smith, G., Chan, R., Gold, L., Sawchenko, P.,
Sleep. Chapman & Hall Medical, London, pp. 175–211. Koob, G., Vale, W., Lee, K.-F., 2000. Mice deficient for corticotrophin-
Ashton, C.H., 2001. Pharmacology and effects of cannabis: a brief releasing hormone receptor2 display anxiety-like behaviour and are
review. Br. J. Psychiatry 178, 101–106. hypersensitive to stress. Nat. Genet. 24, 410–414.
Aston-Jones, G., Shipley, M.T., Chouvet, G., Ennis, M., Van Bockstaele, Bale, T.L., Davis, A.M., Auger, A.P., Dorsa, D.M., McCarthy, M.M.,
E., Pieribone, V., Shiekhattar, R., Akaoka, H., Drolet, G., Astier, B., 2001. CNS region-specific oxytocin receptor expression: importance
Charléty, P., Valentino, R.J., Williams, J.T., 1991. Afferent regulation in regulation of anxiety and sex behavior. J. Neurosci. 21, 2546–2552.
of locus coeruleus neurons: anatomy, physiology and pharmacology. Bale, T.L., Picetti, R., Contarino, A., Koob, G.F., Vale, W.W., Lee, K.F.,
Prog. Brain Res. 88, 47–86. 2002. Mice deficient for both corticotropin-releasing factor receptor1
Aston-Jones, G., Rajowski, J., Cohen, J., 1999. Role of locus coeruleus (CRFR1 ) and CRFR2 have an impaired stress response and display
in attention and behavioral flexibility. Biol. Psychiatry 46, 1309–1320. sexually dichotomous anxiety-like behavior. J. Neurosci. 22, 193–199.
Auclair, A., Cotecchia, S., Glowinski, J., Tassin, J.-P., 2002. d- Ballard, T.M., Sanger, S., Higgins, C.A., 2001. Inhibition of shock-
amphetamine fails to increase extracellular dopamine levels in mice induced foot tapping behaviour in the gerbil by a tachykinin (NK)1
lacking ␣1b -adrenergic receptors: relationship between functional and receptor antagonist. Eur. J. Pharmacol. 412, 255–264.
nonfunctional dopamine release. J. Neurosci. 22, 9150–9154. Bals-Kubik, R., Ableitner, A., Herz, A., Shippenberg, T.S., 1993.
Neuroanatomical sites mediating the motivational effects of opioids
Audinot, V., Beauverger, P., Lahaye, C., Suply, T., Rodriguez, M.,
as mapped by the conditioned place preference paradigm in rats. J.
Ouvry, C., Lamamy, V., Imbert, J., Rique, H., Nahon, J.-L., Galizzi,
Pharmacol. Exp. Ther. 264, 489–495.
J.-P., Canet, E., Levens, N., Fauchère, J.-L., Boutin, J.-A., 2001a.
Bandler, R., Keay, K.A., Floyd, N., Price, J., 2000. Central circuits
Structure–activity relationship studies of melanin-concentrating
mediating patterned autonomic activity during active vs. passive
hormone (MCH)-related peptide ligands at SLC-1, the human MCH
emotional coping. Brain Res. Bull. 53, 95–104.
receptor. J. Biol. Chem. 276, 13554–13562.
Bannon, A.W., Seda, J., Carmouche, M., Francis, J.M., Norman, M.H.,
Audinot, V., Lahaye, C., Suply, T., Beauverger, P., Rodriguez, M.,
Karbon, B., McCaleb, M.L., 2000. Behavioral characterization of
Galizzi, J.-P., Fauchère, J.-L., Boutin, J.A., 2001b. [125 I]-S36057:
neuropeptide Y knock-out mice. Brain Res. 868, 79–87.
a new and highly potent radioligand for the melanin-concentrating
Bantock, R.A., Deakin, J.F.W., Grasby, P.M., 2001. The 5-HT1A receptor
hormone receptor. Br. J. Pharmacol. 133, 371–378.
in schizophrenia: a promising target for novel atypical neuroleptics?
Audinot, V., Fabry, N., Nicolas, J.-P., Beauverger, P., Newman-Tancredi,
J. Psychopharmacol. 15, 37–46.
A., Millan, M.J., Try, A., Bornancin, F., Canet, E., Boutin, J.-A.,
Barann, M., Molderings, G., Brüss, M., Bönisch, H., Urba, B.W.,
2002. Ligand modulation of [35 S]GTP␥S binding at human ␣2A , ␣2B
Göthert, M., 2002. Direct inhibition by cannabinoids of human 5-
and ␣2C -adrenoceptors. Cell. Signal. 14, 829–837.
HT3A receptors: probable involvement of an allosteric modulatory
Aufdembrinke, B., 1998. Abecarnil, a new ␤-carboline, in the treatment site. Br. J. Pharmacol. 137, 589–596.
of anxiety disorders. Br. J. Psychiatry 173, 55–63. Barbara, A., Aceves, J., Arias-Montano, J.-A., 2002. Histamine H1
Avgustinovich, D.F., Alekseyenko, O.V., Koryakina, L.A., 2003. Effects receptors in rat dorsal raphe nucleus: pharmacological characterisation
of chronic treatment with ipsapirone and buspirone on the C57BL/6J and linking to increased neuronal activity. Brain Res. 954, 247–255.
strain mice under social stress. Life Sci. 72, 1437–1444. Barden, N., Reul, J.M.H.M., Hosboer, F., 1995. Do antidepressants
Azmitia, E.C., Segal, M., 1978. An autodiographic analysis of the stabilize mood through actions on the hypothalamic–pituitary–
differential ascending projections of the dorsal and median raphe adrenocortical system? Trends Neurosci. 18, 6–11.
nuclei in the rat. J. Comp. Neurol. 179, 641–688. Baretta, I.P., Assreuy, J., De Lima, T.C.M., 2001. Nitric oxide involvement
Babar, E., Özgünen, T., Melik, E., Polat, S., Akman, H., 2001. Effects in the anxiogenic-like effect of substance P. Behav. Brain Res. 121,
of ketamine on different types of anxiety/fear and related memory 199–205.
in rats with lesions of the median raphe nucleus. Eur. J. Pharmacol. Barili, P., De Carolis, G., Zaccheo, D., Amenta, F., 1998. Sensitivity
431, 315–320. to aging of the limbic dopaminergic system: a review. Mech. Aging
Bagdy, G., Graf, M., Anheuer, Z.E., Modos, E.A., Kantor, S., 2001. Dev. 106, 57–92.
Anxiety-like effects induced by acute fluoxetine, sertraline or mCPP Barker, E.L., Westphal, R.S., Schmidt, D., Sanders-Bush, E., 1994.
treatment are reversed by pretreatment with the 5-HT2C receptor Constitutively active 5-hydroxytryptamine2C receptors reveal novel
antagonist SB-242084 but not the 5-HT1A receptor antagonist WAY- inverse agonist activity of receptor ligands. J. Biol. Chem. 269,
100635. Int. J. Neuropsychopharmacol. 4, 399–408. 11687–11690.
Bagnol, D., Lu, X.Y., Kaelin, C.B., Day, H.E.W., Ollmann, M., Gantz, I., Barnard, E.A., Skolnick, P., Olsen, R.W., Mohler, H., Sieghart, W., Biggio,
Akil, H., Barsh, G.S., Watson, S.J., 1999. Anatomy of an endogenous G., Braestrup, C., Bateson, A.N., Langer, S.Z., 1998. International
antagonist: relationship between AGRP and POMC in brain. J. Union of Pharmacology. XV. Subtypes of ␥-aminobutyric acidA
Neurosci. 19 (RC26), 1–7. receptors: classification on the basis of subunit structure and receptor
Bakshi, V.P., Halin, N.H., 2000. Corticotropin-releasing hormone and function. Pharmacol. Rev. 50, 291–313.
animal models of anxiety: gene-environment interactions. Biol. Barnes, N.M., Sharp, T., 1999. A review of central 5-HT receptors and
Psychiatry 48, 1175–1198. their function. Neuropharmacology 38, 1083–1152.
Bakshi, V.P., Smith-Roe, S., Newman, S.M., Grigoriadis, D.E., Kalin, Barnes, N.M., Costall, B., Domeney, A.M., Gerrard, P.A., Kelly, M.E.,
N.H., 2002. Reduction of stress-induced behavior by antagonism of Krähling, H., Naylor, R.J., Tomkins, D.M., Williams, T.J., 1991.
corticotropin-releasing hormone 2 (CRH2) receptors in lateral septum The effects of umespirone as a potential anxiolytic and antipsychotic
or CRH1 receptors in amygdala. J. Neurosci. 22, 2926–2935. agent. Pharmacol. Biochem. Behav.40, 89–96.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 181

Barnett, S.D., Kramer, M.L., Casat, C.D., Connor, K.M., Davidson, of cholecystokinin in the frontal cortex of the freely moving rat.
J.R.T., 2002. Efficacy of olanzapine in social anxiety disorder: a pilot Neuropharmacology 38, 525–532.
study. J. Psychopharmacol. 16, 365–368. Becker, C., Thiébot, M.-H., Touitou, Y., Hamon, M., Cesselin, F., Benoliel,
Baron, B.M., Harrison, B.L., Kehne, J.H., Schmidt, C.J., Vangiersbergen, J.-J., 2001. Enhanced cortical extracellular levels of cholecystokinin-
P.L.M., White, H.S., Siegel, B.W., Senyah, Y., McCloskey, T.C., like material in a model of anticipation of social defeat in the rat. J.
Fadayel, G.M., Taylor, V.L., Murawsky, M.K., Nyce, P., Salituro, F.G., Neurosci. 21, 262–269.
1997. Pharmacological characterization of MDL-105,519, an NMDA Beckett, S., Marsden, C.A., 1997. The effect of central and systemic
receptor glycine site antagonist. Eur. J. Pharmacol. 323, 181–192. injection of the 5-HT1A receptor agonist 8-OHDPAT and the 5-HT1A
Barrett, J.E., 1992. Studies on the effects of 5-HT1A drugs in the pigeon. receptor antagonist WAY100635 on periaqueductal gray-induced
Drug Dev. Res. 26, 299–317. defence behaviour. J. Psychopharmacol. 11, 35–40.
Barrett, J.E., Gleeson, S., 1991. Anxiolytic effects of 5-HT1A agonists, 5- Beekman, A.T.F., De Beurs, E., Van Balkom, A.J.L.M., Deeg, D.J.H.,
HT3 antagonists and benzodiazepines: conflict and drug discrimination Van Dyck, R., Van Tilburg, W., 2000. Anxiety and depression in later
studies. In: Rodgers, R.J., Cooper, S.J. (Eds.), 5-HT1A Agonists, 5-HT3 life: co-occurrence and communality of risk factors. Am. J. Psychiatry
Antagonists and Benzodiazepines: Their Comparative Behavioural 157, 89–95.
Pharmacology. Wiley, Chichester, pp. 59–105. Beer, B., Chasin, M., Clody, D.E., Vogel, J.R., Horovitz, Z.P., 1972.
Barrett, J.E., Vanover, K.E., 1993. 5-HT receptors as targets for the Cyclic adenosine monophosphate phosphodiesterase in brain: effect
development of novel anxiolytic drug models: mechanisms and future on anxiety. Science 176, 428–430.
directions. Psychopharmacology 112, 1–12. Behan, D.P., De Souza, E.B., Lowry, P.J., Potter, E., Sawchenko, P., Vale,
Barros, M., Tomaz, C., 2002. Non-human primate models for investigating W.W., 1995. Corticotropin releasing factor (CRF) binding protein: a
fear and anxiety. Neurosci. Biobehav. Rev. 26, 187–201. novel regulator of CRF and related peptides. Front. Neuroendocrinol.
Barros, M., De Souza Silva, M.A., Huston, J.P., Tomaz, C., 2002. 16, 362–382.
Anxiolytic-like effects of substance P fragment (SP1–7) in non-human Belcher, S.M., Zsarnovszky, A., 2001. Estrogenic actions in the brain:
primates (Callithrix penicillata). Peptides 23, 967–973. estrogen, phytoestrogens, and rapid intracellular signaling mechanisms.
Bartanusz, V., Aubry, J.-M., Pagliusi, S., Jesova, D., Baffi, J., Kiss, J.Z., J. Pharmacol. Exp. Ther. 299, 408–414.
1995. Stress-induced changes in messenger RNA levels of N-methyl- Belcheva, I., Belcheva, S., Petkov, V.V., Petkov, V.D., 1994. Hippocampal
d-aspartate and AMPA receptor subunits in selected regions of the rat asymmetry in the behavioral responses to the 5-HT1A receptor agonist
hippocampus and hypothalamus. Neuroscience 6, 247–252. 8-OH-DPAT. Brain Res. 640, 223–228.
Bartfai, T., Fisone, G., Langel, U., 1992. Galanin and galanin antagonists: Belcheva, I., Belcheva, S., Petkov, V.V., Petkov, V.D., Hadjiivanova, J.L.,
molecular and biochemical perspectives. Trends Pharmacol. Sci. 13, 1997. Behavioral responses to the 5-HT1A receptor antagonist NAN190
312–317. injected into rat CA1 hippocampal area. Gen. Pharmacol. 28, 435–441.
Bartfai, T., Fisone, G., Langel, U., 1993. Galanin: a neuroendocrine Belelli, D., Casula, A., Ling, A., Lambert, J.J., 2002. The influence of
peptide. Crit. Rev. Neurobiol. 7, 229–274. subunit composition on the interaction of neurosteroids with GABAA
Barton, C.L., Jay, M.T., Meurer, L., Hutson, P.H., 1999. GR205171, a receptors. Neuropharmacology 43, 651–661.
selective NK1 receptor antagonist, attenuates stress-induced increase Belzung, C., 2001. The genetic basis of the pharmacological efects of
of dopamine metabolism in rat medial prefrontal cortex. Br. J. anxiolytics: a review based on rodent models. Behav. Pharmacol. 12,
Pharmacol. 126, 284P. 451–460.
Bartoszyk, G.D., 1998. Anxiolytic effects of dopamine receptor ligands. Belzung, C., Ågmo, A., 1997. Naloxone potentiates the effects of
I. Involvement of dopamine autoreceptors. Life Sci. 62, 649–663. subeffective doses of anxiolytic agents in mice. Eur. J. Pharmacol.
Bartoszyk, G.D., Hegenbart, R., Ziegler, H., 1997. EMD 68843, a 323, 133–136.
serotonin reuptake inhibitor with selective presynaptic 5-HT1A receptor Belzung, C., Dubreuil, D., 1998. Naloxone potentiates the anxiolytic but
agonistic properties. Eur. J. Pharmacol. 322, 147–153. not the amnestic action of chlordiazepoxide to C57BL/6 mice. Behav.
Battey, J., Wada, E., 1991. Two distinct receptor subtypes for mammalian Pharmacol. 9, 691–698.
bombesin-like peptides. Trends Neurosci. 14, 524–528. Belzung, C., Griebel, G., 2001. Measuring normal and pathological
Bauer, E.P., Schale, G.E., LeDoux, J.E., 2002. NMDA receptors anxiety-like behaviour in mice: a review. Behav. Brain Res. 125,
and l-type voltage-gated calcium channels contribute to long-term 141–149.
potentiation and different components of fear memory formation in Benavides, J., Depoortere, H., Sanger, D., Perrault, G., Arbilla, S.,
the lateral amygdala. J. Neurosci. 22, 5239–5249. Langer, S.Z., Zivkovic, B., Scatton, B., 1990. A specific domain (the
Baumgarten, H.G., Grozdanovic, Z., 1997. Anatomy of central omega 1 site) of the GABA(A) receptor may be implicated in the
serotoninergic projection systems. Handbook Exp. Pharmacol. 129, hypnotic effect of zolpidem. Encephale 16, 13–22.
41–89. Benca, R.M., 1994a. Other psychiatric disorders. In: Cooper, R. (Ed.),
Beaufour, C.C., Ballon, N., Le Bihan, C., Hamon, M., Thiébot, M.-H., Sleep. Chapman & Hall Medical, London, pp. 557–583.
1999. Effects of chronic antidepressants in an operant conflict procedure Benca, R.M., 1994b. The affective disorders. In: Cooper, R. (Ed.),
of anxiety in the rat. Pharmacol. Biochem. Behav. 62, 591–599. Sleep. Chapman & Hall Medical, London, pp. 525–556.
Beaufour, C.C., Le Bihan, C., Hamon, M., Thiébot, M.-H., 2001a. Bengtsson, H.J., Kele, J., Johansson, J., Hjorth, S., 2000. Interaction
Extracellular dopamine in the rat prefrontal cortex during reward-, of the antidepressant mirtazapine with ␣2 -adrenoceptors modulating
punishment-, and novelty-associated behaviour. Effects of diazepam. the release of 5-HT in different rat brain regions in vivo. Naunyn
Pharmacol. Biochem. Behav. 69, 133–143. Schmiedebergs Arch. Pharmacol. 362, 406–412.
Beaufour, C.C., Le Bihan, C., Hamon, M., Thiébot, M.-H., 2001b. Benjamin, D., Saiff, E.I., Nevins, T., Lal, H., 1992. Mianserin-induced
Extracellular serotonin is enhanced in the striatum, but not in the 5-HT2 receptor downregulation results in anxiolytic effects in the
dorsal hippocampus or prefrontal cortex, in rats subjected to an elevated plus-maze test. Drug Dev. Res. 26, 287–297.
operant conflict procedure. Behav. Neurosci. 115, 125–137. Benjamin, J., Geraci, M., McCann, U., Greenberg, B.D., Murphy, D.L.,
Beaulieu, E.E., 1997. Neurosteroids: of the nervous system, by the nervous 1999. Attenuated response to mCPP and to pentagastrin after repeated
system, for the nervous system. Recent Prog. Horm. Res. 52, 1–25. mCPP in panic disorder. Psychopharmacology 143, 215–216.
Bechara, A., Damasin, H., Damasio, A.R., 2000. Emotion, decision Benke, T., Bosch, S., Andree, B., 1998. A study of emotional processing
making and the orbitofrontal cortex. Cereb. Cortex 10, 295–307. in Parkinson’s disease. Brain Cogn. 38, 36–52.
Becker, C., Hamon, M., Benoliel, J.-J., 1999. Prevention by 5-HT1A Benke, D., Honer, M., Michel, C., Bettler, B., Möhler, H., 1999. ␥-
receptor agonists of restraint stress- and yohimbine-induced release Aminobutyric acid typeB receptor splice variant proteins GBR1a and
182 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

GBR1b in situ display differential regional and subcellular distribution. Beydoun, A., Uthman, B.M., Sackellares, J.C., 1995. GABApentin:
J. Biol. Chem. 274, 27323–27330. pharmacokinetics, efficacy and safety. Clin. Neuropharmacol. 18,
Bennett, D.A., Amrick, C.L., 1986. 2-Amino-7-phosphonoheptanoic acid 469–481.
(AP7) produces discriminative stimuli and anticonflict effects similar Bezzi, P., Carmignoto, G., Pasti, L., Vesce, S., Rossi, D., Rizzini, B.L.,
to diazepam. Life Sci. 39, 2455–2461. Pozzan, T., Volterra, A., 1998. Prostaglandins stimulate calcium-
Benquet, P., Gee, C.E., Gerber, U., 2002. Two distinct signaling pathways dependent glutamate release in astrocytes. Nature 391, 281–285.
up-regulate NMDA receptor responses via two distinct metabotropic Bhatnagar, A., Willins, D.L., Gray, J.A., Woods, J., Benovic, J.L.,
glutamate receptor subtypes. J. Neurosci. 22, 9679–9686. Roth, B.L., 2001. The dynamin-dependent, arrestin-independent
Bensaid, M., Faucheux, B.A., Hirsch, E., Agid, Y., Soubrié, P., Oury- internalization of 5-hydroxytryptamine2A (5-HT2A ) serotonin receptors
Donat, F., 2001. Expression of tachykinin NK2 receptor mRNA in reveals differential sorting of arrestins and 5-HT2A receptors during
human brain. Neurosci. Lett. 303, 25–29. endocytosis. J. Biol. Chem. 276, 8269–8277.
Bentley, K.R., Barnes, N.M., 1995. Therapeutic potential of serotonin 5- Bhattacharyya, S.K., Sen, A.P., Upadhyay, S.N., Jaiswal, A.K., 1993.
HT3 antagonists in neuropsychiatric disorders. CNS Drugs 3, 363–392. Anxiolytic activity of piracetam, a nootropic agent, following
Bentué-Ferrer, D., Reymann, J.M., Tribut, O., Allain, H., Vasar, E., subchronic administration in rodents. Indian J. Exp. Biol. 31, 902–907.
Bourin, M., 2001. Role of dopaminergic and serotonergic systems on Bhattacharyya, S.K., Chakrabarti, A., Sandler, M., Glover, V., 1996.
behavioural stimulatory effects of low-dose alprazolam and lorazepam. Anxiolytic activity of intraventricularly administered atrial natriuretic
Eur. Neuropsychopharmacol. 11, 41–50. peptide in the rat. Neuropsychopharmacology 15, 199–206.
Benvenga, M.J., Leander, J.D., 1995. Olanzapine, an atypical Bhattacharyya, S.K., Puri, S., Miledi, R., Panicker, M.M., 2002.
antipsychotic, increases rates of punished responding in pigeons. Internalization and recycling of 5-HT2A receptors activated by
Psychopharmacology 119, 133–138. serotonin and protein kinase C-mediated mechanisms. Proc. Natl.
Benvenga, M.J., Ornstein, P.L., Leander, D.J., 1995. Schedule-controlled Acad. Sci. U.S.A. 99, 14470–14475.
behavioral effects of the selective 2-amino-3-(5-methyl-3-hydroxy- Bianchi, M.T., Haas, K.F., MacDonald, R.L., 2002. ␣1 and ␣6
isoxazol-4-yl) propanoic acid antagonist LY293558 in pigeons. J. subunits specify distinct desensitization, deactivation and neurosteroid
Pharmacol. Exp. Ther. 275, 164–170. modulation of GABAA receptors containing the ␦ subunit.
Neuropharmacology 43, 492–502.
Benvenga, M.J., Overshiner, C.D., Monn, J.A., Leander, D.J., 1999.
Bickerdike, M.J., Fletcher, A., Marsden, C.A., 1995. Attenuation of
Disinhibitory effects of LY354740, a new mGluR2 agonist, on
CCK-induced aversion in rats on the elevated x-maze by the selective
behaviors suppressed by electric shock in rats and pigeons. Drug Dev.
5-HT1A receptor antagonists (+)WAY100,135 and WAY100,635.
Res. 47, 37–44.
Neuropharmacology 34, 805–811.
Benya, R.V., Kusui, T., Pradhan, T.K., Battey, J.F., Jensen, R.T., 1995.
Bidzseranova, A., Gueron, J., Toth, G., Penke, B., Varga, J., Teledgy,
Expression and characterization of cloned human bombesin receptors.
G., 1992. Behavioral effects of atrial natriuretic and brain natriuretic
Mol. Pharmacol. 47, 10–20.
peptides in rats. NeuroReport 3, 283–285.
Berg, K.A., Stout, B.D., Maahani, S., Clarke, W.P., 2001. Differences
Biggio, G., Concas, A., Serra, M., Salis, M., Corda, M.G., Nurchi, V.,
in rapid desensitisation of 5-hydroxytryptamine2A and 5-
Crisponi, C., Gessa, G.L., 1984. Stress and ␤-carbolines decrease the
hydroxytryptamine2C receptor-mediated phospholipase C activation. J.
density of low-affinity GABA binding sites: an effect reversed by
Pharmacol. Exp. Ther. 299, 593–602.
diazepam. Brain Res. 305, 13–18.
Bermack, J.E., Debonnel, G., 2001. Modulation of serotonergic
Biggio, G., Follesa, P., Sanna, E., Purdy, R.H., Concas, A., 2001. GABAA -
neurotransmission by short- and long-term treatments with sigma
receptor plasticity during long-term exposure to and withdrawal from
ligands. Br. J. Pharmacol. 134, 691–699.
progesterone. Int. Rev. Neurobiol. 46, 207–241.
Berntson, G.G., Sarter, M., Cacioppo, J.T., 1998. Anxiety and Bignami, G., 1998. Pharmacology and anxiety: inadequacies of current
cardiovascular reactivity: the basal forebrain cholinergic link. Behav. experimental approaches and working models. Pharmacol. Biochem.
Brain Res. 94, 225–248. Behav. 29, 771–774.
Berrandero, F., Maldonado, R., 2002. Involvement of the opioid system Bilkei-Gorzo, A., Ildiko, R., Kerstin, M., Zimmer, A., 2002. Diminished
in the anxiolytic-like effects induced by 9 -tetrahydrocannabinol. anxiety- and depression-related behaviors in mice with selective
Psychopharmacology 163, 111–117. deletion of the Tac1 gene. J. Neurosci. 22, 10046–10052.
Berridge, C.W., Dunn, A.J., 1989. Restraint-stress-induced changes Billinton, A., Upton, N., Bowery, N.G., 1999. GABAB receptors GBR1a
in exploratory behavior appear to be mediated by norepinephrine- and GBR1b, appear to be associated with pre- and post-synaptic
stimulated release of CRF. J. Neurosci. 9, 3513–3521. elements respectively in rat and human cerebellum. Br. J. Pharmacol.
Bert, L., Rodier, D., Bougault, I., Allouard, N., Le Fur, G., Soubrie, P., 126, 1387–1392.
Steinberg, R., 2002. Permissive role of neurokinin NK3 receptors in Billinton, A., Ige, A.O., White, J.H., Marshall, F.H., Emson, P.C., 2001.
NK1 receptor-mediated activation of the locus coeruleus revealed by Advances in the molecular understanding of GABAB receptors. Trends
SR 142801. Synapse 43, 62–69. Neurosci. 24, 277–282.
Berton, O., Aguerre, S., Sarrieau, A., Normede, P., Chaouloff, F., 1998. Bing, O., Möller, C., Engel, J.A., Söderpalm, B., Heilig, M., 1993.
Differential effects of social stress on central serotonergic activity and Anxiolytic-like action of centrally administered galanin. Neurosci.
emotional reactivity in Lewis and spontaneously hypertensive rats. Lett. 164, 17–20.
Neuroscience 82, 147–159. Bing, O., Heilig, M., Kakoulidis, P., Sundblad, C., Wiklund, L., Eriksson,
Bertorelli, R., Calo, G., Ongini, E., Regoli, D., 2000. Nociceptin/ E., 1998. High doses of testosterone increase anticonflict behaviour
orphaninFQ and its receptor: a potential target for drug discovery. in rat. Eur. Neuropsychopharmacol. 8, 321–323.
Trends Pharmacol. Sci. 21, 233–234. Birikh, K.R., Sklan, E.H., Shoham, S., Soreq, H., 2003. Interaction
Bertrand, S., NG, G.Y.K., Purisai, M.G., Wolfe, S.E., Severidt, M.W., of “readthrough” acetylcholinesterase with RACK1 and PKC␤II
Nouel, D., Robitaille, R., Löw, M.J., O’Neill, G.P., Metters, K., correlates with intensified fear-induced conflict behavior. Proc. Natl.
Lacaille, J.-C., Chronwall, B.M., Morris, S.J., 2001. The anticonvulsant, Am. Soc. U.S.A. 100, 283–288.
antihyperalgesic agent GABApentin is an agonist at brain ␥- Birnbaumer, M., 2000. Vasopressin receptors. Trends Endocrinol. Metab.
aminobutyric acid type B receptors negatively coupled to voltage- 11, 406–411.
dependent calcium channels. J. Pharmacol. Exp. Ther. 298, 15–24. Birnir, B., Tierney, M.L., Dalziel, J.E., Cox, G.B., Gage, P.W., 1997. A
Besedovsky, H.O., Del Rey, A., 1996. Immune-neuro-endocrine structural determinant of desensitization and allosteric regulation by
interactions: facts and hypotheses. Endocr. Rev. 17, 64–102. pentobarbitone of the GABAA receptor. J. Membr. Biol. 155, 157–166.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 183

Biro, E., Toth, G., Teledgy, G., 1996. Effect of receptor blockers on brain Blank, T., Nijholt, I., Vollstaedt, S., Spiess, J., 2003b. The corticotropin-
natriuretic peptide and C-type natriuretic peptide caused anxiolytic releasing factor receptor1 antagonist CP-154,526 reverses stress-
state in rats. Neuropeptides 30, 59–65. induced learning deficits in mice. Behav. Brain Res. 138, 207–213.
Bischoff, A., Michel, M.C., 1999. Emerging functions for Y5 neuropeptide Bleakman, D., 1999. Kainate receptor pharmacology and physiology.
Y receptors. Trends Pharmacol. Sci. 20, 104–106. 1999. Cell. Mol. Life Sci. 56, 558–566.
Bischoff, S., Leonhard, S., Reymann, N., Schuler, V., Shigemoto, R., Bleakman, D., Lodge, D., 1998. Neuropharmacology of AMPA and
Kaupmann, K., Bettler, B., 1999. Spatial distribution of GABAB R1 kainate receptors. Neuropharmacology 37, 1187–1204.
receptor mRNA and binding sites in the rat brain. J. Comp. Neurol. Blednov, Y.A., Stoffel, M., Chang, S.R., Harris, R.A., 2001. GIRK2
412, 1–16. deficient mice: evidence for hyperactivity and reduced anxiety. Physiol.
Bitner, R.S., Nikkel, A.L., 2002. Alpha7 nicotinic receptor expression by Behav. 74, 109–117.
two distinct cell types in the dorsal raphe nucleus and locus coeruleus Blednov, Y.A., Stoffel, M., Cooper, R., Wallace, D., Mane, N., Harris,
of rat. Brain Res. 938, 45–54. R.A., 2002. Hyperactivity and dopamine D1 receptor activation in
Bitner, R.S., Nikkel, A.L., Curzon, P., Donnelly-Roberts, D.L., Puufarcken, mice lacking girk2 channels. Psychopharmacology 159, 370–378.
P.S., Namovic, M., Jacobs, I.C., Meyer, M.D., Decker, M.W., 2000. Blednov, Y.A., Jung, S., Alva, H., Wallace, D., Rosahl, T., Whiting,
Reduced nicotinic receptor-mediated antinociception following in vivo P.-J., Harris, R.A., 2003. Deletion of the ␣1 or ␤2 subunit of GABAA
antisense knock-down in rat. Brain Res. 871, 66–74. receptors reduces actions of alcohol and other drugs. J. Pharmacol.
Bitran, D., Kellogg, C.K., Hilvers, R.J., 1993. Treatment with an Exp. Ther. 304, 30–36.
anabolic-androgenic steroid affects anxiety-related behavior and alters Blier, P., De Montigny, C., 1994. Current advances and trends in the
the sensitivity of cortical GABAA receptors in the rat. Horm. Behav. treatment of depression. Trends Pharmacol. Sci. 15, 220–226.
27, 568–583. Bluet-Pajot, M.T., Presse, F., Voko, Z., Hoeger, F., Mounier, J., Epelbaum,
Bitran, D., Dugan, M., Renda, P., Ellis, R., Foley, M., 1999. Anxiolytic J., Nahon, J.L., Bluthé, R.-M., Dantzer, R., Kelley, K.W., 1992.
effects of the neuroactive steroid pregnanolone (3-alpha-OH-5-beta- Effects of interleukin1 receptor antagonist on the behavioural effects
pregnan-20-one) after microinjection in the dorsal hippocampus and of lipopolysaccharide in rat. Brain Res. 573, 318–320.
lateral septum. Brain Res. 850, 217–224. Bluet-Pajot, M.T., Presse, F., Voko, Z., Hoeger, F., Mounier, J.,
Bitsios, P., Szabadi, E., Bradshaw, C.M., 1998. The effects of clonidine Epelbaum, J., Nahon, J.L., 1995. Neuropeptide E1 antagonises the
on the fear-inhibited light reflex. J. Psychopharmacol. 12, 137–145. action of melanin-concentrating hormone on stress induced release
Bittencourt, J.C., Presse, F., Arias, C., Peto, C., Vaughan, J., Nahon, J.-L.,
the adrenocorticotropin in the rat. J. Endocrinol. 7, 297–303.
Vale, W., Sawchenko, P.E., 1992. The melanin-concentrating hormone
Blum, R., Kafitz, K.W., Konnerth, A., 2002. Neurotrophin-evoked depol-
system of the rat brain: an immuno- and hybridization histochemical
arization requires the sodium channel Nav 1.9. Nature 419, 687–693.
characterization. J. Comp. Neurol. 319, 218–245.
Blumcke, I., Behle, K., Malitschek, B., Kuhn, R., Knopfel, T., Wolf,
Bixo, M., Anderson, A., Winblad, B., Purdy, R.H., Blackstrom, T., 1997.
H.K., Wiestler, O.D., 1996. Immunohistochemical distribution of
Progesterone, 5-alpha-pregnane-3,20-dione and 3-alpha-hydroxy-
metabotropic glutamate receptor subtypes mGluR1b, mGluR2,3,
5alpha-pregnane-20-one in specific regions of the human female brain
mGluR4a and mGluR5 in human hippocampus. Brain Res. 736,
in different endocrine state. Brain Res. 764, 173–178.
217–226.
Bjertnaes, A., Block, J.M., Hafstad, P.E., Holte, M., Ottemo, I., Larsen, T.,
Bluthé, R.-M., Dantzer, R., Kelley, K.W., 1992. Effects of interleukin1
Pindek, R.M., Steffensen, K., Stulemeijer, S.M., 1982. A multicentre
receptor antagonist on the behavioural effects of lipopolysaccharide
placebo-controlled trial comparing the efficacy of mianserin and
in rat. Brain Res. 573, 318–320.
chlordiazepoxide in general practice patients with primary anxiety.
Bobker, D.H., Williams, J.T., 1989. Serotonin agonists inhibit synaptic
Acta Psychiatr. Scand. 66, 199–207.
Björklund, M., Sirviö, J., Sallinen, J., Scheinin, M., Kobilka, B.K., potentials in the rat locus ceruleus in vitro via 5-hydroxytryptamine1A
Rieffinen, P., 1999. Alpha2C -adrenoceptor overexpression disrupts and 5-hydroxytryptamine1B receptors. J. Pharmacol. Exp. Ther. 250,
execution of spatial and non-spatial search patterns. Neuroscience 88, 37–43.
1187–1198. Boden, P.R., Woodruff, G.N., Pinnock, R.D., 1991. Pharmacology of
Björklund, M., Siverina, I., Heikkinen, T., Tanila, H., Sallinen, J., Scheinin, a cholecystokinin receptor on 5-hydroxytryptamine neurons in the
M., Riekkinen, P., 2001. Spatial working memory improvement by dorsal raphe of the rat brain. Br. J. Pharmacol. 102, 635–638.
an ␣2 -adrenoceptor agonist dexmedetomidine is not mediated through Bodnoff, S.R., Suranyi-Cadotte, B., Quirion, R., Meaney, M.J., 1989.
␣2C -adrenoceptors. Prog. Neuropsychopharmacol. Biol. Psychiatry 25, A comparison of the effects of diazepam versus typical and atypical
1539–1554. antidepressant drugs in an animal model of anxiety. Psychophar-
Blair, H.T., Schafe, G.E., Bauer, E.P., Rodriges, S.M., LeDoux, J.E., macology 97, 277–279.
2001. Synaptic plasticity in the lateral amygdala: a cellular hypothesis Boehm, S., 1999. Presynaptic ␣2 -adrenoceptors control excitatory, but
of fear conditioning. Learn. Mem. 8, 229–242. not inhibitory, transmission at rat hippocampal synapses. J. Physiol.
Blanchard, D.C., Shepherd, J.K., Rodgers, R.J., Blanchard, R.J., 1992. 519, 439–449.
Evidence for differential effects of 8-OH-DPAT on male and female rats Boess, F.G., Monsma, F.J., Carolo, C., Meyer, V., Rudler, A.,
in the anxiety/defense test battery. Psychopharmacology 106, 531–539. Zwingelstein, C., Sleight, A.J., 1997. Functional and radioligand
Blanchard, D.C., Griebel, G., Blanchard, R.J., 2001a. Mouse defensive binding characterization of rat 5-HT6 receptors stably expressed in
behaviors: pharmacological and behavioral assays for anxiety and HEK293 cells. Neuropharmacology 36, 713–720.
panic. Neurosci. Biobehav. Rev. 25, 205–218. Bojarski, A.J., Kowalski, P., Kowalska, T., Duszynska, B., Charakchieva-
Blanchard, R.J., McKittrick, C.R., Blanchard, D.C., 2001b. Animal Minol, S., Tatarczynska, E., Klodzinska, A., Chojnacka-Wojcik, E.,
models of social stress: effects on behavior and brain neurochemical 2002. Synthesis and pharmacological evaluation of new arylpiperazines.
systems. Physiol. Behav. 73, 261–271. 3-4-[4-(3-Chlorophenyl)-1-piperazinyl]butyl-quinazolidin-4-one—a
Blanchard, D.C., Griebel, G., Blanchard, R.J., 2003. The mouse defense dual serotonin 5-HT1A /5-HT2A receptor ligand with an anxiolytic-like
test battery: pharmacological and behavioral assays for anxiety and activity. Bioorg. Med. Chem. 10, 3817–3827.
panic. Eur. J. Pharmacol. 463, 97–116. Bonaccorso, S., Meltzer, H., Maes, M., 2000. Psychological and
Blank, T., Nijholt, I., Grammatopoulos, D.K., Randeva, H.S., Hillhouse, behavioural effects of interferons. Cur. Opin. Psychiatry 13, 673–677.
E.W., Spiess, J., 2003a. Corticotropin-releasing factor receptors couple Bonanno, G., Raiteri, M., 1994. Pharmacological discrimination between
to multiple G-proteins to activate diverse intracellular signaling ␥-aminobutyric acid type B receptors regulating cholecystokinin
pathways in mouse hippocampus: role in neuronal excitability and and somatostatin release from rat neocortex synaptosomes. Mol.
associative learning. J. Neurosci. 23, 700–707. Pharmacol. 46, 558–562.
184 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Bonaventure, P., Voorn, P., Luyten, W.H.M.L., Jursak, M., Schotte, A., Bouvier, M., 2001. Oligomerization of G-protein-coupled transmitter
Leysen, J.E., 1998. Detailed mapping of serotonin 5-HT1B and 5-HT1D receptors. Nat. Rev. Neurosci. 2, 274–286.
receptor messenger RNA and ligand binding sites in guinea-pig brain Bouwknecht, J.A., Paylor, R., 2002. Behavioral and physiological mouse
and trigeminal ganglion: clues for function. Neuroscience 82, 469–484. assays for anxiety: a survey in nine mouse strains. Behav. Brain Res.
Bonaventure, P., Hall, H., Gommeren, W., Cras, P., Langlois, X., Jurzak, 136, 489–501.
M., Leysen, J.E., 2000. Mapping of serotonin 5-HT4 receptor mRNA Bowers, B.J., Collins, A.C., Tritto, T., Wehner, J.M., 2000. Mice lacking
and ligand binding sites in the post-mortem human brain. Synapse PKC␥ exhibit decreased anxiety. Behav. Genet. 30, 111–121.
36, 35–46. Bowery, N.G., Enna, S.J., 2000. ␥-Aminobutyric acidB receptors: first of
Bonaventure, P., Nepomuceno, D., Kwok, A., Chai, W., Langlois, the functional metabotropic heterodimers. J. Pharmacol. Exp. Ther.
X., Hen, R., Stark, K., Carruthers, N., Lovenberg, T.W., 2002. 292, 2–7.
Reconsideration of 5-hydroxytryptamine (5-HT)7 receptor distribution Bowery, N.G., Bettler, B., Froestl, B.W., Gallagher, J.P., Marshall, F.,
using [3 H]5-carboxamidotryptamine and [3 H]8-hydroxy-2-(di-n- Raiteri, M., Bonner, T.I., Enna, S.J., 2002. International union of
propylamino)tetraline: analysis in brain of 5-HT1A knock-out and pharmacology. XXXIII. Mammalian ␥-aminobutyric acid B receptors:
5-HT1A/1B double-knock-out mice. J. Pharmacol. Exp. Ther. 302, structure and function. Pharmacol. Rev. 54, 247–264.
240–248. Braak, H., Braak, E., 2000. Pathoanatomy of Parkinson’s disease. J.
Bond, A.J., Wingrove, J., Baylis, M., Dalton, J., 2003. Buspirone Neurol. 247, 3–10.
decreases physiological reactivity to unconditioned and conditioned Braas, K.M., Newby, A.C., Wilson, V.S., Snyder, S.H., 1986. Adenosine-
aversive stimuli. Psychopharmacology 165, 291–295. containing neurons in the brain localized by immunocytochemistry. J.
Bonhaus, D.W., Chang, L.K., Kwan, J., Martin, G.R., 1998. Dual Neurosci. 6, 1952–1961.
activation and inhibition of adenylyl cyclase by cannabinoid receptor Bracey, M.H., Hanson, M.A., Masuda, K.R., Stevens, R.C., Cravatt, B.F.,
agonists: evidence for agonist specific trafficking of intracellular 2002. Structural adaptations in a membrane enzyme that terminates
responses. J. Pharmacol. Exp. Ther. 287, 884–888. endocannabinoid signaling. Science 298, 1793–1796.
Bonnet, U., 2002. Moclobemide: evolution, pharmacodynamic, and Brady, C.A., Stanford, I.M., Ali, I., Lin, L., Williams, J.M., Dubin, A.E.,
pharmacokinetic properties. CNS Drug Rev. 8, 283–308. Hope, A.G., Barnes, N.M., 2001. Pharmacological comparison of
Borbély, A.A., 1995. Principal sleep regulation: implications for the effect human homomeric 5-HT3A receptors versus heteromeric 5-HT3A/3B
of hypnotics on sleep. In: Kales, A. (Ed.), Handbook of Experimental receptors. Neuropharmacology 41, 282–284.
Pharmacology. Pharmacology of Sleep, vol. 116. Springer-Verlag, Bradwejn, J., 1993. Neurobiological investigations into the role of
Berlin, Germany, pp. 29–45. cholecystokinin in panic disorder. J. Psychiatry Neurosci. 18, 178–188.
Bordi, F., Ugolini, A., 1999. Group I metabotropic glutamate receptors Bradwejn, J., Koszycki, D., 2001. Cholecystokinin and panic disorder:
implications for brain diseases. Prog. Neurobiol. 59, 55–79. past and future clinical research strategies. Scand. J. Clin. Lab. Invest.
Borjigin, J., Li, X., Snyder, S.H., 1999. The pineal gland and melatonin:
234, 19–27.
molecular and pharmacologic regulation. Annu. Rev. Pharmacol.
Brady, K.T., Sonne, S.C., 1999. The role of stress in alcohol use,
Toxicol. 39, 53–65.
alcoholism treatment, and relapse. Alcohol Res. Health 23, 263–271.
Borowski, T.B., Kokkinidis, L., 1996. Contribution of ventral tegmental
Braga, M.F.M., Aroniadou-Anderjaska, V., Post, R.M., Li, H., 2002.
area dopamine neurons to expression of conditioned fear: effects of
Lamotrigine reduces spontaneous and evoked GABAA receptor-
electrical stimulation, excitotoxin lesions, and quinpirole infusion on
mediated synaptic transmission in the basolateral amygdala:
potentiated startle in rats. Behav. Neurosci. 110, 1349–1364.
implications for its effects in seizure and affective disorders. Neurophar-
Borowski, B., Durkin, M.M., Ogozalek, K., Marzabadi, M.R., DeLeon,
macology 42, 522–529.
J., Heurich, R., Lichtblau, H., Shaposhnik, Z., Daniewska, I.,
Braga, M.F.M., Aroniadou-Anderjaska, V., Xie, J., Li, H., 2003.
Blackburn, T.P., Branchek, T.A., Gerald, C., Vaysse, P.J., Forray, C.,
Bidirectional modulation of GABA release by presynaptic glutamate
2002. Antidepressant, anxiolytic and anorectic effects of a melanin-
receptor 5 kainate receptors in the basolateral amygdala. J. Neurosci.
concentrating hormone1 receptor antagonist. Nat. Med. 8, 825–830.
23, 442–452.
Borsini, F., Podhorna, J., Marazziti, D., 2002. Do animal models of anxiety
Braissant, O., Gotoh, T., Loup, M., Mori, M., Bachmann, C., 1999.
predict anxiolytic-like effects of antidepressants? Psychopharmacology
l-Arginine uptake, the citrulline–NO cycle and arginase II in the rat
163, 121–141.
Borski, R.J., 2000. Nongenomic membrane actions of glucocorticoids in brain: an in situ hybridization study. Mol. Brain Res. 70, 231–241.
vertebrates. Trends Endocrinol. Metab. 11, 427–436. Brambilla, F., Bellodi, L., Perna, G., Bertani, A., Panerai, A., Sacardote,
Boshuisen, M.L., Ter Horst, G.J., Paans, A.M.J., Reinders, A.A.T.S., P., 1994. Plasma interleukin1 beta concentrations in panic disorder.
Den Boer, J.A., 2002. rCBF differences between panic disorder Psychiatry Res. 54, 135–142.
patients and control subjects during anticipatory anxiety and rest. Branchek, T.A., Smith, K.E., Gerald, C., Walker, M.W., 2000. Galanin
Biol. Psychiatry 52, 126–135. receptor subtypes. Trends Pharmacol. Sci. 21, 109–117.
Bouchard, P., Quirion, R., 1997. [3 H]1,3-Di(2-tolyl)guanidine and Brandao, M.L., Cardoso, S.H., Melo, L.L., Motta, V., Coimbra, N.C.,
[3 H](+)pentazocine binding sites in the rat brain: autoradiographic 1994. Neural substrates of defensive behavior in the midbrain tectum.
visualization of the putative ␴1 and ␴2 receptor subtypes. Neuroscience Neurosci. Biobehav. Rev. 18, 339–346.
76, 467–477. Brandao, M.L., Castilho, V.M., Macedo, C.E., 2002. Role of
Boules, M., Warrington, L., Fauq, A., McCormick, D., Richelson, E., benzodiazepine and serotonergic mechanisms in conditioned freezing
2001. A novel neurotensin analog blocks cocaine- and d-amphetamine- and antinociception using electrical stimulation of the dorsal peri-
induced hyperactivity. Eur. J. Pharmacol. 426, 73–76. aqueductal gray as unconditioned stimulus. Soc. Neurosci. Abstr. 370.8.
Bourin, M., 1997. Animal models of anxiety: are they suitable for Brandao, M.L., Troncoso, A.C., de Souza Silva, M.A., Huston, J.P.,
predicting drug action in humans? Pol. J. Pharmacol. 49, 79–84. 2003. The relevance of neuronal substrates of defense in the midbrain
Bourin, M., Hascoët, M., 2003. The mouse light/dark box test. Eur. J. tectum to anxiety and stress: empirical and conceptual considerations.
Pharmacol. 463, 55–65. Eur. J. Pharmacol. 463, 225–233.
Bourin, M., Baker, G.B., Bradwejn, J., 1998. Neurobiology of panic Brann, M.R., Buckley, N.J., Bonner, T.I., 1988. The striatum and cerebral
disorder. J. Psychosom. Res. 44, 163–180. cortex express different muscarinic receptor mRNAs. FEBS Lett. 230,
Bouthillier, A., Blier, P., De Montigny, C., 1991. Flenbuterol, a ␤- 90–94.
adrenoceptor agonist, enhances serotonergic neurotransmission: an Bräuner-Osborne, H., Egebjerg, J., Nielsen, E.O., Madsen, U., Krogsgaard-
electrophysiological study in the rat brain. Psychopharmacology 103, Larsen, P., 2000. Ligands for glutamate receptors: design and
357–365. therapeutic prospects. J. Med. Chem. 43, 2609–2645.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 185

Braunstein-Bercovitz, H., 2000. Is the attentional dysfunction in norepinephrine and serotonin release detection in the behaving animal.
schizotypy related to anxiety? Schizophr. Res. 46, 255–267. Prog. Neuropsychopharmacol. Biol. Psychiatry 22, 353–386.
Breivogel, C.S., Griffin, G., Di Marzo, V., Martin, B.R., 2001. Evidence Brodkin, J., Busse, C., Sukoff, S.J., Varney, M.A., 2002. Anxiolytic-
for a new G protein-coupled cannabinoid receptor in mouse brain. like activity of the mGluR5 antagonist MPEP: a comparison with
Mol. Pharmacol. 60, 155–163. diazepam and buspirone. Pharmacol. Biochem. Behav. 73, 359–366.
Bremner, J.D., Krystal, J.H., Southwick, S.M., Charney, D.S., 1996a. Broersen, L.M., Abbate, F., Feenstra, M.G.P., de Bruin, J.P.C., Heinsbroek,
Noradrenergic mechanisms in stress and anxiety. I. Preclinical studies. R.P.W., Olivier, B., 2000. Prefrontal dopamine is directly involved in
Synapse 23, 28–38. the anxiogenic interoceptive cue of pentylenetetrazol but not in the
Bremner, J.D., Krystal, J.H., Southwick, S.M., Charney, D.S., 1996b. interoceptive cue of chlordiazepoxide in the rat. Psychopharmacology
Noradrenergic mechanisms in stress and anxiety. II. Clinical studies. 149, 366–376.
Synapse 23, 39–51. Broocks, A., Bandelow, B., George, A., Jestrabeck, C., Opitz, M.,
Brennan, J.A., Mauret, E., Comery, T.A., Cheh, M., Kwak, S.P., Rhodes, Bartmann, U., Gleiter, C.H., Meineke, I., Roed, I.-S., Rüther, E.,
K.J., Rosenzweig-Lipson, S., Marquis, K.L., 2002. Mice expressing Hajak, G., 2000. Increased psychological responses and divergent
an inactive KV␤ 1.1 subunit display an anxiolytic phenotype in neuroendocrine responses to m-CPP and ipsapirone in patients with
stress-induced hyperthermia and four-plate models. Soc. Neurosci. panic disorder. Int. Clin. Psychopharmacol. 15, 153–161.
Abstr. 398.12. Broqua, P., Wettstein, J.G., Rocher, M.N., Gauthier-Martin, B., Junien,
Breukel, A.I.M., Weigant, V.M., Lopes da Silva, F.H., Ghijsen, W.E.J.M., J.L., 1995. Behavioral effects of neuropeptide Y receptor agonists in
1998. Presynaptic modulation of cholecystokinin release by protein the elevated plus-maze and fear-potentiated startle procedures. Behav.
kinase C in the rat hippocampus. J. Neurochem. 70, 341–348. Pharmacol. 6, 215–222.
Brinton, R.D., 2002. Selective estrogen receptor modulators (SERM) for Brown, R.E., Stevens, D.R., Haas, H.L., 2001. The physiology of brain
the brain: recent advances and remaining challenges for developing a histamine. Prog. Neurobiol. 63, 637–672.
neuroSERM. Drug Dev. Res. 56, 380–392. Brown, R.E., Sergeeva, O.A., Eriksson, K.S., Haas, H.L., 2002.
Brinton, R.D., Thompson, R.H., Brownson, E.A., 2000. Spatial, cellular Convergent excitation of dorsal raphe serotonin neurons by multiple
and temporal basis of vasopressin potentiation of norepinephrine- arousal systems (orexin/hypocretin, histamine and noradrenaline). J.
induced cAMP formation. Eur. J. Pharmacol. 405, 73–88. Neurosci. 22, 8850–8859.
Brioni, J.D., O’Neill, A.B., Kim, D.J.B., Bucjley, M.J., Decker, M.W., Bruhwyler, J., Chleide, E., Liégeois, J.-F., Delarge, J., Mercier, M.,
Arneric, S.P., 1994. Anxiolytic-like effects of the novel cholinergic 1990. Anxiolytic potential of sulpiride, clozapine and derivatives in
channel activator ABT-418. J. Pharmacol. Exp. Ther. 271, 353–361. the open-field test. Pharmacol. Biochem. Behav. 36, 57–61.
Bristow, L.J., O’Connor, D., Watts, R., Duxon, M.S., Hutson, P.H., 2000. Bruinvels, A.T., Landwehrmeyer, B., Gustafon, E.L., Durkin, M.M.,
Evidence for accelerated desensitisation of 5-HT2C receptors following Mengod, G., Branchek, T.A., Hoyer, D., Palacios, J.M., 1994.
combined treatment with fluoxetine and the 5-HT1A receptor antagonist Localization of 5-HT1B , 5-HT1D␣ , 5-HT1E and 5-HT1F receptor
WAY 100,635 in the rat. Neuropharmacology 39, 1222–1236. messenger RNA in rodent and primate brain. Neuropharmacology 33,
Britton, K.T., Morgan, J., Rivier, J., Vale, W., Koob, G., 1985. 367–386.
Chlordiazepoxide attenuates response suppression induced by corti- Brunello, N., Masotto, C., Steardo, L., Markstein, R., Racagni, G., 1995.
cotropin-releasing factor in the conflict test. Psychopharmacology 86, New insights into the biology of schizophrenia through the mechanism
170–174. of action of clozapine. Neuropsychopharmacology 13, 177–213.
Britton, K.T., Page, M., Baldwin, H., Koob, G.F., 1991. Anxiolytic Brünig, I., Penschuck, S., Berninger, B., Benson, J., Fritschy, J.-M.,
activity of steroid anesthetic alphaxalone. J. Pharmacol. Exp. Ther. 2001. BDNF reduces miniature inhibitory postsynaptic currents by
258, 124–129. rapid downregulation of GABAA receptor surface. Eur. J. Neurosci.
Britton, K.T., McLeod, S., Koob, G.F., Hauger, R., 1992. Pregnane steroid 13, 1320–1328.
alphaxalone attenuates anxiogenic behavioral effects of corticotropin Bryans, J.S., Wustrow, D.J., 1999. 3-Substituted GABA analogs with
releasing factor and stress. Pharmacol. Biochem. Behav. 41, 399–403. central nervous system activity: a review. Med. Res. Rev. 19, 149–177.
Britton, K.T., Southerland, S., Van Uden, E., Kirby, D., Rivier, J., Koob, Buccafusco, J.J., 1992. Neuropharmacologic and behavioural actions
G., 1997. Anxiolytic activity of NPY receptor agonists in the conflict of clonidine: interactions with central neurotransmitters. Int. Rev.
test. Psychopharmacology 132, 6–13. Neurobiol. 33, 55–112.
Britton, K.T., Akwa, Y., Spina, M.G., Koob, G.F., 2000. Neuropeptide Bücheler, M.M., Hadamek, K., Hein, L., 2002. Two ␣2 -adrenergic
Y blocks anxiogenic-like behavioral action of corticotropin-releasing receptor subtypes, ␣2A and ␣2C , inhibit transmitter release in the
factor in an operant conflict test and elevated plus maze. Peptides 21, brain of gene-targeted mice. Neuroscience 109, 819–826.
37–44. Buckley, N.J., Bonner, T.I., Brann, M.R., 1988. Localization of a family
Broca, P., 1878. Anatomie comparée des circonvolutions cérébrales. Le of muscarinic receptor mRNAs in rat brain. J. Neurosci. 8, 4646–4652.
grand lobe limbique et la scissure dans la série des mammifères. Rev. Bugajski, J., Borycz, J., Gadek-Michalska, A., 1998. Involvement of
Anthropol. Paris 2, 285–498. the central noradrenergic system in cholinergic stimulation of the
Brocco, M., Koek, W., Degryse, A.-D., Colpaert, F.C., 1990. Comparative pituitary–adrenal response. J. Physiol. Pharmacol. 49, 285–292.
studies on the anti-punishment effects of chlordiazepoxide, buspirone Buller, R., 1995. Reversible inhibitors of monoamine oxidase A in
and ritanserin in the pigeon, Geller–Seifter and Vogel conflict anxiety disorders. Clin. Neuropharmacol. 18, S38–S44.
procedures. Behav. Pharmacol. 1, 403–418. Bunnemann, B., Fuxe, K., Ganten, D., 1993. The renin–angiotensin
Brocco, M., Dekeyne, A., Denorme, B., Monneyron, S., Gressier, H., system in the brain: an update. Regul. Pept. 46, 487–509.
Veiga, S., Millan, M.J., 1998. Actions of selective antagonists at Burbach, J.P., De Bree, F.M., Terwel, D., Tan, A., Maskova, H.P., Van
serotonin (5-HT)2A , 5-HT2B and 5-HT2C receptors in rodent models der Kleij, A.A., 1993. Properties of aminopeptidase activity involved
of potential anxiolytic properties. Behav. Pharmacol. 9 (Suppl. 1), S18. in the conversion of vasopressin by rat brain membranes. Peptides
Broderick, P.A., 1997. Alprazolam, diazepam, yohimbine, clonidine: in 14, 807–813.
vivo CA1 hippocampal norepinephrine and serotonin release profiles Burbach, J.P., Adan, R.A.H., Lolait, S.J., Van Leeuwen, F.W., Mezey,
under chloral hydrate anesthesia. Prog. Neuropsychopharmacol. Biol. E., Palkovits, M., Barberis, C., 1995. Molecular neurobiology and
Psychiatry 21, 1117–1140. pharmacology of the vasopressin/oxytocin receptor family. Cell. Mol.
Broderick, P.A., Hope, O., Jeannot, P., 1998. Mechanism of triazolo- Neurobiol. 15, 573–595.
benzodiazepine and benzodiazepine action in anxiety and depression: Burbach, J.P., Schoots, O., Hernando, F., 1998. Biochemistry of
behavioural studies with concomitant in vivo CA1 hippocampal vasopressin fragments. Prog. Brain Res. 119, 127–136.
186 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Burgess, N., Maguire, E.A., O’Keefe, J., 2002. The human hippocampus Canli, T., Desmond, J.E., Zhao, Z., Gabrieli, J.D.E., 2002. Sex differences
and spatial and episodic memory. Neuron 35, 625–641. in the neural basis of emotional memories. Proc. Natl. Acad. Sci.
Burnet, P.W.J., Michelson, D., Smith, M.A., Gold, P.W., Sternberg, E.M., U.S.A. 99, 10789–10794.
1994. The effect of chronic imipramine administration on the densities Canto-de-Souza, A., Nunes-de-Souza, R.L., Rodgers, R.J., 2002.
of 5-HT1A and 5-HT2 receptors and the abundancies of 5-HT receptor Anxiolytic-like effect of WAY-100635 microinfusions into the median
and transporter mRNA in the cortex, hippocampus and dorsal raphe (but not dorsal) raphe nucleus in mice exposed to the plus-maze:
of three strains of rat. Brain Res. 638, 311–324. influence of prior test experience. Brain Res. 928, 50–59.
Burnet, P.W.J., Chen, C.P., McGowan, S., Franklin, M., Harrison, P.J., Cao, B.-J., Rodgers, R.J., 1997a. Dopamine D4 receptor and anxiety:
1996. The effects of clozapine and haloperidol on serotonin1A , 2A behavioral profiles of clozapine, L-745,870 and L-741,742, in the
and 2C receptor gene expression and serotonin metabolism in the rat mouse plus-maze. Eur. J. Pharmacol. 335, 117–125.
forebrain. Neuroscience 73, 531–540. Cao, B.-J., Rodgers, R.J., 1997b. Influence of 5-HT1A receptor antagonism
Burnet, P.W.J., Miller, R., Lewis, L.J., Pic, Q., Sharp, T., Harrison, P.J., on plus-maze behaviour in mice. I. Pindolol enantiomers and pindobind
2001. Electroconvulsive shock increases tachykinin NK1 receptors, but 5-HT1A . Pharmacol. Biochem. Behav. 58, 583–591.
not the encoding mRNA, in rat cortex. Eur. J. Pharmacol. 413, 213–219. Cappell, H., Herman, C.P., 1972. Alcohol and tension reduction. Q. J.
Busto, U., Kaplan, H.L., Zawertailo, L., Sellers, E.M., 1994. Stud. Alcohol 33, 33–64.
Pharmacologic effects and abuse liability of bretazenil, diazepam, and Carboni, E., Wieland, S., Lan, N.C., Gee, K.W., 1996. Anxiolytic
alprazolam in humans. Clin. Pharmacol. Ther. 55, 451–463. properties of endogenously occurring pregnanediols in two rodent
Bylund, D.B., Eikenberg, D.C., Hieble, J.-P., Langer, S.Z., Lefkowitz, models of anxiety. Psychopharmacoloy 126, 173–178.
R.J., Minneman, K.P., Molinoff, P.B., Ruffolo, R.R., Trendelenburg, Carboni, E., Silvagni, A., Rolando, M.T.P., Di Chiara, G., 2000.
U., 1994. International Union of Pharmacology nomenclature of Stimulation of in vivo dopamine transmission in the bed nucleus of
adrenoceptors. Pharmacol. Rev. 46, 121–134. stria terminalis by reinforcing drugs. J. Neurosci. 20 (RC102), 1–5.
Caberlotto, L., Fuxe, K., hurd, Y.L., 2000. Characterization of NPY Cardinali, D.P., Gvozdenovich, E., Kaplan, M.R., Fainstein, I., Shifis,
mRNA-expressing cells in the human brain: co-localization with Y2 H.A., Pérez Lloret, S., Albornoz, L., Negri, A., 2002. A double-blind
but not Y1 mRNA in the cerebral cortex, hippocampus, amygdala and placebo-controlled study on melatonin efficacy to reduce anxiolytic
striatum. J. Chem. Neuroanat. 20, 327–337. benzodiazepine use in the elderly. Neuroendocrinol. Lett. 23, 55–60.
Cagetti, E., Liang, J., Spigelman, I., Olsen, R.W., 2003. Withdrawal from Carey, G.J., Varty, G.B., 2002. Nociceptin produces anxiolytic-like effects
chronic intermittent ethanol treatment changes subunit composition, in a rat conditioned lick suppression. Soc. Neurosci. Abstr. 683.10.
Carli, M., Samanin, R., 1988. Potential anxiolytic properties of 8-
reduces synaptic function, and decreases behavioral responses to
hydroxy-2-(di-N-propylamino)tetralin, a selective serotonin1A receptor
positive allosteric modulators of GABAA receptors. Mol. Pharmacol.
agonist. Psychopharmacology 94, 84–91.
63, 53–64.
Carli, M., Prontera, C., Samanin, R., 1989. Evidence that central 5-
Cahill, L., McGaugh, J.L., 1998. Mechanisms of emotional arousal and
hydroxytryptaminergic neurons are involved in the anxiolytic activity
lasting declarative memory. Trends Neurosci. 21, 294–299.
of buspirone. Br. J. Pharmacol. 96, 829–836.
Cai, X., Flores-Hernandez, J., Feng, J., Yan, Z., 2002a. Activity-dependent
Carlson, J.N., Haskew, R., Wacker, J., Maisonneuve, I.M., Glick, S.D.,
bidirectional regulation of GABAA receptor channels by the serotonin
Jerussi, T.P., 2001. Sedative and anxiolytic effects of zopiclone’s
5-HT4 receptor-mediated signalling in rat prefrontal cortical pyramidal
enantiomers and metabolite. Eur. J. Pharmacol. 415, 181–189.
neurons. J. Physiol. 540, 743–749.
Carlsson, M., Carlsson, A., 1990. Interactions between glutamatergic
Cai, X., Gu, Z., Zhong, P., Ren, Y., Yan, Z., 2002b. Serotonin 5-
and monoaminergic systems within the basal ganglia: implications for
HT1A receptors regulate AMPA receptor channels through inhibiting
schizophrenia and Parkinson’s disease. Trends Neurosci. 13, 272–276.
Ca2+ /calmodulin-dependent kinase II in prefrontal cortical pyramidal Carobrez, A.P., Teixeira, K.V., Graeff, F.G., 2001. Modulation of
neurons. J. Biol. Chem. 277, 36553–36562. defensive behavior by periaqueductal gray NMDA/glycine-B receptor.
Cain, C.K., Blouin, A.M., Barad, M., 2002. l-Type voltage-gated Neurosci. Biobehav. Rev. 25, 697–709.
calcium channels are required for extinction, but not for acquisition or Carrasco, G.A., Van de Kar, L., 2003. Neuroendocrine pharmacology of
expression, of conditional fear in mice. J. Neurosci. 22, 9113–9121. stress. Eur. J. Pharmacol. 463, 235–272.
Calo, G., Guerrini, R., Rizzi, A., Salvadori, S., Regoli, E., 2000. Cartmell, J., Schoepp, D.D., 2000. Regulation of neurotransmitter release
Pharmacology of nociceptin and its receptor: a novel therapeutic by metabotropic glutamate receptors. J. Neurochem. 75, 889–907.
target. Br. J. Pharmacol. 129, 1261–1283. Casatti, C.A., Elias, C.F., Sita, L.V., Frigo, L., Furlani, V.C.G., Bauer,
Calvo, N., Volosin, M., 2000. Glucocorticoid and mineralocorticoid J.A., Bittencourt, J.C., 2002. Distribution of melanin-concentrating
receptors are involved in the facilitation of anxiety-like response hormone neurons projecting to the medial mammillary nucleus.
induced by restraint. Neuroendocrinology 73, 261–271. Neuroscience 115, 899–915.
Campbell, A.D., McBride, W.J., 1995. Serotonin3 receptor and ethanol- Cases, O., Seif, I., Grimsby, J., Gaspar, P., Chen, K., et al., 1995.
stimulated dopamine release in the nucleus accumbens. Pharmacol. Aggressive behavior and altered amounts of brain serotonin and
Biochem. Behav. 51, 835–842. norepinephrine in mice lacking MAO A. Science 268, 1763–1766.
Cananzi, A.R., Costa, E., Guidotti, A., 1980. Potentiation by Castilho, V.M., Macedo, C.E., Brandao, M.L., 2002. Role of
intraventricular muscimol of the anticonflict effect of benzodiazepines. benzodiazepine and serotonergic mechanisms in conditioned freezing
Brain Res. 196, 447–453. and antinociception using electrical stimulation of the dorsal
Cancela, L.M., Basso, A.M., Martijena, I.D., Capriles, N.R., Molina, V.A., periaqueductal gray as unconditioned stimulus in rats. Psychophar-
2001. A dopaminergic mechanism is involved in the “anxiogenic-like” macology 165, 77–85.
response induced by chronic amphetamine treatment: a behavioural Castro, M.E., Diaz, A., Del Olmo, E., Pazos, A., 2003. Chronic fluoxetine
and neurochemical study. Brain Res. 909, 179–186. induces opposite changes in G protein coupling at pre- and postsynaptic
Candenas, M.L., Cintado, C.G., Pennefather, J.N., Pereda, M.T., Loizaga, 5-HT1A receptors in rat brain. Neuropharmacology 44, 93–101.
J.M., Maggi, C.A., Pinto, F.M., 2002. Identification of a tachykinin Catterall, W.A., 2000. From ionic currents to molecular mechanisms: the
NK2 receptor splice variant and its expression in human and rat structure and function of voltage-gated sodium channels. Neuron 26,
tissues. Life Sci. 72, 269–277. 13–25.
Cangioli, I., Baidi, E., Mannaioni, P.F., Bucherelli, C., Blandina, P., Cavazzuti, E., Bertolini, A., Vergoni, A.V., Arletti, R., Poggioli, R.,
Passani, M.B., 2002. Activation of histaminergic H3 receptors in the Forgione, A., Benelli, A., 1999. l-sulpiride, at a low, non-neuroleptic
rat basolateral amygdala improves expression of fear memory and dose, prevents conditioned fear stress-induced freezing behavior in
enhances acetylcholine release. Eur. J. Neurosci. 16, 521–528. rats. Psychopharmacology 143, 20–23.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 187

Ceccarelli, I., Casamenti, F., Massafra, C., Pepeu, G., Scali, C., Aloisi, Chaouloff, F., Baudrie, V., Coupry, I., 1994. Effects of chlorisondamine
A.M., 1999. Effects of novelty and pain on behavior and hippocampal and restraint on cortical [3 H]ketanserin binding, 5-HT2A receptor-
extracellular ACh levels in male and female rats. Brain Res. 815, mediated head shakes and behaviours in models of anxiety.
169–176. Neuropharmacology 33, 449–456.
Cecchi, M., Khoshbouei, H., Morilak, D.A., 2002. Modulatory effects Charney, D.S., Deutch, A., 1996. A functional neuroanatomy of anxiety
of norepinephrine, acting on ␣1 receptors in the central nucleus of and fear: implication for the pathophysiology and treatment of anxiety
the amygdala, on behavioral and neuroendocrine responses to acute disorders. Crit. Rev. Neurobiol. 10, 419–446.
immobilization stress. Neuropharmacology 43, 1139–1147. Charney, D.S., Heninger, G.R., Redmond, D.E., 1983. Yohimbine induced
Ceci, A., Smith, M., French, E.D., 1988. Activation of the A10 mesolimbic anxiety and increased noradrenergic function in humans: effects of
system by the ␴-receptor agonist (+)SKF 10,047 can be blocked by diazepam and clonidine. Life Sci. 33, 19–29.
rimcazole, a novel putative antpsychotic. Eur. J. Pharmacol. 154, 5–11. Charney, D.S., Woods, S.W., Goodman, W.K., Heninger, G.R., 1987.
Celada, P., Siuciak, J.A., Tran, T.M., Altar, A., Tepper, J.M., 1996. Serotonin function in anxiety. II. Effects of the serotonin agonist mCPP
Local infusion of brain-derived neurotrophic factor modifies the firing in panic disorder patients and healthy subjects. Psychopharmacology
pattern of dorsal raphé serotonergic neurons. Brain Res. 712, 293–298. 92, 14–24.
Celada, P., Puig, V.M., Casanovas, J.M., Guillazo, G., Artigas, F., 2001. Charrier, D., Dangoumau, L., Hamon, M., Puech, A.J., Thiébot, M.-H.,
Control of dorsal raphe serotonergic neurons by the medial prefrontal 1994. Effects of 5-HT1A receptor ligands on a safety signal withdrawal
cortex: involvement of serotonin1A , GABAA , and glutamate receptors. procedure of conflict in the rat. Pharmacol. Biochem. Behav. 48,
J. Neurosci. 21, 9917–9929. 281–289.
Cenni, B., Picard, D., 1999. Ligand-dependent activation of steroid Charrier, D., Dangoumau, L., Puech, A.J., Hamon, M., Thiébot, M.-H.,
receptors: new roles for old players. Trends Endocrinol. Metab. 10, 1995. Failure of CCK receptor ligands to modify anxiety-related
41–46. behavioural suppression in an operant conflict paradigm in rats.
Centonze, D., Usiello, A., Gubellini, P., Pisani, A., Borrelli, E., Psychopharmacology 121, 127–134.
Bernardi, G., Calabresi, P., 2002. Dopamine D2 receptor-mediated Chatterton, J.E., Awobuluyi, M., Premkumar, L.S., Takahashi, H.,
inhibition of dopaminergic neurons in mice lacking D2L receptors. Talantova, M., Shin, Y., Cui, J., Tu, S., Sevarino, K.A., Nakanishi,
Neuropsychopharmacology 27, 723–726. N., 2002. Excitatory glycine receptors containing the NR3 family of
Cervo, L., Samanin, R., 1995. Presynaptic 5-HT1A receptors mediate NMDA receptor subunits. Nature 415, 793–798.
the effects of ipsapirone on punished responding in rats. Eur. J. Chavigne, C., McLaughlin, J.P., Celver, J.P., 2001. Regulation of opioid
Pharmacol. 284, 249–255. receptor function by chronic agonist exposure: constitutive activity
Cervo, L., Mocaër, E., Bertaglia, A., Samanin, R., 2000. Roles of and desensitisation. Mol. Pharmacol. 60, 20–25.
5-HT1A receptors in the dorsal raphe and dorsal hippocampus in Chazot, P.L., Hann, V., Wilson, C., Lees, G., Thompson, C.L., 2001.
anxiety assessed by the behavioral effects of 8-OH-DPAT and S15535 Immunological identification of the mammalian H3 histamine receptor
in a modified Geller–Seifter conflict model. Neuropharmacology 39, in the mouse brain. NeuroReport 12, 259–262.
1037–1043. Cheeta, S., Kenny, P.J., File, S.E., 2000. Hippocampal and septal injection
Cestari, I.N., Uchida, I., Li, L., Burt, D., Yang, J., 1996. The agonistic of nicotine and 8-OH-DPAT distinguish among different animal tests of
action of pentobarbital on GABAA ␤-subunit homomeric receptors. anxiety. Prog. Neuropsychopharmacol. Biol. Psychiatry 24, 1053–1067.
NeuroReport 7, 943–947. Cheeta, S., Irvine, E., File, S.E., 2001a. Social isolation modifies nicotine’s
Ceulemans, D.L.S., Hoppenbrouwers, M.-L.J.A., Gelders, Y.G., Reyntjens, effects in animal tests of anxiety. Br. J. Pharmacol. 132, 1389–1395.
A.J.M., 1985. The influence of ritanserin, a serotonin antagonist, Cheeta, S., Tucci, S., Sandhu, J., Williams, A.R., Rupniak, N.M.J., File,
in anxiety disorders: a double-blind placebo-controlled study versus S.E., 2001b. Anxiolytic actions of the substance P (NK1 ) receptor
lorazepam. Pharmacopsychiatry 18, 303–305. antagonist L-760735 and the 5-HT1A agonist 8-OH-DPAT in the
Chai, S.Y., Bastias, M.A., Clune, E.F., Matsacos, D.J., Mustafa, T., Lee, social interaction test in gerbils. Brain Res. 915, 170–175.
J.H., McDowall, S.G., Mendelsohn, F.A.O., Albiston, A.L., Paxinos, Chemin, J., Monteil, A., Perez-Reyes, E., Nargeot, J., Lory, P., 2001.
G., 2000. Distribution of angiotensin IV binding sites (AT4 receptor) Direct inhibition of T-type calcium channels by the endogenous
in the human forebrain, midbrain and pons as visualised by in vitro cannabinoid anandamide. EMBO J. 20, 7033–7040.
receptor autoradiography. J. Chem. Neuroanat. 20, 339–348. Chen, X., Herbert, J., 1995. Alterations in sensitivity to intracerebral
Chaki, S., Hirota, S., Funakoshi, T., Suzuki, Y., Suetake, S., Okubo, vasopressin and the effects of a V1a receptor antagonist on cellular,
T., Ishii, T., Nakazato, A., Okuyama, S., 2003a. Anxiolytic-like and autonomic and endocrine responses to repeated stress. Neurosci. 64,
antidepressant-like activities of MCL0129 1-[(S)-2-(4-fluorophenyl)-2- 687–697.
(4-isopropylpiperadin-1-yl)ethyl]-4-[4-(2-methoxynaphthalen-1- yl)but- Chen, Y.Z., Qiu, J., 2001. Possible genomic consequence of nongenomic
yl]piperazine), a novel and potent nonpeptide antagonist of the actions of glucocorticoids in neural cells. News Physiol. Sci. 16,
melanocortin-4 receptor. J. Pharmacol. Exp. Ther. 304, 818–826. 292–296.
Chaki, S., Kawashima, N., Suzuki, Y., Shimazaki, T., Okuyama, S., 2003b. Chen, C., Rainnie, D.G., Greene, R.W., Tonegawa, S., 1994. Abnormal
Cocaine- and amphetamine-regulated transcript peptide produces fear response and aggressive behavior in mutant mice deficient for
anxiety-like behavior in rodents. Eur. J. Pharmacol. 464, 49–54. alpha-calcium-calmodulin kinase II. Science 266, 291–294.
Chakos, M., Lieberman, J., Hoffman, E., Bradford, D., Sheitman, B., Chen, C.L., Yang, Y.R., Chiu, T.H., 1999. Activation of rat locus
2001. Effectiveness of second-generation antipsychotics in patients coeruleus neuron GABA(A) receptors by propofol and its potentiation
with treatment-resistant schizophrenia: a review and meta-analysis of by pentobarbital or alphaxalone. Eur. J. Pharmacol. 386, 201–210.
randomised trials. Am. J. Psychiatry 158, 518–526. Chen, L.-W., Wei, L.-C., Liu, H.-L., Rao, Z.-R., 2000a. Noradrenergic
Chalmers, D.T., Lovenberg, T.W., De Souza, E.B., 1995. Localization of neurons expressing substance P receptor (NK1 ) in the locus coeruleus
novel corticotropin-releasing factor-receptor (CRF2 ) mRNA expression complex: a double immunofluorescence study in the rat. Brain Res.
to specific subcortical nuclei in rat brain: comparison with CRF1 873, 155–159.
receptor mRNA expression. J. Neurosci. 15, 6340–6350. Chen, Y., Brunson, K.L., Müller, M.B., Cariaga, W., Baram, T.Z., 2000b.
Chambers, M.S., Fletcher, S.R., 2000. CCKB antagonists in the control Immunocytochemical distribution of corticotropin-releasing hormone
of anxiety and gastric acid secretion. Prog. Med. Chem. 37, 45–81. receptor type-1 (CRF1 )-like immunoreactivity in the mouse brain:
Chaouloff, F., 1995. Regulation of 5-HT receptors by corticosteroids: light microscopy analysis using an antibody directed against the
where do we stand? Fundam. Clin. Pharmacol. 9, 219–233. C-terminus. J. Comp. Neurol. 420, 305–323.
188 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Chen, Y., Hu, C., Hsu, C.-K., Zhang, Q., Bi, C., Asnicar, M., Hsiung, Clark, M.S., Sexton, T.J., McClain, M., Root, D., Kohen, R., Neumaier,
H.M., Fox, N., Slieker, L.J., Yang, D.D., Heiman, M.L., Shi, Y., J.F., 2002. Overexpression of 5-HT1B receptors in dorsal raphe nucleus
2002a. Targeted disruption of the melatonin-concentrating hormone using herpes simplex virus gene transfer increases anxiety behavior
receptor1 results in hyperphagia and resistance to diet-induced obesity. after inescapable stress. J. Neurosci. 22, 4550–4562.
Endocrinology 143, 2469–2477. Clément, Y., Chapoutier, G., 1998. Biological bases of anxiety. Neurosci.
Chen, Y., Yao, Y., Penington, N.J., 2002b. Effect of pertussis toxin and N- Biobehav. Rev. 22, 623–633.
ethylmaleimide on voltage-dependent and -independent calcium current Clément, Y., Calatayud, F., Belzung, C., 2002. Genetic basis of anxiety-
modulation in serotonergic neurons. Neuroscience 111, 207–214. like behaviour: a critical review. Brain Res. Bull. 57, 57–71.
Cheng, C.H.K., Costall, B., Kelly, M.E., Naylor, R.J., 1994. Actions of Clemett, D.A., Cockett, M.I., Marsden, C.A., Fone, LK.C., 1998.
5-hydroxytryptophan to disinhibit mouse behaviour in the light/dark Antisense oligonucleotide-induced reduction in 5-hydroxytryptamine7
test. Eur. J. Pharmacol. 255, 39–49. receptors in the rat hypothalamus without alteration in exploratory
Cheng, L.L., Wang, S.J., Gean, P.W., 1998. Serotonin depresses excitatory behaviour or neuroendocrine function. J. Neurochem. 71, 1271–1279.
synaptic transmission and depolarization-evoked Ca2+ influx in rat Clemett, D.A., Punhani, T., Duxon, M.S., Blackburn, T.P., Fone, K.C.F.,
basolateral amygdala via 5-HT1A receptors. Eur. J. Neurosci. 10, 2000. Immunohistochemical localization of the 5-HT2C receptor
2163–2172. protein in the rat CNS. Neuropharmacology 39, 123–132.
Cherubini, E., Conti, F., 2001. Generating diversity at GABAergic Clifton, P.G., Lee, M.D., Somerville, E.M., Kennett, G.A., Dourish,
synapses. Trends Neurosci. 24, 155–162. C.T., 2003. 5-HT1B receptor knock-out mice show a compensatory
Chessell, I.P., Black, M.D., Feniuk, W., Humphrey, P.P.A., 1996. reduction in 5-HT2C receptor function. Eur. J. Neurosci. 17, 185–190.
Operational characteristics of somatostatin receptors mediating Cloninger, C.R., 1987. Neurogenetic adaptive mechanisms in alcoholism.
inhibitory actions on rat locus coeruleus neurones. Br. J. Pharmacol. Science 236, 410–416.
117, 1673–1678. Coco, M.L., Kuhn, C.M., Ely, T.D., Kilts, C.D., 1992. Selective
Chizh, B.A., Headley, P.M., Tzschentke, T.M., 2001. NMDA receptor activation of mesoamygdaloid dopamine neurons by conditioned
antagonists as analgesics: focus on the NR2B subtype. Trends stress: attenuation by diazepam. Brain Res. 590, 39–47.
Pharmacol. Sci. 22, 636–642. Coffin, V.L., Spealman, R.D., 1985. Modulation of the behavioral effects
Cho, L., Tsunoda, M., Sharma, R.P., 1999. Effects of endotoxin and of chlordiazepoxide by methylxanthines and analogs of adenosine in
tumor necrosis factor alpha on regional brain neurotransmitters in squirrel monkeys. J. Pharmacol. Exp. Ther. 235, 724–728.
mice. Nat. Toxins 7, 187–195. Cohen, G.A., Doze, V.A., Madison, D.V., 1992. Opioid inhibition
Choi, B.H., Choi, J.-S., Yoon, S.H., Rhie, D.-J., Min, D.S., Jo, Y.-H., of GABA release from presynaptic terminals of rat hippocampal
Kim, M.-S., Hahn, S.J., 2001. Effects of norfluoxetine, the major interneurons. Neuron 9, 325–335.
metabolite of fluoxetine, on the cloned neuronal potassium channel Cohen, C., Perrault, G., Voltz, C., Steinberg, R., Soubrié, P., 2002.
Kv3.1. Neuropharmacology 41, 443–453. SR141716, a central cannabinoid (CB1 ) receptor antagonist, blocks
Choi, D.-S., Wang, D., Dadgar, J., Chang, W.S., Messing, R.O., 2002. the motivational and dopamine-releasing effects of nicotine in rats.
Conditional rescue of protein kinase Cεregulates ethanol preference Behav. Pharmacol. 13, 451–463.
and hypnotic sensitivity in adult mice. J. Neurosci. 22, 9905–9911. Cole, B.J., Hillman, M., Seidelmann, D., Klewer, M., Jones, G.H.,
Chojnacka-Wójcik, E., Klodzińska, A., 1992. Involvement of 5-HT1B 1995a. Effects of benzodiazepine receptor partial inverse agonists in
receptors in the anticonflict effect of m-CPP in rats. J. Neural Transm. the elevated plus maze test of anxiety in the rat. Psychopharmacology
87, 87–96. 121, 118–126.
Chojnacka-Wójcik, E., Przegaliński, E., 1991. Evidence for the Cole, J.C., Burroughs, G.J., Laverty, C.R., Sheriff, N.C., Sparham,
involvement of 5-HT1A receptors in the anticonflict effect of ipsapirone E.A., Rodgers, R.J., 1995b. Anxiolytic-like effects of yohimbine
in rats. Neuropharmacology 30, 703–709. in the murine plus-maze: strain independence and evidence against
Chojnacka-Wójcik, E., Tatarczyńska, E., Dereń-Wesolek, A., 1996a. ␣2 -adrenoceptor mediation. Psychopharmacology 118, 425–436.
Effect of glycine on antidepressant- and anxiolytic-like action of 1- Collado, I., Pedregal, C., Mazon, A., Espinosa, R.F., Blanco-Urgoiti,
aminocyclopropanecarboxylic acid (ACPC) in rats. Pol. J. Pharmacol. J., Schoepp, D.D., Wright, R.A., Johnson, B.G., Kingston, A.E.,
48, 627–629. 2002. (2S),1 S,2 S(3 R)-2-(2 -Carboxy-3 -methylcyclopropyl) glycine
Chojnacka-Wójcik, E., Tatarczýnska, E., Pilc, A., 1996b. Anxiolytic-like is a potent and selective metabotropic group 2 receptor agonist with
effects of metabotropic glutamate antagonist (RS)-alpha-methylserine- anxiolytic properties. J. Med. Chem. 45, 3619–3629.
O-phosphate in rats. Pol. J. Pharmacol. 48, 507–509. Collinson, N., Kuenzi, F.M., Jarolimek, W., Maubahc, K.A., Cothliff, R.,
Chojnacka-Wójcik, E., Tatarczýnska, E., Pilc, A., 1997. The anxiolytic- Sur, C., Smith, A., Out, F.M., Howell, O., Atack, J.R., McKernan,
like effect of metabotropic glutamate receptor antagonists after R.M., Seabrook, G.R., Dawson, G.R., Whiting, P.J., Rosahl, T.W.,
intrahippocampal injection in rats. Eur. J. Pharmacol. 319, 153–156. 2002. Enhanced learning and memory and altered GABAergic synaptic
Chopin, P., Briley, M., 1987. Animal models of anxiety: the effect of transmission in mice lacking the ␣5 subunit of the GABAA receptor.
compounds that modify 5-HT neurotransmission. Trends Pharmacol. J. Neurosci. 22, 5572–5580.
Sci. 8, 383–388. Colombo, G., Agabio, R., Lobina, C., Reali, R., Zocchi, A., Fadda, F.,
Chopin, P., Moret, C., Briley, M., 1994. Neuropharmacology of 5-hydro- Gessa, G.L., 1995. Sardinian alcohol-preferring rats: a genetic animal
xytryptamine1B/1D receptor ligands. Pharmacol. Ther. 62, 385–405. model of anxiety. Physiol. Behav. 57, 1181–1185.
Christopoulos, A., El-Fakahany, E.E., 1999. The generation of nitric Commissaris, R.L., Lyness, W.H., Rech, R.H., 1981. The effects
oxide by G protein-coupled receptors. Life Sci. 64, 1–15. of d-lysergic acid diethylamide (LSD), 2,5-dimethoxy-4-methyl-
Ciliax, B.J., Nash, N., Heilman, C., Sunahara, R., Hartney, A., Tiberi, M., amphetamine (DOM), pentobarbital and methaqualone on punished
Rye, D.B., Caron, M.G., Niznik, H.B., Levey, A.I., 2000. Dopamine responding in control and 5,7-dihydroxytryptamine-treated rats.
D5 receptor immunolocalization in rat and monkey brain. Synapse Pharmacol. Biochem. Behav. 14, 617–623.
37, 125–145. Commissaris, R.L., McCloskey, T.C., Damian, G.M., Brown, B.D.,
Clare, J.J., Tate, S.N., Nobbs, M., Romanos, M.A., 2000. Voltage-gated Barraco, R.A., Altman, H.J., 1990. Antagonism of the anti-conflict
sodium channels as therapeutic targets. Drug Discov. Today 5, 506–520. effects of phenobarbital, but not diazepam, by the A-1 adenosine
Clark, P.B.S., Reuben, M., 1996. Release of [3 H]noradrenaline from agonist I-PIA. Psychopharmacology 102, 283–290.
rat hippocampal synaptosomes by nicotine: mediation by different Commons, K.G., Connolley, K.R., Valentino, R.J., 2003. A
nicotinic receptor subtypes from striatal [3 H]-dopamine release. Br. J. neurochemically distinct dorsal raphe-limbic circuit with a potential
Pharmacol. 117, 595–606. role in affective disorders. Neuropsychopharmacology 28, 206–215.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 189

Concas, A., Santoro, G., Serra, M., Sanna, E., Biggio, G., 1991. Corda, M.G., Orlandi, M., Fratta, W., 1990. Proconflict effect of
Neurochemical action of the general anaesthetic propofol on the ACTH1−−24 : interaction with benzodiazepines. Pharmacol. Biochem.
chloride ion channel coupled with GABAA receptors. Brain Res. 542, Behav. 36, 631–634.
232–255. Cordeaux, Y., Briddon, S.J., Megson, A.E., McDonnell, J., Dickenson,
Condren, R.M., Dinan, T.G., Thakore, J.H., 2002. A preliminary J.M., Hill, S.J., 2000. Influence of receptor number on functional
study of buspirone stimulated prolactin release in generalized responses elicited by agonists acting at the human adenosine A1
social phobia: evidence for enhanced serotonergic responsivity? Eur. receptor: evidence for signalling pathway-dependent changes in agonist
Neuropsychopharmacol. 12, 349–354. potency and relative intrinsic activity. Mol. Pharmacol. 58, 1075–1084.
Conley, R.K., Cumberbatch, M.J., Mason, G.S., Williamson, D.J., Corradetti, R., Laaris, N., Hanoun, N., Laporte, A.-M., Le Poul, E.,
Harrison, T., Locker, K., Swain, C., Maubach, K., O’Donnell, R., Hamon, M., Lanfumey, L., 1998. Antagonist properties of (−)-
Rigby, M., Hewson, L., Smith, D., Rupniak, N.M.J., 2002. Substance pindolol and WAY100,635 at somatodendritic and postsynaptic
P (neurokinin1 ) receptor antagonists enhance dorsal raphe neuronal 5-HT1A receptors in the rat brain. Br. J. Pharmacol. 123, 449–462.
activity. J. Neurosci. 22, 7730–7736. Cornea-Hébert, V., Riad, M., Wu, C., Singh, S.K., Descarries, L., 1999.
Conn, P.J., Pin, J.P., 1997. Pharmacology and functions of metabotropic Cellular and subcellular distribution of the serotonin 5-HT2A receptor in
glutamate receptors. Annu. Rev. Pharmacol. Toxicol. 37, 205–237. the central nervous system of adult rat. J. Comp. Neurol. 409, 187–209.
Connell, J., Graeff, F., Guimaraes, F.S., Hellewell, J.S.E., Hetem, L.A., Corti, C., Aldegheri, L., Somogyl, P., Ferraguti, F., 2002. Distribution
Deakin, J.F.W., 1995. 5-HT2 receptors and anxiety. Behav. Pharmacol. and synaptic localization of the metabotropic glutamate receptor4
6 (Suppl. 1), 35. (mGluR4 ) in the rodent CNS. Neuroscience 110, 403–420.
Connell, J., Sennef, C., Deakin, J.F.W., 1998. The effects of ORG 12962, Costall, B., Naylor, R.J., 1995. Behavioural interactions between 5-
a 5-HT2C receptor agonist, in two models of experimentally induced hydroxytryptophan, neuroleptic agents and 5-HT receptor antagonists
anxiety in healthy female volunteers. In: Proceedings of the XXIst in modifying rodent responding to aversive situations. Br. J. Pharmacol.
CINP Congress, Glasgow, p. 352. 116, 2989–2999.
Connor, T.J., Leonard, B.E., 1998. Depression, stress and immunological Costall, B., Naylor, R.J., 1997. The influence of 5-HT2 and 5-HT4
activation: the role of cytokines in depressive disorders. Life Sci. 62, receptor antagonists to modify drug induced disinhibitory effects in
583–606. the mouse light/dark test. Br. J. Pharmacol. 122, 1105–1118.
Connor, M., Vaughan, C.W., Jennings, E.A., Allen, R.G., Christie, Costall, B., Kelly, M.E., Naylor, R.J., Onaivi, E.S., Tyers, M.B., 1989.
M.J., 1999. Nociceptin, phe1 -nociceptin1–13 , nocistatin and Neuroanatomical sites of action of 5-HT3 receptor agonists and
prepronociception154–181 effects on calcium channel currents and a antagonists for alteration of aversive behaviour in the mouse. Br. J.
potassium current in rat locus coeruleus in vitro. Br. J. Pharmacol. Pharmacol. 96, 325–332.
128, 1779–1787. Costall, B., Domeney, A.M., Gerrard, P.A., Horovitz, Z.P., Kelly, M.E.,
Contarino, A., Gold, L.H., 2002. Targeted mutations of the corticotropin- Naylor, R.J., Tomkins, D.M., 1990. Effects of Captopril and SQ29,852
releasing factor system: effects on physiology and behavior. on anxiety-related behaviours in rodent and marmoset. Pharmacol.
Neuropeptides 26, 103–116. Biochem. Behav. 36, 13–20.
Contarino, A., Dellu, F., Koob, G.F., Smith, G.W., Lee, K.-F., Vale, W., Costall, B., Domeney, A.M., Hughes, J., Kelly, M.E., Naylor, R.J.,
Gold, L.H., 1999a. Reduced anxiety-like and cognitive performance Woodruff, G.N., 1991. Anxiolytic effects of CCK-B antagonists.
in mice lacking the corticotropin-releasing factor receptor 1. Brain Neuropeptides 19, 65–73.
Res. 835, 1–9. Coste, S.C., Kesterson, R.A., Heldwein, K.A., Stevens, S.L., Heard,
Contarino, A., Heinrichs, S.C., Gold, L.H., 1999b. Understanding A.D., Hollis, J.H., Murray, S.E., Hill, J.K., Pantely, G.A., Hohimer,
corticotropin releasing factor neurobiology: contributions from mutant A.R., 2000. Abnormal adaptation to stress and impaired cardiovascular
mice. Neuropeptides 33, 1–12. function in mice lacking corticotropin-releasing hormone receptor2 .
Conti, P., De Amici, M., De Micheli, C., 2002. Selective agonists and Nat. Genet. 24, 403–409.
antagonists for kainate receptors. Mini Rev. Med. Chem. 2, 177–184. Coste, S.C., Murray, S.E., Stenzel-Poore, M.P., 2001. Animals models of
Cooke, L.W., Acharya, A.J., Davis, M.D., 1998. The effects of the ␴1 CRH excess and CRH receptor deficiency display altered adaptations
ligand, igmesine, on neurotransmitter activity in the rat as assessed by to stress. Peptides 22, 733–741.
intracerebral microdialysis. Am. Soc. Neurosci. Abstr. 24, 583–614. Coupland, N.J., Wilson, S.J., Nutt, D., 1996. ␣2 -Adrenoceptors in panic
Cools, A.R., Ellenbroek, B., Heeren, D., Lubbers, L., 1993. Use of high and anxiety disorders. J. Psychopharmacol. 10, 26–34.
and low responders to novelty in rat studies on the role of the ventral Cozzi, N.V., Nichols, D.E., 1996. 5-HT2A receptor antagonists inhibit
striatum in radial maze performance: effects of intra-accumbens potassium-stimulated ␥-aminobutyric acid release in rat frontal cortex.
injections of sulpiride. Can. J. Physiol. Pharmacol. 71, 335–342. Eur. J. Pharmacol. 309, 25–31.
Corbett, R., Dunn, R.W., 1991. Effects of HA-966 on conflict, social Cratty, M.S., Birkle, D.L., 1999. N-Methyl-d-aspartate (NMDA)-mediated
interaction and plus-maze behaviors. Drug Dev. Res. 24, 301–305. corticotropin-releasing factor (CRF) release in cultured rat amygdala
Corbett, R., Dunn, R.W., 1993. Effects of 5,7 dichlorokynurenic neurons. Peptides 20, 93–100.
acid on conflict, social interaction and plus maze behaviors. Cravatt, B.F., Demarest, K., Patricelli, M.P., Bracey, M.H., Giang, D.K.,
Neuropharmacology 32, 461–466. Martin, B.R., Lichtman, A.H., 2001. Supersensitivity to anandamide
Corbett, R., Hartman, H., Kerman, L.L., Woods, A.T., Strupczewski, and enhanced endogenous cannabinoid signaling in mice lacking fatty
J.T., Helsley, G.C., Conway, P.C., Dunn, R.W., 1993. Effects of acid amide hydrolase. Proc. Natl. Acad. Sci. U.S.A. 98, 9371–9376.
atypical antipsychotic agents on social behavior in rodents. Pharmacol. Crawley, J., Glowa, J.R., Maajewska, M.D., Paul, S.M., 1986. Anxiolytic
Biochem. Behav. 45, 9–17. activity of an endogenous adrenal steroid. Brain Res. 398, 382–385.
Corbin, A.E., Meltzer, L.T., Ninteman, F.W., Wiley, J.N., Christoffersen, Crawley, J.N., Mufson, E.J., Hohmann, J.G., Teklemichael, D., Steiner,
C.L., Wustrow, D.J., Wise, L.D., Pugsley, T.A., Heffner, T.G., 2000. R.A., Holmberg, K., Xu, Z.-Q.D., Blakeman, K.H., Xu, X.-J.,
PD 158771, a potential antipsychotic agent with D2 /D3 partial agonist Wiesenfeld-Hallin, Z., Bartfai, T., Hökfelt, T., 2002. Galanin
and 5-HT1A agonist actions. II. Preclinical behavioural effects. overexpressing transgenic mice. Neuropeptides 36, 145–156.
Neuropharmacology 39, 1211–1221. Cremers, T.I.F.H., Wiersma, L.J., Bosker, F.J., Den Boer, J.A., Westerink,
Corda, M.G., Blaker, W.D., Mendelson, W.B., Guidotti, A., Costa, E., B.H.C., Wikström, H.V., 2001. Is the beneficial antidepressant
1983. ␤-Carbolines enhance shock-induced suppression of drinking in effect of coadministration of pindolol really due to somatodendritic
rats. Proc. Natl. Acad. Sci. U.S.A. 80, 2072–2076. autoreceptor antagonism? Biol. Psychiatry 50, 13–21.
190 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Crespi, F., Wright, I.K., Möbius, C., 1992. Isolation rearing of rats alters Cussac, D., Newman-Tancredi, A., Duqueyroix, D., Pasteau, V., Millan,
release of 5-hydroxytryptamine and dopamine in the frontal cortex: M.J., 2002b. Differential activation of Gq/11 and Gi3 proteins at
an in vivo electrochemical study. Exp. Brain Res. 88, 495–501. 5-hydroxytryptamine2C receptors revealed by antibody capture assays:
Crestani, F., Lorez, M., Baer, K., Essrich, C., Benke, D., Laurent, J.P., influence of receptor reserve and relationship to agonist-directed
Belzung, C., Fritschy, J.-M., Lüscher, B., Mohler, H., 1999. Decreased trafficking. Mol. Pharmacol. 62, 578–589.
GABAA -receptor clustering results in enhanced anxiety and a bias Cussac, D., Newman-Tancredi, A., Quentric, Y., Carpentier, N.,
for threat cues. Nat. Neurosci. 2, 833–839. Poissonnet, G., Parmentier, J.-G., Goldstein, S., Millan, M.J., 2002c.
Crestani, F., Martin, J.R., Rudolph, U., 2000. Mechanism of action of Characterization of phospholipase C activity at h5-HT2C compared
the hypnotic zolpidem in vivo. Br. J. Pharmacol. 131, 1251–1254. with h5-HT2B receptors: influence of novel ligands upon membrane-
Crestani, F., Löw, K., Keist, R., Mandelli, M.-J., Möhler, H., Rudolph, bound levels of [3 H]phosphatidylinositols. Naunyn Schmiedebergs
U., 2001. Molecular targets for the myorelaxant action of diazepam. Arch. Pharmacol. 365, 242–252.
Mol. Pharmacol. 59, 442–445. Cyr, M., Charbonneau, C., Morissette, M., Rochford, J., Barden, N., Di
Crestani, F., Assandri, R., Täuber, A., Martin, J.R., Rudolph, U., Paolo, T., 2001. Central 5-hydroxytryptamine2A receptor expression
2002a. Contribution of the ␣1 -GABAA receptor subtype to the in transgenic mice bearing a glucocorticoid receptor antisense.
pharmacological actions of benzodiazepine site inverse agonists. Neuroendocrinology 73, 37–45.
Neuropharmacology 43, 679–684. Czlonkowska, A., Siemiatkowski, M., Plaźnik, A., 1997. Some behavioral
Crestani, F., Keist, R., Fritschy, J.-M., Benke, D., Vogt, K., Prut, effects of AMPA/kainate receptor agonist and antagonists. J. Physiol.
L., Blüthmann, H., Möhler, H., Rudolph, U., 2002b. Trace fear Pharmacol. 48, 479–488.
conditioning involves hippocampal ␣5 GABAA receptors. Proc. Natl. Czlonkowska, A., Sienkiewicz-Jarosz, H., Siemiatkowski, M., Bidzinski,
Acad. Sci. U.S.A. 99, 8980–8985. A., Plaźnik, A., 1999. The effects of neurosteroids on rat behavior
Crick, F., Koch, C., 1998. Constraints on cortical and thalamic projections: and 3 H-muscimol binding in the brain. Pharmacol. Biochem. Behav.
the no-strong-loops hypothesis. Nature 391, 245–251. 63, 639–646.
Crider, A.M., 2002. Recent advances in the development of nonpeptide Czyrak, A., Mackowiak, M., Chocyk, A., Fijal, K., Tokarski, K., Bijak,
somatostatin receptor ligands. Mini Rev. Med. Chem. 2, 507–517. M., Wedzony, K., 2002. Prolonged corticosterone treatment alters
Crissman, A.M., O’Donnell, J.M., 2002. Effects of antidepressants in the responsiveness of 5-HT1A receptors to 8-OH-DPAT in rat CA1
rats trained to discriminate centrally administered isoproterenol. J. hippocampal neurons. Naunyn Schmiedebergs Arch. Pharmacol. 366,
Pharmacol. Exp. Ther. 302, 606–611. 357–367.
Criswell, H.E., Breese, G.R., 1989. A conflict procedure not requiring Dalley, J.W., Theobald, D.E., Eagle, D.M., Passetti, F., Robbins,
deprivation: evidence that chronic ethanol treatment induces tolerance T., 2002. Deficits in impulse control associated with tonically-
to the anticonflict action of ethanol and chlordiazepoxide. Alcohol. elevated serotonergic function in rat prefrontal cortex. Neurop-
Clin. Exp. Res. 13, 680–685. sychopharmacology 26, 716–728.
Criswell, H.E., Ming, Z., Griffith, B.L., Breese, G.R., 2003. Comparison Dallvechia-Adams, S.E., Kuhar, M.J., Smith, Y., 2002. Cocaine-
of effect of ethanol on N-methyl-d-aspartate- and GABA-gated currents and amphetamine-regulated transcript peptide projections in the
from acutely dissociated neurons: absence of regional differences in ventral midbrain: colocalization with ␥-aminobutyric acid, melanin-
sensitivity to ethanol. J. Pharmacol. Exp. Ther. 304, 192–199. concentrating hormone, dynorphin, and synaptic interactions with
Croll, S.D., Suri, C., Compton, D.L., Simmons, M.V., Yancopoulos, dopamine neurons. J. Comp. Neurol. 448, 360–372.
G.D., Lindsay, R.M., Wiegand, S.J., Rudge, J.S., Scharfman, Dalvi, A., Rodgers, R.J., 1996. GABAergic influences on plus-maze
H.E., 1999. Brain-derived neurotrophic factor transgenic mice behaviours in mice. Psychopharmacology 128, 380–397.
exhibit passive avoidance deficits, increased seizure severity and in Damasio, A.R., 1989. The brain binds entities and events by multiregional
vitro hyperexcitability in the hippocampus and entorhinal cortex. activation from convergence zones. Neural Comput. 1, 123–132.
Neuroscience 93, 1491–1506. Damgen, K., Lüddens, H., 1999. Zaleplon displays a selectivity to
Crowder, T.L., Ariwodola, O.J., Weiner, J.L., 2002. Ethanol antagonizes recombinant GABAA receptors different from zolpidem, zopiclone
kainate receptor-mediated inhibition of evoked GABAA inhibitory and benzodiazepines. Neurosci. Res. Commun. 25, 139–148.
postsynaptic currents in the rat hippocampal CA1 region. J. Pharmacol. Danbolt, N.C., 2001. Glutamate uptake. Prog. Neurobiol. 65, 1–105.
Exp. Ther. 303, 937–944. Dani, J.A., 2001. Overview of nicotinic receptors and their roles in the
Csaba, Z., Dournaud, P., 2001. Cellular biology of somatostatin receptors. central nervous system. Biol. Psychiatry 49, 166–174.
Neuropeptides 35, 1–23. Danysz, W., Parsons, C.G., 1998. Glycine and N-methyl-d-aspartate
Cullinan, W.E., 2000. GABAA receptor subunit expression within receptors: physiological significance and possible therapeutic
hypophysiotropic CRH neurons: a dual hybridisation histochemical applications. Pharmacol. Rev. 50, 597–664.
study. J. Comp. Neurol. 419, 344–351. Dascal, N., 1997. Signaling via the G protein-activated K+ channels.
Culpepper, L., 2002. Generalized anxiety disorder in primary care: Cell Signal 9, 551–573.
emerging issues in management and treatment. J. Clin. Psychiatry 63, Daugé, V., Léna, I., 1998. CCK in anxiety and cognitive processes.
35–42. Neurosci. Biobehav. Rev. 22, 815–825.
Cunningham, M.O., Jones, R.S.G., 2000. The anticonvulsant lamotrigine Daugé, V., Sebret, A., Beslot, F., Matsui, T., Roques, B.P., 2001.
decreases spontaneous glutamate release but increases spontaneous Behavioral profile of CCK2 receptor-deficient mice. Neuropsy-
GABA release in the rat enthohinal cortex in vitro. Neuropharmacology chopharmacology 25, 690–698.
39, 2139–2146. Dautzenberg, F.M., Hauger, R.L., 2002. The CRF peptide family and
Cussac, D., Newman-Tancredi, A., Nicolas, J.-P., Boutin, J.-A., Millan, their receptors: yet more partners discovered. Trends Pharmacol. Sci.
M.J., 2000. Antagonist properties of the novel antipsychotic, 23, 71–77.
S16924, at cloned, human serotonin 5-HT2C receptors: a parallel Davidson, C., Stamford, J.A., 2000. Effect of chronic paroxetine treatment
phosphatidylinositol and calcium accumulation comparison with on 5-HT1B and 5-HT1D autoreceptors in rat dorsal raphe nucleus.
clozapine and haloperidol. Naunyn Schmiedebergs Arch. Pharmacol. Neurochem. Int. 36, 91–96.
361, 549–554. Davidson, R.J., Abercrombie, H., Nitschke, J.B., Putnam, K., 1999.
Cussac, D., Duqueyroix, D., Newman-Tancredi, A., Millan, M.J., 2002a. Regional brain function, emotion and disorders of emotion. Curr.
Stimulation by antipsychotic agents of mitogen-activated protein Opin. Neurobiol. 9, 228–234.
kinase (MAPK) coupled to cloned, human (h)serotonin (5-HT)1A Davis, M., 1992a. The role of the amygdala in fear and anxiety. Annu.
receptors. Psychopharmacology 162, 168–177. Rev. Neurosci. 15, 353–375.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 191

Davis, M., 1992b. The role of the amygdala in fear-potentiated startle: defensive-withdrawal behavior in rats. J. Pharmacol. Exp. Ther. 276,
implications for animal models of anxiety. Trends Pharmacol. Sci. 13, 56–64.
35–41. De Gasparo, M., Catt, K.J., Inagami, T., Wright, J.W., Unger, Th.,
Davis, M., Shi, C., 1999. The extended amygdala: are the central nucleus 2000. International union of pharmacology. XXIII. The angiotensin II
of the amygdala and the bed nucleus of the stria terminalis differentially receptors. Pharmacol. Rev. 52, 415–472.
involved in fear versus anxiety? Ann. N. Y. Acad. Sci. 877, 281–291. Degroot, A., Treit, D., 2002. Dorsal and ventral hippocampal cholinergic
Davis, M., Rainnie, D., Cassell, M., 1994. Neurotransmission in the rat systems modulate anxiety in the plus-maze and shock-probe tests.
amygdala related to fear and anxiety. Trends Neurosci. 17, 208–214. Brain Res. 949, 60–70.
Davis, P.A., Hanna, M.C., Hales, T.G., Kirkness, E.F., 1997. A novel Degroot, A., Kashluba, S., Treit, D., 2001. Septal GABAergic and
class of GABAA receptor subunit confers insensitivity to anaesthetic hippocampal cholinergic systems modulate anxiety in the plus-maze
agents. Nature 385, 820–823. and shock-probe tests. Pharmacol. Biochem. Behav. 69, 391–399.
Davis, P.A., Pistis, M., Hanna, M.C., Peters, J.A., Lambert, J.J., Hales, De Groote, L., Olivier, B., Westenberg, H.G.M., 2002. The effects of
T.G., Kirkness, E.F., 1999. The 5-HT3B subunit is a major determinant selective serotonin reuptake inhibitors on extracellular 5-HT levels
of serotonin receptor function. Nature 397, 359–363. in the hippocampus of 5-HT1B receptor knock-out mice. Eur. J.
Davis, S.N., Pertwee, R.G., Riedel, G., 2002. Functions of cannabinoid Pharmacol. 439, 93–100.
receptors in the hippocampus. Neuropharmacology 42, 993–1007. Dekeyne, A., Denorme, B., Monneyron, S., Millan, M.J., 2000a.
Dawson, G.R., Rupniak, N.M.D., Iversen, S.D., Curnow, R., Tye, S., Citalopram reduces social interaction in rats by activation of serotonin
Stanhope, K.J., Tricklebank, M.D., 1995. Lack of effect of CCKB (5-HT2C ) receptors. Neuropharmacology 39, 1114–1117.
receptor antagonists in ethological and conditioned animal screens for Dekeyne, A., Brocco, M., Adhumeau, A., Gobert, A., Millan, M.J., 2000b.
anxiolytic drugs. Psychopharmacology 121, 109–117. The selective serotonin (5-HT)1A receptor ligand, S15535, displays
Dawson, L.A., Nguyen, H.Q., Li, P., 2000. In vivo effects of the 5- anxiolytic-like effects in the social interaction and Vogel models
HT6 antagonist SB-271046 on striatal and frontal cortex extracellular and suppresses dialysate levels of 5-HT in the dorsal hippocampus
concentrations of noradrenaline, dopamine, 5-HT, glutamate and of freely-moving rats. A comparison with other anxiolytic agents.
aspartate. Br. J. Pharmacol. 130, 23–26. Psychopharmacology 152, 55–66.
Dawson, L.A., Nguyen, H.Q., Li, P., 2001. The 5-HT6 receptor antagonist Dekeyne, A., Brocco, M., Papp, M., MacSweeney, C., Peglion, J.-L.,
SB-271046 selectively enhances excitatory neurotransmission in the Millan, M.J., 2001. Actions of the novel naphtoxazine and dopamine
rat frontal cortex and hippocampus. Neuropsychopharmacology 25, D3 /D2 receptor agonist, S32504, in models predictive of antidepressant
662–668. properties. Soc. Neurosci. Abstr. 27, 665.4.
Day, H.E.W., Campeau, S., Watson, S.J., Akil, H., 1997. Distribution of
De Kloet, E.R., Joels, M., Oitzl, M., Sutanto, W., 1991. Implication of
␣1A -, ␣1B -, and ␣1D - adrenergic receptor mRNA in the rat brain and
brain corticosteroid receptor diversity for the adaptation syndrome
spinal cord. J. Chem. Neuroanat. 13, 115–139.
concept. Methods Achiv. Exp. Pathol. 14, 104–132.
Day, H.E.W., Campeau, S., Watson, S.J., Akil, H., 1999a. Expression
De Kloet, E.R., Vreugdenhil, E., Oitzl, M.S., Joels, M., 1998. Brain
of ␣1B adrenoceptor mRNA in corticotropin-releasing-hormone-
corticosteroid receptor balance in health and disease. Endocr. Rev. 19,
containing cells of the rat hypothalamus and its regulation by
269–301.
corticosterone. J. Neurosci. 15, 10098–10106.
Del Arco, A., Mora, F., 2001. Dopamine release in the prefrontal cortex
Day, H.E.W., Curran, E.J., Watson III, S.J., Akil, H., 1999b. Distinct
during stress is reduced by the ionotropic activation of glutamate
neurochemical populations in the rat central nucleus of the amydala
receptors. Brain Res. Bull. 56, 125–130.
and bed nucleus of the stria terminalis: evidence for their selective
Del Arco, A., Segovia, G., Mora, F., 2001. Dopamine release during stress
activation by interleukin-1␤. J. Comp. Neurol. 413, 113–128.
in the prefrontal cortex of the rat decreases with age. NeuroReport
Dazzi, L., Spiga, F., Pira, L., Ladu, S., Vacca, G., Rivano, A., Jentsch,
12, 4019–4022.
J.D., Biggio, G., 2001. Inhibition of stress- or anxiogenic-drug induced
increases in dopamine release in the rat prefrontal cortex by long-term Del Bel, E.A., Guimaraes, F.S., 1997. Social isolation increases
treatment with antidepressant treatment. J. Neurochem. 76, 1212–1220. cholecystokinin mRNA in the central nervous system of rats.
Dazzi, L., Ladu, S., Spiga, F., Vacca, F., Rivano, A., Pira, L., Biggio, G., NeuroReport 8, 3597–3600.
2002. Chronic treatment with imipramine or mirtazapine antagonizes De Ligt, R.A.F., Kourounakis, A.P., Ijzerman, P., 2000. Inverse agonism
stress- and FG7142-induced increase in cortical norepinephrine output at G-protein-coupled receptors: (patho)physiological relevance and
in freely moving rats. Synapse 43, 70–77. implications for drug discovery. Br. J. Pharmacol. 130, 1–12.
Deakin, J.F.W., 1988. 5-HT2 receptors, depression and anxiety. Pharmacol. De Lima, T.C.M., Gavioli, E.C., Duarte, F.S., Rae, G.A., Guerrini, R.,
Biochem. Behav. 29, 819–820. Calo, G., 2002. Are GABAA receptors involved in the behavioural
Deakin, J.F.W., 1991. Depression and 5-HT. Int. Clin. Psychopharmacol. effects of nociceptin and nocistatin C-terminal hexapeptide in mice
6, 23–28. in the plus-maze test? Soc. Neurosci. Abstr. 683.9.
Deakin, J.F.W., Graeff, F.G., 1991. 5-HT and mechanisms of defence. J. De Martinis, N., Rynn, M., Rickels, K., Mandos, L., 2000. Prior
Psychopharmacol. 5, 305–315. benzodiazepine use and buspirone response in the treatment of
De Araújo, J.E., Huston, J.P., Brandão, M.L., 2001. Opposite effects of generalized anxiety disorder. J. Clin. Psychiatry 61, 91–94.
substance P fragments C (anxiogenic) and N (anxiolytic) injected into DeMet, E.M., Chicz-DeMet, A., 2002. Localization of adenosine A2A -
dorsal periaqueductal gray. Eur. J. Pharmacol. 432, 43–51. receptors in rat brain with [3 H]ZM-241385. Naunyn Schmiedebergs
De Boer, S.F., Koolhaas, J.M., 2003. Defensive burying in rodents: Arch. Pharmacol. 366, 478–481.
ethology, neurobiology and psychopharmacology. Eur. J. Pharmacol. De Miranda, J., Panizzutti, R., Foltyn, V.N., Wolosker, H., 2002. Cofactors
463, 145–161. of serine racemase that physiologically stimulate the synthesis of
De Felipe, C., Herrero, J.F., O’Brien, J.A., Palmer, J.A., Doyle, C.A., the N-methyl-d-aspartate (NMDA) receptor co-agonist d-serine. Proc.
Smith, A.J.H., Laird, J.M.A., Belmonte, C., Cervero, F., Hunt, S.P., Natl Acad. Sci. U.S.A. 99, 14542–14547.
1998. Altered nociception, analgesia and aggression in mice lacking Denac, H., Mevissen, M., Scholtysik, G., 2000. Structure, function
the receptor for substance P. Nature 392, 394–397. and pharmacology of voltage-gated sodium channels. Naunyn
De Fonseca, F.R., Rubio, P., Menzaghi, F., Merlo-Pich, E., Rivier, J., Schmiedebergs Arch. Pharmacol. 362, 453–479.
Koob, G.F., Navarro, M., 1996. Corticotropin-releasing factor (CRF) Den Boer, J.A., Westenberg, H.G.M., 1990. Serotonin function in panic
antagonist [d-Phe12 , Nle21,38 , C␣ MeLeu37 ] CRF attenuates the acute disorder: a double-blind placebo-controlled study with fluvoxamine
actions of the highly potent cannabinoid receptor agonist HU-210 on and ritanserin. Psychopharmacology 102, 85–94.
192 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Dennis, T., Dubois, A., Benavides, J., Scatton, B., 1988. Distribution of Ding, Y.-Q., Shigemoto, R., Takada, M., Ohishi, H., Nakanishi, S.,
central omega (benzodiazepine 1) and omega 2 (benzodiazepine 2) Mizuno, N., 1996. Localization of the neuromedin K receptor NK3 in
receptor subtypes in the monkey and human brain. An autoradiographic the central nervous system of the rat. J. Comp. Neurol. 364, 290–310.
study with [3 H]flunitrazepam and the omega 1 selective ligand Dingledine, R., Borges, K., Bowie, D., Traynelis, S.F., 1999. The
[3 H]zolpidem. J. Pharmacol. Exp. Ther. 247, 309–322. glutamate receptor ion channels. Pharmacol. Rev. 51, 7–61.
De Novellis, V., Marabese, I., Palazzo, E., Rossi, F., Berrino, L., Rodella, Dinh, T.P., Carpenter, D., Leslie, F.M., Freund, T.F., Katona, I., Sensi,
L., Bianchi, R., Rossi, F., Maione, S., 2003. Group I metabotropic S.L., Kathuria, S., Piomelli, D., 2002. Brain monoglyceride lipase
glutamate receptors modulate glutamate and ␥-aminobutyric acid participating in endocannabinoid inactivation. Proc. Natl. Acad. Sci.
release in the periaqueductal grey of rats. Eur. J. Pharmacol. 462, 73–81. U.S.A. 99, 10819–10824.
Depoortere, H., Zivkovic, B., Lloyd, K.G., Sanger, D.J., Perrault, Di Piero, V., Ferracuti, S., Sabatini, U., Tombari, D., Di Legge, S.,
G., Langer, S.Z., Bartholini, G., 1986. Zolpidem, a novel Pantano, P., Cruccu, G., Lenzi, G.L., 2001. Diazepam effects on the
nonbenzodiazepine hypnotic. I. Neuropharmacological and behavioral cerebral responses to tonic pain: a SPET study. Psychopharmacology
effects. J. Pharmacol. Exp. Ther. 237, 649–658. 158, 252–258.
Dépôt, M., Caillé, G., Mukherjee, J., Katzman, M.A., Cadieux, A., Dirks, A., Pattij, T., Bouwknecht, J.A., Westphal, T.T., Hijzen, T.H.,
Bradwejn, J., 1999. Acute and chronic role of 5-HT3 neuronal Groenink, L., Oosting, R.S., Hen, R., Geyer, M.A., Olivier, B., 2001.
system on behavioral and neuroendocrine changes induced by 5-HT1B receptor knock-out, but not 5-HT1A receptor knock-out mice,
intravenous cholecystokinin tetrapeptide administration in humans. show reduced startle reactivity and footshock-induced sensitisation, as
Neuropsychopharmacology 20, 177–187. measured with the acoustic startle response. Behav. Brain Res. 118,
De Oliveira, R.M.W., Del Bel, E.A., Mamede-Rosa, M.L.N., Padovan, 169–178.
C.M., Deakin, J.F.W., Guimaraes, F.S., 2000. Expression of neuronal Dirks, A., Groenink, L., Bouwknecht, J.A., Hijzen, T.H., Van Der Gugten,
nitric oxide synthase mRNA in stress-related brain areas after restraint J., Ronken, E., Verbeek, J.S., Veening, J.G., Dederen, P.J.W.C.,
in rats. Neurosci. Lett. 289, 123–126. Korosi, A., Schoolderman, L.F., Roubos, E.W., Olivier, B., 2002.
Desousa, N.J., Wunderlich, G.R., De Cabo, C., Vaccarino, F.J., 1998. Overexpression of corticotropin-releasing hormone in transgenic mice
Individual differences in sucrose intake predict behavioral reactivity in and chronic stress-like autonomic and physiological alterations. Eur.
rodent models of anxiety. Pharmacol. Biochem. Behav. 60, 841–846. J. Neurosci. 16, 1751–1760.
De Souza, E.B., 1995. Corticotropin-releasing factor receptors: physiology, Dixon, A.K., Gubitz, A.K., Sirinathsinghji, D., Richardson, P.J., Freeman,
pharmacology, biochemistry and role in central nervous system and T.C., 1996. Tissue distribution of adenosine receptor mRNA in the
immune disorders. Psychoneuroendocrinology 20, 789–819. rat. Br. J. Pharmacol. 118, 1461–1468.
De Souza, M., Schenberg, L.C., Carobrez, A.P., 1998. NMDA-coupled Djeridane, Y., Touitou, Y., 2001. Chronic diazepam administration
periaqueductal gray glycine receptors modulate anxioselective drug differentially affects melatonin synthesis in rat pineal and harderian
effects on plus-maze performance. Behav. Brain Res. 90, 157–165. glands. Psychopharmacology 154, 403–407.
Devaud, L.L., Fritschy, J.M., Sieghart, W., Morrow, A.L., 1997.
Doble, A., 1999. New insights into the mechanism of action of hypnotics.
Bidirectional alterations of GABAA receptor subunit peptide levels
J. Psychopharmacol. 13, S11–S20.
in rat cortex during chronic ethanol consumption and withdrawal. J.
Dockstader, C.L., Van Der Kooy, D., 2001. Mouse strain differences in
Neurochem. 69, 126–130.
opiate reward learning are explained by differences in anxiety, not
Devi, S.A., 2001. Heterodimerization of G-protein-coupled receptors:
reward or learning. J. Neurosci. 21, 9077–9081.
pharmacology, signaling and trafficking. Trends Pharmacol. Sci. 22,
Doherty, M.D., Pickel, V.M., 2000. Ultrastructural localization of the
532–537.
5-HT2A receptor in dopaminergic neurons in the ventral tegmental
Devoto, P., Flore, G., Pira, L., Diana, M., Gessa, G.L., 2002. Co-release
area. Brain Res. 864, 176–185.
of noradrenaline and dopamine in the prefrontal cortex after acute
morphine and during morphine withdrawal. Psychopharmacology 160, Dolmetsch, R.E., Pajvani, U., Fife, K., Spotts, J.M., Greenberg, M.E.,
220–224. 2001. Signaling to the nucleus by an l-type calcium channel-
DeVries, C.-A., Young, W.-S., Nelson, R.-J., 1997. Reduced aggressive calmodulin complex through the MAP kinase pathway. Science 294,
behaviour in mice with targeted disruption of the oxytocin gene. J. 333–339.
Neuroendocrinol. 9, 363–368. Domyancic, A.V., Morilak, D.A., 1997. Distribution of ␣1A adrenergic
De Vry, J., 1995. 5-HT1A receptor agonists: recent developments and receptor mRNA in the rat brain visualized by in situ hybridization. J.
controversial issues. Psychopharmacology 121, 1–26. Comp. Neurol. 386, 358–378.
Diaz-Cabiale, Z., Narvaez, J.A., Finnman, U.-B., Bellido, I., Ogren, S.O., Dooley, D.J., Donovan, C.M., Pugsley, T.A., 2000a. Stimulus-dependent
Fuxe, K., 2000. Galanin-(−16) modulates 5-HT1A receptors in the modulation of [3 H]norepinephrine release from rat neocortical slices by
ventral limbic cortex of the rat. NeuroReport 11, 515–519. GABApentin and pregabalin. J. Pharmacol. Exp. Ther. 295, 1086–1093.
Diez-Ariza, M., Ramirez, M.J., Lasheras, B., Del Rio, J., 1998. Dooley, D.J., Mieske, C.A., Borosky, S.A., 2000b. Inhibition of K+ -
Differential interaction between 5-HT3 receptors and GABAergic evoked glutamate release form rat neocortical and hippocampal slices
neurons inhibiting acetylcholine release in rat entorhinal cortex slices. by GABApentin. Neurosci. Lett. 280, 107–110.
Brain Res. 81, 228–232. Dooley, D.J., Donovan, C.M., Meder, W.P., Whetzel, S.Z., 2002.
Di Matteo, V., La Grutta, V., Esposito, E., 2001. m-Chlorophenylpiperazine Preferential action of GABApentin and pregagalin at P/Q-type
excites non-dopaminergic neurons in the rat substantia nigra and ventral voltage-sensitive calcium channels: inhibition of K+ -evoked [3 H]nore-
tegmental area by activating serotonin2C receptors. Neuroscience 103, pinephrine release from rat neocortical slices. Synapse 45, 171–190.
111–116. Dorow, R., Horowski, R., Paschelke, G., Amin, M., Braestrup, C.,
DiMarzo, V., Breivogel, C.S., Tao, Q., Bridgen, D.T., Razdan, R.K., 1983. Severe anxiety induced by FG 7142, a ␤-carboline ligand for
Zimmer, A.M., Zimmer, A., Martin, B.R., 2000. Levels, metabolism benzodiazepine receptors. Lancet 9, 98–99.
and pharmacological activity of anandamide in CB1 cannabinoid Dossin, O., Moulédous, L., Baudry, X., Tafani, J.-A.-M., Mazarguil, H.,
receptor knock-out mice: evidence for non-CB1 , non-CB2 receptor- Zajac, J.-M., 2000. Characterization of a new radioiodinated probe for
mediated actions of anandamide in mouse brain. J. Neurochem. 75, the ␣2C adrenoceptor in the mouse brain. Neurochem. Int. 36, 7–18.
2434–2444. Doucet, E., Miquel, M.C., Nosjean, A., Vergé, D., Hamon, M., Emerit,
Dimmock, P.W., Wyatt, K.M., Jones, P.W., O’Brien, P.M., 2000. Efficacy M.B., 2000. Immunolabeling of the rat central nervous system with
of selective serotonin-reuptake inhibitors in premenstrual syndrome: a antibodies partially selective of the short form of the 5-HT3 receptor.
systematic review. Lancet 356, 1131–1136. Neuroscience 95, 881–892.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 193

Dournaud, P., Gu, Y.Z., Schonbrunn, A., Mazella, J., Tanenbaum, G.S., fluoxetine-induced secretion of corticosterone and progesterone. J.
Beaudet, A., 1996. Localization of the somatostatin receptor SST2A Pharmacol. Exp. Ther. 285, 579–587.
in rat brain using a specific anti-peptide antibody. J. Neurosci. 16, Dunn, A.J., Swiergiel, A.H., 1999. Behavioral responses to stress arc
4468–4478. intact in CRF-deficient mice. Brain Res. 845, 14–20.
Dournaud, P., Boudin, H., Schonbrunn, A., Tannenbaum, G.S., Beaudet, Dunn, A.J., Green, E.J., Isaacson, R.L., 1979. Intracerebral adreno-
A., 1998. Interrelationships between somatostatin sst2A receptors and corticotropic hormone mediates novelty-induced grooming in the rat.
somatostatin-containing axons in rat brain: evidence for regulation of Science 203, 281–283.
cell surface receptors by endogenous somatostatin. J. Neurosci. 18, Dunn, R.W., Corbett, R., Fielding, S., 1989. Effects of 5-HT1A receptor
1056–1071. agonists and NMDA receptor antagonists in the social interaction test
Dournaud, P., Slama, A., Beaudet A., Epelbaum, J., 2000. Somatostatin and the elevated plus-maze. Eur. J. Pharmacol. 169, 1–10.
receptors in the central nervous system. In: Quirion, R., Björklund, Dunn, R.W., Flanagan, D.M., Martin, L.L., Kerman, L.L., Woods, A.T.,
A., Hökfelt, T. (Eds.), Handbook of Chemical Neuroanatomy. Peptide Camacho, F., Wilmot, C.A., Corngeldt, M.L., Effland, R.C., Wood,
Receptors, vol. 16, Part I. Elsevier, Amsterdam, pp. 1–43. P.L., Corbett, R., 1992. Stereoselective R-(+) enantiomer of HA-966
Doyle, D.A., Morais Cabral, J., Pfuetzner, R.A., Kuo, A., Gulbis, J.M., displays anxiolytic effects in rodents. Eur. J. Pharmacol. 214, 207–214.
Cohen, S.L., Chait, B.T., MacKinnon, R., 1998. The structure of the Dunn, R.W., Del Vecchio, R.A., LaMarca, S., 1995. The nitric oxide
potassium channel: molecular basis of K+ conduction and selectivity. synthase inhibitors NG -nitro-l-arginine methyl ester and NG -nitro-l-
Science 280, 69–77. arginine are anxiolytic and are devoid of benzodiazepine side-effects
Drake, C.T., Milner, T.A., Patterson, S.L., 1999. Ultrastructural in rats. Soc. Neurosci. Abstr. 21, 1384.
localization of full-lengh trkB immunoreactivity in rat hippocampus Dunn, R.W., Reed, T.A.W., Copemand, P.D., Frye, C.A., 1998. The nitric
suggests multiple roles in modulating activity-dependent synaptic oxide synthase inhibitor 7-nitroindazole displays enhanced anxiolytic
plasticity. J. Neurosci. 19, 8009–8026. efficacy without tolerance in rats following subchronic administration.
Driggers, P.H., Segars, J.H., 2002. Estrogen action and cytoplasmic Neuropharmacology 37, 899–904.
signaling pathways. Part II. The role of growth factors and phosphory- Durand, M., Aguerre, S., Fernandez, F., Edno, L., Combourieu, L.,
lation in estrogen signaling. Trends Endocrinol. Metab. 13, 422–427. Mormede, P., Chaouloff, F., 2000. Strain-dependent neurochemical
Drugan, R.C., Skolnick, P., Paul, S.M., Crawley, J.N., 1986. Low doses and neuroendocrine effects of desipramine, but not fluoxetine or
of muscimol produce anticonflict actions in the lateral septum of the imipramine, in spontaneously hypertensive and Wistar–Kyoto rats.
rat. Neuropharmacology 25, 203–205. Neuropharmacology 39, 2464–2477.
Durel, L.A., Krantz, D.S., Barretti, J.E., 1986. The antianxiety effect of
Drutel, G., Peitsaro, N., Karlstedt, K., Wieland, K., Smit, M.J.,
beta-blockers on punished responding. Pharmacol. Biochem. Behav.
Timmerman, H., Panula, P., Leurs, R., 2001. Identification of rat
25, 371–374.
H3 receptor isoforms with different brain expression and signalling
Dursun, S.M., Deakin, J.F., 2001. Augmenting antipsychotic treatment
properties. Mol. Pharmacol. 59, 1–8.
with lamotrigine or topiramate in patients with treatment-
Dubin, A.E., Huvar, R., D’Andrea, M.R., Pyati, J., Zhu, J.Y., Joy,
resistant schizophrenia: a naturalistic case-series outcome study. J.
K.C., Wilson, S.J., Galindo, J.E., Glass, C.A., Luo, L., Jackson,
Psychopharmacol. 15, 297–301.
M.R., Lovenberg, T.W., Erlander, M.G., 1999. The pharmacological
Dussossoy, D., Carayon, P., Belugou, S., Feraut, D., Bord, A., Goubet,
and functional characteristics of the serotonin 5-HT3A receptor are
C., Roque, C., Vidal, H., Combes, T., Loison, G., Casellas, P., 1999.
specifically modified by a 5-HT3B receptor subunit. J. Biol. Chem.
Colocalization of sterol isomerase and sigma1 receptor at endoplasmic
274, 30799–30810.
reticulum and nuclear envelope level. Eur. J. Biochem. 263, 377–386.
Dubinsky, B., Vaidya, A.H., Rosenthal, D.I., Hochman, C., Crooke,
Duxon, M.S., Flanigan, T.P., Reavley, A.C., Baxter, G.S., Blackburn,
J.J., Deluca, S., Devine, A., Cheo-Isaacs, C.T., Carter, A.R., Jordan,
T.P., Fone, K.C.F., 1997a. Evidence for expression of the 5-
A.D., Rietz, A.B., Shank, R.P., 2002. 5-Ethoxymethyl-7-fluoro-3-
hydroxytryptamine2B receptor protein in the rat central nervous
oxo-1,2,3,5-tetrahydrobenzo[4,5]imidazo[1,2a] pyridine-4-N-(2-fluoro-
system. Neuroscience 76, 323–329.
phenyl)carboxamide (RWJ-51204), a new nonbenzodiazepine anxi- Duxon, M.S., Kennett, G.A., Lightower, S., Blackburn, T.P., Fone, K.C.F.,
olytic. J. Pharmacol. Exp. Ther. 303, 777–790. 1997b. Activation of 5-HT2B receptors in the medial amygdala causes
Duka, T., Millan, M.J., Ulsamer, B., Doenicke, A., 1982. Naloxone anxiolysis in the social interaction test in the rat. Neuropharmacology
attenuates the anxiolytic action of diazepam in man. Life Sci. 31, 36, 601–608.
1833–1836. Duxon, M.S., Starr, K.R., Upton, N., 2000. Latency to paroxetine-induced
Dulawa, S.C., Grandy, D.K., Löw, M.J., Paulus, M.P., Geyer, M.A., 1999. anxiolysis in the rat is reduced by co-administration of the 5-HT1A
Dopamine D4 receptor knock-out mice exhibit reduced exploration of receptor antagonist WAY100635. Br. J. Pharmacol. 130, 1713–1719.
novel stimuli. J. Neurosci. 19, 9550–9556. Ebner, K., Wotjak, C.T., Landgraf, R., Engelmann, M., 2002. Forced
Duman, R.S., Malberg, J., Thome, J., 1999. Neural plasticity to stress swimming triggers vasopressin release within the amygdala to
and antidepressant treatment. Biol. Psychiatry 46, 1181–1191. modulate stress-coping strategies in rats. Eur. J. Neurosci. 15, 384–388.
Dumont, Y., Fournier, A., St-Pierre, S., Quirion, R., 1996a. Autoradi- Echeverry, M.B., Suimaraes, F.S., Oliveira, M.A., De Prado, W.A., Del
ographic distribution of [125 I]Leu31 , Pro[34]PYY and [125 I]PYY3−−36 Bel, E.A., 2002. Delayed stress-induced antinociceptive effect of nitric
binding sites in the rat brain evaluated with two newly developed Y1 oxide synthase inhibition in the dentate gyrus of rats. Pharmacol.
and Y2 receptor radioligands. Synapse 22, 139–158. Biochem. Behav. 74, 149–156.
Dumont, Y., St-Pierre, J.-A., Quirion, R., 1996b. Comparative Eckart, K., Jahn, O., Radulovic, J., Radulovic, M., Blank, T., Stiedl, O.,
autoradiographic distribution of neuropeptide Y Y1 receptors with the Brauns, O., Tezval, H., Zeyda, T., Spiess, J., 2002. Pharmacology and
Y1 receptor agonist [125 I][Leu31 , Pro14 ] PYY and the non-peptide biology of corticotropin-releasing factor (CRF) receptors. Receptors
antagonist [3 H]BIBP3226. NeuroReport 7, 901–904. Channels 8, 163–177.
Dumont, Y., Jacques, D., St.-Pierre, J.-A., Tong, Y., Parker, R., Herzog, Egberongbe, Y.I., Gentleman, S.M., Falkai, P., Bogerts, B., Polak,
H., Quirion, R., 2000. Neuropeptide Y, peptide YY and pancreatic J.M., Roberts, G.W., 1994. The distribution of nitric oxide synthase
polypeptide receptor proteins and mRNAs in mammalian brains. immunoreactivity in the human brain. Neuroscience 59, 561–578.
In: Quirion, R., Björkland, A., Hökfelt, T. (Eds.), Handbook of Eison, A.S., Eison, M.S., 1994. Serotoninergic mechanisms in anxiety.
Chemical Neuroanatomy: Peptide Receptors, vol. 16, Part I. Elsevier, Prog. Neuropsychopharmacol. Biol. Psychiatry 18, 47–62.
Amsterdam, pp. 375–475. Eison, A.S., Eison, M.S., Stanley, M., Riblet, L.A., 1986. Serotonergic
Duncan, G.E., Knapp, D.J., Carson, S.W., Breese, G.R., 1998. Differential mechanisms in the behavioral effects of buspirone and gepirone.
effects of chronic antidepressant treatment on swim stress- and Pharmacol. Biochem. Behav. 24, 701–707.
194 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Elenkov, I.J., Wilder, R.L., Chrousos, G.P., Vizi, E.S., 2000. The Farisse, J., Hery, F., Barden, N., Hery, M., Boulenguez, P., 2000. Central
sympathetic nerve an integrative interface between two supersystems: 5-HT1 and 5-HT2 binding sites in transgenic mice with reduced
the brain and the immune system. Pharmacol. Rev. 52, 595–638. glucocorticoid receptor number. Brain Res. 862, 145–153.
El-Ghundi, M., Fletcher, P.J., Drago, J., Sibley, D.R., O’Dowd, B.F., Fedele, E., Conti, A., Raiteri, M., 1997. The glutamate receptor/NO/cyclic
George, S.R., 1999. Spatial learning deficit in dopamine D(1) receptor GMP pathway in the hippocampus of freely moving rats: modulation by
knock-out mice. Eur. J. Pharmacol. 383, 95–106. cyclothiazide, interaction with GABA and behavioural consequences.
El-Ghundi, M., O’Dowd, B.F., George, S.R., 2001. Prolonged fear Neuropharmacology 36, 1393–1403.
responses in mice lacking dopamine D1 receptors. Brain Res. 892, Feenstra, M.G.P., Botterblom, M.H.A., Mastenbroek, S., 2000. Dopamine
86–93. and noradrenaline efflux in the prefrontal cortex in the light and dark
Emsley, R.A., Oosthuizen, P.P., Joubert, A.F., Roberts, M.C., Stein, D.J., period: effects of novelty and handling and comparison to the nucleus
1999. Depressive and anxiety symptoms in patients with schizophrenia accumbens. Neuroscience 100, 741–748.
and schizophreniform disorder. J. Clin. Psychiatry 60, 747–751. Feenstra, M.G.P., Vogel, M., Botterblom, M.H.A., Joosten, R.N.J.M.A.,
Engel, J.A., Hjorth, S., Svensson, K., Carlsson, A., Liljequist, S., de Bruin, J.P.C., 2001. Dopamine and noradrenaline efflux in the rat
1984. Anticonflict effect of the putative serotonin receptor agonist 8- prefrontal cortex after classical aversive conditioning to an auditory
hydroxy-2(di-n-propylamino)tetralin (8-OH-DPAT). Eur. J. Pharmacol. cue. Eur. J. Neurosci. 13, 1051–1054.
105, 365–368. Feighner, J.P., 1999. Overview of antidepressants currently used to treat
Engel, J.A., Egbe, P., Liljequist, S., Söderpalm, B., 1989. Effects of anxiety disorders. J. Clin. Psychiatry 60 (Suppl. 22), 18–22.
amperozide in two animal models of anxiety. Pharmacol. Toxicol. 64, Fendt, M., Schmid, S., 2002. Metabotropic glutamate receptors are
429–433. involved in amygdaloid plasticity. Eur. J. Neurosci. 15, 1535–1541.
Engel, S.R., Purdy, R.H., Grant, K.A., 2001. Characterization of Fendt, M., Koch, M., Schnitzler, H.-U., 1994. Amygdaloid noradrenaline
discriminative stimulus effects of the neuroactive steroid pregnanolone. is involved in the sensitization of the acoustic startle response in rats.
J. Pharmacol. Exp. Ther. 297, 489–495. Pharmacol. Biochem. Behav. 48, 307–314.
Epelbaum, J., Dournaud, P., Fodor, M., Viollet, C., 1994. The neurobiology Feng, J., Cai, X., Zhao, J., Yan, Z., 2001. Serotonin receptors modulate
of somatostatin. Crit. Rev. Neurobiol. 8, 25–44. GABAA receptor channels through activation of anchored protein
Epperson, C.N., Haga, K., Masn, G.F., Sellers, E., Gueorguieva, R., kinase C in prefrontal cortical neurons. J. Neurosci. 21, 6502–6511.
Zhang, W., Weiss, E., Rothman, D.L., Krystal, J.H., 2002. Cortical Fenton, W.S., 2001. Comorbid conditions in schizophrenia. Curr. Opin.
␥-aminobutyric acid levels across the menstrual cycle in healthy Psychiatry 14, 17–23.
women and those with premenstrual dysphoric disorder. Arch. Gen. Fernández-Guasti, A., López-Rubalcava, C., 1998. Modification of the
Psychiatry 59, 851–858. anxiolytic action of 5-HT1A compounds by GABA-benzodiazepine
Ericsson, E., Ahlenius, S., 1999. Suggestive evidence for inhibitory agents in rats. Pharmacol. Biochem. Behav. 60, 27–32.
effects of galanin on mesolimbic dopaminergic neurotransmission. Ferrara, G., Serra, M., Zammaretti, F., Pisu, M.G., Panzica, G.C., Biggio,
Brain Res. 822, 200–209. G., Eva, C., 2001. Increased expression of the neuropeptide Y receptor
Ertel, E.A., Campbell, K.P., Harpold, M.M., Hofmann, F., Mori, Y., Y 1 gene in the medial amygdala of transgenic mice induced by long-
Perez-Reyes, E., Schwartz, A., Snutch, T.P., Tanabe, T., Birnbaumer, term treatment with progesterone or allopregnanolone. J. Neurochem.
L., Tsien, R.W., Catterali, W.A., 2000. Nomenclature of voltage-gated 79, 417–425.
calcium channels. Neuron 25, 533–535. Ferraro, L., Beani, L., Trist, D., Reggiani, A., Bianchi, C., 1999. Effects of
Espejo, E.F., 1997. Selective dopamine depletion within the medial cholecystokinin peptides and GV 150013, a selective cholecystokininB
prefrontal cortex induces anxiogenic-like effects in rats placed on the receptor antagonist, on electrically evoked endogenous GABA release
elevated plus maze. Brain Res. 762, 281–284. from rat cortical slices. J. Neurochem. 73, 1973–1981.
Evenden, J.L., 1999. Varieties of impulsivity. Psychopharmacology 146, Ferraro, L., Tomasini, M.C., Cassano, T., Bebe, B.W., Siniscalchi, A.,
348–361. O’Connor, W.T., Magee, P., Tanganelli, S., Cuomo, V., Antonelli, T.,
Fahey, J.M., Pritchard, G.A., Grassi, J.M., Pratt, J.S., Shader, R.I., 2001. Cannabinoid receptor agonist WIN 55,212-2 inhibits rat cortical
Greenblatt, D.J., 1999. In situ hybridization histochemistry as a method dialysate ␥- aminobutyric acid levels. J. Neurosci. Res. 66, 298–302.
to assess GABAA receptor subunit mRNA expression following Ferré, S., Cortès, R., Artigas, F., 1994. Dopaminergic regulation of the
chronic alprazolam administration. J. Psychopharmacol. 13, 211–218. serotonergic raphe-striatal pathway: microdialysis studies in freely
Falkenstein, E., Tillmann, H.-C., Christ, M., Feuring, M., Wehling, moving rats. J. Neurosci. 14, 4839–4846.
M., 2000. Multiple actions of steroid hormones—a focus on rapid, Ferrier, I.N., Johnstone, E.C., Crow, T.J., 1985. Clinical effects of
nongenomic effects. Pharmacol. Rev. 52, 513–555. apomorphine in schizophrenia. Br. J. Psychiatry 144, 359–384.
Falls, W.A., Miserendino, M.J.D., Davis, M., 1992. Extinction of fear- Ferris, C.D., Hirsch, D.J., Brooks, B.P., Snowman, A.M., Snyder, S.H.,
potentiated startle: blockade by infusion of an NMDA antagonist into 1990. [3 H]Opipramol labels a novel binding site and sigma1 receptors
the amygdala. J. Neurosci. 12, 854–863. in rat brain membranes. Mol. Pharmacol. 39, 199–204.
Falzone, T.L., Geiman, D.M., Young, J.I., Grandy, D.K., Löw, M.J., Ferris, P., Seward, E., Dawson, G.R., 2001. Interactions between
Rubinstein, M., 2002. Absence of dopamine D4 receptors results in LY354,740, a group II metabotropic agonist and the GABAA -
enhanced reactivity to unconditioned, but not conditioned, fear. Eur. benzodiazepine receptor complex in the rat elevated plus-maze. J.
J. Neurosci. 15, 158–164. Pharmacol. 15, 76–82.
Fanselow, M.S., 1991. The midbrain periaqueductal gray as a coordinator Ferry, B., Magistretti, P.J., Pralong, E., 1997. Noradrenaline modulates
of action in response to fear and anxiety. In: Depaulis, A., Bandler, glutamate-mediated neurotransmission in rat basolateral amygdala in
R. (Eds.), The Midbrain Periaquductal Grey Matter: Functional, vitro. Eur. J. Neurosci. 9, 1356–1364.
Anatomical and Immunohistochemcial Organization. Plenum Press, Feuerstein, T.J., Huber, B., Vetter, J., Aranda, H., Van Velthoven, V.,
New York, pp. 151–173. Limberger, N., 2000. Characterization of the ␣2 -adrenoceptor subtype,
Fanselow, M.S., 1994. Neural organization of the defensive behavior which functions as ␣2 -autoreceptor in human neocortex. J. Pharmacol.
system responsible for fear. Psychon. Bull. Res. 1, 429–438. Exp. Ther. 294, 356–362.
Fanselow, M.S., 2000. Contextual fear, gestalt memories, and the Ffrench-Mullen, J.M.H., 1995. Cortisol inhibition of calcium currents in
hippocampus. Behav. Brain Res. 110, 73–81. guinea pig hippocampal CA1 neurons via G-protein-coupled activation
Faria, M.S., Muscara, M.N., Moreno, H., Texiera, S.A., Dias, H.B., of protein kinase C. J. Neurosci. 15, 903–911.
Oliviera, B.D., Graeff, F.G., Nucci, G.D., 1997. Acute inhibition of Field, M.J., Oles, R.J., Singh, L., 2001. Pregabalin may represent a novel
nitric oxide synthesis induces anxiolysis in the plus maze test. Eur. J. class of anxiolytic agents with a broad spectrum of activity. Br. J.
Pharmacol. 323, 37–43. Pharmacol. 132, 1–4.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 195

Figueiredo, H.F., Dolgas, C.M., Herman, J.P., 2002. Dissociation of Flint, A., Raben, A., Astrup, A., Holst, J.J., 1998. Glucagon-like peptide-
estrogen effects on anxiety-related behavior and neuroendocrine 1 promotes satiety and suppresses energy intake in humans. J. Clin.
responses to stress. Soc. Neurosci. Abstr. 571.14. Invest. 101, 515–520.
File, S.E., 2000. NKP608, an NK1 receptor antagonist, has an anxiolytic Flore, M., Alleva, E., Probert, L., Kollias, G., Angelucci, F., Aloe, L., 1998.
action in the social interaction test in rats. Psychopharmacology 152, Exploratory and displacement behavior in transgenic mice expressing
105–109. high levels of brain TNF-alpha. Physiol. Behav. 63, 571–576.
File, S.E., Clarke, A., 1980. Intraventricular ACTH reduces social Flores, P., Pellón, R., 2000. Antipunishment effects of diazepam on two
interaction in male rats. Pharmacol. Biochem. Behav. 12, 711–715. levels of suppression of schedule-induced drinking in rats. Pharmacol.
File, S.E., Seth, P., 2003. A review of 25 years of the social interaction Biochem. Behav. 67, 207–214.
test. Eur. J. Pharmacol. 463, 35–53. Florio, C., Prezioso, A., Papaioannou, A., Vertua, R., 1998. Adenosine
File, S.E., Hyde, J.R., Odling-Smee, F.J., 1979. Piracetam impairs the A1 receptors modulate anxiety in CDI mice. Psychopharmacology
overshadowing of background stimuli by an informative CS. Physiol. 136, 311–319.
Behav. 23, 827–830. Flügge, G., 1996. Alteration in the central nervous alpha2 -adrenoceptor
File, S.E., Baldwin, H.A., Johnston, A.L., Wilks, L.J., 1988. Behavioral system under chronic psychosocial stress. Neuroscience 75, 187–196.
effects of acute and chronic administration of caffeine in the rat. Flügge, G., Ahrens, O., Fuchs, E., 1997. Monoamine receptors in the
Pharmacol. Biochem. Behav. 30, 809–815. prefrontal cortex of Tupaia belangeri during chronic psychosocial
File, S.E., Andrews, N., Wu, P.Y., Zharkovsky, A., Zangrossi Jr., H., 1992. stress. Cell Tissue Res. 288, 1–10.
Modification of chlordiazepoxide’s behavioural and neurochemical Follesa, P., Cagetti, E., Mancuso, L., Biggio, F., Manca, A., Maciocco,
effects by handling and plus-maze experience. Eur. J. Pharmacol. 218, E., Massa, F., Desole, M.S., Carta, M., Busonero, F., 2001. Increase
9–14. in expression of the GABAA receptor ␣4 subunit gene induced by
File, S.E., Gonzalez, L.E., Andrews, N., 1998. Endogenous acetylcholine withdrawal of, but not by long-term treatment with, benzodiazepine
in the dorsal hippocampus reduces anxiety through actions on nicotinic full or partial agonists. Mol. Brain Res. 92, 138–148.
and muscarinic1 receptors. Behav. Neurosci. 112, 352–359. Follesa, P., Mancuso, L., Biggio, F., Cagetti, E., Franco., Trapani, G.,
File, S.E., Kenny, P.J., Cheeta, S., 2000. The role of dorsal hippocapal Biggio, G., 2002. Changes in GABAA receptor gene expression
serotonergic and cholinergic systems in the modulation of anxiety. induced by withdrawal of, but not by long-term exposure to, zaleplon
Pharmacol. Biochem. Behav. 66, 65–72. or zolpidem. Neuropharmacology 42, 191–198.
File, S.E., Cheeta, S., Irvine, E.F., Tucci, S., Akthar, M., 2002. Fone, K.C.F., Beckett, S.R.G., Topham, I.A., Swettenham, J., Ball,
Conditioned anxiety to nicotine. Psychopharmacology 164, 309–317. M., Maddocks, L., 2002. Long-term changes in social interaction
Filip, M., Baran, L., Siwanowicz, J., Chojnacka-Wójcik, E., Przegaliński, and reward following repeated NMDA administration to adolescent
E., 1992. The anxiolytic-like effects of 5-hydroxytryptamine3 (5-HT3 ) rats without accompanying serotonergic neurotoxicity. Psycho-
receptor antagonists. Pol. J. Pharmacol. Pharm. 44, 261–269. pharmacology 159, 437–444.
Filliol, D., Ghozland, S., Chluba, J., Martin, M., Matthes, H.W.D., Fontana, D.J., Commissaris, R.L., 1988. Effects of acute and chronic
Simonin, F., Befort, K., Gavériaux-Ruff, C., Dierich, A., LeMeur, M., imipramine administration on conflict behavior in the rat: a potential
Valverde, O., Maldonado, R., Kieffer, B.L., 2000. The ␦- and ␮-opioid “animal model” for the study of panic disorder? Psychopharmacology
receptor-deficient mice exhibit opposing alterations of emotional 95, 147–150.
responses. Nat. Genet. 25, 195–200. Fontana, D.J., Carbary, T.J., Commissaris, R.L., 1989a. Effects of acute
Finnerty, R.J., Goldberg, H.L., Nathan, L., Lowrey, C., Cole, J.O., 1976. and chronic anti-panic drug administration on conflict behavior in the
Haloperidol in the treatment of psychoneurotic anxious outpatients. rat. Psychopharmacology 98, 157–162.
Dis. Nerv. Syst. 37, 621–624. Fontana, D.J., McCloskey, T., Jolly, S.K., Commissaris, R.L., 1989b. The
Fink, K., Schmitz, V., Böng, C., Göthert, M., 1995. Stimulation of effects of beta-antagonists and anxiolytics on conflict behavior in the
serotonin release in the rat brain cortex by activation of ionotropic rat. Pharmacol. Biochem. Behav. 32, 807–813.
glutamate receptors and its modulation via ␣2 -heteroceptors. Naunyn Fontana, D.J., Schefke, D.M., Commissaris, R.L., 1990. Acute versus
Schmiedebergs Arch. Pharmacol. 352, 394–401. chronic clonidine treatment effects on conflict behavior in the rat.
Fink, K., Dooley, D.J., Meder, W.P., Suman-Chauhan, N., Duffy, S., Behav. Pharmacol. 1, 201–208.
Clusman, H., Göthert, M., 2002. Inhibition of neuronal Ca2+ influx Forray, M.I., Bustos, G., Gysling, K., 1999. Noradrenaline inhits
by GABApentin and pregabalin in the human neocortex. Neuro- glutamate release in the rat bed nucleus of the stria terminalis: in
pharmacology 42, 229–236. vivo microdialsis studies. J. Neurosci. Res. 55, 311–320.
Finlay, J.M., Zigmond, M.J., 1997. The effects of stress on central Fotuhi, M., Standaert, D.G., Testa, C.M., Penney, J.B., Young, A.D.,
dopaminergic neurons: possible clinical implications. Neurochem. 1994. Differential expression of metabotropic glutamate receptors in
Res. 22, 1387–1394. the hippocampus and entorhinal cortex of the rat. Mol. Brain Res.
Finta, E.P., Regenold, J.T., Illes, P., 1992. Depression by neuropeptide Y 21, 283–292.
of noradrenergic inhibitory postsynaptic potentials of locus coeruleus Foukas, L.C., Daniele, N., Ktori, C., Anderson, K.E., Jensen, J., 2002.
neurones. Naunyn Schmiedebergs Arch. Pharmacol. 346, 472–474. Direct effects of caffeine and theophylline on p110␦ and other
Fish, E.W., De Bold, J.F., Miczek, K.A., 2002. Aggressive behavior phosphoinositide 3-kinases. J. Biol. Chem. 277, 37124–37130.
as a reinforcer in mice: activation by allopregnanolone. Psychophar- Francken, B.J.B., Jurzak, M., Vanhauwe, J.F.M., Luyten, W.H.M.L.,
macology 163, 459–466. Leysen, J.E., 1998. The human 5-HT5A receptor couples to Gi /Go
Fitzgerald, L.W., Ennis, M.D., 2002. 5-HT2C receptor modulators: proteins and inhibits adenylate cyclase in HEK 293 cells. Eur. J.
progress in development of new CNS medicines. Annu. Rep. Med. Pharmacol. 361, 299–309.
Chem. 37, 21–30. Frankland, P.W., Josselyn, S.A., Bradwejn, J., Vaccarino, F.J., Yeomans,
Fletcher, A., Forster, E.A., Bill, D.J., Brown, G., Cliffe, I.A., Hartley, J.E., J.S., 1997. Activation of amygdala cholecystokininB receptors
Jones, D.E., McLenachan, A., Stanhope, K.J., Critchley, D.J., Childs, potentiates the acoustic startle response in the rat. J. Neurosci. 17,
K.J., Middlefell, V.C., Lanfumey, L., Corradetti, R., Laporte, A.M., 1838–1847.
Gozlan, H., Hamon, M., Dourish, C.T., 1996. Electrophysiological, Franowicz, J.S., Arnsten, A.F.T., 2002. Actions of ␣1 -noradrenergic
biochemical, neurohormonal and behavioural studies with WAY100635, agonists on spatial working memory and blood pressure in rhesus
a potent, selective and silent 5-HT1A receptor antagonist. Behav. Brain monkeys appear to be mediated by the same receptor subtype.
Res. 73, 337–353. Psychopharmacology 162, 304–312.
196 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Frantz, K.J., Hansson, K.J., Stouffer, D.G., Parsons, L.H., 2002. 5-HT6 Gadea, A., Lopez-Colomé, A.M., 2001. Glial transporters for glutamate,
receptor antagonism potentiates the behavioural and neurochemical glycine. J. Neurosci. Res. 63, 461–468.
effects of amphetamine but not cocaine. Neuropharmacology 42, Galvez, T., Duthey, B., Kniazeff, J., Blahos, J., Rovelli, G., Bettler, B.,
170–180. Prézeau, L., Pin, J.-P., 2001. Allosteric interactions between GB1 and
Fraser, C.M., Cooke, M.J., Fisher, A., Thompson, I.D., Stone, T.W., 1996. GB2 subunits are required for optimal GABAB receptor function.
Interactions between ifenprodil and dizocilpine on mouse behaviour in EMBO J. 20, 2152–2159.
models of anxiety and working memory. Eur. Neuropsychopharmacol. Gantner, F., Sakai, K., Tusche, M.W., Tusche, M.W., Cruikshank, W.W.,
6, 311–316. Center, D.M., Bacon, K.B., 2002. Histamine H4 and H2 receptors
Frazier, C.J., Rollins, Y.D., Breese, C.R., Leonard, S., Freedman, R., control histamine-induced interleukin-16 release from human CD8+
Dunwiddie, T.V., 1998. Acetylcholine activates an alpha-bungarotoxin- T cells. J. Pharmacol. Exp. Ther. 303, 300–307.
sensitive nicotinic current in rat hippocampal interneurons, but not Gao, X.-B., Van Der Pol, A.N., 2002. Melanin-concentrating hormone
pyramidal cells. J. Neurosci. 18, 1187–1195. depresses L-, N-, and P/Q-type voltage-dependent calcium channels
Fredholm, B.B., Bättig, K., Holmen, J., Nehlig, A., Zvartau, E.E., 1999. in rat lateral hypothalamic neurons. J. Physiol. 542, 273–286.
Actions of caffeine in the brain with special reference to factors that Gao, B., Fritschy, J.-M., Benke, D., Möhler, H., 1993. Neuron-specific
contribute to its widespread use. Pharmacol. Rev. 51, 83–133. expression of GABAA -receptor subtypes: differential association of
Fredholm, B.B., Ijzerman, A.P., Jacobson, K.A., Klotz, K.-N., Linden, J., the ␣1 - and ␣3 -subunits with serotonergic and GABAergic neurons.
2001. International Union of Pharmacology. XXV. Nomenclature and Neuroscience 54, 881–892.
classification of adenosine receptors. Pharmacol. Rev. 53, 527–552. Garcia, F., Buron, E., Martin, M., Navarro, J.F., 2002. Anxiogenic-like
Fremeau, R.T., Burman, J., Qureshi, T., Tran, C.H., Proctor, J., Johnson, J., activity of L-655,708, a selective ligand for the benzodiazepine site of
Zhang, H., Sulzer, D., Copenhagen, D.R., Storm-Mathisen, J., Reimer, GABAA receptors which contain the ␣5 subunit, in the elevated-plus
R.J., Chaudhry, F.A., 2002. The identification of vesicular glutamate maze test. Eur. Neuropsychopharmacol. 12, S333–S334.
transporter 3 suggests novel modes of signaling by glutamate. Proc. Gard, P.R., 2002. The role of angiotensin II in cognition and behaviour.
Natl. Acad. Sci. U.S.A. 99, 14488–14493. Eur. J Pharmacol. 438, 1–14.
French, E.D., 1997. 9 -Tetrahydrocannabinol excites rat VTA dopamine Gardner, C.R., 1985. Pharmacological studies of the role of serotonin in
neurons through activation of cannabinoid CB1 but not opioid animal models of anxiety. In: Green, A.R. (Ed.), Neuropharmacology
receptors. Neurosci. Lett. 226, 159–162. of Serotonin. Oxford University Press, Oxford, pp. 281–325.
Frerking, M., Nicoll, R.A., 2000. Synaptic kainate receptors. Curr. Opin. Gardner, C.R., Piper, D.C., 1982. Effects of adrenoceptor modulation on
Neurobiol. 10, 342–351. drinking conflict in rats. Br. J. Pharmacol. 75, 50P.
Friedman, J.I., Temporini, H., Davis, K.L., 1999. Pharmacologic Garfinkel, D., Zisapel, N., Wainstein, J., Laudon, M., 1999. Facilitation
strategies for augmenting cognitive performance in schizophrenia. of benzodiazepine discontinuation by melatonin. Arch. Intern. Med.
Biol. Psychiatry 45, 1–16. 159, 2456–2460.
Frisch, C., Hasenöhl, R.U., Krauth, J., Huston, J.P., 1998. Anxioytic-like Gargiulo, P.A., Viana, M.B., Graeff, F.G., Silva, M.A., Tomaz, C.,
behavior after lesion of the tuberomammillary nucleus E2-region. 1996. Effects on anxiety and memory of systemic and intra-
Exp. Brain Res. 119, 260–264. amygdala injection of the 5-HT3 receptor antagonist, BRL 46470A.
Frisch, C., Dere, E., Silva, M.A., Godecke, A., Shrader, J., Huston, J.P., Neuropsychopharmacology 33, 189–195.
2000. Superior water maze performance and increase in fear-related Gasior, M., Carter, R.B., Witkin, J.M., 1999. Neuroactive steroids:
behavior in the endothelial nitric oxide synthase-deficient mouse potential therapeutic use in neurological and psychiatric disorders.
together with monoamine changes in cerebellum and ventral striatum. Trends Pharmacol. Sci. 20, 107–112.
J. Neurosci. 20, 6694–6700. Gass, P., Reichardt, H.M., Strekalova, T., Henn, F., Tronche, F., 2001.
Fritschy, J.M., Möhler, H., 1995. GABAA -receptor heterogeneity in the Mice with targeted mutations of glucocorticoid and mineralocorticoid
adult rat brain: differential regional and cellular distribution of seven receptors: models for depression and anxiety. Physiol. Behav. 73,
major subunits. J. Comp. Neurol. 359, 154–194. 811–825.
Fritschy, J.M., Weinmann, O., Wenzel, A., Benke, D., 1998. Synapse- Gavioli, E.C., Canteras, N.S., De Lima, T.C.M., 1999. Anxiogenic-like
specific localization of NMDA and GABAA receptor subunits revealed effect induced by substance P injected into the lateral septal nucleus.
by antigen-retrieval immunohistochemistry. J. Comp. Neurol. 390, NeuroReport 10, 3399–3403.
194–210. Gavioli, E.C., Canteras, N.S., De Lima, T.C.M., 2002a. The role of
Froger, N., Gardier, A.M., Moratalia, R., Alberti, I., Lena, I., Boni, lateral septal NK1 receptors in mediating anxiogenic effects induced
C., De Felipe, C., Rupniak, N.M.J., Hunt, S.P., Jacquot, C., Hamon, by intracerebroventricular injection of substance P. Behav. Brain Res.
M., Lanfumey, L., 2001. 5-Hydroxytryptamine (5-HT)1A autoreceptor 134, 411–415.
adaptive changes in substance P (neurokinin1 ) receptor knock-out Gavioli, E.C., Rae, G.A., Calo, G., Guerrini, R., De Lima, T.C.M.,
mice mimic antidepressant-induced desensitisation. J. Neurosci. 21, 2002b. Central injections of nocistatin or its C-terminal hexapeptide
8188–8197. exert anxiogenic-like effect on behaviour of mice in the plus-maze
Frussa-Filho, R., Barbosa, H., Silva, R.H., Da Cunba, C., Mello, C.F., test. Br. J. Pharmacol. 136, 764–772.
1999. Naltrexone potentiates the anxiolytic effects of chlordiazepoxide Gazzara, R.A., Compan, V., Mullen, T., Leahy, C., Breuer, M., Manzano,
in rats exposed to novel environments. Psychopharmacology 147, M., Douglas, A., Brunner, D., Cockayne, D.A., Martin, R.S., Hen, R.,
168–173. 2002. Behavioral characterization of 5-HT4 receptor knock-out mice.
Fuchs, E., Flügge, G., 2002. Social stress in tree shrews: effects on Soc. Neurosci. Abstr. 398.14.
physiology, brain function, and behavior of subordinate individuals. Ge, J., Barnes, N.M., 1996. 5-HT4 receptor mediated modulation of
Pharmacol. Biochem. Behav. 73, 247–258. 5-HT release in the rat hippocampus in vivo. Br. J. Pharmacol. 117,
Fulford, A.J., Marsden, C.A., 1997. Social isolation in the rat enhances 1474–1480.
alpha2 autoreceptor function in the hippocampus in vivo. Neuroscience Gee, N.S., Brown, J.P., Dissanayake, U.K., Offord, J., Thurlow, R.,
77, 57–64. Woodruff, G.N., 1996. The novel anticonvulsant drug, GABApentin
Gacsalyi, I., Schmidt, E., Gyertyan, I., Vasar, E., Lang, A., Haapalinna, (neurontin), binds to the ␣2␦ subunit of a calcium channel. J. Biol.
A., Fekete, M., Hietala, J., Syvälahti, E., Tuomainen, P., Männistö, Chem. 271, 5768–5776.
P.T., 1997. Receptor binding profile and anxiolytic-type activity of Gehlert, D.R., Morin, S., Hipskind, P., Zink, C., Gackenheimer, S.,
deramciclane (EGIS-3886) in animals models. Drug Dev. Res. 40, Gackenheimer, S., Shax, J., Song, N., Large, T., McKinzie, D.,
333–348. Kandasamy, R., Dougherty, R., Sajdyk, T., Shekhar, A., 2002.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 197

Characterization of corticotropin releasing factor (CRF) and urocortin receptors but not by sympathomimetic effects. Neuropharmacology
(UCN) responses in behavioural tests of anxiety: the role of CRF1 . 33, 457–465.
Soc. Neurosci. Abstr. 398.6. Gieschke, R., Cluydts, R., Dingemanse, J., De Roeck, J., De Cock, W.,
Geiger, J.R., Melcher, T., Koh, D.S., Sakmann, B., Seeburg, P.H., 1994. Effects of bretazenil vs. zolpidem and placebo on experimentally
Jonas, P., Monyer, H., 1995. Relative abundance of subunit mRNAs induced sleep disturbance in healthy volunteers. Methods Find.
determines gating and Ca2+ permeability of AMPA receptors in Exp. Clin. Pharmacol. 16, 667–675.
principal neurons and interneurons in rat CNS. Neuron 15, 193–204. Gifkins, A., Greba, Q., Kokkinidis, L., 2002. Ventral tegmental area
Geller, I., Seifter, J., 1960. The effects of meprobamate, barbiturate, dopamine neurons mediate the shock sensitization of acoustic startle:
d-amphetamine and promazine on experimentally-induced conflict in a potential site of action for benzodiazepine anxiolytics. Behav.
the rat. Psychopharmacologia 1, 482–492. Neurosci. 116, 785–794.
Gendreau, P.L., Petitto, J.M., Schnauss, R., Frantz, K.J., Van Hartesveldt, Gilad, G.M., 1987. The stress-induced reponse of the septohippocampal
C., Gariépy, J.-L., Lewis, M.H., 1997. Effects of the putative dopamine cholinergic system. A vectorial outcome of psychoendocrinological
D3 receptor antagonist PNU 99194A on motor behavior and emotional interactions. Psychoneuroendocrinology 12, 167–184.
reactivity in C57Bl/6J mice. Eur. J. Pharmacol. 337, 147–155. Gilligan, P.J., Robertson, D.W., Zaczek, R., 2000. Corticotropin releasing
Gendreau, P.L., Petitto, J.M., Gariépy, J.-L., Lewis, M.H., 1998. D2 - factor (CRF) receptor modulators: progress and opportunities for new
like dopamine receptor mediation of social–emotional reactivity therapeutic agents. J. Med. Chem. 43, 1641–1660.
in a mouse model of anxiety: strain and experience effects. Giménez-Llort, L., Fernandez-Teruel, A., Escorihuela, R.M., Fredholm,
Neuropsychopharmacology 18, 210–221. B.B., Tobena, A., Pekny, M., Johansson, B., 2002. Mice lacking the
Gendreau, P.L., Petitto, J.M., Petrova, A., Gariépy, J.-L., Lewis, M.H., adenosine A1 receptor are anxious and aggressive, but are normal
2000. D3 and D2 dopamine receptor agonists differentially modulate learners with reduced muscle strength and survival rate. Eur. J.
isolation-induced social–emotional reactivity in mice. Behav. Brain Neurosci. 16, 547–550.
Res. 114, 107–117. Gimpl, G., Fahrenholz, F., 2001. The oxytocin receptor system: structure,
Gentil, V., Gorenstein, C., Camargo, C.H.P., Singer, J.M., 1989. Effects function and regulations. Physiol. Rev. 81, 629–683.
of flunitrazepam on memory and their reversal by two antagonists. J. Giovannini, M.G., Rakovska, A., Benton, R.S., Pazzagli, M., Bianchi, L.,
Clin. Psychopharmacol. 9, 191–197. Pepeu, G., 2001. Effects of novelty and habituation on acetylcholine,
George, T.P., Verrico, C.D., Picciotto, M.R., Roth, R.H., 2000. Nicotinic GABA and glutamate release from the frontal cortex and hippocampus
modulation of mesoprefrontal dopamine neurons: pharmacologic and of freely moving rats. Neuroscience 106, 43–53.
neuroanatomic characterization. J. Pharmacol. Exp. Ther. 295, 58–66. Girdler, S.S., Straneva, P.A., Light, K.C., Pedersen, C.A., Morrow, A.L.,
George, T.P., Picciotto, M.R., Verrico, C.D., Roth, R.H., 2001. Effects 2001. Allopregnanolone levels and reactivity to mental stress in
of nicotine pretreatment on dopaminergic and behavioral responses premenstrual dysphoric disorder. Biol. Psychiatry 49, 788–797.
to conditioned fear stress in rats: dissociation of biochemical and Giuffrida, A., Beltramo, M., Piomelli, D., 2001. Mechanisms of
behavioral effects. Biol. Psychiatry 49, 300–306. endocannabinoid inactivation: biochemistry and pharmacology. J.
George, S.R., O’Dowd, B.F., Lee, S.P., 2002. G-protein-coupled receptor Pharmacol. Exp. Ther. 298, 7–14.
oligomerization and its potential for drug discovery. Nat. Rev. Drug Giusti, P., Ducic, I., Puia, G., Arban, R., Walser, A., Guidotti, A., Costa,
Discov. 1, 808–820. E., 1993. Imidazenil: a new partial positive allosteric modulator of ␥-
Georgiev, V., Stancheva, S., Kambourova, T., Getova, D., 1990. Effect aminobutyric acid (GABA) action at GABAA receptors. J. Pharmacol.
of angiotensin II on the Vogel conflict paradigm and on the content Exp. Ther. 266, 1018–1028.
of dopamine and noradrenaline in rat brain. Acta Physiol. Pharmacol. Glass, M.J., Huang, J., Aicher, S.A., Milner, T.A., Pickel, V.M., 2001.
Bulg. 16, 32–37. Subcellular localization of alpha2A -adrenergic receptors in the rat
Geraci, M., Anderson, T.S., Slate-Cothren, S., Post, R.M., McCann, U.D., medial nucleus tractus solitarius: regional targeting and relationship
2002. Pentagastrin-induced sleep panic attacks: panic in the absence with catecholaminergic neurons. J. Comp. Neurol. 433, 193–207.
of elevated baseline arousal. Biol. Psychiatry 52, 1183–1189. Glass, M.J., Colago, E.E.O., Pickel, V.M., 2002. Alpha2A -adrenergic
Gérard, C., El Mestikawy, S., Lebrand, C., Adrien, J., Ruat, M., Traiffort, receptors are present on neurons in the central nucleus of the amygdala
E., Hamon, M., Martres, M.P., 1996. Quantitative RT-PCR distribution that project to the dorsal vagal complex in the rat. Synapse 46, 258–268.
of serotonin 5-HT6 receptor mRNA in the central nervous system of Glatt, C.E., Freimer, N.B., 2002. Association analysis of candidate genes
control or 5,7-dihydroxytryptamine-treated rats. Synapse 23, 164–173. for neuropsychiatric disease: the perpetual campaign. Trends Genet.
Gerevich, Z., Tretter, L., Adam-Vizi, V., Branyi, M., Kiss, J.-P., Zelles, 18, 307–312.
T., Vizi, E.S., 2001. Analysis of high intracellular [Na+ ]-induced Gleeson, S., Ahlers, S.T., Mansbach, R.S., Foust, J.M., Barrett, J.E., 1989.
release of [3 H]noradrenaline in rat hippocampal slices. Neuroscience Behavioral studies with anxiolytic drugs. VI. Effects on punished
104, 761–768. responding of drugs interacting with serotonin receptor subtypes. J.
Gerhardt, C.C., Heerikhuizen, H., 1997. Functional characteristics of Pharmacol. Exp. Ther. 250, 809–817.
heterologously expressed 5-HT receptors. Eur. J. Pharmacol. 334, 1–23. Glover, V., Medvedev, A., Sandler, M., 1995. Isatin is a potent endogenous
Getova, D., Georgiev, V., Ivanova, V., 1990. Investigation of 2-hydro- antagonist of guanylate cyclase-coupled atrial natriuretic peptide
xylamino-5-ethyl-5-sec-pentylbarbituric acid for anxiolytic action. receptors. Life Sci. 57, 2073–2079.
Acta Physiol. Pharmacol. Bulg. 16, 42–47. Gobert, A., Millan, M.J., 1999a. Modulation of dialysate levels of
Gewirtz, J.C., McNish, K.A., Davis, M., 2000. Is the hippocampus dopamine, noradrenaline, and serotonin (5-HT) in the frontal cortex
necessary for contextual fear conditioning? Behav. Brain Res. 110, of freely-moving rats by (−)-pindolol alone and in association with
83–95. 5-HT reuptake inhibitors: comparative roles of ␤-adrenergic, 5-HT1A ,
Ghersi, C., Bonfanti, A., Manzari, B., Feligioni, M., Raiteri, M., and 5-HT1B receptors. Neuropsychopharmacology 21, 268–284.
Pittaluga, A., 2003. Pharmacological heterogeneity of release- Gobert, A., Millan, M.J., 1999b. Serotonin (5-HT)2A receptor activation
regulating presynaptic AMPA/kainate receptors in the rat brain: study enhances dialysate levels of dopamine and noradrenaline, but not
with receptor antagonists. Neurochem. Int. 42, 283–292. 5-HT, in the frontal cortex of freely-moving rats. Neuropharmacology
Giardino, L., Zanni, M., Pozza, M., Benelli, C., Covelli, V., 1998. 38, 315–317.
Dopamine receptors in the striatum of rats exposed to repeated restraint Gobert, A., Rivet, J.-M., Lejeune, F., Newman-Tancredi, A., Adhumeau-
stress and alprazolam treatment. Eur. J. Pharmacol. 344, 143–147. Auclair, A., Nicolas, J.-P., Cistarelli, L., Melon, C., Millan, M.J., 2000.
Gibson, E.L., Barnfizeld, A.M.C., Curzon, G., 1994. Evidence that mCPP- Serotonin2C receptors tonically suppress the activity of mesocortical
induced anxiety in the plus-maze is mediated by postsynaptic 5-HT2C dopaminergic and adrenergic, but not serotonergic, pathways: a
198 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

combined dialysis and electrophysiological analysis in the rat. Synapse Gower, A.J., Tricklebank, M.D., 1988. Alpha-2-adrenoceptor antagonist
36, 205–221. activity may account for the effects of buspirone in an anticonflict
Gobert, A., Di Cara, B., Cistarelli, L., Millan, M.J., 2003. Piribedil test in the rat. Eur. J. Pharmacol. 155, 129–137.
enhances frontocortical and hippocampal release of acetylcholine Grachev, I.D., Fredrickson, B.E., Abkarian, A.V., 2002. Brain chemistry
in freely-moving rats by blockade of ␣2 -adrenoceptors: a dialysis reflexts dual states of pain and anxiety in chronic low back pain. J.
comparison to other antiparkinson agents in the absence of Neuronal Transm. 109, 1309–1334.
acetylcholinesterase inhibitors. J. Pharmacol. Exp. Ther. 305, 1–9. Graeff, F.G., 1990. Brain defense systems and anxiety. In: Roth, M.,
Goddard, A.W., Woods, S.W., Money, R., Pane, A.C., Charney, D.S., Burrow, G.D., Noyes, R. (Eds.), Handbook of Anxiety, vol. 3. Elsevier,
Goodman, W.K., Heninger, G.R., Price, L.H., 1999. Effects of the New York, pp. 307–354.
CCK(B) antagonist CI-988 on responses to mCPP in generalized Graeff, F.G., 2002. On serotonin and experimental anxiety. Psycho-
anxiety disorder. Psychiatry Res. 85, 225–240. pharmacology 163, 467–476.
Goddard, A.W., Mason, G.F., Almai, A., Rothman, D.L., Behar, K.L., Graeff, F.G., Zuardi, A.W., Giglio, J.S., Lima Filhoe, E.C., Karniol, I.G.,
Petroff, A.C., Charney, D.S., Krystal, J.H., 2001. Reductions in 1985. Effect of metergoline on human anxiety. Psychopharmacology
occipital cortex GABA levels in panic disorder detected with 3 H- 86, 334–338.
magnetic resonance spectroscopy. Arch. Gen. Psychiatry 58, 556–561.
Graeff, F.G., Silveira, M.C.L., Nogueira, R.L., Audi, E.A., Oliveira,
Gogos, J.A., Morgan, M., Luine, V., Santha, M., Ogawa, S., Plaff,
R.M.W., 1993. Role of the amygdala and periaqueductal gray in
D., Karayiorgou, M., 1998. Catechol-O-methyltransferase-deficient
anxiety and panic. Behav. Brain Res. 58, 123–131.
mice exhibit sexually dimorphic changes in catecholamine levels and
Graeff, F.G., Guimarães, F.S., de Andrade, T.G.C.S., Deakin, J.F.W.,
behavior. Proc. Natl. Acad. Sci. U.S.A. 95, 9991–9996.
1996a. Anxiety and role of 5-HT in stress and anxiety and depression.
Gold, P.W., Chrousos, G.P., 2002. Organization of the stress system and
Pharmacol. Biochem. Behav. 54, 129–141.
its dysregulation in melancholic and atypical depression: high vs. low
CRH/NE states. Mol. Psychiatry 7, 254–275. Graeff, F.G., Viana, M.B., Mora, P.O., 1996b. Opposed regulation by
Goldstein, D.J., Detke, M., Lu, Y., Demitrack, M.A., 2002. The dorsal raphe nucleus 5-HT pathways of two types of fear in the
antidepressant duloxetine reduces anxiety symptom severity. Eur. elevated T-maze. Pharmacol. Biochem. Behav. 53, 171–177.
Neuropsychopharmacol. 12, S213. Graeff, F.G., Ferreira Netto, C., Zangrossi Jr., H., 1998. The elevated
Golombek, D.A., Martini, M., Cardinali, D.P., 1993. Melatonin as an T-maze as an experimental model of anxiety. Neurosci. Biobehav.
anxiolytic in rats: time dependence and interaction with the central Rev. 23, 237–246.
GABAergic system. Eur. J. Pharmacol. 237, 231–236. Graeff, F.G., Silva, M., Del Ben, C.M., Zuardi, A.W., Hetem,
Golombek, D.A., Pévet, P., Cardinali, D.P., 1996. Melatonin effects L.A., Guimaraes, F.S., 2001. Comparison between two models of
on behavior: possible mediation by the central GABAergic system. experimental anxiety in healthy volunteers and panic disorder patients.
Neurosci. Biobehav. Rev. 20, 404–412. Neurosci. Biobehav. Rev. 25, 753–759.
Gomeza, J., Shannon, H., Kostenis, E., Felder, C., Zhang, L., Brodkin, J., Grailhe, R., Waeber, C., Dulawa, S.C., Hornung, J.P., Zhuang, X.,
Grinberg, A., Sheng, H., Wess, J., 1999. Pronounced pharmacologic Brunner, D., Geyer, M.A., Hen, R., 1999. Increased exploratory
deficits in M2 muscarinic acetylcholine receptor knock-out mice. activity and altered response to LSD in mice lacking the 5-HT5A
Proc. Natl. Acad. Sci. U.S.A. 96, 1692–1697. receptor. Neuron 22, 581–591.
Gomez, C., Saldivar-Gonzalez, A., Delgado, G., Rodriguez, R., 2002. Grailhe, R., Grabtree, G.W., Hen, R., 2001. Human 5-HT5 receptors:
Rapid anxiolytic activity of progesterone and pregnanolone in male the 5-HT5A receptor is functional but the 5-HT5B receptor was lost
rats. Pharmacol. Biochem. Behav. 72, 543–550. during mammalian evolution. Eur. J. Pharmacol. 418, 157–167.
Gong, H.C., Hang, J., Kohler, W., Su, T.-Z., 2001. Tissue-specific Grant, K.A., Shively, C.A., Nader, M.A., Ehrenkaufer, R.L., Line, S.W.,
expression and GABApentin-binding properties of calcium channel Morton, T.E., Gage, H.D., Mach, R.H., 1998. Effect of social status on
␣2␦ subunit subtypes. J. Membr. Biol. 184, 35–43. striatal DA D2 receptor binding characteristics in cynomolgus monkeys
Gonzalez, L.E., Andrews, N., File, S.E., 1996a. 5-HT1A and benzo- assessed with positron emission tomography. Synapse 29, 80–83.
diazepine receptors in the basolateral amygdala modulate anxiety in Gras, C., Herzog, E., Bellenchi, G.C., Bernard, V., Ravassard, P., Pohl,
the social interaction test, but not in the elevated plus-maze test. Brain M., Gasnier, B., Giros, B., El Mestikawy, S., 2002. A third vesicular
Res. 732, 145–153. glutamate transporter expressed by cholinergic and serotoninergic
Gonzalez, M.I., Vaziri, S., Wilson, C.A., 1996b. Behavioral effects of neurons. J. Neurosci. 22, 5442–5451.
␣-MSH and MCH after central administration in the female rat. Gratacos, M., Nadal, M., Martin-Santos, R., Pujana, M.A., Gago, J.,
Peptides 17, 171–177. Peral, B., Armengol, L., Ponsa, I., Miro Bulbena, A., Estivill, X., 2001.
Gonzalez-Alvear, G.M., Werling, L.L., 1995. Sigma receptor regulation
A polymorphic genomic duplication on human chromosome 15 is a
of norepinephrine release from rat hippocampal slices. Brain Res.
susceptibility factor for panic and phobic disorders. Cell 106, 367–379.
673, 61–69.
Gray, J.A., 1971. Sex differences in emotional behaviour in mammals
Gooney, M., Lynch, M.A., 2001. Long-term potentiation in the dentate
including man: endocrine bases. Acta Psychol. 35, 29–46.
gyrus of the rat hippocampus is accompanied by brain-derived
Gray, J.A., 1987. The Neuropsychology of Anxiety. Oxford University
neurotrophic factor-induced activation of TrkB. J. Neurochem. 77,
Press, Oxford.
1198–1207.
Gorman, J.M., 2002. Treatment of generalized anxiety disorder. J. Clin. Gray, T.S., Magnuson, D.J., 1992. Peptide immunoreactive neurons in
Psychiatry 63 (Suppl. 8), 17–23. the amygdala and the bed nucleus of the stria terminalis project to
Gorman, J.M., Liebowitz, M.R., Fyer, A.J., Stein, J., 1989. A neuro- the midbrain central gray in the rat. Peptides 13, 451–460.
anatomical hypothesis for panic disorder. Am. J. Psychiatry 146, Gray, J.A., McNaughton, N., 2000. The Neuropsychology of Anxiety:
148–161. An Enquiry into the Functions of the Septo-Hippocampal System,
Gorman, J.M., Kent, J.M., Sullivan, G.M., Coplan, J.D., 2000. second ed. Oxford University Press, Oxford.
Neuroanatomical hypothesis of panic disorder, revised. Am. J. Greba, Q., Kokkinidis, L., 2000. Peripheral and intraamygdalar
Psychiatry 157, 493–505. administration of the dopamine D1 receptor antagonist SCH 23390
Gould, E., Tanapat, P., 1999. Stress and hippocampal neurogenesis. Biol. blocks fear-potentiated startle but not shock reactivity or the shock
Psychiatry 46, 1472–1479. sensitization of acoustic startle. Behav. Neurosci. 114, 262–272.
Gould, R.J., Murphy, K.M., Snyder, S.H., 1985. Autoradiographic Greba, Q., Gifkins, A., Kokkinidis, L., 2001. Inhibition of amygdaloid
localization of calcium channel antagonists receptors in rat brain with dopamine D2 receptors impairs emotional learning measured with
[3 H]nitrendipine. Brain Res. 330, 217–223. fear-potentiated startle. Brain Res. 899, 218–226.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 199

Green, A.R., McGregor, I.S., 2002. On the anxiogenic and anxiolytic receptor antagonist, SSR149415, suggest an innovative approach for
nature of long-term cerebral 5-HT depletion following MDMA. the treatment of stress-related disorders. Proc. Natl. Acad. Sci. U.S.A.
Psychopharmacology 162, 448–450. 99, 6370–6375.
Greenshaw, A.J., Siverstone, P.H., 1997. The non-antiemetic uses of Griebel, G., Simiand, J., Steinberg, R., Jung, M., Gully, D., Roger, P.,
serotonin 5-HT3 receptor antagonists. Drugs 53, 20–39. Geslin, M., Scatton, B., Maffrand, J.-P., Soubrié, P., 2002b. 4-(2-
Greist, J.H., Mundt, J.C., Kobak, K., 2002. Factors contributing to failed Chloro-4-metoxy-5-methylphenyl)-N-[(1S)-2-cyclopropyl-1-(3-fluoro-
trials of new agents: can technology prevent some problems? J. Clin. 4-methylphenyl)ethyl]5-methyl-N-(2-propynyl) - 1,3 - thiazol - 2 - amine
Psychiatry 63, 8–13. hydrochloride (SSR125543A), a potent and selective corticotrophin-
Grevert, P., Goldstein, A., 1978. Endorphins: naloxone fails to alter releasing factor1 receptor antagonist. II. Characterization in rodent
experimental pain or mood in humans. Science 199, 1093–1095. models of stress-related disorders. J. Pharmacol. Exp. Ther. 301,
Griebel, G., 1999a. Is there a future for neuropeptide receptor ligands in 333–345.
the treatment of anxiety disorders? Pharmacol. Ther. 82, 1–61. Griffin, L.D., Mellon, S.H., 1999. Selective serotonin reuptake inhibitors
Griebel, G., 1999b. 5-HT1A receptor blockers as potential drug candidates directly alter activity of neurosteroidogenic enzymes. Proc. Natl.
for the treatment of anxiety disorders. Drug News Perspect. 12, Acad. Sci. U.S.A. 96, 13512–13517.
484–490. Griffiths, J.L., Lovick, T.A., 2002. Co-localization of 5-HT2A -receptor-
Griebel, G., Saffroy-Spittler, M., Misslin, R., Remmy, D., Vogel, E., and GABA-immunoreactivity in neurones in the periaqueductal grey
Bourguignon, J.-J., 1991. Comparison of the behavioural effects matter of the rat. Neurosci. Lett. 326, 151–154.
of an adenosine A1 /A2 -receptor antagonist, CGS15943A, and an Griffond, B., Baker, B.I., 2002. Cell and molecular cell biology of
A1 -selective antagonist, DPCPX. Psychopharmacology 103, 541–544. melanin-concentrating hormone. Int. Rev. Cytol. 213, 233–277.
Griebel, G., Moreau, J.-L., Jenck, F., Misslin, R., Martin, J.R., Grillon, C., 2002. Startle reactivity and anxiety disorders: aversive
1994. Acute and chronic treatment with 5-HT reuptake inhibitors conditioning, context, and neurobiology. Biol. Psychiatry 52, 958–975.
differentially modulate emotional responses in anxiety models in Grimsby, J., Toth, M., Chen, K., Kumazawa, T., Klaidman, L., Adams,
rodents. Psychopharmacology 113, 463–470. J.D., Karoum, F., Gal, J., Shih, J.C., 1997. Increased stress response
Griebel, G., Perrault, G., Sanger, D.J., 1997a. A comparative study of and beta-phenylethylamine in MAOB-deficient mice. Nat. Genet. 17,
the effects of selective and non-selective 5-HT2 receptor subtype 206–210.
antagonists in rat and mouse models of anxiety. Neuropharmacology Grobin, A.C., Deutch, A.Y., 1998. Dopaminergic regulation of
36, 793–802. extracellular GABA levels in the prefrontal cortex. J. Pharmacol.
Griebel, G., Perrault, G., Sanger, D.J., 1997b. CCK receptor antagonists Exp. Ther. 285, 350–357.
Grobin, A.C., Papadeas, S.T., Morrow, A.L., 2000. Regional vairations
in animal models of anxiety: comparison between exploration tests,
in the effects of chronic ethanol administration on GABA(A) receptor
conflict procedures and a model based on defensive behaviours.
expression: potential mechanisms. Neurochem. Int. 37, 453–461.
Behav. Pharmacol. 8, 549–560.
Groenewegen, H.J., Uylings, H.B.M., 2000. The prefrontal cortex and
Griebel, G., Curet, O., Perrault, G., Sanger, D.J., 1998a. Behavioral
the integration of sensory, limbic and autonomic information. Prog.
effects of phenelzine in an experimental model for screening anxiolytic
Brain Res. 126, 3–28.
and anti-panic drugs: correlation with changes in monoamine-oxidase
Groenink, L., Joordens, J.E., Hijzen, T.H., Dirks, A., Olivier, B., 2000.
activity and monoamine levels. Neuropharmacology 37, 927–935.
Infusion of flesinoxan into the amygdala blocks the fear-potentiated
Griebel, G., Perrault, G., Sanger, D.J., 1998b. Characterization of the
startle. NeuroReport 11, 2285–2289.
behavioural profile of the non-peptide CRF receptor antagonist CP-
Groenink, L., Dirks, A., Verdouw, P.M., Schipholt, M.L., Veening, J.G.,
154,526 in anxiety models in rodents: comparison with diazepam and
Van der Gugten, J., Olivier, B., 2002. HPA Axis dysregulation in
buspirone. Psychopharmacology 138, 55–66.
mice overexpressing corticotropin releasing hormone. Biol. Psychiatry
Griebel, G., Perrault, G., Letang, V., Granger, P., Avenet, P., Shoemaker,
51, 875–881.
H., Sanger, D.J., 1999a. New evidence that the pharmacological effects Groenink, L., Pattij, T., De Jongh, R., Van Der Gugten, J., Oosting, R.S.,
of benzodiazepine receptor ligands can be associated with activities at Dirks, A., Olivier, B., 2003. 5-HT1A receptor knock-out mice and
different BZ (␻) receptor subtypes. Psychopharmacology 146, 205–213. mice overexpressing corticotrophin-releasing hormone in models of
Griebel, G., Perrault, G., Sanger, D.J., 1999b. orphaninFQ, a novel anxiety. Eur. J. Pharmacol. 463, 185–197.
neuropeptide with anti-stress-like activity. Brain Res. 836, 221–224. Gross, C., Zhuang, X., Stark, K., Ramboz, S., Oosting, R., Kirby, L.,
Griebel, G., Perrault, G., Tan, S., Shoemaker, H., Sanger, D.J., 1999c. Santarelli, L., Beck, S., Hen, R., 2002. Serotonin1A receptor acts
Comparison of the pharmacological properties of classical and novel during development to establish normal anxiety-like behaviour in the
BZ-␻ receptor ligands. Behav. Pharmacol. 10, 483–495. adult. Nature 416, 396–400.
Griebel, G., Rodgers, J.R., Perrault, G., Sanger, D.J., 2000. The effects Grove, K.L., Campbell, R.E., Ffrench-Mullen, J.M., Cowley, M.A., Smith,
of compounds varying in selectivity as 5-HT1A receptor antagonists M.S., 2000. Neuropeptide Y Y5 receptor protein in the cortical/limbic
in three rat models of anxiety. Neuropharmacology 39, 1848–1857. system and brainstem of the rat: expression on ␥-aminobutyric acid and
Griebel, G., Moindrot, N., Aliaga, C., Simiand, J., Soubrié, P., 2001a. corticotropin-releasing hormone neurons. Neuroscience 100, 731–740.
Characterization of the profile of neurokinin-2 and neurotensin receptor Gu, Q., Moss, R.L., 1996. 17␤-Estradiol potentiates kainate-induced
antagonists in the mouse defense test battery. Neurosci. Behav. Rev. currents via activation of the camp cascade. J. Neurosci. 16, 3620–3629.
25, 619–626. Guardiola-Lemaı̂tre, B., Lenegre, A., Porsolt, R.D., 1992. Combined
Griebel, G., Perrault, G., Simiand, J., Cohen, C., Granger, P., Decobert, effects of diazepam and melatonin in two tests for anxiolytic activity
M., Françon, D., Avenet, P., Depoortere, H., Tan, S., Oblin, A., in the mouse. Pharmacol. Biochem. Behav. 41, 405–408.
Shoemaker, H., Evanno, Y., Sevrin, M., George, P., Scatton, B., 2001b. Guidotti, A., Costa, E., 1998. Can the antidysphoric and anxiolytic profiles
SL651498: an anxioselective compound with functional selectivity for of selective serotonin reuptake inhibitors be related to their ability
␣2 - and ␣3 -containing ␥-aminobutyric acidA (GABAA ) receptors. J. to increase brain 3␣,5␣-tetrahydroprogesterone (allopregnanolone)
Pharmacol. Exp. Ther. 298, 753–768. availability? Biol. Psychiatry 44, 865–873.
Griebel, G., Perrault, G., Soubrié, P., 2001c. Effects of SR48968, a Guimarães, F.S., Mbaya, P.S., Deakin, J.F.W., 1997. Ritanserin facilitates
selective non-peptide NK2 receptor antagonist on emotional processes anxiety in a simulated public-speaking paradigm. J. Neuropharmacol.
in rodents. Psychopharmacology 158, 241–251. 11, 225–231.
Griebel, G., Simiand, J., Serradeil-Le Gal, C., Wagnon, J., Pascal, Guldner, J., Trachsel, L., Kratschmayr, C., Roth, B., Holsboer, F., Steiger,
M., Scatton, B., Maffrand, J.-P., Soubrié, P., 2002a. Anxiolytic- A., 1995. Bretazenil modulates sleep EEG and nocturnal hormone
and antidepressant-like effects of the non-peptide vasopressin V1b secretion in normal men. Psychopharmacology 122, 115–121.
200 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Gulinello, M., Gong, Q.H., Li, X., Smith, S.S., 2001. Short-term exposure Halbreich, U., Petty, F., Yonkers, K., Kramer, G.L., Rush, A.J., Bibi,
to a neuroactive steroid increases alpha4 GABA(A) receptor subunit K.W., 1996. Low plasma ␥-aminobutyric acid levels during the late
levels in association with increased anxiety in the female rat. Brain luteal phase of women with premenstual dysphoric disorder. Am. J.
Res. 910, 55–66. Psychiatry 153, 718–720.
Gulinello, M., Gong, Q.H., Smith, S.S., 2002. Progesterone withdrawal Hallbeck, M., Hermanson, O., Blomqvist, A., 1999. Distribution of
increases the ␣4 subunit of the GABAA receptor in male rats in preprovasopressin mRNA in the rat central nervous system. J.
association with anxiety and altered pharmacology—a comparison Comp. Neurol. 411, 181–200.
with female rats. Neuropharmacology 43, 701–714. Haller, J., 2001. The link between stress and the efficacy of anxiolytics.
Günther, U., Benson, J., Benke, D., Fritschy, J., Reyes, G., Knoflach, F., A new avenue of research. Physiol. Behav. 73, 337–342.
Crestani, F., Aguzzi, A., Argoni, M., Lang, Y., Bluethmann, H., Mohler, Haller, J., Halasz, J., 2000. The effect of two acute stressors on the
H., Luscher, B., 1995. Benzodiazepine-insensitive mice generated by anxiolytic efficacy of chlordiazepoxide. Psychopharmacology 151, 1–6.
targeted disruption of the ␥2 subunit gene of ␥-aminobutyric acid Haller, J., Halasz, J., Makara, G.B., Kruk, M.R., 1998. Acute effects
type A receptors. Proc. Natl. Acad. Sci. U.S.A. 92, 7749–7753. of glucocorticoids: behavioural and pharmacological perspectives.
Gur, E., Dremencov, E., Lerer, B., Newman, M.E., 2001. Functional Neurosci. Biobehav. Rev. 23, 337–344.
effects of corticosterone on 5-HT1A and 5-HT1B receptor activity in rat Haller, J., Leveleki, C., Halasz, J., Baranyi, J., Makara, G.B., 2001.
brain: in vivo microdialysis studies. Eur. J. Pharmacol. 411, 115–122. The effect of glucocorticoids on the anxiolytic efficacy of buspirone.
Gutierrez, A., Khan, Z.U., De Blas, A.L., 1994. Immunocytochemical Psychopharmacology 157, 388–394.
localization of ␥2 short and ␥2 long subunits of the GABAA receptor Haller, J., Bakos, N., Rodriguiz, R.M., Caron, M.G., Wetsel, W.C.,
in rat brain. J. Neurosci. 14, 7168–7179. Liposits, Z., 2002a. Behavioral responses to social stress in
Gutkowska, J., Nemer, M., 1989. Structure, expression, and function of noradrenaline transporter knock-out mice: effects on social behavior
atrial natriuretic factor in extraatrial tissues. Endocr. Rev. 10, 519–536. and depression. Brain Res. Bull. 58, 279–284.
Haas, H., Panula, P., 2003. The role of histamine and the tuberomamillary Haller, J., Bakos, N., Szirmay, M., Ledent, C., Freund, T.F., 2002b.
nucleus in the nervous system. Nat. Neurosci. 4, 121–130. The effects of genetic and pharmacological blockade of the CB1
Habib, K.E., Weld, K.P., Rice, K.C., Pushkas, J., Champoux, M., cannabinoid receptor on anxiety. Eur. J. Neurosci. 16, 1395–1398.
Listwak, S., Webster, E.L., Atkinson, A.J., Schulkin, J., Contoreggi, Hamilton, M.E., Weddige, F.K., Freeman, A.S., 2001. Effects of forebrain
C., Chrousos, G.P., McCann, S.M., Suomi, S.J., Higley, J.D., microinjection of cholecystokinin on dopamine cell firing rate.
Gold, P.W., 2000. Oral administration of a corticotropin-releasing Peptides 22, 1063–1069.
hormone receptor antagonist significantly attenuates behavioural, Hammock, E.A.D., Young, L.J., 2002. Variation in the vasopressin
neuroendocrine, and autonomic responses to stress in primates. Proc. V1a receptor promoter and expression: implications for inter- and
Natl. Acad. Sci. U.S.A. 97, 6079–6084. intraspecific variation in social behaviour. Eur. J. Neurosci. 16,
399–402.
Hadingham, K.L., Wingrove, P.B., Wafford, K.A., Bain, C., Kemp, J.A.,
Hammack, S.E., Richey, K.J., Schmid, M.J., LoPresti, M.L., Watkins,
Palmer, K.J., Wilson, A.W., Wilcox, A.S., Sikela, J.M., Ragan, C.I.,
L.R., Maier, S.F., 2002. The role of corticotropin-releasing hormone
1993. Role of beta subunit in determining the pharmacology of human
in the dorsal raphe nucleus in mediating the behavioral consequences
␥-aminobutyric acid type A receptors. Mol. Pharmacol. 44, 1211–1218.
of uncontrollable stress. J. Neurosci. 22, 1020–1026.
Haddad, J.J., 2002. Cytokines and related receptor-mediated signaling
Hammack, S.E., Schmid, M.J., LoPresti, M., Der-Avakian, A.,
pathways. Biochem. Biophys. Res. Commun. 297, 700–713.
Pellymounter, M.A., Foster, A.C., Watkins, L.R., Maier, S.F., 2003.
Haddjeri, N., Blier, P., De Montigny, C., 1998. Long-term antidepressant
Corticotropin releasing hormone type 2 receptors in the dorsal raphe
treatments result in a tonic activation of forebrain 5-HT1A receptors.
nucleus mediate the behavioral consequences of uncontrollable stress.
J. Neurosci. 18, 10150–10156.
J. Neurosci. 23, 1019–1025.
Haefely, W., Facklam, M., Schoch, P., Martin, J.R., Bonetti, E.P., Moreau, Hamon, M., Doucet, E., Lefèvre, K., Miquel, M.-C., Lanfumey, L., Insusti,
J.L., Jenck, F., Richards, J.G., 1992. Partial agonists of benzodiazepine R., Frechilla, D., Del Rio, J., Vergé, D., 1999. Antibodies and antisense
receptors for the treatment of epilepsy, sleep, and anxiety disorders. oligonucleotide for probing the distribution and putative functions of
Adv. Biochem. Psychopharmacol. 47, 379–394. central 5-HT6 receptors. Neuropsychopharmacology 21, 68S–76S.
Hagan, J.J., Price, G.W., Jeffrey, P., Deeks, N.J., Stean, T., Piper, D., Hancock, A.A., 1996. ␣1 Adrenoceptor subtypes: a synopsis of their
Smith, M.I., Upton, N., Medhurst, A.D., Middlemiss, D.N., Riley, G.J., pharmacology and molecular biology. Drug Dev. Res. 39, 54–107.
Lovell, P.J., Bromidge, S.M., Thomas, D.R., 2000. Characterization of Handley, S.L., 1995. 5-Hydroxytryptamine pathways in anxiety and its
SB-269970-A, a selective 5-HT7 receptor antagonist. Br. J. Pharmacol. treatment. Pharmacol. Ther. 66, 103–148.
130, 539–548. Handley, S.L., Mithani, S., 1984. Effects of alpha-adrenoceptor agonists
Hahn, M.K., Bannon, M.J., 1999. Stress-induced c-fos expression in the and antagonists in a maze-exploration model of “fear”-motivated
rat locus coeruleus is dependent on neurokinin1 receptor activation. behaviour. Naunyn Schmiedebergs Arch. Pharmacol. 327, 1–5.
Neuroscience 94, 1183–1188. Haney, M., Hart, C.L., Ward, A.S., Foltin, R.W., 2003. Nefazodone
Hajos, N., Katona, I., Naiem, S.S., Ledent, C., Mody, I., Freund, T.F., decreases anxiety during marijuana withdrawal in humans.
2000a. Cannabinoids inhibit hippocampal GABAergic transmission Psychopharmacology 165, 157–165.
and network oscillation. Eur. J. Neurosci. 12, 3239–3249. Hanna, M.C., Davies, P.A., Hales, T.G., Kirkness, E.F., 2000. Evidence
Hajos, N., Musser, Z., Rancz, E.A., Freund, T.F., Mody, I., 2000b. Cell for expression of heteromeric sertonin 5-HT(3) receptors in rodents.
type and synapse-specific variability in synaptic GABAA receptor J. Neurochem. 75, 240–247.
occurency. Eur. J. Neurosci. 12, 810–818. Hanner, M., Moebius, F.F., Flandorfer, A., Knaus, H.G., Striessnig, J.,
Hajszan, T., Zabroszky, L., 2002. Direct catecholaminergic–cholinergic Kempner, E., Glossman, H., 1996. Purification, molecular cloning,
interactions in the basal forebrain. III. Adrenergic innervation of choline and expression of the mammalian sigma1 -binding site. Proc. Natl.
acetyltransferase-containing neurons in the rat. J. Comp. Neurobiol. Acad. Sci. U.S.A. 93, 8072–8077.
449, 141–157. Hara, M., Kai, Y., Ikemoto, Y., 1993. Propofol activates GABAA receptor-
Halasy, K., Miettinen, R., Szabat, E., Freund, T.F., 1992. GABAergic chloride ionophore complex in dissociated hippocampal pyramidal
interneurons are the major postsynaptic targets of median raphe neurons of the rat. Anesthesiology 79, 781–788.
afferents in the rat dentate gyrus. Eur. J. Neurosci. 4, 144–153. Haraguchi, H., Kuribara, H., 1991. Behavioral effects of adenosine
Halbreich, U., 1997. Role of estrogen in postmenopausal depression. agonists: evaluation by punishment, discrete shuttle avoidance and
Neurobiology 48, S16–S19. activity tests in mice. Jpn. J. Pharmacol. 55, 303–310.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 201

Hariri, A.R., Mattay, V.S., Tessitore, A., Fera, F., Smith, W.G., Haskell-Luevano, C., Chen, P., Li, C., Chang, K., Smith, M.S., Cameron,
Weinberger, D.R., 2002a. Dextroamphetamine modulates the response J.L., Cone, R.D., 1999. Characterization of the neuroanatomical
of the human amygdala. Neuropsychopharmacology 27, 1036–1040. distribution of agouti-related protein immunoreactivity in the rhesus
Hariri, A.R., Mattay, V.S., Tessitore, A., Kolachana, B., Fera, F., Goldman, monkey and the rat. Endocrinology 140, 1408–1415.
D., Egan, M.F., Weinberger, D.W., 2002b. Serotonin transporter Hasue, B.H., Shammah-Lagnado, S.J., 2002. Origin of the dopaminergic
genetic variation and the response of the human amygdala. Science innervation of the central extended amygdala and accumbens shell:
297, 400–403. a combined retrograde tracing and immunohistochemical study in the
Harkin, A.J., Bruce, K.H., Craft, B., Paul, I.A., 1999. Nitric oxide rat. J. Comp. Neurol. 454, 15–33.
synthase inhibitors have antidepressant-like properties in mice. I. Hauger, R.L., Grigoriadis, D.E., Dallaman, M.F., Plotsky, P.M., Vale,
Acute treatments are active in the forced swim test. Eur. J. Pharmacol. W.W., Dautzenberg, F.M., 2003. International union of pharmacology.
372, 207–213. XXXVI. Current status of the nomenclature of receptors for corti-
Harris, B.D., Wong, G., Moody, E.J., Skolnick, P., 1995. Different subunit cotropin-releasing factor and their ligands. Pharmacol. Rev. 55, 21–26.
requirements for volatile and non-volatile anesthetics at ␥-aminubutyric Haurie, P.J., Friedman, M., Ravaris, C.L., 1989. Sleep in patients with
acid type A receptors. Am. Soc. Pharmacol. Exp. Ther. 47, 363–367. spontaneous panic attacks. Sleep 12, 323–337.
Harrison, Y.E., Jenkins, J.A., Rochat, B.A., Lytle, D.A., Jung, M.E., Hawes, B.E., Graziano, M.P., Lambert, D.G., 2000a. Cellular actions of
Oglesby, M.W., 1998. Discriminative stimulus effects of diazepam, nociceptin: transduction mechanisms. Peptides 21, 961–967.
ketamine and their mixture: ethanol substitution patterns. Behav. Hawes, B.E., Kil, E., Green, B., O’Neill, K., Fried, S., Graziano,
Pharmacol. 9, 31–40. M.P., 2000b. The melanin-concentrating hormone receptor couples to
Harro, J., 2002. Long-term partial 5-HT depletion: interference of anxiety multiple G-proteins to activate diverse intracellular signaling pathways.
and impulsivity? Psychopharmacology 164, 433–434. Endocrinology 141, 4524–4532.
Harro, J., Kiivet, R.A., Lang, A., Vasar, E., 1990. Rats with anxious or non- Hawkins, M.F., Uzelac, S.M., Baumeister, A.A., Hearn, J.K., Broussard,
anxious type of exploratory behaviour differ in their brain CCK8 and J.I., Guillot, T.S., 2002. Behavioral responses to stress following
benzodiazepine receptor characteristics. Behav. Brain Res. 18, 63–71. central and peripheral injection of the 5-HT2 agonist DOI. Pharmacol.
Hart, S., Sarter, M., Berntson, G.G., 1999. Cholinergic inputs to the rat Biochem. Behav. 73, 537–544.
medial prefrontal cortex mediate potentiation of the cardiovascular Hayashi, T., Su, T.P., 2001. Regulating ankyrin dynamics: roles of
defensive response by the anxiogenic benzodiazepine receptor partial sigma-1 receptors. Proc. Natl. Acad. Sci. U.S.A. 98, 491–496.
inverse agonist FG 7142. Neuroscience 94, 1029–1038. Hayashi, T., Maurice, T., Su, T.-P., 2000a. Ca2+ signaling via ␴1 -
Hartmann, M., Heumann, R., Lessmann, V., 2001. Synaptic secretion receptors: novel regulatory mechanism affecting intracellular Ca2+
of BDNF after high-frequency stimulation of glutamatergic synapses. concentration. J. Pharmacol. Exp. Ther. 293, 788–798.
EMBO J. 20, 5887–5897. Hayashi, Y., Shi, S.H., Esteban, J.A., Piccini, A., Poncer, A., Malinow, R.,
Harvey, J., Wedley, S., Findlay, J.D., Sidell, M.R., Pullar, I.A., 1996. 2000b. Driving AMPA receptors into synapses by LTP and CaMKII:
␻-Agatoxin identifies a single calcium channel subtype which requirement for GluR1 and PDZ domain interaction. Science 287,
contributes to the potassium-induced release of acetylcholine, 5- 2262–2267.
hydroxytryptamine, dopamine, ␥-aminobutyric acid and glutamate Heading, C., 2001. Siramesine. Curr. Opin. Invest. Drugs 2, 266–270.
from rat brain slices. Neuropharmacology 35, 385–392. Heal, D.J., Butler, S.A., Hurst, E.M., Buckett, W.R., 1989. Antidepressant
Hascoët, M., Bourin, M., Todd, K.G., Coüetoux du Tertre, A., 1994. treatments, including sibutramine hydrochloride and electroconvulsive
Anti-conflict effect of 5-HT1A agonists in rats: a new model for shock, decrease ␤1 - but not ␤2 -adrenoceptors in rat cortex. J.
evaluating anxiolytic-like activity. J. Psychopharmacol. 8, 227–237. Neurochem. 53, 1019–1025.
Hascoët, M., Bourin, M., Colombel, M.C., Fiocco, A.J., Baker, G.B., 2000. Heilig, M., 1995. Antisense inhibition of neuropeptide Y (NPY)-Y1
Anxiolytic-like effects of antidepressants after acute administration in receptor expression blocks the anxiolytic-like action of NPY in
a four-plate test in mice. Pharmacol. Biochem. Behav. 65, 339–344. amygdala and paradoxically increases feeding. Regul. Pept. 59,
Haseneder, R., Rammes, G., Zieglgänsberger, W., Kocks, E., Hapfelmeier, 201–205.
G., 2002. GABAA receptor activation and open-channel block by Heilig, M., Söderpalm, B., Engel, J.A., Widerlöv. 1989. Centrally
volatile anaesthetics: a new principle of receptor modulation. Eur. J. administered neuropeptide Y (NPY) produces anxiolytic-like effects
Pharmacol. 451, 43–50. in animal anxiety models. Psychopharmacology 98, 524–529.
Hasenöhrl, R.U., Jentjens, O., De Souza Silva, M.A., Tomaz, C., Heilig, M., McLeod, S., Koob, G.K., Britton, K.T., 1992. Anxiolytic-like
Huston, J.P., 1998. Anxiolytic-like action of neurokinin substance P effect of neuropeptide Y (NPY), but not other peptides in an operant
administered systematically or into the nucleus basalis magnocellularis conflict test. Regul. Pept. 41, 61–69.
region. Eur. J. Pharmacol. 354, 123–133. Heilig, M., McLeod, S., Brot, M., Heinrichs, S.C., Menzaghi, F., Koob,
Hasenöhrl, R.U., Weth, K., Huston, J.P., 1999. Intraventricular infusion G.F., Britton, K.T., 1993. Anxiolytic-like action of neuropeptide Y:
of the histamine H1 receptor antagonist chlorpheniramine improves mediation by Y1 receptors in amygdala, and dissociation from food
maze performance and has anxiolytic-like effects in aged hybrid intake effects. Neuropsychopharmacology 8, 357–363.
Fischer 344Xbrown Norway rats. Exp. Brain Res. 128, 435–440. Heilig, M., Koob, G.F., Ekman, R., Britton, K.T., 1994. Corticotropin-
Hashimoto, H., Nogi, H., Mori, K., Ohishi, H., Shigemoto, R., releasing factor and neuropeptide Y: role in emotional integration.
Yamamoto, K., Matsuda, T., Mikuno, N., Nagata, S., Baba, A., 1996. Trends Neurosci. 17, 80–85.
Distribution of the mRNA for a pituitary adenylate cyclase-activating Heim, C., Nemeroff, C.B., 1999. The impact of early adverse experiences
polypeptide receptor in the rat brain: an in situ hybridization study. on brain systems involved in the pathophysiology of anxiety and
J. Comp. Neurol. 371, 567–577. affective disorders. Biol. Psychiatry 46, 1509–1522.
Hashimoto, H., Shintani, N., Tanaka, K., Mori, W., Hirose, M., Matsuda, Hein, L., 2001. Transgenic models of ␣2 -adrenergic receptor subtype
T., Sakaue, M., Miyazaki, J., Niwa, H., Tashiro, F., Yamamoto, K., function. Rev. Phys. Biochem. Pharmacol. 142, 161–185.
Koga, K., Tomimoto, S., Kunugi, A., Suetake, S., Baba, A., 2001. Hein, L., Altman, J.D., Kobilka, B.K., 1999. Two functionally distinct ␣2 -
Altered psychomotor behaviors in mice lacking pituitary adenylate adrenergic receptors regulate sympathetic neurotransmission. Nature
cyclase-activating polypeptide (PACAP). Proc. Natl. Acad. Sci. U.S.A. 402, 181–184.
98, 13355–13360. Heinrichs, S.C., Lapsansky, J., Lovenberg, T.W., De Souza, E.B.,
Haskell-Luevano, C., Monck, E.K., 2001. Agouti-related protein functions Chalmers, D.T., 1997a. Corticotropin-releasing factor CRF1 , but not
as an inverse agonist at a constitutively active brain melanocortin-4 CRF2 , receptors mediate anxiogenic-like behavior. Regul. Pept. 71,
receptor. Regul. Pept. 99, 1–7. 15–21.
202 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Heinrichs, S.C., Min, H., Tamraz, S., Carmouche, M., Boehme, S.A., Vale, Hermann, H., Marsicano, G., Lutz, B., 2002. Coexpression of the
W.W., 1997b. Anti-sexual and anxiogenic behavioral consequences of cannabinoid receptor type I with dopamine and serotonin receptors
corticotropin-releasing factor overexpression are centrally mediated. in distinct neuronal subpopulations of the adult mouse frebrain.
Psychoneuroendocrinology, 215–224. Neuroscience 109, 451–460.
Heinrichs, S.C., De Souza, E.B., Schulteis, G., Lapsansky, J.L., Hernando, F., Schoots, O., Lolait, S.J., Burbach, J.P., 2001.
Grigoriadis, D.E., 2002. Brain penetrance, receptor occupancy and Immunohistochemical localization of the vasopressin V1b receptor in
antistress in vivo efficacy of a small molecule corticotropin releasing the rat brain and pituitary gland: anatomical support for its involvement
factor type I receptor selective antagonist. Neuropsychopharmacology in the central effects of vasopressin. Endocrinology 142, 1659–1668.
27, 194–202. Hervieu, G.J., Emson, P.C., 1998. The localisation of somatostatin
Heisler, L.K., Bajwa, L.K., Tecott, L.H., 1998a. Altered anxiety-like receptor I (sst1) immunoreactivity in the rat brain using an N-terminal
behavior in 5-HT2C receptor null mutant mice. Soc. Neurosci. Abstr. specific antibody. Neurosci. 85, 1263–1284.
24, 602. Hervieu, G.J., Cluderay, J.E., Harrison, D., Meakin, J., Maycox, P., Nasir,
Heisler, L.K., Chu, H.M., Brennan, T.J., Danao, J.A., Bajwa, P., Parsons, S., Leslie, R.A., 2000. The distribution of the mRNA and protein
L.H., Tecott, L.H., 1998b. Elevated anxiety and antidepressant-like products of the melanin-concentrating hormone (MCH) receptor gene,
responses in serotonin 5-HT1A receptor mutant mice. Proc. Natl. slc-1, in the central nervous system of the rat. Eur. J. Neurosci. 12,
Acad. Sci. U.S.A. 95, 15049–15054. 1194–1216.
Hellewell, S.B., Bruce, A., Feinstein, G., Orringer, J., Williams, W., Herz, A., Akil, H., Simon, E. (Eds.), 1993. The Opioids. Handbook of
Bowen, W.D., 1994. Rat liver and kidney contain high densities of ␴1 Experimental Pharmacology, vol. 104. Springer, Berlin, p. 841.
and ␴2 receptors: characterization by ligand binding and photoaffinity Herzog, E., Bellenchi, G.C., Gras, C., Bernard, V., Ravassard, P., Bedet,
labeling. Eur. J. Pharmacol. 268, 9–18. C., Gasnier, B., Giros, B., El Mestikawy, S., 2001. The existence of
Hellewell, J.S.E., Guimaraes, F.S., Wang, M., Deakin, J.F.W., 1999. a second vesicular glutamate transporter specifies subpopulations of
Comparison of buspirone with diazepam and fluvoxamine on aversive glutamatergic neurones. J. Neurosci. 21 (RC181), 1–6.
classical conditioning in humans. J. Psychopharmacol. 13, 122–127. Hettema, J.M., Neale, M.C., Kendler, K.S., 2001. A review and meta-
Helmeste, D.M., Tang, S.W., 1998. The role of calcium in the etiology analysis of the genetic epidemiology of anxiety disorders. Am. J.
of the affective disorders. Jpn. J. Pharmacol. 77, 107–116. Psychiatry 158, 1568–1578.
Helmeste, D.M., Tang, S.W., 2000. Dopamine D4 receptors. Jpn. J. Hevers, W., Lüddens, H., 1998. The diversity of GABAA receptors. Mol.
Pharmacol. 82, 1–14. Neurobiol. 18, 35–86.
Hidalgo, R.B., Barnett, S.D., Davidson, J.R.T., 2001. Social anxiety
Helmreich, D.L., Watkins, L.R., Deak, T., Maier, S.F., Akil, H., Watson,
disorder in review: two decades of progress. Int. J. Neuropsychopha-
S.J., 1999. The effect of stressor controllability on stress-induced
rmacol. 4, 279–298.
neuropeptide mRNA expression within the paraventricular nucleus of
Hieble, J.P., 2000. Adrenoceptor subclassification: an approach to
the hypothalamus. J. Neuroendocrinol. 11, 121–128.
improved cardiovascular therapeutics. Pharm. Acta Helv. 74, 163–171.
Helton, D.R., Tizzano, J.P., Monn, J.A., Schoepp, D.D., Kallman, M.J.,
Hieble, J.P., Bondinell, W.E., Ruffolo, R.R., 1995. Alpha- and beta-
1998. Anxiolytic and side-effect profile of LY354740: a potent, highly
adrenoceptors: from the gene to the clinic. 1. Molecular biology and
selective, orally active agonist for group II metabotropic glutamate
adrenoceptor subclassification. J. Med. Chem. 38, 3415–3444.
receptors. J. Pharmacol. Exp. Ther. 284, 651–660.
Higelin, J., Py-Lang, G., Paternoster, C., Ellis, G., Patel, A., Dautzenberg,
Hemrick-Lubecke, S.K., Evans, D.C., 2002. Comparison of the potency
F., 2001. [125 I]antisauvagine-30 a novel and specific high-affinity
of MDL 100,907 and SB 242,084 in blocking the serotonin (5-
radioligand for the characterization of corticotropin factor type 2
HT)2 receptor agonist-induced increases in rat serum corticosterone
receptors. Neuropharmacology 40, 114–122.
concentration: evidence for 5-HT2A receptor mediation of the HPA Higgins, G.A., Bradbury, A.J., Jones, B.J., Oakley, N.R., 1988.
axis. Neuropharmacology 42, 162–169. Behavioural and biochemical consequences following activation of
Hemrick-Lubecke, S.K., Bymaster, F.P., Evans, D.C., Wess, J., 5-HT1 -like and GABA receptors in the dorsal raphé nucleus of the
Felder, C.C., 2002. Muscarinic agonist-mediated increases in serum rat. Neuropharmacology 27, 993–1001.
corticosterone levels are abolished in M2 muscarinic acetylcholine Higgins, G.A., Jones, B.J., Oakley, N.R., Tyers, M.B., 1991. Evidence
receptor knock-out mice. J. Pharmacol. Exp. Ther. 303, 99–103. that the amygdala is involved in the disinhibitory effects of 5-HT3
Hendrie, C.A., Dourish, C.T., 1990. Anxiolytic profile of the chole- receptor antagonists. Psychopharmacology 104, 545–551.
cystokinin antagonist devazepide in mice. Br. J. Pharmacol. 99, 138P. Higgins, G.A., Jones, B.J., Oakley, N.R., 1992. Effect of 5-HT1A
Henneberger, C., Jüttner, R., Rothe, T., Grantyn, R., 2002. Postsynaptic receptor agonists in two models of anxiety after dorsal raphe injection.
action of BDNF on GABAergic synaptic transmission in the superficial Psychopharmacology 106, 261–267.
layers of the mouse superior colliculus. J. Neurophysiol. 88, 595–603. Higgins, G.A., Grottick, A.J., Ballard, T.M., 2001. Influence of the
Hensler, J.G., 2002. Differential regulation of 5-HT1A receptor-G protein selective ORL1 receptor agonist, Ro64-6198, on rodent neurological
interactions in brain following chronic antidepressant administration. function. Neuropharmacology 41, 97–107.
Neuropsychopharmacology 26, 565–573. Higgins, G.A., Ballard, T.M., Huwyler, J., Kemp, J.A., Gill, R., 2003.
Hensman, H., Guimarães, F.S., Wang, M., Deakin, J.F.W., 1991. Evaluation of the NR2B-selective NMDA receptor antagonist Ro63-
Effects of ritanserin on aversive classical conditioning in humans. 1908 on rodent behaviour: evidence for an involvement of NR2B
Psychopharmacology 104, 220–224. NMDA receptors in response inhibition. Neuropharmacology 44,
Herbison, A.E., Fénelon, V.S., 1995. Estrogen regulation of GABAA 324–341.
receptor subunit mRNA expression in preoptic area and bed nucleus Hill, D.R., Suman-Chauban, N., Woodruff, G.N., 1993. Localization of
of the stria terminalis of female rat brain. J. Neurosci. 15, 2328–2337. [3 H]GABApentin to a novel site in rat brain: autoradiographic studies.
Herman, J.P., 1999. Neurocircuit control of the hypothalamus–pituitary– Eur. J. Pharmacol. 244, 303–309.
adrenocortical axis during stress. Curr. Opin. Endocrinol. Diabetes 6, Hill, S.J., Ganellin, C.R., Timmerman, H., Schwartz, J.C., Shankley, N.P.,
3–9. Young, J.M., Schunack, W., Levi, R., Haas, H.L., 1997. International
Herman, J.P., Thomas, G.J., Gash, D.M., 1986. Behavioral characteristics union of pharmacology. XIII. Classification of histamine receptors.
of Roman high avoidance rats homozygous for diabetes insipidus Pharmacol. Rev. 49, 253–278.
(RHA:di/di). Behav. Brain Res. 20, 27–38. Hirschmann, S., Dannon, P.N., Iancu, I., Dolberg, O.T., Zohar, J.,
Herman, J.P., Cullinan, W.E., Ziegler, D.R., Tasker, J.G., 2002. Role of Grunhaus, L., 2000. Pindolol augmentation in patients with treatment-
the paraventricular nucleus microenvironment in stress integration. resistant panic disorder: a double-blind, placebo-controlled trial. J.
Eur. J. Neurosci. 16, 381–385. Clin. Psychopharmacol. 20, 556–559.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 203

Hjorth, S., Engel, J.A., Carlsson, A., 1986. Anticonflict effects of low Holmes, A., Yang, R.J., Vishwanath, J., Li, Q., Ma, L., Saavedra, M.C.,
doses of the dopamine agonist apomorphine in the rat. Pharmacol. Innerfield, C.E., Jacoby, A.S., Shine, J., Lismaa, T.E., Crawley, J.N.,
Biochem. Behav. 24, 237–240. 2002b. The GAL-R1 galanin receptor subtype modulates anxiety-like
Hjorth, S., Carlsson, A., Engel, J.A., 1987a. Anxiolytic-like action of the 3- behavior in mice. Soc. Neurosci. Abstr. 82.18.
PPP enantiomers in the Vogel conflict paradigm. Psychopharmacology Holsboer, F., 1999. The rationale for corticotropin-releasing hormone
92, 371–375. receptor (CRH-R) antagonists to treat depression and anxiety. J.
Hjorth, S., Söderpalm, B., Engel, J.A., 1987b. Biphasic effect of l,5-HTP Psychiat. Res. 33, 181–214.
in the Vogel conflict model. Psychopharmacology 92, 96–99. Homberg, J.R., Van Den Akker, M., Raase, H.S., Wardeh, G., Binnekade,
Hjorth, S., Bengtsson, H.J., Kullberg, A., Carlzon, D., Peilot, H., R., Schoffelmeer, A.N.M., De Vries, T.J., 2002. Enhanced motivation
Auerbach, S.B., 2000. Serotonin autoreceptor function and antide- to self-administer cocaine is predicted by self-grooming behaviour
pressant drug action. J. Psychopharmacol. 14, 177–185. and relates to dopamine release in the rat medial prefrontal cortex
Ho, S.P., Takahashi, L.K., Livanov, V., spencer, K., Lesher, T., Macing, C., and amygdala. Eur. J. Pharmacol. 15, 1542–1550.
2001. Attenuation of fear conditioning by antisense inhibition of brain Homanics, G.E., Harrison, N.L., Quinlan, J.J., Krasowski, M.D., Rick,
corticotropin releasing factor-2 receptor. Mol. Brain Res. 89, 29–40. C.E.M., de Blas, A.L., Mehta, A.K., Kist, F., Mihalek, R.M., Aul, J.J.,
Ho, Y.-J., Eichendorff, J., Schwarting, R.K.W., 2002. Individual response
Firestone, L3.L., 1999. Normal electrophysiological and behavioral
profiles of male Wistar rats in animal models for anxiety and
responses to ethanol in mice lacking the long splice variant of the ␥2
depression. Behav. Brain Res. 136, 1–12.
subunit of the ␥-aminobutyric type A receptor. Neuropharmacology
Hobom, M., Dai, S., Marais, E., Lacinová, L., Hofman, F., Klugbauer,
38, 253–255.
N., 2000. Neuronal distribution and functional characterization of the
Hong, W., Werling, L.L., 2000. Evidence that the ␴1 receptor is not
calcium channel ␣2␦ -2 subunit. Eur. J. Neurosci. 12, 1217–1226.
directly coupled to G proteins. Eur. J. Pharmacol. 408, 117–125.
Hodges, H., Green, S., Glenn, B., 1987. Evidence that the amygdala
is involved in benzodiazepine and serotonergic effects on punished Hood, S.D., Argyropoulos, V., Nutt, D.J., 2000. Agents in development
responding but not on discrimination. Psychopharmacology 92, 491– for anxiety disorders. CNS Drugs 13, 421–431.
504. Hooks, M.S., Kalivas, P.W., 1995. The role of mesoaccumbens–pallidal
Hodges, C.W., Mehmert, K.K., Kelley, S.P., McMahon, T., Haywood, A., circuitry in novelty-induced behavioral activation. Neuroscience 64,
Olive, M.F., Wang, D., Sanchez-Perez, A.M., Messing, R.O., 1999. 587–597.
Supersensitivity to allosteric GABAA receptor modulators and alcohol Hopkins, S.J., Rothwell, N.J., 1995. Cytokines and the nervous system.
in mice lacking PKCε. Nat. Neurosci. 2, 997–1002. I. Expression and cognition. Trends Neurosci. 18, 83–88.
Hodges, C.W., Raber, J., McMahon, T., Walter, H., Sanchez-Perez, Hopwood, S.E., Stamford, J.A., 2001a. Multiple 5-HT1 autoreceptor
A.M., Olive, M.F., Mehmert, K., Morrow, A.L., Messing, R.O., subtypes govern serotonin release in dorsal and median raphe nuclei.
2002. Decreased anxiety-like behavior, reduced stress hormones, and Neuropharmacology 40, 508–519.
neurosteroid supersensitivity in mice lacking protein kinase Cε. J. Hopwood, S.E., Stamford, J.A., 2001b. Noradrenergic modulation of
Clin. Invest. 110, 1003–1010. serotonin release in rat dorsal and median raphe nuclei via ␣1 and
Hoffman, A.F., Lupica, C.R., 2000. Mechanisms of cannabinoid inhibition ␣2A adrenoceptors. Neuropharmacology 41, 433–442.
of GABA(A) synaptic transmission in the hippocampus. J. Neurosci. Hough, L.B., 2001. Genomics meets histamine receptors: new subtypes,
20, 2470–2479. new receptors. Mol. Pharmacol. 59, 415–419.
Hofmman, F., Lacinova, L., Klugbauer, N., 1999. Voltage-dependent Hsu, S.Y., Hsueh, A.J.W., 2001. Human stresscopin and stresscopin-
calcium channels: from structure to function. Rev. Physiol. Biochem. related peptide are selective ligands for the type 2 corticotropin-
Pharmacol. 139, 33–87. releasing hormone receptor. Nat. Med. 7, 605–611.
Hökfelt, T., Xu, Z., Shi, T., Holmberg, K., Zhang, X., 1998. Galanin in Hsu, F.-C., Smith, S.S., 2003. Progesterone withdrawal reduces paired-
ascending systems: focus on coexistence with 5-hydroxytryptamine pulse inhibition in rat hippocampus: dependence on GABAA receptor
and noradrenaline. Ann. N. Y. Acad. Sci. 863, 252–263. alpha4 subunit up-regulation. J. Neurophysiol. 89, 186–198.
Hökfelt, T., Broberger, C., Xu, Z.Q.D., Sergeyev, V., Ubink, R., Diez, Hsu, D.T., Chen, F.-L., Takahashi, L.K., Kalin, N.H., 1998. Rapid stress-
M., 2000. Neuropeptides—an overview. Neuropharmacology 39, induced elevations in corticotropin-releasing hormone mRNA in rat
1337–1356. central amygdala and hypothalamic paraventricular nucleus: an in situ
Hollander, E., DeCaria, C., Gully, R., Nitescu, A., Suckow, R.F., Gorman, hybridization analysis. Brain Res. 788, 305–310.
J.M., Klein, D.F., Liebowitz, M.R., 1991. Effects of chronic fluoxetine
Hu, G., Duffy, P., Swanson, C., Behnam Ghasemzadeh, M., Kalivas,
treatment on behavioural and neuroendocrine responses to meta-
P.W., 1999. The regulation of dopamine transmission by metabotropic
chloro-phenylpiperazine in obsessive–compulsive disorder. Psychiatry
glutamate receptors. J. Pharmacol. Exp. Ther. 289, 412–416.
Res. 36, 1–17.
Huang, Y.Y., Kandel, E.R., 1996. Modulation of both the early and
Hollmann, M., Heinemann, S., 1994. Cloned glutamate receptors. Annu.
the late phase of mossy fiber LTP by the activation of ␤-adrenergic
Rev. Neurosci. 17, 31–108.
receptors. Neuron 16, 611–617.
Hollon, T.R., Bek, M.J., Lachowicz, J.E., Ariano, M.A., Mezey, E.,
Ramachandran, R., Wersinger, S.R., Soares-da-Silva, P., Liu, Z.F., Huang, C.C., Lin, C.H., Gean, P.W., 1998. Potentiation of N-methyl-
Grinberg, A., Drago, J., Young, W.S., Westphal, H., Jose, P.A., Sibley, d-aspartate currents by isoproterenol in the acutely dissociated rat
D.R., 2002. Mice lacking D5 dopamine receptors have increased amygdala neurons. Neurosci. Lett. 253, 9–12.
sympathetic tone and are hypertensive. J. Neurosci. 22, 10801–10810. Huidobro-Toro, J.P., Valenzuela, C.F., Harris, R.A., 1996. Modulation of
Holmes, A., 2001. Targeted gene mutation approaches to the study of GABAA receptor function by G protein-coupled 5-HT2C receptors.
anxiety-like behavior in mice. Neurosci. Biobehav. Rev. 25, 261–273. Neuropharmacology 35, 1355–1363.
Holmes, P., Blanchard, D., Blanchard, R., Brady, L., Crawley, I., 1995. Hurley, K.M., Herbert, H., Moga, M.M., Saper, C.B., 1991. Efferent
Chronic social stress increases levels of preprogalanin mRNA in the projections of the infralimbic cortex of the rat. J. Comp. Neurol. 308,
rat locus coeruleus. Pharmacol. Biochem. Behav. 50, 655–660. 249–276.
Holmes, A., Hollon, T.R., Gleason, T.C., Liu, Z., Dreiling, J., Sibley, Hyytia, P., Koob, G.F., 1995. GABAA receptor antagonism in the
D.R., Crawley, J.N., 2001. Behavioral characterization of dopamine extended amygdala decreases ethanol self-administration in rats. Eur.
D5 receptor null mutant mice. Behav. Neurosci. 115, 1129–1144. J. Pharmacol. 283, 151–159.
Holmes, A., Yang, R.J., Crawley, J.N., 2002a. Evaluation of an anxiety- Ichikawa, J., Dai, J., Meltzer, H.Y., 2000. Acetylcholinesterase inhibitors
related phenotype in galanin overexpressing transgenic mice. J. Mol. are neither necessary nor desirable for microdialysis studies of brain
Neurosci. 18, 151–165. acetylcholine. Curr. Sep. 19, 37–44.
204 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Ichikawa, J., Dai, J., O’Laughlin, I.A., Fowler, W.L., Meltzer, H.Y., Izzo, E., Auta, J., Impagnatiello, F., Pesold, C., Guidotti, A., Costa, E.,
2002. Atypical, but not typical, antipsychotic drugs increase cortical 2001. Glutamic acid decarboxylase and glutamate receptor changes
acetylcholine release without an effect in the nucleus accumbens or during tolerance and dependence to benzodiazepines. Proc. Natl.
striatum. Neuropsychopharmacology 26, 325–339. Acad. Sci. U.S.A. 98, 3483–3488.
Ichiki, T., Labosky, P.A., Shiota, C., Okuyama, S., Imagawa, Y., Fogo, Jackson, D.M., Westlind-Danielsson, A., 1994. Dopamine receptors:
A., Niimura, F., Hogan, B.L., Inagami, T., 1995. Effects on blood molecular biology, biochemistry and behavioural aspects. Pharmacol.
pressure and exploratory behaviour of mice lacking angiotensin II Ther. 64, 291–370.
type-2 receptor. Nature 377, 748–750. Jackson, A., Mead, A.N., Stephens, D.N., 2000. Behavioural effects of ␣-
Ihalainen, J.A., Tanila, H., 2002. In vivo regulation of dopamine and amino-3-hydroxy-5-methyl-4-isoxazolepropionate-receptor antagonists
noradrenaline release by ␣2A -adrenoceptors in the mouse prefrontal and their relevance to substance abuse. Pharmacol. Ther. 88, 59–76.
cortex. Eur. J. Neurosci. 15, 1789–1794. Jacob, C.A., Cabral, A.H.C.L., Almeida, L.P., Magierek, V., Ramos,
Illes, P., Finta, E.P., Nieber, K., 1993. Neuropeptide Y potentiates via Y2 - P.L., Zanoveli, J.M., Landeira-Fernandez, J., Zangrossi, H., Nogueira,
receptors the inhibitory effect of noradrenaline in rat locus coeruleus R.L., 2002. Chronic imipramine enhances 5-HT1A and 5-HT2
neurones. Naunyn Schmiedebergs Arch. Pharmacol. 348, 546–548. receptor-mediated inhibition of panic-like behavior in the rat dorsal
Imaizumi, M., Onodera, K., 1993. The behavioural and biochemical periaqueductal gray. Pharmacol. Biochem. Behav. 72, 761–766.
effects of thioperamide, a histamine H3 , receptor antagonist, in a Jahn, H., Montkowski, A., Knaudt, K., Ströhle, A., Ströhle, A., Holsboer,
light/dark test measuring anxiety in mice. Life Sci. 53, 1675–1683. F., Wiedemann, D., 1997. CNP exerts anxiogenic effects in rats:
Impagnatiello, F., Pesold, C., Longone, P., Caruncho, H., Fritschy, J.M., evidence that CRH may be involved. Pharmacopsychiatry 30, 184.
Costa, E., Guidotti, A., 1996. Modifications of ␥-aminobutyric acidA Jahn, H., Montkowski, A., Knaudt, K., Ströhle, A., Kiefer, F., Schick,
receptor subunit expression in rat neocortex during tolerance to M., Wiedemann, K., 2001. ␣-Helical-corticotropin-releasing hormone
diazepam. Mol. Pharmacol. 49, 822–831. reverses anxiogenic effects of C-type natriuretic peptide in rats. Brain
Impagnatiello, F., Bastia, E., Ongini, E., Monopoli, A., 2000. Adenosine Res. 893, 21–28.
receptors in neurological disorders. Emerg. Ther. Target 4, 635–664. Jain, N., Kemp, N., Adeyemo, O., Buchanan, P., Stone, T.W., 1995.
Imura, H., Nakao, K., Itoh, H., 1992. The natriuretic peptide system in Anxiolytic activity of adenosine receptor activation in mice. Br. J.
the brain: implications in the central control of cardiovascular and Pharmacol. 116, 2127–2133.
neuroendocrine functions. Front. Neuroendocrinol. 13, 217–249.
Jakab, R.L., Goldman-Rakic, P.S., 2000. Segregation of serotonin 5-HT2A
Inglis, F.M., Moghaddam, B., 1999. Dopaminergic innervation of the
and 5-HT3 receptors in inhibitory circuits of the primate cerebral
amygdala is highly responsive to stress. J. Neurochem. 72, 1088–1094.
cortex. J. Comp. Neurol. 417, 337–348.
Inui, A., Okita, M., Nakajima, M., Momose, K., Ueno, N., Teranishi,
Jakeman, L.B., To, Z.P., Eglen, R.M., Wong, E.H., Bonhaus, D.W., 1994.
A., Miura, M., Hirosue, Y., Sano, K., Sato, M., Waranabe, M., Sakai,
Quantitative autoradiography of 5-HT4 receptors in brains of three
T., Watanabe, T., Ishida, K., Silver, J., Baba, S., Kasuga, M., 1998.
species using two structurally distinct radioligands, [3 H]GR113808
Anxiety-like behavior in transgenic mice with brain expression of
and [3 H]BIMU-1. Neuropharmacology 33, 1027–1038.
neuropeptide Y. Proc. Assoc. Am. Phys. 110, 171–182.
Jayakumar, A.R., Sujatha, R., Paul, V., Asokan, C., Govindasamy, S.,
Irvine, E.E., Bagnalasta, M., Marcon, C., Motta, C., Tessari, M.,
Jayakumar, R., 1999. Role of nitric oxide on GABA, glutamic acid,
File, S.E., Cjiamulera, C., 2001. Nicotine self-administration and
activities of GABA-T and GAD in rat brain cerebral cortex. Brain
withdrawal: modulation of anxiety in the social interaction test in
Res. 837, 229–235.
rats. Psychopharmacology 153, 315–320.
Jedema, H.P., Moghaddam, B., 1996. Characterization of excitatory
Ishida, Y., Hashiguchi, H., Takeda, R., Ishizuka, Y., Mitsuyama, Y.,
amino acid modulation of dopamine release in the prefrontal cortex
Kannan, H., Nishimori, T., Nakahara, D., 2002. Conditioned-fear stress
of conscious rats. J. Neurochem. 66, 1448–1453.
increases Fos expression in monoaminergic and GABAergic neurons
of the locus coeruleus and dorsal raphe nuclei. Synapse 45, 46–51. Jenck, F., Broekkamp, C.L.E., Van Delft, A.M.L., 1989. Opposite control
Ishida-Tokuda, K., Ohno, Y., Sakamoto, H., Ishibashi, T., Wakabayashi, J., mediated by central 5-HT1A and non-5-HT1A (5-HT1B or 5-HT1C )
Tojima, R., Morita, T., Nakamura, M., 1996. Evaluation of perospirone receptors on periaqueductal gray aversion. Eur. J. Pharmacol. 161,
(SM-9018), a novel serotonin-2 and dopamine-2 receptor antagonist, 219–221.
and other antipsychotics in the conditioned fear stress-induced freezing Jenck, F., Broekkamp, C.L.E., Van Delft, A.M.L., 1990. 5-HT1C receptors
behavior model in rats. Jpn. J. Pharmacol. 72, 119–126. in the serotonergic control of periaqueductal gray induced aversion in
Isogawa, K., Akiyoshi, J., Hikichi, T., Yamamoto, Y., Tsutsumi, T., rats. Psychopharmacology 100, 372–376.
Nagayama, H., 2000. Effect of corticotropin releasing factor receptor Jenck, F., Moreau, J.-L., Mutel, V., Martin, J.R., 1994. Brain 5-HT1C
1 antagonist on extracellular norepinephrine, dopamine and serotonin receptors and antidepressants. Prog. Neuropsychopharmacol. Biol.
in the hippocampus and prefrontal cortex in vivo. Neuropeptides 34, Psychiatry 18, 563–574.
234–239. Jenck, F., Moreau, J.-L., Martin, J.R., Kilpatrick, G., Reinscheid, R.K.,
Ito, C., Shen, H., Toyota, H., Kubota, Y., Sakurai, E., Watanabe, T., Monsma, F.J., Nothacker, H.-P., Civelli, O., 1997. orphaninFQ acts as
Sato, M., 1999. Effects of the acute and chronic restraint stresses on an anxiolytic to attenuate behavioral responses to stress. Proc. Natl.
the central histaminergic neuron system of Fischer rat. Neurosci. Lett. Acad. Sci. U.S.A. 94, 14854–14858.
262, 143–145. Jenck, F., Moreau, J.-L., Berendsen, H.H.G., Boes, M., Broekkamp,
Izumi, J., Washizuka, M., Hayashi-Kuwabara, Y., Yoshinaga, K., Tanaka, C.L.E., Martin, J.R., Wichmann, J., Van Delft, A.M.L., 1998.
Y., Ikeda, Y., Kiuchi, Y., Oguchi, K., 1996. An attenuated ␣1 - Antiaversive effects of 5-HT2C receptor agonists and fluoxetine in a
potentiation of ␤-adrenoceptor-stimulated cyclic AMP formation after model of panic-like anxiety in rats. Eur. Neuropsychopharmacol. 8,
repeated saline injections in Fischer 344 strain rats. Life Sci. 59, 33–42. 161–168.
Izumi, J., Washizuka, M., Hayashi-Kuwabara, Y., Yoshinaga, K., Tanaka, Jenck, F., Martin, J.R., Moreau, J.L., 1999. The 5-HT1A receptor agonist
Y., Ikeda, Y., Kiuchi, Y., Oguchi, K., 1997. Protective effect of flesinoxan increases aversion in a model of panic-like anxiety in rats.
citalopram against the attenuation of the alpha-1-potentiation of cAMP J. Psychopharmacol. 13, 166–170.
formation in Fischer 344 strain rats. Behav. Brain Res. 83, 209–212. Jenck, F., Wichmann, J., Dautzenberg, F.M., Moreau, J.-L., Ouagazzal,
Izumi, T., Suzuki, K., Inoue, T., Li, X.B., Maki, Y., Muraki, I., Kitaichi, A.M., Martin, J.R., Lundstrom, K., Cesura, A.M., Poli, S.M., Roever,
Y., Hashimoto, S., Koyama, T., 2002. Long-lasting change in 5-HT2A S., Kolczewski, S., Adam, G., Kilpatrick, G., 2000. A synthetic agonist
receptor-mediated behavior in rats after a single footshock. Eur. J. at the orphaninFQ/nociceptin receptor ORL1: anxiolytic profile in the
Pharmacol. 452, 199–204. rat. Proc. Natl. Acad. Sci. U.S.A. 97, 4938–4943.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 205

Jenkins, A., Andreasen, A., Trudell, J.R., Harrison, N.L., 2002. Joseph, J.D., Wang, Y.M., Miles, P.R., Budygin, E.A., Picetti, R.,
Tryptophan scanning mutagenesis in TM4 of the GABAA receptor ␣1 Gainetdinov, R.R., Caron, M.G., Wightman, R.M., 2002. Dopamine
subunit: implications for modulation by inhaled anesthetics and ion autoreceptor regulation of release and uptake in mouse brain slices in
channel structure. Neuropharmacology 43, 669–678. the absence of D3 receptors. Neuroscience 112, 39–49.
Jensen, R.T., Coy, D.H., 1991. Progress in the development of potent Joyce, J.N., 2001. Dopamine D3 receptor as a therapeutic target for
bombesin receptor antagonists. Trends Pharmacol. Sci. 12, 13–19. antipsychotic and antiparkinsoninan drugs. Pharmacol. Ther. 90,
Jessa, M., Nazar, M., Plaznik, A., 1995. Anxiolytic-like action of 231–259.
intra-hippocampally administered NMDA antagonists in rats. Pol. J. Jung, M., Michaud, J.C., Steinberg, R., Barnouin, M.C., Hayar, A.,
Pharmacol. 47, 81–84. Mons, G., Souilhac, J., Edmonds-Alt, X., Soubrié, P., Le Fur, G.,
Jessa, M., Nazar, M., Bidzinski, A., Plaznik, A., 1996a. The effects of 1996. Electrophysiological, behavioural and biochemical evidence for
repeated administration of diazepam, MK-801 and CGP37849 on rat activation of brain noradrenergic systems following neurokinin NK3
behavior in two models of anxiety. Eur. Neuropsychopharmacol. 6, receptor stimulation. Neuroscience 74, 403–414.
55–61. Jung, M., Lal, H., Gatch, M.B., 2002. The discriminative stimulus effects
Jessa, M., Nazar, M., Plaznik, A., 1996b. Effect of intra-accumbens of pentylenetetrazol as a model of anxiety: recent developments.
blockade of NMDA receptors in two models of anxiety, in rats. Neurosci. Biobehav. Rev. 26, 429–439.
Neurosci. Res. Commun. 19, 19–25. Kable, J.W., Murrin, L.C., Bylund, D.B., 2000. In vivo gene modification
Jin, S.L., Han, V.K., Simmons, J.G., Towle, A.C., Lauder, J.M., Lund, elucidates subtype-specific functions of alpha(2) -adrenergic receptors.
P.K., 1988. Distribution of glucagon-like peptide-1 (GLP-1), glucagon, J. Pharmacol. Exp. Ther. 293, 1–7.
and glicentin in the rat brain: an immunocytochemical study. J. Kachaturian, H., Lewis, M.E., Schafer, M.K.-H., Watson, S.J., 1985.
Comp. Neurol. 271, 519–532. Anatomy of the CNS opioid systems. Trends Neurosci. 8, 111–119.
Jin, L.-Q., Wang, H.-Y., Friedman, E., 2001. Stimulated D1 dopamine Kaczorowski, G.J., Garcia, M.L., 1999. Pharmacology of voltage-gated
receptors couple to multiple G␣ proteins in different brain regions. J. and calcium-activated potassium channels. Curr. Opin. Chem. Biol. 3,
Neurochem. 78, 981–990. 448–458.
Joëls, M., Vreugdenhil, E., 1998. Corticosteroids in the brain. Mol. Kagaya, A., Katagiri, H., Tawara, Y., Kugaya, A., Inagaki, M., Miyoshi,
Neurobiol. 17, 87–108. I., Jitsuiki, H., Kozuru, T., Kouhata, S., Uchitomi, Y., Morinobu, S.,
Johansson, B., Halldner, L., Dunwiddie, T.V., Masino, S.A., Poelchen, Yamawaki, S., 2001. Effect of sub-chronic treatment with duloxetine
W., Giménez-Llort, L., Escorihuela, R.M., Fernández-Teruel, A., on serotonin-2A receptor function in vivo and in vitro. Biogenic
Wiesenfeld-Hallin, Z., Xu, X.-J., Hårdemark, A., Betsholtz, C., Amines 16, 541–554.
Herlenius, E., Fredholm, B.B., 2001. Hyperalgesia, anxiety, and Kahaya, T., Yonaga, M., Furuya, Y., Hashimoto, T., Kuroki, J., Nishizawa,
decreased hypoxic neuroprotection in mice lacking the adenosine A1 Y., 1996. Dopamine D3 agonists disrupt social behavior in rats. Brain
receptor. Proc. Natl. Acad. Sci. U.S.A. 98, 9407–9412. Res. 721, 229–232.
Johnston, A.L., File, S.E., 1988. Yohimbine’s anxiogenic action: evidence Kahn, R.S., Wetzler, S., 1991. m-Chlorophenylpiperazine as a probe of
for noradrenergic and dopaminergic sites. Pharmacol. Biochem. Behav. serotonin function. Biol. Psychiatry 30, 1139–1166.
32, 151–156. Kaiser, F.C., Palmer, G.C., Wallace, A.V., Carr, R.D., Fraser-Rae,
Johnston, A.L., File, S.E., 1991. Sex differences in animal tests of L., Hallam, C., 1992. Antianxiety properties of the angiotensin
anxiety. Physiol. Behav. 49, 245–250. II antagonist, DUP 753, in the rat using the elevated plus-maze.
Jolas, T., Schreiber, R., Laporte, A.M., Chastanet, M., De Vry, J., Glaser, NeuroReport 3, 922–924.
T., Adrien, J., Hamon, M., 1995. Are postsynaptic 5-HT1A receptors Kalin, N.H., Shelton, S.E., Davidson, R.J., Kelley, A.E., 2001. The primate
involved in the anxiolytic effects of 5-HT1A receptor agonists and in amygdala mediates acute fear but not the behavioral and physiological
their inhibitory effects on the firing of serotonergic neurons in the rat. components of anxious temperament. J. Neurosci. 21, 2067–2074.
J. Pharmacol. Exp. Ther. 272, 920–929. Kalivas, P.W., Duffy, P., 1995. Selective activation of dopamine
Jonas, J.M., Cohon, M.S., 1993. A comparison of the safety and efficacy transmission in the shell of the nucleus accumbens by stress. Brain
of alprazolam versus other agents in the treatment of anxiety, panic and Res. 675, 325–328.
depression: a review of the literature. J. Clin. Psychiatry 54, 25–45. Kamei, H., Kameyama, T., Nabeshima, T., 1994. (+)-SKF-10,047 reverses
Jones, R.S.G., Mondadori, C., Olpe, H.-R., 1985. Neuronal sensitivity to stress-induced motor suppression: interaction with dopaminergic
substance P is increased after repeated treatment with tranylcypromine, system. Eur. J. Pharmacol. 260, 39–46.
carbamazepine or oxaprotiline, but decreased after repeated Kamei, H., Kameyama, T., Nabeshima, T., 1996. (+)-SKF-10,047 and
electroconvulsive shock. Neuropharmacology 24, 627–633. dextromethorphan ameliorate conditioned fear stress through the
Joshi, B.N., Troiani, M.E., Milin, J., Nurnburger, F., Reiter, R.J., 1986. activation of phenytoin-regulated ␴1 sites. Eur. J. Pharmacol. 299,
Adrenal-mediated depression of N-acetyl-transferase activity and 21–28.
melatonin levels in the rat pineal gland. Life Sci. 38, 1573–1580. Kamei, H., Noda, Y., Kameyama, T., Nabeshima, T., 1997. Role of (+)-
Jones, B.J., Costall, B., Domeney, A.M., Kelly, M.E., Naylor, R.J., Oakley, SKF-10,047-sensitive sub-population of ␴1 receptors in amelioration
N.R., Tyers, M.B., 1988. The potential anxiolytic activity of GR38032F, of conditioned fear stress in rats: association with mesolimbic
a 5-HT3 -receptor antagonist. Br. J. Pharmacol. 93, 985–993. dopaminergic systems. Eur. J. Pharmacol. 319, 165–172.
Jones, N., Duxon, M.S., King, S.M., 2002a. 5-HT2C receptor mediation Kamikawa, H., Hori, T., Nakane, H., Aou, S., Tashiro, N., 1998. L1␤
of unconditioned escape behaviour in the unstable elevated exposed increases norepinephrine level in rat frontal cortex: involvement of
plus maze. Psychopharmacology 164, 214–220. prostanoids, NO, interleukin1B and glutamate. Am. J. Physiol. 275,
Jones, K., King, S.M., Duxon, M.S., 2002b. Further evidence for the R803–R810.
predictive validity of the unstable elevated exposed plus-maze, a Kamisaki, Y., Hamahashi, T., Hamada, T., Maeda, K., Itoh, T., 1992.
behavioural model of extreme anxiety in rats: differential effects of Presynaptic inhibition by clonidine of neurotransmitter amino acid
fluoxetine and chlordiazepoxide. Behav. Pharmacol. 13, 525–535. release in various brain regions. Eur. J. Pharmacol. 217, 57–63.
Jorgensen, H., Knigge, U., Kjaer, A., Vadsholt, T., Warberg, J., 1998. Kang, I., Thompson, M.L., Heller, J., Miller, L.G., 1991. Persistent
Serotonergic involvement in stress-induced ACTH release. Brain Res. elevation in GABAA receptor subunit mRNAs following social stress.
811, 10–20. Brain Res. Bull. 26, 809–812.
Jorgensen, H., Kjaer, A., Warberg, J., Knigge, U., 2001. Differential Kang, W., Wilson, S.P., Wilson, M.A., 2000. Overexpression of
effect of serotonin 5-HT1A receptor antagonists on the secretion of proenkephalin in the amygdala potentiates the anxiolytic effects of
corticotrophin and prolactin. Neuroendocrinology 73, 322–333. benzodiazepines. Neuropsychopharmacology 22, 77–88.
206 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Kaplan, D.R., Miller, F.D., 2000. Neurotrophin signal transduction in the Kask, A., Harro, J., Von Hörsten, S., Redrobe, J.P., Dumont, Y.,
nervous system. Curr. Opin. Neurobiol. 10, 381–391. Quirion, R., 2002. The neurocircuitry and receptor subtypes mediating
Karasinska, J.M., George, S.R., El-Ghundi, M., Fletcher, P.J., O’Dowd, anxiolytic-like efects of neuropeptide Y. Neurosci. Biobehav. Rev. 26,
B.F., 2000. Modification of dopamine D1 receptor knock-out 259–283.
phenotype in mice lacking both dopamine D1 and D3 receptors. Eur. Kasper, S., Resinger, E., 2001. Panic disorder: the place of
J. Pharmacol. 399, 171–181. benzodiazepines and selective serotonin reuptake inhibitors. Eur.
Karcz-Kubicka, M., Jessa, M., Nazar, M., Plaznik, A., Hartmann, S., Neuropsychopharmacol. 11, 307–321.
Parsons, C.G., Danysz, W., 1997. Anxiolytic activity of glycine-B Kasper, S., Blagden, M., Seghers, S., Veerman, A., Volz, H.P., Geniaux,
antagonists and partial agonists—no relation to intrinsic activity in A., Strub, N., Marchand, A., Maisonobe, P., 2002. A placebo-
the patch clamp. Neuropharmacology 36, 1355–1367. controlled study of pregabalin and venlafaxine treatment of GAD.
Karkanias, G.B., Etgen, A.M., 1993. Estradiol attenuates ␣2 -adrenoceptor- Eur. Neuropsychopharmacol. 12, S341–S342.
mediated inhibition of hypothalamic norepinephrine release. J. Katagiri, H., Kagaya, A., Nakae, S., Morinobu, S., Yamawaki, S., 2001.
Neurosci. 13, 3448–3455. Modulation of serotonin2A receptor function in rats after repeated
Karkanias, G.B., Li, C.S., Etgen, A.M., 1997. Estradiol reduction of ␣2 - treatment with dexamethasone and l-type calcium channel antagonist
adrenoceptor binding in female rat cortex is correlated with decreases nimodipine. Prog. Neuropsychopharmacol. Biol. Psychiatry 25, 1269–
in ␣2A/D adrenoceptor messenger RNA. Neuroscience 81, 593–597. 1281.
Karolyi, I.J., Burrows, H.L., Ramesh, T.M., Nakajima, M., Lesh, J.S., Kataoka, Y., Shibata, K., Miyazaki, A., Inoue, Y., Tominaga, K.,
Seong, E., Camper, S.A., Seasholtz, A.F., 1999. Altered anxiety Koizumi, S., Ueki, S., Niwa, M., 1991. Involvement of the dorsal
and weight gain in corticotropin-releasing hormone-binding protein- hippocampus in mediation of the antianxiety action of tandospirone, a
deficient mice. Proc. Natl. Acad. Sci. U.S.A. 96, 11595–11600. 5-hydroxytryptamine1A agonistic anxiolytic. Neuropharmacology 30,
Karreman, M., Moghaddam, B., 1996. Effect of a pharmacological stressor 475–480.
on glutamate efflux in the prefrontal cortex. Brain Res. 716, 180–182. Kathmann, M., Bauer, U., Schlicker, E., Göthert, M., 1999. Cannabinoid
Karschin, C., Dissmann, E., Stuhmer, W., Karschin, A., 1996. IRK(1– CB1 receptor-mediated inhibition of NMDA- and kainate-stimulated
3) and GIRK(1–4) inwardly rectifying K+ channel mRNAs are noradrenaline and dopamine release in the brain. Naunyn
differentially expressed in the adult rat brain. J. Neurosci. 16, 3559– Schmiedebergs Arch. Pharmacol. 359, 466–470.
3570. Katona, I., Sperlagh, B., Ski, A., Käfalvi, A., Vizi, E.S., Mackie, K.,
Karst, H., Karten, Y.J., Reichardt, H.M., De Kloet, E.R., Schütz, G., Joels, Freund, T.F., 1999. Presynaptically located CB1 cannabinoid receptors
M., 2000. Corticosteroid actions in hippocampus require DNA binding regulate GABA release from axon terminals of specific hippocampal
of glucocorticoid receptor homodimers. Nat. Neurosci. 3, 977–978. interneurones. J. Neurosci. 19, 4544–4558.
Karten, Y.J.G., Nair, S.M., Van Essen, L., Sibug, R., Joëls, M., 1999. Katona, I., Rancz, E.A., Acsady, L., Ledent, C., Mackie, K., Hajos, N.,
Long-term exposure to high corticosterone levels attenuates serotonin Freund, T.F., 2001. Distribution of CB1 cannabinoid receptors in the
responses in rat hippocampal CA1 neurons. Proc. Natl. Acad. Sci. amygdala and their role in the control of GABAergic transmission. J.
U.S.A. 96, 13456–13461. Neurosci. 21, 9506–9518.
Kaschka, W., Feistel, H., Ebert, D., 1995. Reduced benzodiazepine Katz, R.J., Landau, P.S., Lott, M., Bystritsky, A., Diamond, B., Hoehn-
receptor binding in panic disorders measured by iomazenil SPECT. J. Saric, R., Rosenthal, M., Weise, C., 1993. Serotonergic (5-HT2 )
Psychiatry Res. 29, 427–434. mediation of anxiety—therapeutic effect of serazepine in generalized
Kash, S.F., Tecott, L.H., Hodge, C., Baekkeskov, S., 1999. Increased anxiety disorder. Biol. Psychiatry 34, 41–44.
anxiety and altered responses to anxiolytics in mice deficient in the Katz, M.M., Halbreich, U.M., Bowden, C.L., Frazer, A., Pinder, R.M.,
65-kDa isoform of glutamic acid decarboxylase. Proc. Natl. Acad. Rush, A.J., Wheatley, D.P., Lebowitz, B.D., 2002. Enhancing the
Sci. U.S.A. 96, 1698–1703. technology of clinical trials and the trials model to evaluate newly
Kask, A., Rägo, L., Harro, J., 1997. Alpha-helical CRF(9–41) prevents developed, targeted antidepressants. Neuropsychopharmacology 27,
anxiogenic-like effect of NPY Y1 receptor antagonist BIBP3226 in 319–328.
rats. NeuroReport 8, 3645–3647. Kaupmann, K., Huggel, K., Heid, J., Flor, P.J., Bischoff, S., Mickel,
Kask, A., Rägo, L., Harro, J., 1998a. Anxiolytic-like effect of neuropeptide S.J., McMaster, G., Angst, C., Bittiger, H., Froesti, W., Bettler, B.,
Y (NPY) and NPY13−−36 microinjected into the vicinity of locus 1997. Expression cloning of GABAB receptors uncovers similarity to
coeruleus in rats. Brain Res. 788, 345–348. metabotropic glutamate receptors. Nature 386, 239–246.
Kask, A., Rägo, L., Harro, J., 1998b. NPY Y1 receptors in the dorsal Kawahara, H., Kawahara, Y., Westerink, B.H.C., 2000. The role of
periaqueductal gray matter regulate anxiety in the social interaction afferents to the locus coeruleus in the handling stress-induced increase
test. NeuroReport 9, 2713–2716. in the release of noradrenaline in the medial prefrontal cortex: a
Kask, A., Kivastik, T., Rägo, L., Harro, J., 1999. Neuropeptide Y Y1 dual-probe microdialysis study in the rat brain. Eur. J. Pharmacol.
receptor antagonist BIBP3226 produces conditioned place aversion in 387, 279–286.
rats. Prog. Neuropsychopharmacol. Biol. Psychiatry 23, 705–711. Kawai, H., Zago, W., Berg, D.K., 2002. Nicotinic ␣7 receptor clusters
Kask, A., Eller, M., Oreland, L., Harro, J., 2000. Neuropeptide Y on hippocampal GABAergic neurons: regulation by synaptic activity
attenuates the effects of locus coeruleus denervation by DSP-4 and neurotrophins. J. Neurosci. 22, 7903–7912.
treatment on social behaviour in the rat. Neuropeptides 34, 58–61. Keck, M.E., Welt, T., Wigger, A., Renner, U., Engelmann, M., Holsboer,
Kask, A., Nguyen, H.P., Pabst, R., Von Hörsten, S., 2001a. Neuropeptide F., Landgraf, R., 2001. The anxiolytic effect of the CRH1 receptor
Y Y1 receptor-mediated anxiolysis in the dorsocaudal lateral septum: antagonist R121919 depends on innate emotionality in rats. Eur. J.
functional antagonism of corticotropin-releasing hormone-induced Neurosci. 13, 373–380.
anxiety. Neuroscience 104, 799–806. Keck, M.E., Wigger, A., Welt, T., Müller, M.B., Gesing, A., Reul,
Kask, A., Pabst, N.R., Von Horstin, S., 2001b. Neuropeptide Y Y1 J.M.H.M., Holsboer, F., Landgraf, R., Neumann, I.D., 2002.
receptor-mediated anxiolysis in the dorsocaudal lateral septum: Vasopressin mediates the response of the combined dexametha-
functional antagonism of corticotrophin-releasing hormone-induced sone/CRH test in hyper-anxious rats: implications for pathogenesis of
anxiety. Neuroscience 104, 799–806. affective disorders. Neuropsychopharmacology 26, 94–105.
Kask, A., Vasar, E., Heidmets, L.-T., Allikmets, L., Wikberg, J.E.S., Kehne, J.H., McCloskey, C., Baron, B.M., Chi, E.M., Harrison, B.L.,
2001c. Neuropeptide Y Y5 receptor antagonist CGP71683A: the Whitten, J.P., Palfreyman, M.G., 1991. NMDA receptor complex anta-
effects on food intake and anxiety-related behavior in the rat. Eur. J. gonists have potential anxiolytic effects as measured with separation-
Pharmacol. 414, 215–224. induced ultrasonic vocalization. Eur. J. Pharmacol. 193, 283–292.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 207

Kehr, J., Yoshitake, T., Wang, F.-H., Razani, H., Gimenez-Llort, L., Kennett, G.A., Bright, F., Trail, B., Blackburn, T.P., Sanger, G.J., 1997b.
Jansson, A., Yamaguchi, M., Ogren, S.O., 2002a. Galanin is a potent Anxiolytic-like actions of the selective 5-HT4 receptor antagonists SB
in vivo modulator of mesencephalic serotonergic neurotransmission. 204070A and SB 207266A in rats. Neuropharmacology 36, 707–712.
Neuropsychopharmacology 27, 341–356. Kennett, G.A., Trail, B., Bright, F., 1998. Anxiolytic-like actions of BW
Kehr, J., Yoshitake, T., Wang, F.-H., Razani, H., Madjid, N., 2002b. 723C86 in the rat Vogel Conflict Test are 5-HT2B receptor mediated.
Bidirectional interaction between galanin and 5-HT1A receptors in the Neuropharmacology 37, 1603–1610.
regulation of hippocampal 5-HT release. Eur. Neuropsychopharmacol. Kennett, G.A., Lightowler, S., Trail, B., Bright, F., Bromidge, S., 2000.
12, S203. Effects of RO 60 0175, a 5-HT2C receptor agonist, in three animal
Keim, S.R., Shekhar, A., 1996. The effects of GABAA receptor blockade models of anxiety. Eur. J. Pharmacol. 387, 197–204.
in the dorsomedial hypothalamic nucleus on corticotrophin (ACTH) Kenny, P.J., Cheeta, S., File, S.E., 2000. Anxiogenic effects of nicotine
and corticosterone secretion in male rats. Brain Res. 739, 46–51. in the dorsal hippocampus are mediated by 5-HT1A and not by
Keim, S.R., Sajdyk, T.J., Gehlert, D.R., Shekhar, A., 2002. The anxiogenic muscarinic M1 receptors. Neuropharmacology 39, 300–307.
effect of urocortin microinjections into the bed nucleus of the stria Kent, P., Anisman, H., Merali, Z., 1998. Are bombesin-like peptides
terminals (BNST). Soc. Neurosci. Abstr. 398.5. involved in the mediation of the stress response? Life Sci. 62, 103–114.
Kellendonk, C., Gass, P., Kretz, O., Schütz, G., Tronche, F., 2002. Kent, J.M., Coplan, J.D., Lombardo, I., Hwang, D.-R., Huang, Y.,
Corticosteroid receptors in the brain: gene targeting studies. Brain Mawlwi, O., Van Heertum, R.L., Slifstein, M., Abi-Dargham, A.,
Res. Bull. 57, 73–83. Gorman, J.M., Laruelle, M., 2002a. Occupancy of brain serotonin
Keller, M.B., 2002. The long-term clinical course of generalized anxiety transporters during treatment with paroxetine in patients with social
disorder. J. Clin. Psychiatry 63, 77–116. phobia: a positron emission tomography study with [11 C]McN 5652.
Kelley, S.P., Bratt, A.M., Hodge, C.W., 2003. Targeted gene deletion Psychopharmacology 164, 341–348.
of the 5-HT3A receptor subunit produces an anxiolytic phenotype in Kent, J.M., Mathew, S.J., Gorman, J.M., 2002b. Molecular targets in the
mice. Eur. J. Pharmacol. 461, 19–25. treatment of anxiety. Biol. Psychiatry 52, 1008–1030.
Kellner, M., Diehl, I., Knaudt, K., Schüle, C., Jahn, H., Wiedemann, Keogh, E., Cochrane, M., 2002. Anxiety sensitivity, cognitive biases and
K., 1997. C-type natriuretic peptide exerts opposite effects to strial pain. J. Pain 3, 320–329.
natriuretic hormone on the corticotropin-releasing hormone-induced Kerr, D.I.B., Ong, J., Made Puspawati, N., Prager, R.H., 2002.
release of hormones in normal man. Eur. J. Endocrinol. 136, 388–393. Arylalkylamines are a novel class of positive allosteric modulators at
Kellner, M., Jahn, H., Naber, D., Wiedemann, K., 2002. Natriuretic GABAB receptors in rat neocortex. Eur. J. Pharmacol. 451, 69–77.
peptides modulate the psychometric and endocrine effects of Kessler, R.C., Wittchen, H.-U., 2002. Patterns and correlates of generalized
cholecystokinin tetrapeptide in man. Eur. Psychopharmacol. 17, S32.2. anxiety disorder in community samples. J. Clin. Psychiatry 63, 4–10.
Kelly, M.J., Wagner, E.J., 1999. Estrogen modulation of G-protein- Ketelaars, C.E.J., Bollen, E.L., Rigter, H., Bruinvels, J., 1988. GABA-B
coupled receptors. Trends Endocrinol. Metab. 10, 369–374. receptor activation and conflict behaviour. Life Sci. 42, 933–942.
Kelly, M.A., Rubinstein, M., Phillips, T.J., Lessov, C.N., Burkhart-Kazch, Kew, J.N.C., Koester, A., Moreau, J.-L., Jenck, F., Ouagazzal, A.-
S., Zhang, G., Bunzow, J.R., Fang, Y., Gerhardt, G.A., Grandy, M., Mutel, V., Richards, J.G., Trube, G., Fischer, G., Montkowski,
D.K., Löw, M.J., 1998. Locomotor activity in D2 dopamine receptor- A., Hundt, W., Reinscheid, R.K., Pauly-Evers, M., Kemp, J.A.,
deficient mice is determined by gene dosage, genetic background, and Bluethmann, H., 2000. Functional consequences of reduction in
developmental adaptations. J. Neurosci. 18, 3470–3479. NMDA receptor glycine affinity in mice carrying targeted point
Kenakin, T., 1995. Agonist-receptor efficacy. II. Agonist trafficking of mutations in the glycine binding site. J. Neurosci. 20, 4037–4049.
receptor signals. Trends Pharmacol. Sci. 16, 232–238. Khan, A., Khan, S., Brown, W.A., 2002. Are placebo controls necessary
Kenakin, T., 2002. Efficacy at G-protein-coupled receptors. Nat. Rev. to test new antidepressants and anxiolytics? Int. J. Neuropsycho-
Drug Discov. 11, 103–110. pharmacol. 5, 193–197.
Kenis, G., Maes, M., 2002. Effects of antidepressants on the production Khoshbouei, H., Cecchi, M., Dove, S., Javors, M., Morilak, D.A.,
of cytokines. Int. J. Neuropsychopharmacol. 5, 401–412. 2002a. Behavioral reactivity to stress: amplification of stress-induced
Kennedy, J.L., Neves-Pereira, M., King, N., Lizak, M.V., Basile, V.S., noradrenergic activation elicits a galanin-mediated anxiolytic effect in
Chartier, M.J., Stein, M.B., 2001. Dopamine system genes not linked central amygdala. Pharmacol. Biochem. Behav. 71, 407–417.
to social phobia. Psychiatry Genet. 11, 213–217. Khoshbouei, H., Cecchi, M., Morilak, D.A., 2002b. Modulatory
Kennett, G.A., Shah, K., Curzon, G., 1989. Anxiogenic-like effects of effects of galanin in the lateral bed nucleus of the stria terminalis
mCPP and TFMPP in animal models are opposed by 5-HT1C receptor on behavioral and neuroendocrine responses to acute stress.
antagonists. Eur. J. Pharmacol. 164, 445–454. Neuropsychopharmacology 27, 25–34.
Kennett, G.A., Lightowler, S., De Biasi, V., Stevens, N.C., Wood, M.D., Kieffer, B.L., Gavériaux-Ruff, C., 2002. Exploring the opioid system by
Tulloch, I.F., Blackburn, T.P., 1994. Effect of chronic administration of gene knock-out. Prog. Neurobiol. 66, 285–306.
selective 5-hydroxytryptamine and noradrenaline uptake inhibitors on Kiianman, K., Nurmi, M., Nykanen, I., Sinclaire, J.D., 1995. Effect
a putative index of 5-HT2C/2B receptor function. Neuropharmacology of ethanol on extracellular dopamine in the nucleus accumbens of
33, 1581–1588. alcohol preferring AA and alcohol-avoiding ANA rats. Pharmacol.
Kennett, G.A., Bailey, F., Piper, D.C., Blackburn, T.P., 1995. Effect of Biochem Behav. 52, 29–34.
SB 200646A, a 5-HT2C /5-HT2B receptor antagonist, in two conflict Kim, J.J., Diamond, D.M., 2002. The stressed hippocampus, synaptic
models of anxiety. Psychopharmacology 118, 178–182. plasticity and lost memories. Nat. Rev. Neurosci. 3, 453–462.
Kennett, G.A., Bright, F., Trail, B., Baxter, G.S., Blackburn, T.P., 1996a. Kim, J.J., Fanselow, M.S., 1992. Modality-specific retrograde amnesia
Effects of the 5-HT2B receptor agonist, BW 723C86, on three rat of fear. Science 25, 675–677.
models of anxiety. Br. J. Pharmacol. 117, 1443–1448. Kim, Y., Oh, S., 2002. Changes of GABAA receptor binding and subunit
Kennett, G.A., Wood, M.D., Bright, F., Cilia, J., Piper, D.C., Gager, mRNA level in rat brain by infusion of NOS inhibitor. Brain Res.
T., Thomas, D., Baxter, G.S., Forbes, I.T., Ham, P., Blackburn, 952, 246–256.
T.P., 1996b. In vitro and in vivo profile of SB 206553, a potent 5- Kim, M., Campeau, S., Falls, W.A., Davis, M., 1993. Infusion of the
HT2C /5-HT2B receptor antagonist with anxiolytic-like properties. Br. non-NMDA receptor antagonist CNQX into the amygdala blocks the
J. Pharmacol. 117, 427–434. expression of fear-potentiated startle. Behav. Neural Biol. 59, 5–8.
Kennett, G.A., Bright, F., Trail, B., 1997a. SB 242084, a selective and Kim, J.J., Shih, J.C., Chen, K., Chen, L., Bao, S., Maren, S., Anagnostaras,
brain penetrant 5-HT2C receptor antagonist. Neuropharmacology 36, S.G., Fanselow, M.S., De Maeyer, E., Seif, I., Thompson, R.F.,
609–620. 1997. Selective enhancement of emotional, but not motor, learning in
208 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

monoamine oxidase A-deficient mice. Proc. Natl. Acad. Sci U.S.A. Klodzińska, A., Chojnacka-Wójcik, E., 2000. Anticonflict effect
94, 5929–5933. of the glycineB receptor partial agonist, d-cycloserine, in rats.
Kim, J.J., Han, J.S., Packard, M.G., 2001. Amygdala is critical for Pharmacological analysis. Psychopharmacology 152, 224–228.
stress-induced modulation of hippocampal long-term potentiation and Klodzińska, A., Chojnacka-Wójcik, E., Palucha, A., Branski, P., Popik,
learning. J. Neurosci. 21, 5222–5228. P., Pilc, A., 1999. Potential anti-anxiety, anti-addictive effects of
King, C.M.F., Gommans, J., Joordens, R.J.E., Hijzen, T.H., Maes, LY354740, a selective group II glutamate metabotropic receptors
R.A.A., Olivier, B., 1997. Effects of 5-HT1A receptor ligands in a agonist in animal models. Neuropharmacology 38, 1831–1839.
modified Geller–Seifter conflict model in the rat. Eur. J. Pharmacol. Klodzińska, A., Tatarczyńska, E., Chojnacka-Wójcik, E., Pilc, A., 2000.
325, 121–128. Anxiolytic-like effects of group I metabotropic glutamate antagonist
Kinsey, A.M., Wainwright, A., Heavens, R., Sirinathsinghji, D.J.S., 2-methyl-6-(phenylethyl)-pyridine (MPEP) in rats. Pol. J. Pharmacol.
Oliver, K.R., 2001. Distribution of 5-HT5A , 5-HT5B , 5-HT6 and 5-HT7 52, 463–466.
receptor mRNAs in the rat brain. Mol. Brain Res. 88, 194–198. Klugbauer, N., Lacinova, L., Marais, E., Hobom, M., Hofman, F., 1999.
Kinzig, K.P., D’Alessio, D.A., Seeley, R.J., 2002. The diverse roles Molecular diversity of the calcium channel alpha2 delta subunit. J.
of specific GLP-1 receptors in the control of food intake and the Neurosci. 19, 684–691.
response to visceral illness. J. Neurosci. 22, 10470–10476. Knauber, J., Müller, W.E., 2000. Decreased exploratory activity and
Kirby, L.G., Pernar, L., Valentino, R.J., Beck, S.G., 2003. Distinguishing impaired passive avoidance behaviour in mice deficient for the
characteristics of serotonin and non-serotonin-containing cells in the ␣1b -adrenoceptor. Eur. Neuropsychopharmacol. 10, 423–427.
dorsal raphe nucleus: electrophysiological and immunohistochemical Kniazeff, J., Galvez, T., Labesse, G., Pin, J.-P., 2002. No ligand binding in
studies. Neuroscience 116, 669–683. the GB2 subunit of the GABAB receptor is required for activation and
Kirk, I.J., 1998. Frequency modulation of hippocampal theta by the allosteric interaction between the subunits. J. Neurosci. 22, 7352–7361.
supramammillary nucleus, and other hypothalamo-hippocampal Kniest, A., Wiesenberg, C., Weber, B., Colla, M., Heuser, I., Deuschle,
interactions: mechanisms and functional implications. Neurosci. M., 2001. The glutamate antagonist riluzole and its effects upon basal
Biobehav. Rev. 22, 291–302. and stress-induced activity of the human hypothalamus–pituitary–
Kishi, T., Aschkenasi, C.J., Lee, C.E., Mountjoy, K.G., Saper, C.B., adrenocortical system in elderly subjects. Neuropsychobiology 43,
Elmquist, J.K., 2003. Expression of melanocortin4 receptor mRNA in 91–95.
the central nervous system of the rat. J. Comp. Neurol. 457, 213–235.
Knobelman, D.A., Hen, R., Lucki, I., 2001. Regional patterns of
Kishimoto, K., Koyama, S., Akaike, N., 2000a. Presynaptic modulation
compensation following the genetic deletion of either the 5-hydroxy-
of synaptic ␥-aminobutyric acid transmission by tandospirone in rat
tryptamine1A or 5-hydroxytryptamine1B receptor in the mouse. J.
basolateral amygdala. Eur. J. Pharmacol. 407, 257–265.
Pharmacol. Exp. Ther. 298, 1092–1100.
Kishimoto, T., Radulovic, J., Radulovic, M., Lin, C.R., Schrick, C.,
Kobayashi, T., Ikeda, K., Ichikawa, T., Abe, S., Togashi, S., Kumanichi,
Hooshmand, F., Hermanson, O., Rosenfeld, M.G., Spiess, J., 2000b.
T., 1995. Molecular cloning of a mouse G-protein-activated K+
Deletion of cfht2 reveals an anxiolytic role for corticotropin-releasing
channel (MGIRK1) and distinct distributions of three GIRK (GIRK1
hormone receptor-2. Nat. Genet. 24, 415–419.
2 and 3) mRNAs in mouse brain. Biochem. Biophys. Res. Commun
Kishimoto, K., Koyama, S., Akaike, N., 2001. Synergistic ␮-opioid and
208, 1166–1173.
5-HT1A presynaptic inhibition of GABA release in rat periaqueductal
Kobayashi, S., Ohta, M., Miyasaka, K., Funakoshi, A., 1996. Decrease in
gray neurons. Neuropharmacology 41, 529–538.
exploratory behavior in naturally occurring cholecystokinin (CCK)-A
Kitahama, K., Nagatsu, I., Geffard, M., Maeda, T., 2000. Distribution
receptor gene knock-out rats. Neurosci. Lett. 214, 61–64.
of dopamine-immunoreactive fibers in the rat brainstem. J. Chem.
Kobayashi, S., Washiyama, K., Ikeda, K., 2003. Inhibition of G protein-
Neuroanat. 18, 1–9.
activated inwardly rectifying K+ channels by fluoxetine (Prozac). Br.
Kitaichi, K., Hori, T., Srivastava, L.K., Quirion, R., 1999. Antisense
J. Pharmacol. 138, 1119–1128.
oligodeoxynucleotides against the muscarinic m2, but not m4, receptor
supports its role as autoreceptors in the rat hippocampus. Mol. Brain Koek, W., Colpaert, F.C., 1991. Use of conflict procedure in pigeons
Res. 67, 98–106. to characterize anxiolytic drug activity: evaluation of N-methyl-d-
Kitaichi, K., Chabot, J.-G., Moebius, F.F., Flandorfer, A., Glossmann, H., aspartate antagonists. Life Sci. 49, PL37–PL42.
Quirion, R., 2000. Expression of the purported sigma1 (␴1 ) receptor Koenig, J.I., 2001. Estrogen and brain function. Trends Endocrinol.
in the mammalian brain and its possible relevance in deficits induced Metab. 12, 4–6.
by antagonism of the NMDA receptor complex as revealed using an Köhnke, M.D., Zabetian, C.P., Anderson, G.M., Kolb, W., Gaertner,
antisense strategy. J. Chem. Neuroanat. 20, 375–387. I., Buchkremer, G., Vonthein, R., Schick, S., Lutz, U., Köhnke,
Kitayama, M., Hirota, K., Kudo, M., Kudo, T., Ishihara, H., Matsuki, A.M., Cubells, J.F., 2002. A genotype-controlled analysis of plasma
A., 2002. Inhibitory effects of intravenous anaesthetic agents on K+ - dopamine ␤-hydroxylase in healthy and alcoholic subjects: evidence
evoked glutamate release from rat cerebrocortical slices. Involvement for alcohol-related differences in noradrenergic function. Biol.
of voltage-sensitive Ca2+ channels and GABAA receptors. Naunyn Psychiatry 52, 1151–1158.
Schmiedebergs Arch. Pharmacol. 366, 246–253. Kõks, S., Manisto, P.T., Bourin, M., Shlik, J., Vasar, V., Vasar, E., 2000.
Kittler, J.T., McAinsh, K., Moss, S.J., 2002. Mechanisms of GABAA Cholecystokinin-induced anxiety in rats: relevance of pre-experimental
receptor assembly and trafficking. Mol. Neurobiol. 26, 251–268. stress and seasonal variations. J. Psychiatry Neurosci. 25, 33–42.
Kiyama, H., Maeno, H., Tohyama, M., 1993. Substance P receptor Kõks, S., Beljajev, S., Koovit, I., Abramov, U., Bourin, M., Vasar,
(NK-1) in the central nervous system: possible functions from a E., 2001a. 8-OH-DPAT, but not deramciclane, antagonizes the
morphological aspect. Regul. Pept. 46, 114–123. anxiogenic-like action of paroxetine in an elevated plus-maze.
Kjelstrup, K.G., Tuvnes, F.A., Steffenach, H.-A., Murison, R., Moser, Psychopharmacology 153, 365–372.
E.I., Moser, M.-B., 2002. Reduced fear expression after lesions of the Kõks, S., Volke, V., Veraksits, A., Rünkorg, K., Sillat, T., Abramov,
ventral hippocampus. Proc. Natl. Acad. Sci. U.S.A. 99, 10825–10830. U., Bourin, M., Männistö, P.T., Matsui, T., Vasar, E., 2001b.
Klein, D.F., Thase, M.E., Endicott, J., Adler, L., Glick, I., Kalali, A., Cholecystokinin2 receptor-deficient mice display alterated function of
Leventer, S., Mattes, J., Ross, P., Bystritsky, A., 2002. Improving brain dopaminergic system. Psychopharmacology 158, 198–204.
clinical trials. Arch. Gen. Psychiatry 59, 272–278. Koller, K.J., Lowe, D.G., Bennet, G.L., Minamino, N., Kangawa, K.,
Kleven, M.S., Koek, W., 1999. Effects of different classes of partial Matsuo, H., Goeddel, D.V., 1991. Selective activation of the B
benzodiazepine agonists on punished and unpunished responding in natriuretic peptide receptor by C-type natriuretic peptide (CNP).
pigeons. Psychopharmacology 144, 405–410. Science 252, 120–123.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 209

Köning, M., Zimmer, A.M., Steiner, H., Holmes, P.V., Crawley, J.N., Koylu, E.O., Couceyro, P.R., Lambert, P.D., Kuhar, M.J., 1998. Cocaine-
Brownstein, M.J., Zimmer, A., 1996. Pain responses, anxiety and and amphetamine-regulated transcript peptide immunohistochemical
aggression in mice deficient in preproenkephalin. Nature 383, 535–538. localization in the rat brain. J. Comp. Neurol. 391, 115–132.
Koob, G.F., Wall, T.L., Schafer, J., 1987. Rapid induction of tolerance Kozicz, T., 2002. Met-enkephalin immunoreactive neurons recruited by
to the antipunishment effects of ethanol. Alcohol 4, 481–484. acute stress are innervated by axon terminals immunopositive for
Koob, G.F., Heinrichs, S.C., Menzahghi, F., Pich, E.M., Britton, K.T., tyrosine hydroxylase and dopamine-␣-hydroxylase in the anterolateral
1994. Corticotropin releasing factor, stress, and behavior. Semin. division of bed nuclei of the stria terminalis in the rat. Eur. J.
Neurosci. 6, 221–229. Neurosci. 16, 823–835.
Kopchia, K.L., Altman, H.J., Commissaris, R.L., 1992. Effects of lesions Kralic, J.E., Korpi, E.R., O’Buckley, T.K., Homanics, G.E., Morrow, A.L.,
of the central nucleus of the amygdala on anxiety-like behaviors in 2002a. Molecular and pharmacological characterization of GABAA
the rat. Pharmacol. Biochem. Behav. 43, 453–461. receptor ␣1 subunit knock-out mice. J. Pharmacol. Exp. Ther. 302,
Kopp, C., Vogel, E., Rettori, M.-C., Delagrange, P., Guardiola-Lemaı̂tre, 1037–1045.
B., Misslin, R., 1999a. Effects of melatonin on neophobic responses Kralic, J.E., O’Buckley, T.K., Khisti, R.T., Hodge, C.W., Homanics, G.E.,
in different strains of mice. Pharmacol. Biochem. Behav. 63, 521–526. Morrow, A.L., 2002b. GABAA receptor alpha-1 subunit deletion alters
Kopp, C., Vogel, E., Rettori, M.-C., Delagrange, P., Renard, P., Lesieur, receptor subtype assembly, pharmacological and behavioral responses
D., Misslin, R., 1999b. Antagonistic effects of S33153, a new MT1 to benzodiazepines and zolpidem. Neuropharmacology 43, 685–694.
and MT2 receptor ligand, on the neophobia-reducing properties of Kramer, M.S., Cutler, N.R., Ballenger, J.C., Patterson, W.M., Mendels, J.,
melatonin in BALB/c mice. Pharmacol. Biochem. Behav. 64, 131–136. Chenault, A., Shrivastava, R., Matzurawolfe, D., Lines, C., Reines, S.,
Kopp, C., Vogel, E., Rettori, M.-C., Delagrange, P., Misslin, R., 2000. 1995. A placebo-controlled trial of L-365–260, a CCKB antagonist,
Anxiolytic-like properties of melatonin receptor agonists in mice: in panic disorder. Biol. Psychiatry 37, 462–466.
involvement of MT1 and/or MT2 receptors in the regulation of Kramer, M.S., Cutler, N., Feighner, J., Shrivastava, R., Carman, J., Sramek,
emotional responsiveness. Neuropharmacology 39, 1865–1871. J.J., Reines, S.A., Liu, G., Snavely, D., Wyatt-Knowles, E., Hale,
Korpi, E.R., Gründer, G., Lüddens, H., 2002. Drug interactions at J.J., Mills, S.G., MacCoss, M., Swain, C.J., Horrison, T., Hill, R.G.,
GABAA receptors. Prog. Neurobiol. 67, 113–159. Hefti, F., Scolnick, E.M., Cascieri, M.A., Chicchi, G.G., Sadowski, S.,
Korte, S.M., 2001. Corticosteroids in relation to fear, anxiety and Williams, A.R., Hewson, L., Smith, D., Carlson, E.J., Hargreaves, R.J.,
psychophathology. Neurosci. Behav. Rev. 25, 117–142. Rupniak, N.M.J., 1998. Distinct mechanism for antidepressant activity
Korte, S.M., De Boer, S.F., De Kloet, E.R., Bohus, B., 1995. Anxiolytic- by blockade of central substance P receptors. Science 281, 1640–1645.
like effects of selective mineralocorticoid and glucocorticoid Krasowski, M.D., Koltchine, V.V., Rick, C.E., Ye, Q., Finn, S.E.,
antagonists on fear-enhanced behavior in the elevated plus-maze. Harrison, J., 1998. Propofol and other intravenous anesthetics have
Psychoneuroendocrinology 20, 385–394. sites of action on the ␥-aminobutyric acid type A receptor distinct
Korte, S.M., Buwalda, B., De Kloet, E.R., Bohus, B., 1996. Adrenaline from that for isoflurane. Mol. Pharmacol. 53, 530–538.
release by the 5-HT1A receptor agonist 8-OH-DPAT is partly Krishek, B.J., Xie, X., Blackstone, C., Huganir, R.L., Moss, S.J., Smart,
responsible for pituitary activation. Eur. J. Pharmacol. 309, 281–286. T.G., 1994. Regulation of GABAA receptor function by protein kinase
Kosofsky, B.E., Mollinver, M.E., 1987. The serotoninergic innervation of C phosphorylation. Neuron 12, 1081–1095.
cerebral cortex: different classes of axon terminal arise from dorsal Krishnan, K.R.R., 2002. Clinical experience with substance P receptor
and median raphe nuclei. Synapse 1, 153–168. (NK1 ) antagonists in depression. J. Clin. Psychiatry 63, 25–29.
Köster, A., Montkowski, A., Schulz, S., Stübe, E.-M., Dnaudt, K., Jenck, Kronfol, Z., Remick, 2000. Cytokines and the brain: implications for
F., Moreau, J.L., Nothacker, H.-P., Civelli, O., Reinscheid, R.K., 1999. clinical psychiatry. Am. J. Psychiatry 157, 683–694.
Targeted disruption of the orphaninFQ/nociceptin gene increases stress Kroog, G.S., Jensen, R.T., Battey, J.F., 1995. Mammalian bombesin
susceptibility and impairs stress adaptation in mice. Proc. Natl. Acad. receptors. Med. Res. Rev. 15, 389–417.
Sci. U.S.A. 96, 10444–10449. Krugers, H.J., Koolhaas, J.M., Bohus, B., Korf, J., 1993. A single social
Kostowski, W., Plaznik, A., Stefanski, R., 1989. Intra-hippocampal stress-experience alters glutamate receptor-binding in rat hippocampal
buspirone in animal models of anxiety. Eur. J. Pharmacol. 168, CA3 area. Neurosci. Lett. 154, 73–77.
393–396. Krysiak, R., Obuchowicz, E., Herman, Z.S., 2000. Conditioned fear-
Kostowski, W., Plaznik, A., Bidzinski, A., Rosnowska, E., Jessa, induced changes in neuropeptide Y-like immunoreactivity in rats: the
M., Nazar, M., 1994. Studies on antidepressant actions of a new effect of diazepam and buspirone. Neuropeptides 34, 148–157.
oxazolidinone derivative AS-8. Pol. J. Pharmacol. 46, 15–20. Krystal, J.H., McDougle, C.J., Woods, S.W., Price, L.H., Heninger, G.R.,
Kotlinska, J., Liljequist, S., 1998a. A characterization of anxiolytic-like Charney, D.S., 1992. Dose–response relationship for oral idazoxan
actions induced by the novel NMDA/glycine site antagonist, L701,324. effects in healthy human subjects: comparison with oral yohimbine.
Psychopharmacology 135, 175–181. Psychopharmacology 108, 313–319.
Kotlinska, J., Liljequist, S., 1998b. The putative AMPA receptor Krystal, J.H., D’Souza, C., Petrakis, I.L., Belger, A., Berman,
antagonist, LY326325, produces anxiolytic-like effects without altering R.M., Charney, D.S., Abi-Saab, W., Madonick, S., 1999. NMDA
locomotor activity in rats. Pharmacol. Biochem. Behav. 60, 119–124. agonists and antagonists as probes of glutamatergic dysfunction and
Kovacs, K.J., Földes, A., Sawchenko, P.E., 2000. Glucocortical negative pharmacotherapies in neuropsychiatric disorders. Harv. Rev. Psychiatry
feedback selectively targets vasopressin transcription in paracellular 7, 125–143.
neurosecretory neurons. J. Neurosci. 20, 3843–3852. Kubo, T., Okatani, H., Kanaya, T., Hagiwara, Y., Fukumori, R., Goshima,
Koyama, S., Kubo, C., Rhee, J.-S., Akaike, N., 1999. Presynaptic Y., 2003. Cholinergic mechanism in the lateral septal area is involved
serotonergic inhibition of GABAergic synaptic transmission in in the stress-induced blood pressure increase in rats. Brain Res. Bull.
mechanically dissociated rat basolateral amygdala neurons. J. Physiol. 59, 359–364.
518, 525–538. Kuhar, M.J., DallVechia, S.E., 1999. CART peptides: novel addiction-
Koyama, S., Matsumoto, N., Kubo, C., Akaike, N., 2000. Presynaptic and feeding-related neuropeptides. Trends Neurosci. 22, 316–320.
5-HT3 receptor-mediated modulation of synaptic GABA release in Kuiper, G.G., Carlsson, R., Grandien, K., Enmark, E., Haggblad, J.,
the mechanically dissociated rat amygdala neurons. J. Physiol. 529, Nilsson, S., Gustafson, J.A., 1997. Comparison of the ligand binding
373–383. specificity and transcript tissue distribution of estrogen receptors alpha
Koyama, S., Matsumoto, N., Murakami, N., Kubo, C., Nakebura, J., and beta. Endocrinology 138, 863–870.
Akaike, N., 2002. Role of presynaptic 5-HT1A and 5-HT3 receptors Kukkonen, J.P., Jansson, C.C., Akerman, K.E.O., 2001. Agonist trafficking
in modulation of synaptic GABA transmission in dissociated rat of Gi/o -mediated ␣2A -adrenoceptor responses in HEL 92.1.7 cells. Br.
basolateral amygdala neurons. Life Sci. 72, 375–387. J. Pharmacol. 132, 1477–1484.
210 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Kuner, R., Kohr, G., Grunewald, S., Eisenhardt, G., Bach, A., Kornau, Langub, M.C., Watson, R.E., Herman, J.P., 1995. Distribution of
H.C., 1999. Role of heteromer formation in GABAB receptor function. natriuretic peptide precursor mRNA in the rat brain. J. Comp. Neurol.
Science 283, 74–77. 356, 183–199.
Kung, A.W.C., 1999. Cytokines and hormonal regulations. Curr. Opin. Lanneau, C., Green, A., Hirst, W.D., Wise, A., Brown, J.T., Donnier,
Endocrinol. Diabetes 6, 77–83. E., Charles, K.J., Wood, M., Davies, C.H., Pangalos, M.N., 2001.
Kunishima, N., Shimada, Y., Tsuji, Y., Sato, T., Yamamoto, M., GABApentin is not a GABAB receptor agonist. Neuropharmacology
Kumasaka, T., Nakanishi, S., Jingami, H., Morikawa, K., 2000. 41, 965–975.
Structural basis of glutamate recognition by a dimeric metabotropic Lapiz, M.D.S., Mateo, Y., Durkin, S., Parker, T., Marsden, C.A., 2001.
glutamate receptor. Nature 407, 971–977. Effects of central noradrenaline depletion by the selective neurotoxin
Kuoppamaki, M., Palvimaki, E.-P., Hietala, J., Syvalahti, E., 1995. DSP-4 on the behaviour of the isolated rat in the elevated plus maze
Differential regulation of rat 5-HT2A and 5-HT2C receptors after and water maze. Psychopharmacology 155, 251–259.
chronic treatment with clozapine, chlorpromazine and three putative Largent, B.L., Gundlach, A.L., Snyder, S.H., 1986. Pharmacological
atypical antipsychotic drugs. Neuropsychopharmacology 13, 139–150. and autoradiographic discrimination of sigma and phencyclidine
Kuribara, H., Fujiwara, S., Yasuda, H., Tadokoro, S., 1990. The receptor binding sites in brain with (+)-[3 H]SKF-10,047, (+)-
anticonflict effect of MK-801, an NMDA antagonist: investigation by [3 H]-3-[3-hydroxyphenyl]-N-(1-propyl)piperidine and [3 H]-1-[1-(2-
punishment procedure in mice. Jpn. J. Pharmacol. 54, 250–252. thienyl)cyclohexyl]piperidine. J. Pharmacol. Exp. Ther. 238, 739–748.
Kustova, Y., Sei, Y., Morse, H.C., Basile, A.S., 1998. The influence of
Latini, S., Pedata, F., 2001. Adenosine in the central nervous system:
a targeted deletion of the IFNgamma gene on emotional behaviors.
release mechanisms and extracellular concentrations. J. Neurochem.
Brain Behav. Immun. 12, 308–324.
79, 463–484.
Kyuhou, S., Gemba, H., 1999. Injection of orphaninFQ/nociceptin into
the periaqueductal gray suppresses the forebrain-elicited vocalization Laughlin, T.M., Tram, K.V., Wilcox, G.L., Birnbaum, A.K., 2002.
in the guinea pig. Neurosci. Lett. 260, 113–160. Comparison of antiepileptic drugs tiagabine, lamotrigine, and
Laakman, G., Schüle, C., Lorkowski, G., Baghai, T., Kuhn, K., Ehrentraut, GABApentin in mouse models of acute, prolonged, and chronic
S., 1998. Buspirone and lorazepam in the treatment of generalized nociception. J. Pharmacol. Exp. Ther. 302, 1168–1175.
anxiety disorder in outpatients. Psychopharmacology 136, 357–366. Laurie, D.J., Seeburg, P.H., Wisden, W., 1992. The distribution of 13
Laaris, N., Le Poul, E., Laporte, A.M., Hamon, M., Lanfumey, L., GABAA receptor subunit mRNAs in the rat brain. II. Olfactory bulb
1999. Differential effects of stress on presynaptic and postsynaptic and cerebellum. J. Neurosci. 12, 1063–1076.
5-hydroxytryptamine-1A receptors in the rat brain: an in vitro Lavine, N., Ethier, N., Oak, J.N., Pei, L., Liu, F., Trieu, P., Rebois, R.V.,
electrophysiological study. Neuroscience 91, 947–958. Bouvier, M., Hébert, T.E., Van Tol, H.H.M., 2002. G protein-coupled
Labarca, C., Schwarz, J., Deshpande, P., Schwarz, S., Nowak, M.W., receptors form stable complexes with inwardly rectifying potassium
Fonck, C., Nashmi, R., Kofuji, P., Dang, H., Shi, W., Fidan, M., channels and adenylyl cyclase. J. Biol. Chem. 48, 46010–46019.
Khakh, B.S., Chen, Z., Bowers, B.J., Boulter, J., Wehner, J.M., Lester, Lax, P., Fucile, S., Eusebi, F., 2002. Ca2+ permeability of human
H.A., 2001. Point mutant mice with hypertensive alpha 4 nicotinic heteromeric nAChRs expressed by transfection in human cells. Cell
receptors show dopaminergic deficits and increased anxiety. Proc. Calcium 32, 53–58.
Natl. Acad. Sci. U.S.A. 98, 2786–2791. Lê, A.D., Quan, B., Juzytch, W., Fletcher, P.J., Joharchi, N., Shaham,
Laburthe, M., Couvineau, A., 2002. Molecular pharmacology and structure Y., 1998. Reinstatement of alcohol-seeking by priming injections
of VPAC receptors for VIP and PACAP. Regul. Pept. 108, 165–173. of alcohol and exposure to stress in rats. Psychopharmacology 135,
Lacerra, C., Martijena, I.D., Bustos, S.G., Molina, V.A., 1999. 169–174.
Benzodiazepine withdrawal falicitates the subsequent onset of escape Ledent, C., Vaugeois, J.M., Schiffman, S.N., Pedrazzini, T., El Yacoubi,
failures and anhedonia: influence of different antidepressant drugs. M., Vanderhaeghen, J.J., Costentin, J., Heath, J.K., Vassart, G.,
Brain Res. 819, 40–47. Parmentier, M., 1997. Aggressiveness, hypoalgesia and high blood
Lader, M., Scotto, J.C., 1998. Multicentre double-blind comparison of pressure in mice lacking the adenosine A2A receptor. Nature 388,
hydroxyzine, buspirone and placebo in patients with generalized 674–678.
disorder. Psychopharmacology 139, 402–406. LeDoux, J.E., 2000. Emotion circuits in the brain. Annu. Rev. Neurosci.
Lähdesmäki, J., Sallinen, J., MacDonald, E., Kobilka, B.K., Fagerholm, V., 23, 155–184.
Scheinin, M., 2002. Behavioral and neurochemical characterization of Lee, J.J., Croucher, M.J., 2003. Actions of group I and group II
␣2A -adrenergic receptor knock-out mice. Neuroscience 113, 289–299. metabotropic glutamate receptor ligands on 5-hydroxytryptamine
Lai, N.L., Bowen, W.D., Matsumoto, R.R., Thurkauf, A., Rice, K.C.,
release in the rat cerebral cortex in vivo: differential roles in the
Walker, J.M., 1989. Anxiogenic effects of two selective sigma ligands
regulation of central serotonergic neurotransmission. Neuroscience
in the rat. Am. Soc. Neurosci. Abstr. 15, 270.9.
117, 671–679.
Lambert, J.J., Belelli, D., Hill-Venning, C., Peters, J.A., 1995.
Lee, A., Rosin, D.L., Van Bockstaele, E.J., 1998. Ultrastructural evidence
Neurosteroids and GABA receptor function. Trends Pharmacol. Sci.
for prominent postsynaptic localization of ␣2C -adrenergic receptors
16, 295–303.
La Marca, S., Dunn, R.W., 1994. The ␣2 -antagonists idazoxan and in catecholaminergic dendrites in the rat nucleus locus coeruleus. J.
rauwolscine but not yohimbine or piperoxan are anxiolytic in the vogel Comp. Neurol. 394, 218–229.
lick-shock conflict paradigm following intravenous administration. Lee, H.K., Barbarosie, M., Kameyama, K., Bear, M.F., Huganir, R.L.,
Life Sci. 54, PL179–184. 2000. Regulation of distinct AMPA receptor phosphorylation sites
Lan, N.C., Gee, K.W., 1997. Epalons as promising therapeutic agents. during bidirectional synaptic plasticity. Nature 405, 955–959.
Drug News Perspect. 10, 604–611. Lees, G., 2000. Pharmacology of AMPA/kainate receptor ligands and
Landen, M., Eriksson, O., Sundblad, C., Andersch, B., Naessen, T., their therapeutic potential in neurological and psychiatric disorders.
Eriksson, E., 2001. Compounds with affinity for serotonergic receptors Drugs 59, 33–78.
in the treatment of premenstrual dysphoria: a comparison of buspirone, Léger, L., Gay, N., Cespuglio, R., 2002. Neurokinin NK1 - and NK3 -
nefazodone and placebo. Psychopharmacology 155, 292–298. immunoreactive neurons in serotonergic cell groups in the rat brain.
Landgraf, R., Gerstberger, R., Montkowski, A., Brobst, J.C., Wotjak, Neurosci. Lett. 323, 146–150.
C.T., Holsboer, F., Engelmann, M., 1995. V1 vasopressin receptor Lei, Q., Talley, E.M., Bayliss, D.A., 2001. Receptor-mediated inhibition
antisense oligodeoxynucleotide into septum reduces vasopressin of G protein-coupled inwardly rectifying potassium channels involves
binding, social discrimination abilities, and anxiety-related behavior G␣q family subunits, phospholipase C, and a readily diffusible
in rats. J. Neurosci. 15, 4250–4258. messenger. J. Biol. Chem. 276, 16720–16730.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 211

Lejeune, F., Millan, M.J., 2000. Pindolol excites dopaminergic and Li, X.B., Inoue, T., Hashimoto, S., Koyama, T., 2001c. Effect of chronic
adrenergic neurons, and inhibits serotonergic neurons, by activation administration of flesinoxan and fluvoxamine on freezing behavior
of 5-HT1A receptors. Eur. J. Neurosci. 12, 3265–3275. induced by conditioned fear. Eur. J. Pharmacol. 425, 43–50.
Lejeune, F., Millan, M.J., 2003. The novel corticotropin releasing factor Li, S., Ohgami, Y., Dai, Y., Quock, R.M., 2003. Antagonism of
(CRF)1 antagonist, DMP695, abolishes activation of locus coeruleus nitrous oxide-induced anxiolytic-like behavior in the mouse light/dark
(LC) noradrenergic perikarya by CRF in anesthetized rats. Eur. J. exploration procedure by pharmacologic disruption of endogenous
Pharmacol. 464, 127–133. nitric oxide function. Psychopharmacology 166, 366–372.
Lejeune, F., Gobert, A., Millan, M.J., 2002. The selective neurokinin Liang, F., Hatanaka, Y., Saito, H., Yamamori, T., Hashikawa, T., 2000.
(NK)1 antagonist, GR205,171, stereospecifically enhances mesocortical Differential expression of ␥-aminobutyric acid type B receptor-1␣ and
dopaminergic transmission in the rat: a combined dialysis and -␤ mRNA variants in GABA and non-GABAergic neurons of the rat
electrophysiological study. Brain Res. 935, 134–139. brain. J. Comp. Neurol. 416, 475–495.
Lelas, S., Rowlett, J.K., Spealman, R.D., Cook, J.M., Ma, C., Li, X., Liang, L., Valla, J., Sefidvash-Hockley, S., Rogers, J., Li, R., 2002.
Yin, W., 2002. Role of GABAA /benzodiazepine receptors containing Effects of estrogen treatment on glutamate uptake in cultured human
␣1 and ␣5 subunits in the discriminative stimulus effects of triazolam astrocytes derived from cortex of alzheimer’s disease patients. J.
in squirrel monkeys. Psychopharmacology 161, 180–188. Neurochem. 80, 807–814.
Le Moal, M., Simon, H., 1991. Mesocorticolimbic dopaminergic network: Liberzon, I., Zubieta, J.K., Fig. L.M., Phan, K.L., Koeppe, R.A., Taylor,
functional and regulatory roles. Physiol. Rev. 71, 155–234. S.F., 2002. ␮-Opioidreceptors and limbic responses to aversive
Le Mellédo, J.-M., Bradwejn, J., Koszycki, D., Bichet, D.G., Bellavance, emotional stimuli. Proc. Natl. Acad. Sci. 99, 7084–7089.
F., 1998. The role of the ␤-noradrenergic system in cholecystokinin- Liebowitz, M.R., Stein, M.B., Tancer, M., Carpenter, D., Oakes, R., Pitts,
tetrapeptide-induced panic symptoms. Biol. Psychiatry 44, 364–366. C.D., 2002. A randomised, double-blind, fixed-dose comparison of
Lenkei, A., Palkovits, M., Corvol, P., Llorens-Cortès, C., 1997. Expression paroxetine and placebo in the treatment of generalized social anxiety
of angiotensin type-1 (AT1 ) and type-2 (AT2 ) receptor mRNAs disorder. J. Clin. Psychiatry 63, 66–74.
in the adult rat brain: a functional neuroanatomical review. Front. Liebsch, G., Landgraf, R., Gerstberger, R., Probst, J.C., Wotjak, C.T.,
Neuroendocrinol. 18, 383–439. Engelmann, M., Holsboer, F., Montkowski, A., 1995. Chronic infusion
Leonard, B.E., Song, C., 2002. Changes in the immune system in rodent of a CRH1 receptor antisense oligodeoxynucleotide into the central
models of depression. Int. J. Neuropsychopharmacol. 5, 345–356. nucleus of the amygdala reduced anxiety-related behavior in socially
Lerma, J., Paternain, A.V., Rodriguez-Moreno, A., Lopez-Garcia, J.C., defeated rats. Regul. Pept. 59, 229–239.
2001. Molecular physiology of kainate receptors. Physiol. Rev. 81, Liebsch, G., Wotjak, C.T., Landgraf, R., Engelmann, M., 1996. Septal
971–998. vasopressin modulates anxiety-related behaviour in rats. Neurosci.
Lesch, K.P., 2001. Molecular foundation of anxiety disorders. J. Neural Lett. 217, 101–104.
Transm. 108, 717–746. Liebsch, G., Montkowski, A., Holsboer, F., Landgraf, R., 1998.
Lesch, K.P., Bengel, D., Heils, A., Sabol, S.Z., Greenberg, B.D., Petri, Behavioural profiles of two Wistar rat lines selectively bred for high
S., Benjamin, J., Müller, C.R., Hamer, D.H., Murphy, D.L., 1996. or low-anxiety-related behaviour. Behav. Brain Res. 94, 301–310.
Association of anxiety-related traits with a polymorphism in the Liebsch, G., Landgraf, R., Lorscher, P., Holsboer, F., 1999. Differential
serotonin transporter gene regulatory region. Science 274, 1527–1531. behavioural effects of chronic infusion of CRH 1 and CRH 2 receptor
Leurs, R., Blandine, P., Tedford, C., Timmerman, H., 1998. Therapeutic antisense oligonucleotides into the rat brain. Psychiatry Res. 33,
potential of histamine H3 receptor agonists and antagonists. Trends 153–163.
Pharmacol. Sci. 19, 177–183. Liechti, M.E., Vollenweider, F.X., 2001. Which neuroreceptors mediate the
Levant, B., 1997. The D3 dopamine receptor: neurobiology and potential subjective effects of MDMA in humans? A summary of mechanistic
clinical relevance. Pharmacol. Rev. 49, 231–252. studies. Hum. Psychopharmacol. 16, 589–598.
Levey, A.I., Kitt, C.A., Simonds, W.F., Price, D.L., Brann, M.R., 1991. Lightowler, S., Easton, N., Kennett, G.A., 2000. Investigation of the
Identification and localization of muscarinic acetylcholine receptor contribution of 5-HT2B receptor activation to the activity of Ro60
proteins in brain with subtype-selective antibodies. J. Neurosci. 11, 0175 in the rat social interaction model of anxiety. Br. J. Pharmacol.
3218–3226. 129, 153P.
Levita, L., Dalley, J.W., Robbins, T.W., 2002. Nucleus accumbens Liljequist, S., Engel, J.A., 1984a. Reversal of the anti-conflict action of
dopamine and learned fear revisited: a review and some new findings. valproate by various GABA and benzodiazepine antagonists. Life Sci.
Behav. Brain Res. 137, 115–127. 34, 2525–2533.
Lewin, G.R., Barde, Y.-A., 1996. Physiology of neurotrophins. Annu. Liljequist, S., Engel, J.A., 1984b. The effects of GABA and
Rev. Neurosci. 19, 289–317. benzodiazepine receptor antagonists on the anti-conflict actions of
Lewis, K., Li, C., Perrin, M.H., 2001. Identification of urocortin III, an diazepam or ethanol. Pharmacol. Biochem. Behav. 21, 521–525.
additional member of the corticotropin-releasing factor (CRF) family Lima, M., Trejo, E., Urbina, M., 1995. Serotonin turnover rate,
with high affinity for the CRF2 receptor. Proc. Natl. Acad. Sci. U.S.A. [3 H]paroxetine binding sites, and 5-HT1A receptors in the hippocampus
98, 7570–7575. of rats subchronically treated with clozapine. Neuropharmacology 34,
Lewohl, J.M., Wilson, W.R., Mayfield, R.D., Brozowski, S.J., Morrisett, 1327–1333.
R.A., Harris, R.A., 1999. G-protein-coupled inwardly rectifying Lin, D., Parsons, L.H., 2002. Anxiogenic-like effect of serotonin1B
potassium channels are targets of alcohol action. Nat. Neurosci. 2, receptor stimulation in the rat elevated plus-maze. Pharmacol.
1084–1090. Biochem. Behav. 71, 581–587.
Li, H.S., Xu, X.Z., Montell, C., 1999. Activation of a TRPC3-dependent Lin, Y.F., Angelotti, T.P., Dudek, E.M., Browning, M.D., Macdonald,
cation current through the neurotrophin BDNF. Neuron 24, 261–273. R.L., 1996. Enhancement of recombinant alpha1 beta1 ␥2L , ␥-
Li, A.H., Yeh, T., Tan, P.P., Hwang, H., Wang, H., 2001a. Neurotensin aminobutyric acid A receptor whole-cell currents by protein kinase C
excitation of serotonergic neurons in the rat nucleus raphe magnus: is mediated through phosphorylation of both beta1 and ␥2L subunits.
ionic and molecular mechanisms. Neuropharmacology 40, 1073–1083. Mol. Pharmacol. 50, 185–195.
Li, R., Nishijo, H., Wang, Q., Uwano, T., Tamura, R., Ohtani, O., Ono, Lin, C.H., Huang, Y.C., Tsai, J.J., Gean, P.W., 2001. Modulation of
T., 2001b. Light and electron microscopic study of cholinergic and voltage-dependent calcium currents by serotonin in acutely isolated
noradrenergic elements in the basolateral nucleus of the rat amygdala: rat amygdala neurons. Synapse 41, 351–359.
evidence for interactions between the two systems. J. Comp. Neurol. Lindblom, J., Schioth, H.B., Larsson, A., Wikberg, J.E., Bergstrom,
439, 411–425. L., 1998. Autoradiographic discrimination of melanocortin receptors
212 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

indicates that the MC3 subtype dominates in the medial rat brain. Lolait, S.J., O’Carroll, A.M., Shepard, E., Ginns, E.I., Young, W.S., 2000.
Brain Res. 810, 161–171. Characterization of a vasopressin V1b receptor knock-out mouse. Soc.
Lindheim, S.R., Legro, R.S., Bernstein, L., Stanczyk, F.Z., Vijod, Neurosci. Abstr. 26, 2406.
M.A., Presser, S.C., Lobo, R.A., 1992. Behavioral stress responses Lopez-Rubalcava, C., Fernandez-Guasti, A., 1994. Noradrenaline–
in premenopausal and postmenopausal women and the effects of serotonin interactions in the anxiolytic effects of 5-HT1A agonists.
estrogen. Am. J. Obstet. Gynecol. 167, 1831–1836. Behav. Pharmacol. 5, 42–51.
Lindia, J.A., Abbadie, C., 2003. Distribution of the voltage gated sodium Lopez, L.C., Frazier, M.L., Su, C.J., Kumar, A., Saunders, G.F., 1983.
channel NA, 1.3-like immunoreactivity in the adult rat central nervous Mammalian pancreatic preproglucagon contains three glucagon-related
system. Brain Res. 960, 132–141. peptides. Proc. Natl. Acad. Sci. U.S.A. 80, 5485–5489.
Lindvall, O., Björklund, A., 1984. General organisation of cortical Lopez, J.F., Chalmers, D.T., Little, K.Y., Watson, S.J., 1998. Regulation
monoamine systems. In: Descartes, L., Reader, T., Jasper, H. (Eds.), of serotonin1A , glucocorticoid, and mineralocorticoid receptor in
Monoamine Innervation of Cerebral Cortex. Liss, New York, pp. 9–40. rat and human hippocampus: implications for the neurobiology of
Lines, C., Challenor, J., Traub, M., 1995. Cholecystokinin and anxiety depression. Biol. Psychiatry 43, 547–573.
in normal volunteers: an investigation of the anxiogenic properties Lopez-Giménez, J.F., Mengod, G., Palacios, J.M., Vilaró, M.T.,
of pentagastrin and reversal by the cholecystokinin receptor subtypeB 1997. Selective visualization of rat brain 5-HT2A receptors by
antagonist L-365,260. Br. J. Clin. Pharmacol. 39, 235–242. auroradiography with [3 H]MDL 100,907. Naunyn Schmiedbergs Arch.
Linner, L., Arborelius, L., Nomikos, G.G., Bertilsson, L., Svensson, T.H., Pharmacol. 356, 446–454.
1999. Locus coeruleus neuronal activity and noradrenaline availability Lopez-Giménez, J.F., Vilaro, M.T., Palacios, J.M., Mengod, G., 1998.
in the frontal cortex of rats chronically treated with imipramine: effect [3 H]MDL 100,907 labels 5-HT2A serotonin receptors selectively in
of ␣2 -adrenoceptor blockade. Biol. Psychiatry 46, 766–774. primate brain. Neuropharmacology 37, 1147–1158.
Linthorst, A.C.E., Flachskamm, C., Reul, J.M.H., Holsboer, F., Oshima, Lopez-Giménez, J.F., Mengod, G., Palacios, J.M., Vilaró, M.T., 2001.
A., 2002a. Altered serotonergic neurotransmission in mice after Regional distribution and cellular localization of 5-HT2C receptor
chronic treatment with the corticotrophin-releasing hormone receptor mRNA in monkey brain: comparison with [3 H]mesulergine binding
type 1 antagonist R121919. Soc. Neurosci. Abstr. 75.3. sites and choline acetyltransferase mRNA. Synapse 42, 12–26.
Linthorst, A.C.E., Penalva, R.G., Flachskamm, C., Holsboer, F., Reul, Lotarski S.M., Kinsora, J.J., Taylor, C.T., Baron, S.P., 2002.
J.M.H.M., 2002b. Forced swim stress activates rat hippocampal Characterization of the anxiolytic-like activity and side-effects of
serotonergic neurotransmission involving a corticotropin-releasing pregabalin in mice. Soc. Neurosci. Abstr. 396.8.
hormone receptor-dependent mechanism. Eur. J. Neurosci. 16, 2441– Lovick, T.A., 2000. Panic disorder—a malfunction of multiple transmitter
2452. control systems within the midbrain periaqueductal gray matter.
Lister, R.G., 1985. The amnesic action of benzodiazepines in man. Neuroscientist 6, 48–59.
Neurosci. Biol. Behav. 9, 87–94. Löw, K., Crestani, F., Keist, R., Benke, D., Brünig, I., Benson, J.A.,
Littleton-Kearney, M.T., Ostrowski, N.L., Cox, D.A., Rossberg, M.I., Fritschy, J.-M., Rülicke, T., Bluethmann, H., Möhler, H., Rudolph, U.,
Hurn, P.D., 2002. Selective estrogen receptor modulators: tissue 2000. Molecular and neuronal substrate for the selective attenuation
actions and potential for CNS protection. CNS Drug Rev. 8, 309–330. of anxiety. Science 290, 131–134.
Liu, M., Glowa, J.R., 1999. Alterations of GABAA receptor subunit Lowy, M.T., Gault, L., Yamamoto, B.K., 1993. Adrenalectomy attenuates
mRNA levels associated with increases in punished responding stress-induced elevations in extracellular glutamate concentrations in
induced by acute alprazolam administration: an in situ hybridization the hippocampus. J. Neurochem. 61, 1957–1960.
study. Brain Res. 882, 8–16. Lowy, M.T., Wittenberg, L., Yamamoto, B.K., 1995. Effect of acute
Liu, R., Jolas, T., Aghajanian, G.K., 2000. Serotonin 5-HT2 receptors stress on hippocampal glutamate levels and spectrin proteolysis in
activate local GABA inhibitory inputs to serotonergic neurons of the young and aged rats. J. Neurochem. 65, 268–274.
dorsal raphe nucleus. Brain Res. 873, 34–45. Lu, Y., Grady, S., Marks, M.J., Piccitto, M., Changeux, J.P., Collins,
Liu, F., Wan, Q., Pristupa, Z., Yu, X.M., Wang, Y.T., Niznik, H.B., 2001. A.C., 1998. Pharmacological characterization of nicotinic receptor-
Direct protein-protein coupling enables cross-talk between dopamine stimulated GABA release from mouse brain synaptosomes. J.
D5 and ␥-aminobutyric acid A receptors. Nature 403, 274–280. Pharmacol. Exp. Ther. 287, 648–657.
Liu, R., Ding, Y., Aghajanian, G.K., 2002. Neurokinins activate local Lu, X.-Y., Ghasemzadeh, M.B., Kalivas, P.W., 1999. Regional distribution
glutamatergic inputs to serotonergic neurons of the dorsal raphe and cellular localization of ␥-aminobutyric acid subtype 1 receptor
nucleus. Neuropsychopharmacology 27, 329–340. mRNA in the rat brain. J. Comp. Neurol. 407, 166–182.
Llorca, P.-M., Spadone, C., Sol, O., Danniau, A., Bougerol, T., Corruble, Lu, S., Carey, G.J., Varty, G.B., 2002. Nociceptin produces anxiolytic-like
E., Faruch, M., Macher, J.-P., Sermet, E., Servant, D., 2002. Efficacy effects in a rat conditioned lick suppression model. Soc. Neurosci.
and safety of hydroxyzine in the treatment of generalized anxiety Abstr. 28, 683.10.
disorder: a 3-month double-blind study. J. Clin. Psychiatry 63, Lucas, G., Debonnel, G., 2002. 5-HT4 receptors exert a frequency-related
1020–1027. facilitatory control on dorsal raphe nucleus 5-HT neuronal activity.
Lloyd, G.K., Williams, M., 2000. Neuronal nicotinic receptors as novel Eur. J. Neurosci. 16, 817–822.
drug targets. J. Pharmacol. Exp. Ther. 292, 461–467. Lyons, W.E., Mamounas, L.A., Ricaurte, G.A., Coppola, V., Reid, S.W.,
Loftis, J.M., Janowsky, A., 2003. The N-methyl-d-aspartate receptor Bora, S.H., Wihler, C., Koliatsos, V.E., Tessarollo, L., 1999. Brain-
subunit NR2B: localization, functional properties, regulation, and derived neurotrophic factor-deficient mice develop aggressiveness and
clinical implications. Pharmacol. Ther. 97, 55–85. hyperphagia in conjunction with brain serotonergic abnormalities.
Loke, W.H., Hinrichs, J.V., Ghonheim, M.M., 1985. Caffeine and Proc. Natl. Acad. Sci. U.S.A. 96, 15239–15244.
diazepam: separate and combined effects on mood, memory and Ma, X.M., Levy, A., Lightman, S.L., 1997. Rapid changes in heteronuclear
psychomotor performance. Psychopharmacology 87, 344–350. RNA for corticotrophin-releasing hormone and arginine vasopressin
Lolait, S.J., O’Carroll, A.M., McBride, O.W., Konig, M., Morel, A., in response to acute stress. J. Endocrinol. 152, 81–89.
Bowbstein, M.J., 1992. Cloning and characterization of a vasopressin Maccarone, M., Valverde, O., Barbaccia, M.L., Castané, A., Maldonado,
V2 receptor and possible link to nephrogenic diabetes insipidus. R., Ledent, C., Parmentier, M., Finazzi-Agro, A., 2002. Age-related
Nature 357, 336–339. changes of anandamide metabolism in CB1 cannabinoid receptor
Lolait, S.J., O’Carroll, A.M., Mahan, L.C., Felder, C.C., Button, D.C., knock-out mice: correlation with behaviour. Eur. J. Neurosci. 15,
Young Jr., W.S.J, Mezey, E., Browstein, M.J., 1995. Extrapituitary 1178–1186.
expression of the rat V1b vasopressin receptor gene. Proc. Natl. Acad. MacDonald, R.L., Olsen, R.W., 1994. GABAA receptor channels. Annu.
Sci. U.S.A. 92, 6783–6787. Rev. Neurosci. 17, 569–602.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 213

MacDonald, E., Haapalinna, A., Virtanen, R., Lammintausta, R., 1988. Maisonnette, S.S., Villela, A.P., Carotti, A.P., Landeira-Fernandez, J.,
Medetomidine, a potent and highly selective ␣2 -adrenoceptor agonist 2000. Microinfusion of nefazodone into the basolateral nucleus of the
has anxiolytic properties in rodents. FASEB J. 2, 7366. amygdala enhances defensive behavior induced by NMDA stimulation
Macedo, C.E., Castilho, V.M., de Souza e Silva, M.A., Brandao, M.L., of the inferior colliculus. Physiol. Behav. 70, 243–247.
2002. Dual 5-HT mechanisms in basolateral and central nuclei of Maj, J., Bijak, M., Dziedzicka-Wasylewska, M., Rogoz, R., Skuza, G.,
amygdala in the regulation of the defensive behavior induced by Tokarski, T., 1996. The effects of paroxetine given repeatedly on the
electrical stimulation of the inferior colliculus. Brain Res. Bull. 59, 5-HT receptor subpopulations in the rat brain. Psychopharmacology
189–195. 127, 73–82.
MacIver, M.B., 1997. General anesthetic actions on transmission at Maj, J., Rogoz, Z., Dlaboga, D., Dzieddzicka-Wasylewska, M., 2000.
glutamate and GABA synapses. In: Biebuyck, J.F., Lynch, C., Maze, Pharmacological effects of milnacipran, a new antidepressant, given
M., Saidman, L.J., Yaksh, T.L., Zapol, W.M. (Eds.), Anesthesia: repeatedly on the ␣1 -adrenergic and serotonergic 5-HT2A systems. J.
Biologic Foundations. Lippincott–Raven, New York, pp. 277–286. Neural Transm. 107, 1345–1349.
Mackowiak, M., O’Neill, M.J., Hicks, C.A., Bleakman, D., Skolnick, Makara, G.B., Haller, J., 2001. Non-genomic effects of glucocorticoids
P., 2002. An AMPA receptor potentiator modulates hippocampal in the neural system. Evidence, mechanisms and implications. Prog.
expression of BDNF: an in vivo study. Neuropharmacology 43, 1–10. Neurobiol. 65, 367–390.
Maki, Y., Inoue, T., Izumi, T., Muraki, I., Ito, K., Kitaichi, Y., Li, X.,
MacLean, P.D., 1949. Psychosomatic disease and the visceral brain.
Koyama, T., 2000. Monoamine oxidase inhibitors reduce conditioned
Recent developments bearing on the Papez theory of emotion.
fear stress-induced freezing behavior in rats. Eur. J. Pharmacol. 406,
Psychosom. Med. 11, 338–353.
411–418.
MacLean, P.D., 1952. Some psychiatric implications of physiological
Makino, S., Shibasaki, T., Yamauchi, N., Nishioka, T., Mimoto, T.,
studies on the frontotemporal portion of the limbic system.
Wakabayashi, I., Gold, P.W., Hashimoto, K., 1999. Psychological
Electroencephalogr. Clin. Neurophysiol. 4, 407–418.
stress increased corticotropin-releasing hormone mRNA and content
MacNeil, D.J., Howard, A.D., Guan, X., Fong, T.M., Nargund, R.P., in the central nucleus of the amygdala but not in the hypothalamic
Bednarek, M.A., Goulet, M.T., Weinberg, D.H., Strack, A.M., Marsh, paraventricular nucleus in the rat. Brain Res. 850, 136–143.
D.J., Chen, H.Y., Shen, C.-P., Chen, A.S., Rosenblum, C.I., MacNeil, Makino, S., Baker, R.A., Smith, R.A., Gold, P.W., 2000. Differential
T., Tota, M., MacIntyre, E.D., Van der Ploeg, L.H.T., 2002. The role regulation of neuropeptide Y mRNA expression in the arcuate nucleus
of melanocortins in body weight regulation: opportunities for the and locus coeruleus by stress and antidepressants. J. Neuroendocrinol.
treatment of obesity. Eur. J. Pharmacol. 440, 141–157. 12, 387–395.
MacQueen, G.M., Ramakrishnan, K., Croll, S.D., Siuciak, J.A., Yu, G., Makino, S., Hashimoto, K., Gold, P.W., 2002a. Multiple feedback
Young, L.T., Pahnestock, M., 2001. Performance of heterozygous brain- mechanisms activating corticotropin releasing hormone system in the
derived neurotrophic factor knock-out mice on behavioral analogues of brain during stress. Pharmacol. Biochem. Behav. 73, 147–158.
anxiety, nociception, and depression. Behav. Neurosci. 115, 1145–1153. Makino, S., Smith, M.A., Gold, P.W., 2002b. Regulatory role of
Madeira, M.D., Paula-Barbosa, M.M., 1999. Effects of alcohol on the glucocorticoids and glucocorticoid receptor mRNA levels on tyrosine
synthesis and expression of hypothalamic peptides. Brain Res. Bull. hydroxylase gene expression in the locus coeruleus during repeated
48, 3–22. immobilization stress. Brain Res. 943, 216–223.
Madhav, T.R., Pei, Q., Zetterström, T.S.C., 2001. Serotonergic cells of Malberg-Aiello, P., Ipponi, A., Bartolini, A., Schunack, W., 2002. Mouse
the rat raphe nuclei express mRNA of tyrosine kinaseB (trkB), the light/dark box test reveals anxiogenic-like effects by activation of
high-affinity receptor for brain-derived neurotrophic factor (BDNF). histamine H1 receptors. Pharmacol. Biochem. Behav. 71, 313–318.
Mol. Brain Res. 93, 56–63. Malgorzata, F., Baran, L., Siwanowicz, J., Chojnacka-Wojcik, E.,
Maes, M., Van Gastel, A., Delmeire, L., Kenis, G., Bosmans, E., Song, Przegaliński, E., 1992. The anxiolytic-like effects of 5-hydroxy-
C., 2002. Platelet ␣2 -adrenoceptor density in humans: relationship to tryptamine3 (5-HT3 ) receptor antagonists. Pol. J. Pharmacol. Pharm.
stress-induced anxiety, psychasthenic constitution, gender and stress- 44, 261–269.
induced changes in the inflammatory response system. Psychol. Med. Malizia, A.L., 2002. Receptor binding and drug modulation in anxiety.
32, 919–928. Eur. Neuropsychopharmacol. 12, 567–574.
Magarinos, A.M., McEwen, B.S., 1995. Stress induced atrophy of apical Malleret, G., Hen, R., Guillou, J.L., Segu, L., Buhot, M.C., 1999. 5-
dendrites of hippocampal CA3 neurons: involvement of glucocorticoid HT1B receptor knock-out mice exhibit increased exploratory activity
secretion and excitatory amino acid receptors. Neuroscience 69, 89–98. and enhanced spatial memory performance in the Morris water maze.
Maggi, C.A., 1995. The mammalian tachykinin receptors. Gen. Pharmacol. J. Neurosci. 19, 6157–6168.
26, 911–944. Man, M.-S., Young, A.H., McAllister-Williams, H., 2002. Corticosterone
modulation of somatodendritic 5-HT1A receptor function in mice. J.
Maggio, R., Vogel, Z., Wess, J., 1993. Coexpression studies with mutant
Psychopharmacol. 16, 245–252.
muscarinic/adrenergic receptors provide evidence for intermolecular
Manns, I.D., Alonso, A., Jones, B.E., 2003. Rhythmically discharging
“cross-talk” between G-protein linked receptors. Proc. Natl. Acad.
basal forebrain units comprise cholinergic, GABAergic, and putative
Sci. U.S.A. 90, 3103–3107.
glutamatergic cells. J. Neurophysiol. 89, 1057–1066.
Maier, W., Falkai, P., 1999. The epidemiology of comorbidity between Mansbach, R.S., Harrod, C., Hoffmann, S.M., Nader, M.A., Lei, Z.,
depression, anxiety disorders and somatic diseases. Int. Clin. Witkin, J.M., Barrett, J.E., 1988. Behavioral studies with anxiolytic
Psychopharmacol. 14, S1–S6. drugs. V. Behavioral and in vivo neurochemical analyses in pigeons
Mailliet, F., Galloux, P., Poisson, D., 2001. Comparative effects of of drugs that increase punished responding. J. Pharmacol. Exp. Ther.
melatonin, zolpidem and diazepam on sleep, body temperature, blood 246, 114–120.
pressure and heart rate measured by radiotelemetry in Wistar rats. Mansour, A., Khachaturian, H., Lewis, M.E., Akil, H., Watson, S.J., 1988.
Psychopharmacology 156, 417–426. Anatomy of CNS opioid receptors. Trends Neurosci. 11, 308–314.
Maines, L.W., Keck, B.J., Smith, J.E., Lakoski, J.M., 1999. Corticosterone Mansour, A., Fox, C.A., Akil, H., Watson, S.J., 1995. Opioid receptor
regulation of serotonin transporter and 5-HT1A receptor expression in mRNA expression in the rat CNS: anatomical and functional
the aging brain. Synapse 32, 58–66. implications. Trends Neurosci. 18, 22–29.
Maione, S., Palazzon, E., De Novellis, V., Stella, L., Leyva, J., Rossi, F., Mantella, R.C., Amico, J.A., 2002. Female oxytocin deficient mice
1998. Metabotropic glutamate receptors modulate serotonin release display anxiety-like behavior. Soc. Neurosci. Abstr. 571.21.
in the rat periaqueductal gray matter. Naunyn Schmiedebergs Arch. Mantyh, P.W., 2002. Neurobiology of substance P and the NK1 receptor.
Pharmacol. 358, 411–417. J. Clin. Psychiatry 63, 6–10.
214 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Mantyh, P.W., Hunt, S.P., Maggio, J.E., 1984. Substance P receptors Martin, J.R., Ballard, T.M., Higgins, G.A., 2002a. Influence of the 5-
localization by light microscopic autoradiography in rat brain using HT2C receptor antagonist, SB-242,084 in tests of anxiety. Pharmacol.
[3 H]SP as the radioligand. Brain Res. 307, 147–165. Biochem. Behav. 71, 615–625.
Manzaneque, J.M., Brain, P.F., Navarro, J.F., 2002. Effect of low doses Martin, M., Ledent, C., Parmentier, M., Maldonado, R., Valverde, O.,
of clozapine on behaviour of isolated and group-housed male mice 2002b. Involvement of CB1 cannabinoid receptors in emotional
in the elevated plus-maze test. Prog. Neuropsychopharmacol. Biol. behaviour. Psychopharmacology 159, 379–387.
Psychiatry 26, 349–355. Martinez, M., Phillips, P.J., Herbert, J., 1998. Adaptation in patterns of
Manzoni, O.J., Manabe, T., Nicoll, R.A., 1994. Release of adenosine c-fos expression in the brain associated with exposure to either single
by activation of NMDA receptors in the hippocampus. Science 238, or repeated social stress in male rats. Eur. J. Neurosci. 10, 20–33.
2098–2101. Martinez, D., Hwang, D.-R., Mawlawi, O., Slifstein, M., Kent, J.,
Marais, E., Klugbauer, N., Hofmann, F., 2001. Calcium channel ␣2␦ Simpson, N., Parsey, R.V., Hashimoto, T., Huang, Y., Shinn, A.,
subunits-structure and GABApentin binding. Mol. Pharmacol. 59, Van Heertum, R., Abi-Dargham, A., Caltabiano, S., Malizia, A.,
1243–1248. Cowley, H., Mann, J.J., Laruelle, M., 2001. Differential occupancy
of somatodendritic and postsynaptic 5-HT1A receptors by pindolol: a
Marco, N., Thirion, A., Mons, G., Bougault, I., Le Fur, G., Soubrié,
dose-occupancy study with [11 C]WAY100,635 and positron emission
P., Steinberg, R., 1998. Activaton of dopaminergic and cholinergic
tomography in humans. Neuropsychopharmacology 24, 209–229.
neurotransmission by tachykinin NK3 receptor stimulation: an in vivo
Martinez, G., Ropero, C., Funes, A., Flores, E., Blotta, C., Landa, A.I.,
microdialysis approach in guinea pig. Neuropeptides 32, 481–488.
Gargiulo, P.A., 2002. Effects of selective NMDA and non-NMDA
Margeta-Mitrovic, M., Mitrovic, I., Riley, R.C., Jan, L.Y., Basbaum, A.I.,
blockade in the nucleus accumbens on the plus-maze test. Physiol.
1999. Immunohistochemical localizaton of GABAB receptors in the
Behav. 76, 219–224.
rat central nervous system. J. Comp. Neurol. 405, 299–321.
Martin-Ruiz, R., Puig, V.M., Celada, P., Shapiro, D.A., Roth, B.L.,
Marin, S., Marco, E., Biscaia, M., Fernandez, B., Rubio, M., Guaza, Mengod, G., Artigas, F., 2001. Control of serotonergic function
C., Schmidhammer, H., Viveros, M.P., 2003. Involvement of the in medial prefrontal cortex by serotonin-2A receptors through a
␬-opioidreceptor in the anxiogenic-like effect of CP-55,940 in male glutamate-dependent mechanism. J. Neurosci. 21, 9856–9866.
rats. Pharmacol. Biochem. Behav. 74, 649–656. Martins, A.P., Marras, R.A., Guimaraes, F.S., 1997. Anxiogenic effect
Marinelli, M., Piazza, P.V., 2002. Interaction between glucocorticoid of corticotropin-releasing hormone in the dorsal periaqueductal grey.
hormone, stress and psychostimulant drugs. Eur. J. Neurosci. 16, NeuroReport 8, 3601–3604.
387–394. Martin-Schild, S., Gerall, A.A., Kastin, A.J., Zadina, J.E., 1999.
Marinus, J., Leenjens, A.F.G., Visser, M., Stiggelbout, A.M., Van Hilten, Differential distribution of endomorphin1 - and endomorphin2 -like
J.J., 2002. Evaluation of the hospital anxiety and depression scale in immunoreactivities in the CNS of the rodent. J. Comp. Neurol. 405,
patients with Parkinson’s disease. Clin. Neuropharmacol. 25, 318–324. 450–471.
Mark, M.D., Herlitze, S., 2000. G-protein mediated gating of inward- Marx, C.E., VanDoren, M.J., Duncan, G.E., Lieberman, J.A., Morrow,
rectifier K+ channels. Eur. J. Biochem. 267, 5830–5836. A.L., 2003. Olanzapine and clozapine increase the GABAergic
Marksitzer, R., Benke, D., Fritschy, J.M., Trzeciak, A., Bannwarthn, neuroactive steroid allopregnanolone in rodents. Neuropsychopharma-
W., Möhler, H., 1993. GABAA -receptors: drug binding profile and cology 28, 1–13.
distribution of receptors containing the ␣2 -subunit in situ. J. Receptor Mascia, M.P., Biggio, F., Mancuso, L., Cabras, S., Cocco, P.L., Gorini,
Res. 13, 467–477. G., Manca, A., Mara, C., Purdy, R.H., Follesa, P., Biggio, L., 2002.
Marsh, D.J., Weingarth, D.T., Novi, D.E., Chen, H.Y., Trumbauer, M.E., Changes in GABAA receptor gene expression induced by withdrawal
Chen, A.S., Guan, X.-M., Jiang, M.M., Feng, Y., Camacho, R.E., of, but not by long-term exposure to, ganaxolone in cultured rat
Shen, Z., Frazier, E.G., Yu, H., Metzger, J.M., Kuca, S.J., Shearman, cerebellar granule cells. J. Pharmacol. Exp. Ther. 303, 1014–1020.
L.P., Gopal-Truter, S., Macnell, D.J., Strack, A.M., Macintyre, D.E., Mash, D.C., Zabetian, C.P., 1992. Sigma receptors are associated with
Van der Ploeg, L.H.T., Qian, S., 2002. Melanin-concentrating hormone cortical limbic areas in the primate brain. Synapse 12, 195–205.
1 receptor-deficient mice are lean, hyperactive, and hyperphagic and Masino, S.A., Diao, L., Illes, P., Zahniser, N.R., Larson, G.A.,
have altered metabolism. Proc. Natl. Acad. Sci. U.S.A. 99, 3240–3245. Johansson, B., Fredholm, B.B., Dunwiddle, T.V., 2002. Modulation
Marshall, F.H., 2001. Heterodimerization of G-protein-coupled receptors of hippocampal glutamatergic transmission by ATP is dependent on
in the CNS. Curr. Opin. Pharmacol. 1, 40–44. adenosine A1 receptors. J. Pharmacol. Exp. Ther. 303, 356–363.
Mason, P., Skinner, J., Luttinger, D., 1987. Two tests in rats for
Marsicano, G., Wotjak, C.T., Azad, S.C., Bisogno, T., Rammes, G.,
antianxiety effects of clinically anxiety-attenuating antidepressants.
Casclo, M.G., Hermann, H., Tang, J., Hofmann, C., Zieglgänsberger,
Psychopharmacology 92, 30–34.
W., Di Marzo, V., Lutz, B., 2002. The endogenous cannabinoid system
Masood, A., Banerjee, B., Vitayan, V.K., Ray, A., 2003. Modulation of
controls extinction of aversive memories. Nature 418, 530–534.
stress-induced neurobehavioral changes by nitric oxide in rats. Eur. J.
Martijena, I.D., Rodriguez Manzanares, P.A., Lacerra, C., Molina, V.A.,
Pharmacol. 458, 135–139.
2002. Gabaergic modulation of the stress response in frontal cortex Mateo, Y., Fernandez-Pastor, B., Meana, J.J., 2001. Acute and chronic
and amygdala. Synapse 45, 86–94. effects of desipramine and clorgyline on ␣2 -adrenoceptors regulating
Marti, M., Stocchi, S., Paganini, F., Mela, F., De Risi, C., Calo, noradrenergic transmission in the rat brain: a dual-probe microdialysis
G., Guerrini, R., Barnes, T.A., Lambert, D.G., Beani, L., Bianchi, study. Br. J. Pharmacol. 133, 1362–1370.
C., Morari, M., 2003. Pharmacological profiles of presynaptic Matheson, G.K., Guthrie, D., Bauer, C., Knowles, A., White, G., Ruston,
nociceptin/orphaninFQ receptors modulating 5-hydroxytryptamine and C., 1991. Sigma receptor ligands alter concentrations of corticosterone
noradrenaline release in the rat neocortex. Br. J. Pharmacol. 138, 91–98. in plasma in the rat. Neuropharmacology 30, 79–87.
Martin, J.R., Schöch, P., Jenck, F., Moreau, J.-L., Haefely, W.E., 1993. Matheus, M.G., Guimaraes, F.S., 1997. Antagonism of non-NMDA
Pharmacological characterization of benzodiazepine receptor ligands receptors in the periaqueductal grey induces anxiolytic effect in the
with intrinsic efficacies ranging from high to zero. Psychopharmacology elevated plus maze. Psychopharmacology 132, 14–18.
111, 415–422. Matheus, M.G., Nogueira, R.L., Carobrez, A.P., Graeff, F.G., Guimaraes,
Martin, J.R., Bös, M., Jenck, F., Moreau, J.-L., Mutel, V., Sleight, A.J., F.S., 1994. Anxiolytic effect of glycine antagonists microinjected into
Wichmann, J., Andrews, J.S., Berendsen, H.H.G., Broekkamp, C.L.E., the dorsal periaqueductal grey. Psychopharmacology 113, 565–569.
Ruigt, G.S.F., Köhler, C., Van Delft, A.M.L., 1998. 5-HT2C receptor Matsumo, K., Matsunaga, K.H., Mita, S., 1995. Acute effects of sigma
agonists: pharmacological characteristics and therapeutic potential. J. ligands on the extracellular DOPAC level in rat frontal cortex and
Pharm. Exp. Ther. 286, 913–924. striatum. Neurochem. Res. 20, 233–238.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 215

Matsumoto, Y., Kataoka, Y., Watanabe, Y., Miyazaki, A., Taniyama, McDonnell, D.O., 1999. The molecular pharmacology of SERMs. Trends
K., 1994. Antianxiety actions of Ca2+ channel antagonists with Endocrinol. Metab. 10, 301–311.
Vogel-type conflict test in rats. Eur. J. Pharmacol. 264, 107–110. McElroy, J.F., Ward, K.A., Zeller, K.L., Jones, K.W., Gilligan, P.J., He,
Matsumoto, M., Togashi, H., Mori, K., Ueno, K.-I., Miyamoto, A., L., Lelas, S., 2002. The CRF1 receptor antagonist DMP696 produces
Yoshioka, M., 1999. Characterization of endogenous serotonin- anxiolytic effects and inhibits the stress-induced hypothalamic–
mediated regulation of dopamine release in the rat prefrontal cortex. pituitary–adrenal axis activation without sedation or ataxia in rats.
Eur. J. Pharmacol. 383, 39–48. Psychopharmacology 165, 86–92.
Matsuo, M., Kataoka, Y., Mataki, S., Kato, Y., Oi, K., 1996. Conflict McGaugh, J.L., McIntyre, C.K., Power, A.E., 2002. Amygdala modulation
situation increases serotonin release in rat dorsal hippocampus: in vivo of memory consolidation: interaction with other brain systems.
study with microdialysis and Vogel test. Neurosci. Lett. 215, 197–200. Neurobiol. Learn. Memory 78, 539–552.
Matsuo, M., Ayuse, T., Oi, K., Kataoka, Y., 1997. Propofol produces McIntyre, C.K., Hatfield, T., McGaugh, J.L., 2002a. Amygdala
anticonflict action by inhibiting 5-HT release in rat dorsal hippocampus. norepinephrine levels after training predict inhibitory avoidance
NeuroReport 8, 3087–3090. retention performance in rats. Eur. J. Neurosci. 16, 1223–1226.
Matsuyama, S., Nei, K., Tanaka, C., 1997. Regulation of GABA release McIntyre, K.L., Porter, D.M., Henderson, L.P., 2002b. Anabolic
via NMDA and 5-HT1A receptors in guinea pig dentate gyrus. Brain androgenic steroids induce age-, sex-, and dose-dependent changes
Res. 761, 105–112. in GABAA receptor subunit mRNAs in the mouse forebrain.
Maubach, K.A., Martin, K., Chicchi, G., Harrison, T., Wheeldon, A., Neuropharmacology 43, 634–645.
Swain, C.J., Cumberbatch, M.J., Rupniak, N.M.J., Seabrook, G.R., McKell-Carter, R., Hofstötter, C., Tsuchiya, N., Koch, C., 2003. Working
2002. Chronic substance P (NK1 ) receptor antagonist and conventional memory and fear conditioning. Proc. Natl. Acad. Sci. U.S.A. 100,
antidepressant treatment increases burst firing of monoamine neurones 1399–1404.
in the locus coeruleus. Neuroscience 109, 609–617. McKernan, R.M., 2002. The role of GABAA receptors subtypes in the
Maurel-Remy, S., Schreiber, R., Dalmus, M., De Vry, J., 1996. effects of benzodiazepine. Eur. Neuropsychopharmacol. 12, S155–
Somatodendritic 5-HT1A receptors are critically involved in the S156.
anxiolytic effects of 8-OH-DPAT. Psychopharmacology 125, 89–91. McKernan, R.M., Rosahal, T.W., Reynolds, D.S., Sur, C., Wafford,
Maurice, T., Phan, V.L., Urani, A., Kamei, H., Noda, Y., Nabeshima, K.A., Atack, J.R., Farrar, S., Myers, J., Cook, G., Ferris, P., Garrett,
T., 1999. Neuroactive neurosteroids as endogenous effectors for L., Bristow, L., Marshall, G., Macaulay, A., Brown, N., Howell,
the sigma1 (␴1 ) receptor: pharmacological evidence and therapeutic O., Moore, K.W., Carling, R.W., Street, L.J., Castro, J.L., Ragan,
opportunities. Jpn. J. Pharmacol. 81, 125–155. C.I., Dawson, G.R., Whiting, P.J., 2000. Sedative but not anxiolytic
Maurice, T., Urani, A., Phan, V.-L., Romieu, P., 2001. The interaction properties of benzodiazepines are mediated by the GABAA receptor
between neuroactive steroids and the ␴1receptor function: behavioral ␣1 subtype. Nat. Neurosci. 3, 587–592.
consequences and therapeutic opportunities. Brain Res. Rev. 37, McKinney, M., 1993. Muscarinic receptor subtype-specific coupling to
116–132. second messengers in neuronal systems. Prog. Brain Res. 98, 333–340.
Maurin, Y., Arbilla, S., Langer, S.Z., 1985. Inhibition by yohimbine of McKittrick, C.R., Blanchard, D.C., Blanchard, R.J., McEwen, B.S.,
the calcium-dependent evoked release of [3 H]GABA in rat and mouse Sakai, R.R., 1995. Serotonin receptor binding in a colony model of
brain slices in vitro. Eur. J. Pharmacol. 111, 37–48. chronic social stress. Biol. Psychiatry 37, 383–393.
Mayberg, H.S., Silva, J.A., Brannan, S.K., Tekell, J.L., Mahurin, R.K., McLay, R.N., Freeman, S.M., Harlan, R.E., Ide, C.F., Kastin, A.J.,
McGinnis, S., Jerabek, P.A., 2002. The functional neuroanatomy of Zadina, J.E., 1997. Aging in the hippocampus: interrelated actions
the placebo effect. Am. J. Psychiatry 159, 728–737. of neurotrophins and glucocorticoids. Neurosci. Biobehav. Rev. 21,
Mazzucchelli, C., Pannacci, M., Nonno, R., Lucini, V., Fraschini, F., 615–629.
Stankov, B.M., 1996. The melatonin receptor in the human brain: McLeod, M., Pralong, D., Copolov, D., Dean, B., 2002. The heterogeneity
cloning experiments and distribution studies. Brain Res. Mol. Brain of central benzodiazepine receptor subtypes in the human hippocampal
Res. 39, 117–126. formation, frontal cortex and cerebellum using [3 H]flumazenil and
McAskill, R., Mir, S., Taylor, D., 1998. Pindolol augmentation of zolpidem. Mol. Brain Res. 104, 203–209.
antidepressant therapy. Br. J. Psychiatry 173, 203–208. McMahon, L.R., France, C.P., 2003. Discriminative stimulus effects
McCall, R.B., Clement, M.E., 1994. Role of serotonin1A and serotonin2 of positive GABAA modulators and other anxiolytics, sedatives,
receptors in the central regulation of the cardiovascular system. and anticonvulsants in untreated and diazepam-treated monkeys. J.
Pharmacol. Rev. 46, 231–243. Pharmacol. Exp. Ther. 304, 109–120.
McCarthy, M.M., McDonald, C.H., Brooks, P.J., Goldman, D., 1996. An McMahon, L.R., Gerak, L.R., France, C.P., 2001. Potency of positive
anxiolytic action of oxytocin is enhanced by estrogen in the mouse. ␥-amynobutiric acidA modulators to substitute for a midazolam
Physiol. Behav. 60, 1209–1215. discriminative stimulus in untreated monkeys dose not predict potency
McCloskey, T.C., Damian, G.M., Brown, B.D., Barraco, R.A., Altman, to attenuate a flumazenil discriminative stimulus in diazepam-treated
H.J., 1990. Antagonism of the anti-conflict effects of phenobarbital, but monkeys. J. Pharmacol. Exp. 298, 1227–1235.
not diazepam, by the A1 adenosine agonist l-PIA. Psychopharmacology McMahon, B.M., Boules, M., Warrington, L., Richelson, E., 2002.
102, 283–290. Neurotensin analogs: indications for use as potential antipsychotic
McCullough, L.D., Salamone, J.D., 1992. Anxiogenic drugs beta-CCE compounds. Life Sci. 70, 1101–1119.
and FG 7142 increase extracellular dopamine levels in nucleus McMillan, D.E., Hardwick, W.C., DeCosta, B.R., Rice, K.C., 1991.
accumbens. Psychopharmacology 109, 379–382. Effects of drugs that bind to PCP and sigma receptors on punished
McDonald, A.J., Mascagni, F., 2002. Immunohistochemical charac- responding. J. Pharmacol. Exp. Ther. 258, 1015–1018.
terization of somatostatin containing interneurons in the rat basolateral McQuade, R., Sharp, T., 1997. Functional mapping of dorsal and median
amygdala. Brain Res. 943, 237–244. raphe 5-hydroxytryptamine pathways in forebrain of the rat using
McDonald, A.J., Pearson, J.C., 1989. Coexistence of GABA and microdialysis. J. Neruochem. 69, 791–796.
peptide immunoreactivity in non-pyramidal neurons of the basolateral McQuade, R., Creton, D., Stanford, S.C., 1999. Effect of novel
amygdala. Neurosci. Lett. 100, 53–58. environmental stimuli on rat behaviour and central noradrenaline
McDonald, B.J., Amato, A., Connolly, C.N., Benke, D., Moss, S.J., function measured by in vivo microdialysis. Psychopharmacology
Smart, T.G., 1998. Adjacent phosphorylation sites on GABAA receptor 145, 393–400.
␤ subunits determine regulation by camp-dependent protein kinase. Meadows, H.J., Kumar, C.S., Pritchett, D.B., Blackburn, T.P., Benham,
Nat. Neurosci. 1, 23–28. C.D., 1998. SB-205384: a GABAA receptor modulator with a novel
216 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

mechanism of action that shows subunit selectivity. J. Pharmacol. Messaoudi, E., Ying, S.-W., Kahnhema, T., Croll, S.D., Bramham,
123, 1253–1259. C.R., 2002. Brain-derived neurotrophic factor triggers transcription-
Meijer, O.C., De Kloet, E.R., 1998. Corticosterone and serotonergic dependent, late phase long-term potentiation in vivo. J. Neurosci. 22,
neurotransmission in the hippocampus: functional implications of 7453–7461.
central corticosteroid receptor diversity. Crit. Rev. Neurobiol. 12, 1–20. Meyer, D.A., Carta, M., Partridge, L.D., Covey, D.F., Valenzuela,
Meis, S., Pape, H.-C., 2001. Control of glutamate and GABA release by C.F., 2002. Neurosteroids enhance spontaneous glutamate release in
nociceptin/orphaninFQ in the rat lateral amygdala. J. Physiol. 532.3, hippocampal neurons. J. Biol. Chem. 277, 28725–28732.
701–712. Micheau, J., Van Marrewijk, B., 1999. Stimulation of 5-HT1A receptors
Meller, R., Harrison, P.J., Elliott, J.M., Sharp, T., 2002. In vitro evidence by systemic or medial septum injection induces anxiogenic-like effects
that 5-hydroxytryptamine increases efflux of glial glutamate via and facilitates acquisition of a spatial discrimination task in mice.
5-HT2A receptor activations. J. Neurosci. Res. 67, 399–405. Prog. Neuropsychopharmacol. Biol. Psychiatry 23, 1113–1133.
Mellman, T.A., Uhde, T.W., 1989a. Electroencephalographic sleep in Michel, M.C., Beck-Sickinger, A., Cox, H., Doods, H.N., Herzog,
panic disorder. Arch. Gen. Psychiatry 46, 178–184. H., Larhammar, D., Quirion, R., Schwartz, T., Westfall, T., 1998.
Melo, L.L., Brandao, M.L., 1995. Role of 5-HT1A and 5-HT2 receptors XVI. International Union of Pharmacology recommendations for
in the aversion induced by electrical stimulation of inferior colliculus. the nomenclature of neuropeptide Y, peptide YY, and pancreatic
Pharmacol. Biochem. Behav. 51, 317–321. polypeptide receptors. Pharmacol. Rev. 50, 143–150.
Meltzer, H.Y., 1995. Atypical antipsychotic drugs. In: Bloom, F.E., Miczek, K.A., Lau, P., 1975. Effects of scopolamine, physostigmine and
Kupfer, D.J. (Eds.), Psychopharmacology: The Fourth Generation of chlordiazepoxide on punished and extinguished water consumption in
Progress. Raven, New York, pp. 1277–1286. rats. Psychopharmacologia 42, 263–269.
Meltzer, H.Y., 1999. The role of serotonin in antipsychotic drug action. Miczek, K.A., Fish, E.W., De Bold, J.F., De Almeida, R.M.M.,
Neuropsychopharmacology 21, 106S–115S. 2002. Social and neural determinants of aggressive behavior:
Menard, J., Treit, D., 1998. The septum and the hippocampus differentially pharmacotherapeutic targets at serotonin, dopamine and ␥-aminobutyric
mediate anxiolytic effects of R(+)-8-OH-DPAT. Behav. Pharmacol. 9, acid systems. Psychopharmacology 163, 434–458.
93–101. Migita, K., Loewy, A.D., Ramabhadran, T.V., Krause, J.E., Waters, S.M.,
Menard, J., Treit, D., 1999. Effects of centrally administered anxiolytic 2001. Immunohistochemical localization of the neuropeptide Y Y1
compounds in animal models of anxiety. Neurosci. Biobehav. Rev. receptor in rat central nervous system. Brain Res. 889, 23–37.
23, 591–613. Mihailescu, S., Guzmin-Marin, R., Del Carmen Frias Dominguez, M.,
Menard, J., Treit, D., 2000. Intra-septal infusions of excitatory amino acid Drucker-Colin, R., 2002. Mechanisms of nicotine actions on dorsal
receptor antagonists have differential effects in two animal models of raphe serotoninergic neurons. Eur. J. Pharmacol. 452, 77–82.
anxiety. Behav. Pharmacol. 11, 99–108. Mihalek, R.M., Banerjee, P.K., Korpi, E.R., Quinlan, J.J., Firestone, L.L.,
Mendlin, A., Martin, F.J., Jacobs, B.L., 1999. Dopaminergic input is Mi, Z.P., Lagenaur, C., Tretter, V., Sieghart, W., Anagnostaras, S.G.,
required for increases in serotonin output produced by behavioural Sage, J.R., Fanselow, M.S., Guidotti, A., Spigelman, I., Li, Z., DeLorey,
activation: an in vivo microdialysis study in rat forebrain. Neuroscience T.M., Olsen, R.W., Homanics, G.E., 1999. Attenuated sensitivity to
93, 897–905. neuroactive steroids in ␥-aminobutyrate type A receptor delta subunit
Meneses, A., 1999. 5-HT system and cognition. Neurosci. Biobehav. knock-out mice. Proc. Natl. Acad. Sci. U.S.A. 96, 12905–12910.
Mihalek, R.M., Bowers, B.J., Wehner, J.M., Kralic, J.E., VanDoren, M.J.,
Rev. 23, 1111–1125.
Morrow, A.L., Homanics, G.E., 2001. GABAA-receptor ␦ subunit
Meneses, A., Hong, E., 1993. Modification of the anxiolytic effects of
knock-out mice have multiple defects in behavioral responses to
5-HT1A agonists by shock intensity. Pharmacol. Biochem. Behav. 46,
ethanol. Alcohol. Clin. Exp. Res. 25, 1708–1718.
569–573.
Mihic, S.J., 1999. Acute effects of ethanol on GABAA and glycine
Menza, M.A., Robertson-Hoffman, D.E., Bonapace, A.S., 1993.
receptor function. Neurochem. Int. 35, 115–123.
Parkinson’s disease and anxiety: comorbidity with depression. Biol.
Miklös, I.H., Kovacs, K.J., 2002. GABAergic innervation of corticotropin-
Psychiatry 34, 465–470.
releasing hormone (CRH)-secreting parvocellular neurons and its
Menzaghi, F., Behan, D.P., Chalmers, D.T., 2002. Constitutively activated
plasticity as demonstrated by quantitative immunoelectron microscopy.
G-protein-coupled receptors: a novel approach to CNS drug discovery.
Neuroscience 113, 581–592.
Curr. Drug Targets CNS Neurol. Disord. 1, 105–121.
Miksys, S.L., Tyndale, R.F., 2002. Drug-metabolizing cytochrome P450s
Merali, Z., Lacosta, S., Anisman, H., 1997. Effects of interleukin1␤ and
in the brain. Rev. Psychiatry Neurosci. 27, 406–415.
mild stress on alterations of norepinephrine, dopamine and serotonin Mileusnic, D., Lee, J.M., Magnuson, D.J., Hejna, M.J., Krause, J.E.,
neurotransmission: a regional microdialysis study. Brain Res. 761, Lorens, J.B., 1999. Neurokinin-3 receptor distribution in rat and human
225–235. brain: an immuhistochemical study. Neuroscience 89, 1269–1290.
Merali, Z., McIntosch, J., Kent, P., Michaud, D., Anisman, H., 1998. Miljanich, G.P., Ramachandran, J., 1995. Antagonists of neuronal calcium
Aversive and appetitive events evoke the release of corticotropin- channels: structure, function, and therapeutic implications. Annu. Rev.
releasing hormone and bombesin-like peptides at the central nucleus Pharmacol. Toxicol. 35, 707–734.
of the amygdala. J. Neurosci. 18, 4758–4766. Millan, M.J., 1986. Multiple opioid systems and pain. Pain 27, 303–349.
Merchenthaler, I., Lopez, F.J., Negro-Vilar, A., 1993. Anatomy and Millan, M.J., 1990. Kappa-opioid receptors and analgesia. Trends
physiology of central galanin-containing pathways. Prog. Neurobiol. Pharmacol. Sci. 11, 70–76.
40, 711–769. Millan, M.J., 1999. The induction of pain: an integrative review. Prog.
Merchenthaler, I., Lane, M., Shughrue, P., 1999. Distribution of pre-pro- Neurobiol. 57, 1–164.
glucagon and glucagon-like peptide-1 receptor messenger RNAs in Millan, M.J., 2000. Improving the treatment of schizophrenia: focus on
the rat central nervous system. J. Comp. Neurol. 403, 261–280. serotonin (5-HT)1A receptors. J. Pharmacol. Exp. Ther. 295, 853–861.
Merighi, A., 2002. Costorage and coexistence of neuropeptides in the Millan, M.J., 2002a. Descending control of pain. Prog. Neurobiol. 66,
mammalian CNS. Prog. Neurobiol. 66, 161–190. 355–474.
Merikangas, K.R., Methta, R.L., Molnar, B.E., Walters, E.E., Swendsen, Millan, M.J., 2002b. N-Methyl-d-aspartate receptor-coupled glycineB
J.D., Aguilar-Gaziola, S., Bijl, R., Borges, G., Caraveo-Anduaga, receptors in the pathogenesis and treatment of schizophrenia: a critical
J.J., DeWit, D.J., Kolody, B., Vega, W.A., Wittchen, H.-U., Kessler, review. Curr. Drug Targets CNS Neurol. Disord. 1, 191–213.
R.C., 1998. Comorbidity of substance use disorders with mood and Millan, M.J., Brocco, M., 2003. The Vogel Conflict Test of anxiolytic
anxiety disorders: results of the international consortium in psychiatric properties: procedural aspects, GABAergic, glutamatergic and
epidemiology. Addict. Behav. 23, 893–907. serotonergic mechanisms. Eur. J. Pharmacol. 463, 67–96.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 217

Millan, M.J., Duka, T., 1981. Anxiolytic properties of opiates and Millan, M.J., Brocco, M., Gobert, A., Dorey, G., Casara, P., Dekeyne, A.,
endogenous opiod peptides and their relationship to the actions of 2001a. Anxiolytic properties of the selective non-peptidergic CRF1
benzodiazepines. Mod. Probl. Pharmacopsychiatry 17, 123–141. antagonists, CP-154,526 and DMP695: a comparison to other classes
Millan, M.J., Canton, H., Lavielle, G., 1992. Targeting multiple serotonin of anxiolytic agent. Neuropsychopharmacology 25, 585–600.
receptors: mixe 5-HT1A agonists/5-HT1C/2 antagonists as therapeutic Millan, M.J., Dekeyne, A., Brocco, M., Gobert, A., 2001b. Contrasting
agents. Drug News Perspect. 5, 397–406. influence of triazolobenzodiazepines as compared with benzodiazepines
Millan, M.J., Hjorth, S., Samanin, R., Schreiber, R., Jaffard, R., De upon dialysate levels of noradrenaline, dopamine and serotonin in
Ladonchamps, B., Veiga, S., Goument, B., Peglion, J.-L., Spedding, frontal cortex of freely-moving rats. Neuropsychopharmacology 11,
M., Brocco, M., 1997. S15535, a novel benzodioxopiperazine ligand S300.
of serotonin (5-HT1A ) receptors. II. modulation of hippocampal Millan, M.J., Dekeyne, A., Papp, M., Drieu La Rochelle, C., MacSweeny,
serotonin release in relation to potential anxiolytic properties. J. C., Peglion, J.-L., Brocco, M., 2001c. S33005, a novel ligand at both
Pharmacol. Exp. Ther. 282, 148–161. serotonin and norepinephrine transporters. II. Behavioural profile in
Millan, M.J., Gobert, A., Newman-Tancredi, A., Audinot, V., Lejeune, comparison to venlafaxine, reboxetine, citalopram and clomipramine.
F., Rivet, J.-M., Cussac, D., Nicolas, J.-P., Muller, O., Lavielle, G., J. Pharmacol. Exp. Ther. 298, 581–591.
1998. S16924 ((R)-2-{1-[2-(2,3-dihydro-benzo[1,4]dioxin-5-yloxy)- Millan, M.J., Gobert, A., Lejeune, F., Newman-Tancredi, A., Rivet, J.-M.,
ethyl]-pyrrolidin-3yl}-1-(4-fluoro-phenyl)-ethanone), a novel, potential Auclair, A., Peglion, J.-L., 2001d. S33005, a novel ligand at both
antipsychotic with marked serotonin (5-HT)1A agonist properties. I. serotonin and norepinephrine transporters. II. Behavioural profile in
Receptorial and neurochemical profile in comparison with clozapine comparison to venlafaxine, reboxetine, citalopram and clomipramine.
and haloperidol. J. Pharmacol. Exp. Ther. 286, 1341–1355. J. Pharmacol. Exp. Ther. 298, 565–580.
Millan, M.J., Brocco, M., Gobert, A., Schreiber, R., Dekeyne, A., Millan, M.J., Lejeune, F., De Nanteuil, G., Gobert, A., 2001e. Selective
1999a. S-16924 ((R)-2-{1-[2-(2,3-dihydro-benzo[1,4] dioxin-5-yloxy)- blockade of neurokinin (NK)1 receptors facilitates the activity
ethyl]-pyrrolidin-3yl}-1-(4-fluorophenyl)-ethanone), a novel, potential of adrenergic pathways projecting to frontal cortex and dorsal
antipsychotic with marked serotonin1A (5-HT1A ) agonist properties. hippocampus in rats. J. Neurochem. 76, 1949–1954.
III. Anxiolytic actions in comparison to clozapine and haloperidol. J. Millan, M.J., Newman-Tancredi, A., Duqueyroix, D., Cussac, D., 2001f.
Pharmacol. Exp. Ther. 288, 1002–1014. Agonist properties of pindolol at h5-HT1A receptors coupled to
Millan, M.J., Gobert, A., Audinot, V., Dekeyne, A., Newman-Tancredi,
mitogen-activated protein kinase. Eur. J. Pharmacol. 424, 13–17.
A., 1999b. Inverse agonists and serotonergic transmission: from
Millan, M.J., Maiofiss, L., Cussac, D., Audinot, V., Boutin, J.-A.,
recombinant, human serotonin (5-HT)1B receptors to G-protein
Newman-Tancredi, A., 2002. Differential actions of antiparkinson
coupling and function in corticolimbic structures in vivo. Neuropsycho-
agents at multiple classes of monoaminergic receptor. I. A multivariate
pharmacology 21, 61S–67S.
analysis of the binding profiles of 14 drugs at 21 native and cloned
Millan, M.J., Brocco, M., Rivet, J.-M., Audinot, V., Newman-Tancredi,
human receptor subtypes. J. Pharmacol. Exp. Ther. 303, 791–804.
A., Maiofiss, L., Queriaux, S., Despaux, N., Peglion, J.-L., Dekeyne,
Milner, T.A., Lee, A., Aicher, S.A., Rosin, D.L., 1998. Hippocampal
A., 2000a. S18327 (1-{2-[4-(6-fluoro 1,2-benzisoxazol-3-yl) piperid-1-
␣2A -adrenergic receptors are located predominantly presynaptically
yl] ethyl} 3-phenyl imidazolin-2-one), a novel, potential antipsychotic
but are also found postsynaptically and in selective astrocytes. J.
displaying marked antagonist properties at ␣1 - and ␣2 -adrenergic
Comp. Neurol. 395, 310–327.
receptors. II. Functional profile and a multiparametric comparison
Miquel, M.-C., Emerit, M.B., Nosjean, A., Simon, A., Rumajogee,
with haloperidol, clozapine, and 11 other antipsychotic agents. J.
P., Brisorgueil, M.-J., Doucet, E., Hamon, M., Vergé, D., 2002.
Pharmacol. Exp. Ther. 292, 54–66.
Millan, M.J., Dekeyne, A., Newman-Tancredi, A., Cussac, D., Audinot, Differential subcellular localization of the 5-HT3 -As receptor subunit
V., Milligan, G., Duqueyroix, D., Girardon, S., Mulot, J., Boutin, in the rat central nervous system. Eur. J. Neurosci. 15, 449–457.
J.A., Nicolas, J.P., Renouard-Try, A., Lacoste, J.M., Cordi, A., 2000b. Mineka, S., Öhman, A., 2002. Phobias and preparedness: the selective,
S18616, a highly potent spiroimidazoline agonist at ␣2 -adrenoceptors. automatic, and encapsulated nature of fear. Biol. Psychiatry 52,
I. Receptor profile, antinociceptive and hypothermic actions in 927–937.
comparison with dexmedetomidine and clonidine. J. Pharmacol. Misane, I., Razani, H., Wang, F.-H., Jansson, A., Fuxe, K., Ögren, S.O.,
Exp. Ther. 295, 1192–1205. 1998. Intraventricular galanin modulates a 5-HT1A receptor-mediated
Millan, M.J., Gobert, A., Rivet, J.-M., Adhumeau-Auclair, A., Newman- behavioural reponse in the rat. Eur. J. Neurosci. 10, 1230–1240.
Tancredi, A., Dekeyne, A., Nicolas, J.-P., Lejeune, F., 2000c. Mitler, M.M., 2000. Nonselective and selective benzodiazepine receptor
Mirtazapine enhances frontocortical dopaminergic and corticolimbic agonists—where are we today? Sleep 23, S39–S47.
adrenergic, but not serotonergic, transmission by blockade of ␣2 - Miu, P., Jarvie, K.R., Radhakrishnan, B., Gates, M.R., Ogden, A.,
adrenergic and serotonin2C receptors: a comparison with citalopram. Ornstein, P.L., Zarrinmayeh, H., Ho, K., Peters, D., Grabell, J., Gupsa,
Eur. J. Neurosci. 12, 1079–1095. A., Zimmerman, D.M., Bleakman, D., 2001. Novel AMPA receptor
Millan, M.J., Lejeune, F., Gobert, A., 2000d. Reciprocal autoreceptor and potentiators LY392098 and LY404187: effects on recombinant human
heteroreceptor control of serotonergic, dopaminergic and noradrenergic AMPA receptors in vitro. Neuropharmacology 40, 976–983.
transmission in the frontal cortex: relevance to the actions of Miura, H., Naoi, M., Nakahara, D., Ohta, T., Nagatsu, T., 1996. Effects
antidepressant agents. J. Psychopharmacol.14, 114–138. of moclobemide on forced-swimming stress and brain monoamine
Millan, M.J., Lejeune, F., Gobert, A., Brocco, M., Auclair, A., Bosc, C., levels in mice. Pharmacol. Biochem. Behav. 53, 469–475.
Rivet, J.-M., Lacoste, J.-M., Cordi, A., Dekeyne, A., 2000e. S18616, Miura, H., Qiao, H., Ohta, T., 2002. Influence of aging and social
a highly potent spiroimidazoline agonist at ␣2 -adrenoceptors. II. isolation on changes in brain monoamine turnover and biosynthesis
Influence on monoaminergic transmission, motor function, and anxiety of rats elicited by novelty stress. Synapse 46, 116–124.
in comparison with dexmedetomidine and clonidine. J. Pharmacol. Miyasaka, K., Kobayashi, S., Ohta, M., Kanai, S., Yoshida, Y., Nagata,
Exp. Ther. 295, 1206–1222. A., Matsui, T., Noda, T., Takiguchi, S., Takata, Y., Kawanami, T.,
Millan, M.J., Newman-Tancredi, A., Audinot, V., Cussac, D., Lejeune, Funakoshi, A., 2002. Anxiety-related behaviors in cholecystokinin-A,
F., Nicolas, J.-P., Cogé, F., Galizzi, J.-P., Boutin, J.A., Rivet, J.-M., B, and AB receptor gene knock-out mice in the plus-maze. Neurosci.
Dekeyne, A., Gobert, A., 2000f. Agonist and antagonist actions of Lett. 335, 115–118.
yohimbine as compared to fluparoxan at ␣2 -adrenergic receptors Mizoule, J., Gauthier, A., Uzan, A., Renault, C., Dubroeucq, M.C.,
(AR)s, serotonin (5-HT)1A , 5-HT1B , 5-HT1D and dopamine D2 Guérémy, C., Le Fur, G., 1985. Opposite effects of two ligands
and D3 receptors: significance for the modulation of frontocortical for peripheral type benzodiazepine binding sites, PK 11195 and
monoaminergic transmission and depressive states. Synapse 35, 79–95. R05-4864, in a conflict situation in the rat. Life Sci. 36, 1059–1068.
218 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Mizuki, Y., Suetsugi, M., Ushijima, I., Yamada, M., 1997. Differential Montgomery, S.A., Mahé, V., Haudiquet, V., Hackett, D., 2002.
aspects of dopaminergic drugs on anxiety and arousal in healthy Effectiveness of venlafaxine, extended release formulation, in the
volunteers with high and low-anxiety. Prog. Neuropsychopharmacol. short-term and long-term treatment of generalized anxiety disorder:
Biol. Psychiatry 21, 573–590. results of a survival analysis. J. Clin. Psychopharmacol. 22, 561–567.
Mody, I., 2001. Distinguishing between GABAA receptors responsible Montkowski, A., Barden, N., Wotjak, C., Stec, I., Ganster, J., Meaney,
for tonic and phasic conductances. Neurochem. Res. 26, 907–913. M., Engelmann, M., Reul, J.M., Landgraf, R., Holsboer, F., 1995.
Moebius, F.F., Reiter, R.J., Hanner, M., Glossmann, H., 1997. High Long-term antidepressant treatment reduces behavioural deficits in
affinity of sigma1 -binding sites for sterol isomerization inhibitors: transgenic mice with impaired glucocorticoid receptor function. J.
evidence for a pharmacological relationship with the yeast sterol Neuroendocrinol. 7, 841–845.
C8 –C7 isomerase. Br. J. Pharmacol. 121, 1–6. Montkowski, A., Jahn, H., Ströhle, A., Poettig, M., Holsboer, F.,
Moghaddam, B., 2002. Stress activation of glutamate neurotransmission in Wiedemann, K., 1998. C-type natriuretic peptide exerts effects
the prefrontal cortex: implications for dopamine-associated psychiatric opposing those of atrial natriuretic peptide on anxiety-related behavior
disorders. Biol. Psychiatry 51, 775–787. in rats. Brain Res. 972, 358–360.
Moghaddam, B., Bolinao, M.L., Stein-Behrens, B., Sapolsky, R., 1994. Monzon, M.E., De Barioglio, S.R., 1999. Response to novelty after i.c.v.
Glucocorticoids mediate the stress-induced extracellular accumulation injection of melanin-concentrating hormone (MCH) in rats. Physiol.
of glutamate. Brain Res. 655, 251–254. Behav. 67, 813–817.
Mogil, J.S., Pasternak, G.W., 2001. The molecular and behavioral Monzon, M.E., Varas, M.M., De Barioglio, S.R., 2001. Anxiogenesis
pharmacology of the orphaninFQ/nociceptin peptide and receptor induced by nitric oxide synthase inhibition and anxiolytic effect of
family. Pharmacol. Rev. 53, 381–415. melanin-concentrating hormone (MCH) in rat brain. Peptides 22,
Möhler, H., Crestani, F., Rudolph, U., 2001. GABAA receptor subtypes: 1043–1047.
a new pharmacology. Curr. Opin. Pharmacol. 1, 22–25. Moore, N.A., Rees, G., Sanger, G., Tye, N.C., 1994. Effects of olanzapine
Möhler, H., Fritschy, J.M., Rudolph, U., 2002. A new benzodiazepine and other antipsychotic agents on responding maintained by a conflict
pharmacology. J. Pharmacol. Exp. Ther. 300, 2–8. schedule. Behav. Pharmacol. 5, 196–202.
Mora, P.O., Netto, C.F., Graeff, F.G., 1997. Role of 5-HT2A and 5-HT2C
Molchanov, M.L., Guimarães, F.S., 2002. Anxiolytic-like effects of
receptor subtypes in the two types of fear generated by the elevated
AP7 injected into the dorsolateral or ventrolateral columns of the
T-maze. Pharmacol. Biochem. Behav. 58, 1051–1057.
periaqueductal gray of rats. Psychopharmacology 160, 30–38.
Moraes-Ferreira, V.M., Morato, G., 1997. d-Cycloserine blocks the
Molewijk, H.E., Van Der Poel, A.M., Olivier, B., 1995a. The ambivalent
effects of ethanol and HA-966 in rats tested in the elevated plus-maze.
behaviour “stretched approach posture” in the rat as a paradigm to
Alcohol. Clin. Exp. Res. 21, 1638–1642.
characterize anxiolytic drugs. Psychopharmacology 121, 81–90.
Moragues, N., Ciofi, P., Tramu, G., Garret, M., 2002. Localization of
Molewijk, H.E., van der Poel, A.M., Mos, J., van der Heyden, J.A.,
GABAA receptor ε-subunitin cholinergic and aminergic neurones
Olivier, B., 1995b. Conditioned ultrasonic distress vocalizations in
and evidence for co-distribution with the ␪-subunitin rat brain.
adult male rats as a behavioural paradigm for screening anti-panic
Neuroscience 111, 657–669.
drugs. Psychopharmacology 117, 32–40.
Morales, M., Bloom, F.E., 1997. The 5-HT3 recepor is present in different
Molderings-Godlewski, G., Kwolek, G., Malinowska, B., Barann, M.,
subpopulations of GABAergic neurons in the rat telencephalon. J.
Golderings, G.J., Urban, B.W., Gothert, M., 2002. Inhibitory effects
Neurosci. 17, 3157–3167.
of cannabinoid receptor agonists on the 5-HT3 receptor-mediated
Morales, M., Wang, S.-D., 2002. Differential composition of 5-
Barzold–Jarisch reflex in rats and human 5-HT3A receptors in HEK
hydroxytryptamine3 receptors synthesized in the rat CNS and
293 cells. Naunyn Schmiedebergs Arch. Pharmacol. 365, R28.
peripheral nervous system. J. Neurosci. 22, 6732–6741.
Möller, C., Bing, O., Heilig, M., 1994. c-fos Expression in the amygdala: Moret, C., Briley, M., 2000. The possible role of 5-HT1B/D receptors in
in vivo antisense modulation and role in anxiety. Cell. Mol. Neurobiol. psychiatric disorders and their potential as a target for therapy. Eur.
14, 415–423. J. Pharmacol. 404, 1–12.
Möller, C., Wiklund, L., Sommer, W., Thorsell, A., Heilig, M., 1997. Moret, C., Briley, M., 2001. Where are the new therapies for anxiety
Decreased experimental anxiety and voluntary ethanol consumption and obsessive–compulsive disorders? Drugs 4, 1031–1042.
in rats following central but not basolateral amygdala lesions. Brain Morimoto, S., Sigmund, C.D., 2002. Angiotensin mutant mice: a focus
Res. 760, 94–101. on the brain renin–angiotensin system. Neuropeptides 36, 194–200.
Möller, C., Sommer, W., Thorsell, A., Heilig, M., 1999. Anxiogenic- Morimoto, M., Morita, N., Ozawa, H., Yokoyama, K., Kawata, M., 1996.
like action of galanin after intra-amygdala administration in the rat. Distribution of glucocorticoid receptor immunoreactivity and mRNA
Neuropsychopharmacology 21, 507–512. in the rat brain: an immunohistochemical and in situ hybridization
Möller, H.-J., Volz, H.-P., Reimann, I.W., Stoll, K.-D., 2001. Opipramol study. Neuroscience 26, 235–269.
for the treatment of generalized anxiety disorder: a placebo-controlled Morimoto, I., Yamamoto, S., Kai, K., Fujihira, T., Morita, E., Eto,
trial including an alprazolam-treated group. J. Clin. Psychopharmacol. S., 2000. Centrally administered murine-leptin stimulates the
21, 59–65. hypothalamus–pituitary–adrenal axis through arginine–vasopressin.
Möller, C., Sommer, W., Thorsell, A., Rimondini, R., Heilig, M., Neuroendocrinology 71, 366–374.
2002. Anxiogenic-like action of centrally administered glucagon-like Morin-Surun, M.P., Collin, T., Denavit-Saubié, M., Baulieu, E.-E.,
peptide-1 in a punished drinking test. Prog. Neuropsychopharmacol. Monnet, F.P., 1999. Intracellular ␴1 receptor modulates phospholipase
Biol. Psychiatry 26, 119–122. C and protein kinase C activities in the brainstem. Proc. Natl. Acad.
Mongeau, R., 1998. Differential effects of neurokinin-1 receptor activation Sci. U.S.A. 96, 8196–8199.
in subregions of the periaqueductal gray matter on conditional and Morisset, S., Rouleau, A., Ligneau, X., Gbahou, F., Tardivel-Lacombe,
unconditional fear behaviors in rats. Behav. Neurosci. 112, 1125–1135. J., Stark, H., Schunack, W., Ganellin, C.R., Schwartz, J.C., Arrang,
Monleon, S., Urquiza, A., Arenas, M.C., Vinader-Caerols, C., Parra, A., J.M., 2000. High constitutive activity of native H3 receptors regulates
2002. Chronic administration of fluoxetine impairs inhibitory avoidance histamine neurons in brain. Nature 408, 860–864.
in male but not female mice. Behav. Brain Res. 136, 483–488. Morrison, C.F., 1969. The effects of nicotine on punished behaviour.
Monnet, F.P., De Costa, B.R., Bowen, W.D., 1996. Differentiation of Psychopharmacology 14, 221–232.
(ligand-activated receptor subtypes that modulate NMDA-evoked Morrow, B.A., Elsworth, J.D., Rasmusson, A.M., Roth, R.H., 1999.
[3 H]noradrenaline release in rat hippocampal slices. Br. J. Pharmacol. The role of mesoprefrontal dopamine neurons in the acquisition and
119, 65–72. expression of conditioned fear in the rat. Neurosci. 92, 553–564.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 219

Morrow, B.A., Elsworth, J.D., Lee, E.J.K., Roth, R.H., 2000. Divergent Naftolin, F., Guadalupe Stanbury, M., 2002. Phytoestrogens: are they
effects of putative anxiolytics on stress-induced Fos expression in the really estrogen mimics? Fertil. Steril. 77, 15–17.
mesoprefrontal system of the rat. Synapse 36, 143–154. Naguib, M., Samarkandi, A.H., 2000. The comparative dose–response
Morrow, A.L., VanDoren, M.J., Penland, S.N., Matthews, D.B., 2001. The effects of melatonin and midazolam for premedication of adult
role of GABAergic neuroactive steroids in ethanol action, tolerance patients: a double-blinded, placebo-controlled study. Anesth. Analg.
and dependence. Brain Res. Rev. 37, 98–109. 91, 473–479.
Moses, E.L., Drevets, W.C., Smith, G., Mathis, C.A., Kalro, B.N., Nagy, J.I., Geiger, J.D., Staines, W.A., 1990. Adenosine deaminase and
Butters, M.A., Leondires, M.P., Greer, P.J., Lopresti, B., Loucks, T.L., purinergic neuroregulation. Neurochem. Int. 16, 211–221.
Berga, S.L., 2000. Effects of estradiol and progesterone administration Nahon, J.-L., 1994. The melanin-concentrating hormone: from the peptide
on human serotonin2A receptor binding: a PET study. Biol. Psychiatry to the gene. Crit. Rev. Neubiol. 8, 221–262.
48, 854–859. Nakajima, M., Inui, A., Azakawa, A., Momose, K., Ueno, N., Teranishi,
Mössner, R., Silke, D., Dietmar, A., Heils, A., Okladnova, O., Schmitt, A., Baba, S., Kasuga, M., 1998. Neuropeptide Y produces anxiety
A., Lesch, K.-P., 2000. Serotonin transporter function is modulated via Y2-type receptors. Peptides 19, 359–363.
by brain-derived neurotrophic factor (BDNF) but not nerve growth Nakamura, K., Kurasawa, M., 2001. Anxiolytic effects of aniracetam
factor (NGF). Neurochem. Int. 36, 197–202. in three different mouse models of anxiety and the underlying
Mothet, J.P., Parent, A.T., Wolosker, H., Brady, R.O., Linden, D.J., mechanism. Eur. J. Pharmacol. 420, 33–43.
Ferris, C.D., Rogawski, M.A., Snyder, S.H., 2000. d-Serine is an Nakaya, Y., Kaneko, T., Shigemoton, R., Nakanishi, S., Mizuno, N.,
endogenous ligand for the glycine site of the N-methyl-d-aspartate 1994. Immunohistochemical localization of substance P receptor in the
receptor. Proc. Natl. Acad. Sci. U.S.A. 97, 4926–4931. central nervous system of the adult rat. J. Comp. Neurol. 347, 249–274.
Motta, V., Maisonnette, S., Morato, S., Castrechini, P., Brandao, M.L., Nakazi, M., Bauer, U., Nickel, T., Kathman, M., Schlicker, E., 2000.
1992. Effects of blockade of 5-HT2 receptors and activation of 5- Inhibition of serotonin release in the mouse brain via presynaptic
HT1A receptors on the exploratory activity of rats in the elevated cannabinoid CB1 receptors. Naunyn Schnmiedebergs Arch. Pharmacol.
plus-maze. Psychopharmacology 107, 135–139. 361, 19–24.
Mountjoy, K.G., Mortrud, M.T., Löw, M.J., Simerly, R.B., Cone, Nalepa, I., Kreiner, G., Kowalska, M., Sanak, M., Zelek-Molik, A.,
R.D., 1994. Localization of the melanocortin-4 receptor (MC4-R) Vetulani, J., 2002. Repeated imipramine and electroconvulsive shock
in neuroendocrine and automatic control circuits in the brain. Mol. increase ␣1A -adrenoceptor mRNA level in rat prefrontal cortex. Eur.
Endocrinol. 8, 1298–1308. J. Pharmacol. 444, 151–159.
Mrzlkak, L., Bergson, C., Pappy, M., Huff, R., Levenson, R., Goldman- Nankai, M., Yamada, S., Muneoka, K., Toru, M., 1995. Increased
Rakic, P.S., 1996. Localization of dopamine D4 receptors in 5-HT2 receptor-mediated behavior 11 days after shock in learned
GABAergic neurons of the primate brain. Nature 381, 245–248. helplessness. Eur. J. Pharmacol. 281, 123–130.
Mueller, N.K., Beck, S.G., 2000. Corticosteroids alter the 5-HT1A Narita, M., Yoshizawa, K., Nomura, M., Aoki, K., Suzuki, T., 2001. Role
receptor-mediated response in CA1 hippocampal pyramidal cells. of the NMDA receptor subunit in the expression of the discriminative
Neuropsychopharmacology 23, 419–427. stimulus effect induced by ketamine. Eur. J. Pharmacol. 423, 41–46.
Muglia, L.J., Jacobson, L., Weninger, S.C., Karalis, K.P., Jeong, D., Nava, F., Carta, G., 2001. Melatonin reduces anxiety induced by
Majzoub, J.A., 2001. The physiology of corticotropin-releasing lipopolysaccharide in the rat. Neurosci. Lett. 307, 57–60.
hormone deficiency in mice. Peptides 22, 725–731. Navarro, J.F., Maldonado, E., 2002. Acute and subchronic effects of
Müller, N., Ackenheil, M., 1998. Psychoneuroimmunology and cytokine MDMA (“ecstasy”) on anxiety in male mice tested in the elevated plus-
actions in the CNS: implications for psychiatric disorders. Prog. maze. Prog. Neuropsychopharmacol. Biol. Psychiatry 26, 1151–1154.
Neuropsychopharmacol. Biol. Psychiatry 22, 1–33. Navarro, M., Fernandez-Ruiz, J.J., de Miguel, R., Hernandez, M.L.,
Müller, M.B., Holsboer, F., Keck, M.E., 2002. Genetic modification of Cebeira, M., Ramos, J.A., 1993. An acute dose of delta 9-
corticosteroid receptor signalling: novel insights into pathophysiology tetrahydrocannabinol affects behavioral and neurochemical indices of
and treatment strategies of human affective disorders. Neuropeptides mesolimbic dopaminergic activity. Behav. Brain Res. 57, 37–46.
36, 117–131. Navarro, M., Hernandez, E., Munoz, R.M., Del Arco, I., Villanua,
Munro, L.J., Kokkinidis, L., 1997. Infusion of quinpirole and muscimol M.A., Carrera, M.R.A., Rodriguez de Fonseca, M.R., 1997. Acute
into the ventral tegmental area inhibits fear-potentiated startle: administration of the CB1 cannabinoid receptor antagonist SR 141716A
implications for the role of dopamine in fear expression. Brain Res. induces anxiety-like responses in the rat. NeuroReport 8, 491–496.
746, 231–238. Navarro, J.F., Buron, E., Martin-Lopez, M., 2002. Anxiogenic-like activity
Murakami, N., Ishibashi, H., Katsurabayashi, S., Akaike, N., 2002. of L-655,708, a selective ligand for the benzodiazepine site of GABAA
Calcium channel subtypes on single GABAergic presynaptic terminals receptors which contain the alpha-5 subunit, in the elevated plus-maze
projecting to rat hippocampal neurons. Brain Res. 951, 121–129. test. Prog. Neuropsychopharmacol. Biol. Psychiatry 26, 1389–1392.
Murphy, B.E., 1997. Antiglucocorticoid therapies in major depression: a Naveilhan, P., Neveu, I., Arenas, E., Ernfors, P., 1998. Complementary and
review. Psychoneuroendocrinology 22, 125–132. overlapping expression of Y1 , Y2 and Y5 receptors in the developing
Murrin, L.C., Gerety, M.E., Happe, H.K., Bylund, D.B., 2000. Inverse and adult mouse nervous system. Neuroscience 87, 289–302.
agonism at ␣2 -adrenoceptors in native tissue. Eur. J. Pharmacol. 398, Naveilhan, P., Canals, J.M., Valjakka, A., Vartiainen, J., Arenas, E.,
185–191. Emfors, P., 2001. Neuropeptide Y alters sedation through a hypo-
Murugaiah, K.D., O’Donnell, J.M., 1995. Beta adrenergic receptors thalamic Y1 -mediated mechanism. Eur. J. Neurosci. 13, 2241–2246.
facilitate norepinephrine release from rat hypothalamic and hippo- Nazar, M., Jessa, M., Plaznik, A., 1997. Benzodiazepine-GABAA receptor
campal slices. Res. Commun. Mol. Pathol. Pharmacol. 90, 179–190. complex ligands in two models of anxiety. J. Neural Transm. 104,
Myers, K.M., Davis, M., 2002. Behavioral and neural analysis of 733–746.
extinction. Neuron 36, 567–584. Nazar, M., Siemiatkowski, M., Bidzinski, A., Czlonkowska, A.,
Nabeshima, T., Kamei, H., Kameyama, T., 1988. A role played by Sienkiewicz-Jarosz, H., Plaźnik, A., 1999a. The influence of serotonin
sigma receptors in the conditioned suppression of motility in mice. depletion on rat behavior in the vogel test and brain 3 H-zolpidem
Psychopharmacology 94, 515. binding. J. Neural Transm. 106, 355–368.
Nadal, A., Ropero, A.B., Laribi, O., Maillet, M., Fuentes, E., Soria, B., Nazar, M., Siemiatkowski, M., Czlonkowska, A., Sienkiewicz-Jarosz,
2000. Nongenomic actions of estrogens and xenoestrogens by binding H., Plaźnik, A., 1999b. The role of hippocampus and 5-HT/GABA
at a plasma membrane receptor unrelated to estrogen receptor alpha and interaction in the central effects of benzodiazepine receptor ligands.
estrogen receptor beta. Proc. Natl. Acad. Sci. U.S.A. 97, 11603–11606. J. Neural Transm. 106, 369–381.
220 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Neal, C.R., Mansour, A., Reinscheid, R., Nothacker, H.P., Civelli, suggests an integral role in central neuroendocrine control. J. Neurosci.
O., Akil, H., Watson, S.J., 1999a. Opioid receptor-like (ORL1) 19, 10295–10304.
receptor distribution in the rat central nervous system: comparison of Nicholas, A.P., Hökfelt, T., Pieribone, V.A., 1996. The distribution and
ORL1 receptor mRNA expression with [(125)]-[(14)Tyr]-orphaninFQ significance of CNS adrenoceptors examined with in situ hybridization.
binding. J. Comp. Neurol. 112, 563–605. Trends Pharmacol. Sci. 18, 210–211.
Neal, C.R., Mansour, A., Reinscheid, R., Nothacker, H.P., Civelli, O., Nickell, P.V., Uhde, T.W., 1994. Dose–response effects of intravenous
Watson, S.J., 1999b. Localization of orphaninFQ (nociceptin) peptide caffeine in normal volunteers. Anxiety 1, 161–168.
and messenger RNA in the central nervous system of the rat. J. Nickola, T.J., Ignatowski, T.A., Spengler, R.N., 2000. Antidepressant
Comp. Neurol. 406, 503–547. drug administration midifies the interactive relationship between
Nedergaard, M., Takano, T., Hansen, A.J., 2002. Beyond the role of alpha(2) -adrenergic sensitivity and levels of TNF in the rat brain. J.
glutamate as a neurotransmitter. Nat. Rev. 3, 748–755. Neuroimmunol. 107, 50–58.
Nestler, E.J., Barrot, M., DiLeone, R.J., Eisch, A.J., Gold, S.J., Monteggia, Nieoullon, A., 2002. Dopamine and the regulation of cognition and
L.M., 2002. Neurobiology of depression. Neuron 34, 13–25. attention. Prog. Neurobiol. 67, 53–83.
Netter, P., Hennig, J., Huwe, S., Olbrich, R., 1998. Personality Nikbakht, M.-R., Stone, T.W., 2001. Suppression of presynaptic responses
related effects of nicotine, mode of application, and expectancies to adenosine by activation of NMDA receptors. Eur. J. Pharmacol.
on performance, emotional states, and desire for smoking. 427, 13–25.
Psychopharmacology 135, 52–62. Nikolaus, S., Huston, J.P., Hasenöhrl, R.U., 2000. Anxiolytic-like effects
Netto, M.S., Silveira, R., Cysne Coimbra, N., Joca, S.R.L., Silveira in rats produced by ventral pallidal injection of both N- and C-terminal
Guimaraes, F., 2002. Anxiogenic effect of median raphe nucleus fragments of substance P. Neurosci. Lett. 283, 37–40.
lesion in stressed rats. Prog. Neuropsychopharmacol. Biol. Psychiatry Ninan, P.T., Rush, J., Crits-Christoph, P., Kornstein, S.G., Manber, R.,
26, 1135–1141. Thase, M.E., Trivedi, M.H., Rothbaum, B.O., Zajecka, J., Borian,
Neumaier, J.F., Sexton, T.J., Yracheta, J., Diaz, A.M., Brownfield, F.E., Keller, M.B., 2002. Symptomatic and syndromal anxiety in
M., 2001. Localization of 5-HT7 receptors in rat brain by chronic forms of major depression: effects of nefazodone, cognitive
immunocytochemistry: in situ hybridization and agonist stimulated behavioural analysis system of psychotherapy, and their combination.
c-fos expression. J. Chem. Neuroanat. 21, 63–73. J. Clin. Psychiatry 63, 434–441.
Nevins, M.E., Anthony, E.W., 1994. Antagonists at the serotonin-3 Ninteman, E.W., Corbin, A.E., Johnston, D.J., Wiley, J.N., Christoffersen,
receptor can reduce the fear-potentiated startle response in the rat: C.L., Meltzer, L.T., Heffner, T.G., 1996. Behavioral profile of the
evidence for different types of anxiolytic activity? J. Pharmacol. D2 /D3 /5-HT1A partial agonist, PD 158771: a potential antipsychotic.
Exp. Ther. 268, 248–268.
Soc. Neurosci. Abstr. 22, 1187.
Newman, M.E., Gur, E., Shapira, B., Lerer, B., 1998. Neurochemical
Nishioka, T., Anselmo-Franci, J.A., Li, P., Callahan, M.F., Morris, M.,
mechanisms of action of ECS: evidence from in vivo studies. J.
1998. Stress increases oxytocin release within the hypothalamic
Electroconvulsive Ther. 14, 153–171.
paraventricular nucleus. Brain Res. 781, 56–60.
Newman-Tancredi, A., Gavaudan, S., Conte, C., Chaput, C., Audinot, V.,
Njung’e, K., Critchley, M.A.E., Handley, S.L., 1993. Effects of beta-
Millan, M.J., 1998. Agonist and antagonist actions of antipsychotic
adrenoceptor ligands in the elevated X-maze “anxiety” model
agents at 5-HT1A receptors: a [35 S]GTP␥S binding study. Eur. J.
and antagonism of the “anxiogenic” response to 8-OH-DPAT. J.
Pharmacol. 355, 245–256.
Psychopharmacol. 7, 173–180.
Newman-Tancredi, A., Cussac, D., Audinot, V., Millan, M.J., 1999.
Noble, F., Roques, B.P., 1999. CCK-B receptor: chemistry, molecular
Actions of roxindole at recombinant human dopamine D2 , D3 and
biology, biochemistry and pharmacology. Prog. Neurobiol. 58, 349–
D4 and serotonin 5-HT1A , 5-HT1B and 5-HT1D receptors. Naunyn
379.
Schmiedebergs Arch. Pharmacol. 359, 447–453.
Noble, F., Roques, B.P., 2002. Phenotypes of mice with invalidation
Newman-Tancredi, A., Chaput, C., Touzard, M., Millan, M.J., 2001.
of cholecystokinin (CCK1 or CCK2 ) receptors. Neuropeptides 36,
Pindolol antagonises G-protein activation at both pre- and postsynaptic
serotonin 5-HT1A receptors: a [35 S]GTP␥S autoradiography study. 157–170.
Naunyn Schmiedebergs Arch. Pharmacol. 363, 391–398. Noble, S., Langtry, H.D., Lamb, H.M., 1998. Zopiclone. An update of
Newman-Tancredi, A., Cussac, D., Marini, L., Millan, M.J., its pharmacology, clinical efficacy and tolerability in the treatment of
2002a. Antibody capture assay reveals bell-shaped concentration- insomnia. Drugs 55, 277–302.
response isotherms for h5-HT1A receptor-mediated G␣i3 activation: Nobre, M.J., Ribeiro dos Santos, N., Aguiar, M.S., Brandao, M.L., 2000.
conformational selection by high-efficacy agonists, and relationship Blockade of ␮- and activation of ␬-opioid receptors in the dorsal
to trafficking of receptor signalling. Mol. Pharmacol. 62, 590–601. periaqueductal gray matter produce defensive behavior in rats in the
Newman-Tancredi, A., Quentric, Y., Touzard, M., Verrièle, L., Carpentier, elevated plus-maze. Eur. J. Pharmacol. 404, 145–151.
N., Millan, M.J., 2002b. Differential actions of antiparkinson agents at Nocjar, C., Roth, B.L., Pehek, E.A., 2002. Localization of 5-HT2A
multiple classes of monoaminergic receptor. III. Agonist and antagonist receptors on dopamine cells in subnuclei of the midbrain A10 cell
properties at serotonin, 5-HT1 and 5-HT2 , receptor subtypes. J. group. Neuroscience 111, 163–176.
Pharmacol. Exp. Ther. 303, 815–822. Nogueira, R.L., Graeff, F.G., 1995. Role of 5-HT receptor subtypes
Ng, G.Y.K., Bertrand, S., Sullivan, R., Ethier, N., Wang, J., Yergey, J., in the modulation of dorsal periaqueductal gray generated aversion.
Belley, M., Trimble, L., Bateman, K., Alder, L., Smith, A., McKerman, Pharmacol. Biochem. Behav. 52, 1–6.
R., Metters, K., O’Neill, G.P., Lacaille, J.-C., Hébert, T.E., 2001. Nomoto, S., Miyake, M., Ohta, M., Funakoshi, A., Miyasaka, K., 1999.
␥-Aminobutyric acid type B receptors with a specific heterodimer Impaired learning and memory in OLETF rats without cholecystokinin
composition and postsynaptic actions in hippocampal neurons are (CCK)-A receptors. Physiol. Behav. 66, 869–872.
targets of anticonvulsant GABApentin action. Mol. Pharmacol. 59, North, R.A., 1989. Drug receptors and the inhibition of nerve cells. Br.
144–152. J. Pharmacol. 98, 13–28.
Nibuya, M., Takahashi, M., Russel, D.S., Duman, R.S., 1999. Repeated Novakovic, S.D., Eglen, R.M., Hunter, J.C., 2001. Regulation of Na+
stress increases catalytic TrkB mRNA in rat hippocampus. Neurosci. channel distribution in the nervous system. Trends Neurosci. 24,
Lett. 267, 81–84. 473–478.
Nichol, K.A., Morey, A., Couzens, M.H., Shine, J., Herzog, H., Nukina, I., Glavin, G.B., LaBella, F.S., 1987. Acute cold-restraint stress
Cunningham, A.M., 1999. Conservation of expression of neuropeptide affects alpha2 -adrenoceptors in specific brain regions of the rat. Brain
Y5 receptor between human and rat hypothalamus and limbic regions Res. 401, 30–33.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 221

Nunes-de-Souza, R., Canto-de-Souza, A., Da-Costa, M., Fornari, R.V., Okuyama, Y., Ishiguro, H., Nankai, M., Shibuya, H., Watanabe, A.,
Graeff, F.G., Pela, I.R., 2000. Anxiety-induced antinociception in Arinami, T., 2000. Identification of a polymorphism in the promoter
mice: effects of systemic and intra-amygdala administration of 8-OH- region of DRD4 associated with the human novelty seeking personality
DPAT and midazolam. Psychopharmacology 150, 300–310. trait. Mol. Psychiatry 5, 64–69.
Nunes-de-Souza, R., Canto-de-Souza, A., Rodgers, R.J., 2002. Effects of Olah, M.E., Stiles, G.L., 1985. The role of receptor structure in
intra-hippocampal infusion of WAY-100635 on plus-maze behavior in determining adenosine receptor activity. Pharmacol. Ther. 85, 55–75.
mice: influence of site of injection and prior test experience. Brain Olah, M.E., Stiles, G.L., 1995. Adenosine receptor subtypes: charac-
Res. 927, 87–96. terization and therapeutic regulation. Annu. Rev. Pharmacol. Toxicol.
Nusser, Z., Sieghart, W., Benke, D., Fritschy, J.-M., Somogyi, P., 1996. 35, 581–606.
Differential synaptic localization of two major ␥-aminobutyric acid Oleskewich, S., Lacaille, J.-C., 1992. Reduction of GABAB inhibitory
type A receptor ␣ subunits on hippocampal pyramidal cells. Proc. postsynaptic potentials by serotonin via pre- and postsynaptic mech-
Natl. Acad. Sci. U.S.A. 93, 11939–11944. anisms in CA3 pyramidal cells of rat hippocampus in vitro. Synapse
Nutt, D., 1996. Early action of nefazodone in anxiety associated with 12, 173–188.
depression. J. Psychopharmacol. 1, 18–21. Olive, M.F., Mehmert, K.K., Messing, R.O., Hodge, C.W., 2000. Reduced
Nutt, D., 2000. Treatment of depression and concomitant anxiety. Eur. operant ethanol self-administration and in vivo mesolimbic dopamine
Neuropsychopharmacol. 10, S433–S437. responses to ethanol in PKCε-deficient mice. Eur. J. Neurosci. 12,
Nutt, D., Malizia, A.L., 2001. New insights into the role of the GABAA - 4131–4140.
benzodiazepine receptor in psychiatric disorder. Br. J. Psychiatry 179, Olive, M.F., Mehmert, K.K., Koenig, H.N., Camarii, R., Kim, J.A.,
390–396. Nannini, M.A., Ou, C.J., Hodge, C.W., 2003. A role for corticotropin
Nyiri, G., Freund, T.F., Somogyi, P., 2001. Input-dependent synaptic releasing factor (CRF) in ethanol consumption, sensitivity, and reward
targeting of ␣2 -subunit-containing GABAA receptors in synapses of as revealed by CRF-deficient mice. Psychopharmacology 165, 181–187.
hippocampal pyramidal cells in the rat. Eur. J. Neurosci. 13, 428–442. Oliver, K.R., Kinsey, A.M., Wainwright, A., Sirinathsinghji, D.J.S., 2000.
Oak, J.N., Oldenhof, J., Van Tol, J.H.M., 2000. The dopamine D4 Localization of 5-HT5A receptor-like immunoreactivity in the rat
receptor: one decade of research. Eur. J. Pharmacol. 405, 303–327. brain. Brain Res. 867, 131–142.
Obara, Y., Nakahata, N., 2002. The signaling pathway of neurotrophic Oliveira-Dos-Santos, A.J., Matsumoto, G., Snow, B.E., Bai, D., Houston,
factor biosynthesis. Drug News Perpect. 15, 290–298. F.P., Whishaw, I.Q., Mariathasan, S., Sasaki, T., Wadeham, A., Ohashi,
Ochiishi, T., Chen, L., Yukawa, A., Saitoh, Y., Sekino, Y., Arai, T., Nakata, P.S., Roder, J.C., Barnes, C.A., Siderovski, D.P., Penninger, J.M.,
H., Miyamoto, H., 1999. Cellular localization of adenosine A1 receptors 2000. Regulation of T cell activation, anxiety, and female aggression
in rat forebrain: immunohistochemical analysis using adenosine A1 by RGS2. Proc. Natl. Acad. Sci. U.S.A. 97, 12272–12277.
receptor-specific monoclonal antibody. J. Comp. Neurol. 411, 301–316. Olivier, B., Molewijk, H.E., Van Der Heyden, J.A., Van Oorschot, R.,
O’Donnell, D., Ahmad, S., Wahlestedt, C., Walker, P., 1999. Expression Ronken, E., Mos, J., Miczek, K.A., 1998. Ultrasonic vocalization in
of the novel galanin receptor subtype GALR2 in the adult rat CNS: rat pups: effects of serotonergic ligands. Neurosci. Biobehav. Rev. 23,
distinct distribution from GALR1. J. Comp. Neurol. 409, 469–481. 215–227.
Ogawa, S., Lubahn, D.B., Korach, K.S., Pfaff, D.W., 1997. Behavioral Olivier, B., Soudijn, W., Van Wijngaarden, I., 1999. The 5-HT1A receptor
effects of estrogen receptor gene disruption in male mice. Proc. Natl. and its ligands: structure and function. Prog. Drug Res. 52, 103–165.
Acad. Sci. U.S.A. 94, 1476–1481. Olivier, B., Van Wijngaarden, I., Soudijn, W., 2000. 5-HT3 receptor
Ohata, H., Arai, K., Shibasaki, T., 2002. Effect of chronic administration antagonists and anxiety: a preclinical and clinical review. Eur.
of a CRF1 receptor antagonist, CRA1000, on locomotor activity and Neuropsychopharmacol. 10, 77–95.
endocrine responses to stress. Eur. J. Pharmacol. 457, 201–206. Olivier, B., Pattij, T., Wood, S.J., Oosting, R., Samyai, Z., Toth, M.,
Ohki-Hamazaki, H., 2000. NeuromedinB . Prog. Neurobiol. 62, 297–312. 2001. The 5-HT1A receptor knock-out mouse and anxiety. Behav.
Ohki-Hamazaki, H., Wada, E., Matsui, K., Wada, K., 1997. Cloning Pharmacol. 12, 439–450.
and expression of the neuromedinB receptor and the third subtype of Onaivi, E.S., Green, M.R., Martin, B.R., 1990. Pharmacological
bombesin receptor genes in the mouse. Brain Res. 762, 165–172. characterization of cannabinoids in the elevated plus maze. J.
Ohl, F., Roedel, A., Storch, C., Holsboer, F., Landgraf, R., 2002. Pharmacol. Exp. Ther. 253, 1002–1009.
Cognitive performance in rats differing in their inborn anxiety. Behav. Onaivi, E.S., Bishop-Robinson, C., Darmani, N.A., Sander-Bush, E.,
Neurosci. 116, 464–471. 1995. Behavioral effects of (±)-1-(2,5-dimethoxy-4-iodophenyl)-2-
Ohl, F., Roedel, A., Binder, E., Holsboer, F., 2003. Impact of high- and aminopropane (DOI) in the elevated plus-maze test. Life Sci. 57,
low-anxiety on cognitive performance in a modified hole board test 2455–2466.
in C57BL/6 and DBA/2 mice. Eur. J. Pharmacol. 17, 128–136. Ordway, G.A., Gambarana, C., Frazer, A., 1988. Quantitative
Ohno-Shosaku, T., Maejima, T., Kano, M., 2001. Endogenous autoradiography of central beta adrenoceptor subtypes: comparison
cannabinoids mediate retrograde signals from depolarized postsynaptic of the effects of chronic treatment with desipramine or centrally
neurons to presynaptic terminals. Neuron 29, 729–738. administered I-isoproterenol. J. Pharmacol. Exp. Ther. 247, 379–389.
Oitzl, M.S., Reichardt, H.M., Joels, M., De Kloet, E.R., 2001. Point Ordway, G.A., Gambarana, C., Tejani-Butt, S.M., Areso, P., Hauptmann,
mutation in the mouse glucocorticoid receptor preventing DNA M., Frazer, A., 1991. Preferential reduction of binding of 125 I-
binding impairs spatial memory. Proc. Natl. Acad. Sci. U.S.A. 98, iodopindolol to beta1 adrenoceptors in the amygdala of rat after
12790–12795. antidepressant treatments. J. Pharmacol. Exp. Ther. 257, 681–690.
Okada, M., Nutt, D.J., Murakami, T., Zhu, G., Kamata, A., Kawata, Y., Oshima, T., Kasuya, Y., Terazawa, E., Nagase, K., Saitoh, Y., Dohi, S.,
Kaneko, S., 2001. Adenosine receptor subtypes modulate two major 2001. The anxiolytic effects of the 5-hydroxytryptamine1A agonist
functional pathways for hippocampal serotonin release. J. Neurosci. tandospirone before otolaryngologic surgery. Anesth. Anal. 93, 1214–
21, 628–640. 1216.
Okuyama, S., Chaki, S., Kawawhima, N., Suzuki, Y., Ogawa, S.-I., Osterlund, M.K., Hurd, Y.L., 1998. Acute 17␤-estradiol treatment down-
Nakazato, A., Kumagai, T., Okubo, T., Tomisawa, K., 1999a. Receptor regulates serotonin 5-HT1A receptor mRNA expression in the limbic
binding, behavioral, and electrophysiological profiles of nonpeptide system of female rats. Mol. Brain Res. 55, 169–172.
corticotropin-releasing factor subtype 1 receptor antagonists CRA1000 Otano, A., Frechilla, D., Cobreros, A., Cruz-Orive, L.M., Insausti,
and CRA1001. J. Pharmacol. Exp. Ther. 289, 926–935. A., Insausti, R., Hamon, M., Del Rio, J., 1999. Anxiogenic-like
Okuyama, S., Sakagawa, T., Chaki, S., Imagawa, Y., Ichiki, T., Inagami, effects and reduced stereological counting of immunolabelled 5-
T., 1999b. Anxiety-like behavior in mice lacking the angiotensin II hydroxytryptamine6 receptors in rat nucleus accumbens by antisense
type-2 receptor. Brain Res. 82, 150–159. oligonucleotides. Neuroscience 92, 1001–1009.
222 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Otsuka, M., Yoshioka, K., 1993. Neurotransmitter functions of mammalian Panocka, I., Massi, M., Lapo, I., Swiderski, T., Kowalczyk, M., Sadowski,
tachykinins. Physiol. Rev. 73, 229–294. B., 2001. Antidepressant-type effect of the NK3 tachykinin receptor
Otto, C., Martin, M., Wolfer, D.P., Lipp, H.P., Maldonado, R., Schütz, G., agonist aminosenktide in mouse lines differing in endogenous opioid
2001. Altered emotional behavior in PACAP-type-I-receptor-deficient system activity. Peptides 22, 1037–1042.
mice. Mol. Brain Res. 92, 78–84. Panula, P., Pirvola, U., Auvinen, S., Airaksinen, M.S., 1989. Histamine-
Ouagazzal, A.-M., Kenny, P.J., File, S.E., 1999. Stimulation of nicotinic immunoreactive nerve fibers in the rat brain. Neuroscience 28, 585–610.
receptors in the lateral septal nucleus increases anxiety. Eur. J. Papadeas, S., Grobin, A.C., Morrow, A.L., 2001. Chronic ethanol
Neurosci. 11, 3957–3962. consumption differentially alters GABAA receptor ␣1 and ␣4 subunit
Owens, M.J., Bissette, G., Nemeroff, C.B., 1989. Acute effects of peptide expression and GABAA receptor-mediated 36 Cl− uptake in
alprazolam and adinazolam on the concentrations of corticotropin- mesocorticolimbic regions of rat brain. Alcohol. Clin. Exp. Res. 25,
releasing factor in the rat brain. Synapse 4, 196–202. 1270–1275.
Owens, M.J., Vargas, A., Knight, D.L., Nemeroff, C.B., 1991. The Papadimitriou, G.N., Kerkhofs, M., Kempenaers, C., Mendlewicz, J.,
effects of alprazolam on corticotropin-releasing factor neurons in the 1988a. EEG sleep in patients with generalized anxiety disorder.
rat brain: acute time course, chronic treatment and abrupt withdrawal. Psychiatry Res. 26, 183–190.
J. Pharmacol. Exp. Ther. 258, 349–356. Papez, J.W., 1937. A proposed mechanism of emotion. Arch. Neurol.
Ozawa, S., Kamiya, H., Tzuzuki, K., 1998. Glutamate receptors in the Psychiatry 38, 725–743.
mammalian central nervous system. Prog. Neurobiol. 54, 581–618. Pardon, M.-C., Gould, G.G., Garcia, A., Phillips, L., Cook, M.C.,
Page, M.E., Abercrombie, E.D., 1997. An analysis of the effects of Miller, S.A., Mason, P.A., Morilak, D.A., 2002. Stress reactivity of
acute and chronic fluoxetine on extracellular norepinephrine in the rat the brain noradrenergic system in three rat strains differing in their
hippocampus during stress. Neuropsychopharmacology 16, 419–425. neuroendocrine and behavioral responses to stress: implications for
Page, M.E., Lucki, I., 2002. Effects of acute and chronic reboxetine susceptibility to stress-related neuropsychiatric disorders. Neuroscience
treatment on stress-induced monoamine efflux in the rat frontal cortex. 115, 229–242.
Neuropsychopharmacology 27, 237–247. Park-Chung, M., Wu, F.-S., Purdy, R.H., Malayes, A.A., Gibbs, T.T.,
Page, M.E., Cryan, J.F., Sullivan, A., Dalvi, A., Saucy, B., Manning, Farb, D.H., 1997. Distinct sites for inverse modulation of N-methyl-
D.R., Lucki, I., 2002. Behavioral and neurochemical effects of 5-{4- d-aspartate receptor by sulphated steroids. Mol. Pharmacol. 52,
[5 - cyano - 3-indolyl)-butyl]-1-piperazinyl}-benzofuran-2-carboxamide 1113–1123.
(EMD 68843): a combined selective inhibitor of serotonin reuptake Parker, R.M., Herzog, H., 1999. Regional distribution of Y-receptor
and 5-hydroxytryptamine 1A receptor partial agonist. J. Pharmacol. subtype mRNAs in rat brain. Eur. J. Neurosci. 4, 1431–1448.
Exp. Ther. 302, 1220–1227. Paronis, C.A., Cox, E.D., Cook, J.M., Bergman, J., 2001. Different types
of GABAA receptors may mediate the anticonflict and response rate-
Pain, L., Gobaille, S., Schleef, C., Aunis, D., Oberling, P., 2002. In vivo
decreasing effects of zaleplon, zolpidem and midazolam in squirrel
dopamine measurements in the nucleus accumbens after nonaesthetic
monkeys. Psychopharmacology 156, 461–468.
and anesthetic dose propofol in rats. Anesth. Analg. 95, 915–919.
Parrott, A.C., 1995. Smoking cessation leads to reduced stress, but why?
Paine, T.A., Jackman, S.L., Olmstead, M.C., 2002. Cocaine-induced
Int. J. Addict. 30, 1509–1516.
anxiety: alleviation by diazepam, but not buspirone, dimenhydrinate
Parrott, A.C., 2000. Human psychopharmacology of ecstasy (MDMA):
or diphenhydramine. Behav. Pharmacol. 13, 511–523.
a review of 15 years of empirical research. Hum. Psychopharmacol.
Palanza, P., 2001. Animal models of anxiety and depression: how are
16, 557–577.
females different? Neurosci Biobehav. Rev. 25, 219–233.
Pashkov, V.N., Hemmings, H.C., 2002. The effects of general anesthetics
Palij, P., Stamford, J.A., 1996. Rauwolscine potentiates the effect of
on norepinephrine release from isolated rat cortical nerve terminals.
desipramine on limbic noradrenaline efflux. NeuroReport 7, 1121–
Anesth. Analg. 95, 1274–1281.
1124.
Paslawski, T., Treit, D., Baker, G.B., George, M., Coutts, R.T., 1996. The
Palmiter, R.D., Erickson, J.C., Hollopeter, G., Baraban, S.C., Schwartz, antidepressant drug phenelzine produces antianxiety effects in the plus-
M.W., 1998. Life without neuropeptide Y. Recent Prog. Horm. Res. maze and increases rat brain GABA. Psychopharmacology 127, 19–24.
53, 163–199. Pasqualetti, M., Ori, M., Nardi, I., Castagna, M., Cassano, G.B., Marazitti,
Paluchowska, M.H., Duszynska, B., Klodzinska, A., Tatarczynska, E., D., 1998. Distribution of the 5-HT5A serotonin receptor mRNA in the
2000. Influence of the aliphatic spacer length on the 5-HT1A receptor human brain. Mol. Brain Res. 56, 1–8.
activity of new arylpiperazines with an indazole system. Pol. J. Patapoutian, A., Reichardt, L.F., 2001. Trk receptors: mediators of
Pharmacol. 52, 209–216. neurotrophin action. Curr. Opin. Neurobiol. 11, 272–280.
Pan, W.-X., McNaughton, N., 2002. The role of the medial Patat, A., Trocjerie, S., Thebaut, J.J., Rosenzweig, P., Dubruc, C.,
supramammillary nucleus in the control of hippocampal theta activity Bianchetti, G., Court, L.A., Morselli, P.L., 1994. EEG profile of
and behaviour in rats. Eur. J. Neurosci. 16, 1797–1809. intravenous zolpidem in healthy volunteers. Psychopharmacology 114,
Pan, Z.Z., Williams, J.T., 1989. GABA- and glutamate-mediated synaptic 138–146.
potentials in rat dorsal raphe neurons in vitro. J. Neurophysiol. 61, Patchev, V.K., Shoaib, M., Holsboer, F., Almeida, O.F.X., 1994. The
719–726. neurosteroid tetrahydroprogesterone counteracts corticotropin-releasing
Pan, Y.-X., Bolan, E., Pasternak, G.W., 2002. Dimerization of morphine hormone-induced anxiety and alters the release and gene expression of
and orphaninFQ/nociceptin receptors: generation of a novel opioid corticotropin-releasing hormone in the rat hypothalamus. Neuroscience
receptor subtype. Biochem. Biophys. Res. Commun. 297, 659–663. 62, 265–271.
Pande, A.C., Davidson, J.R.T., Jefferson, J.W., Janney, C.A., Katzelnick, Patel, S., Naeem, S., Kesingland, A., Froestl, W., Capogna, M., Urban, L.,
D.J., Weisler, R.H., Greist, J.H., Sutherland, S.M., 1999a. Treatment Fox, A., 2001. The effects of GABAB agonists and GABApentin on
of social phobia with GABApentin: a placebo-controlled study. J. mechanical hyperalgesia in models of neuropathic and inflammatory
Clinical Psychopharmacology 19, 341–348. pain in the rat. Pain 90, 217–226.
Pande, A.C., Greiner, M., Adams, J.B., Lydiard, R.B., Pierce, M.V., Paterson, D., Nordberg, A., 2000. Neuronal nicotinic receptors in the
1999b. Placebo-controlled trial of the CCK-B antagonist, CI-988, in human brain. Prog. Neurobiol. 61, 75–111.
panic disorder. Biol. Psychiatry 46, 860–862. Pattij, T., Groenink, L., Hijzen, T.H., Oosting, R.S., Maes, R.A.A.,
Pande, A.C., Crockatt, J.G., Janney, C., Feltner, D.E., 2000. Pregabalin Van Der Gugten, J., Olivier, B., 2002a. Autonomic changes
treatment of generalized anxiety disorder (GAD): three randomized, associated with enhanced anxiety in 5-HT1A receptor knock-out mice.
placebo-controlled trials. Eur. J. Psychiatry 15, 244S. Neuropsychopharmacology 27, 380–390.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 223

Pattij, T., Groenink, L., Oosting, R.S., Van Der Gugten, J., Maes, Phillis, J.W., 2001. Adenosine A2A receptor ligands: effects of neuronal
R.A.A., Olivier, B., 2002b. GABAA -benzodiazepine receptor complex excitability. Drug Dev. Res. 52, 331–336.
sensitivity in 5-HT1A receptor knock-out mice on a 129/Sv background. Piascik, M.T., Perez, D.M., 2001. ␣1 -adrenergic receptors: new insights
Eur. J. Pharmacol. 447, 67–74. and directions. J. Pharmacol. Exp. Ther. 298, 403–410.
Paudice, P., Raiteri, M., 1991. Cholecystokinin release mediated by Piazza, P.V., Deminière, J.M., Le Moal, M., Simon, H., 1990. Stress-
5-HT3 receptors in rat cerebral cortex and nucleus accumbens. Br. J. and pharmacologically-induced behavioural sensitization increases
Pharmacol. 103, 1790–1794. vulnerability to acquisition of amphetamine self-administration. Brain
Pawlak, R., Magarinos, A.M., Melchor, J., McEwen, B., Strickland, S., Res. 514, 22–26.
2003. Tissue plasminogen activator in the amygdala is critical for Piazza, P.V., Rouge-Pont, F., Deroche, V., Maccari, S., Simon, H.,
stress-induced anxiety-like behavior. Nat. Neurosci. 6, 168–174. Le Moal, M., 1996. Glucocorticoids have state-dependent stimulant
Paylor, R., Nguyen, M., Crawley, J.-N., Patrick, J., Beaudet, A., Orr- effects on the mesencephalic dopaminergic transmission. Proc. Natl.
Urtreger, A., 1998. ␣7 Nicotinic receptor subunits are not necessary for Acad. Sci. U.S.A. 93, 8716–8720.
hippocampal-dependent learning or sensorimotor gating: a behavioural Picazo, O., Fernandez-Guasti, A., 1995. Anti-anxiety effects of
characterization of ␣7 -deficient mice. Learn. Mem. 5, 302–316. progesterone and some of its reduced metabolites: an evaluation using
Pecknold, J.C., 1997. A risk-benefit assessment of buspirone in the the burying behavior test. Brain Res. 680, 135–141.
treatment of anxiety disorders. Drug Safety 16, 118–132.
Picciotto, M.R., 1999. Knock-out mouse models used to study
Pehek, E.A., McFarlane, H.G., Maguschak, K., Price, B., Pluto, C.P.,
neurobiological systems. Crit. Rev. Neurobiol. 13, 103–149.
2001. M100,907, a selective 5-HT2A antagonist, attenuates dopamine
Pich, E.M., Samanin, R., 1986. Disinhibitory effects of buspirone and low
release in the rat medial prefrontal cortex. Brain Res. 888, 51–59.
doses of sulpiride and haloperidol in two experimental anxiety models
Pelleymounter, M.A., Joppa, M., Ling, N., Foster, A.C., 2002.
in rats: possible role of dopamine. Psychopharmacology 89, 125–130.
Pharmacological evidence supporting a role for central corticotropin-
releasing factor2 receptors in behavioural, but not endocrine, response Piepponen, T.P., Kiianmaa, K., Ahtee, L., 2002. Effects of ethanol on the
to environmental stress. J. Pharmacol. Exp. Ther. 302, 145–152. accumbal output of dopamine, GABA and glutamate in alcohol-tolerant
Penalva, R.G., Flachskamm, C., Zimmermann, S., Wurst, W., Holsboer, and alcohol-nontolerant rats. Pharmacol. Biochem. Behav. 74, 21–30.
F., Reul, J.M., Linthorst, A.C., 2002. Corticotropin-releasing hormone Pieribone, V.A., Nicholas, A.P., Dagerlind, A., Hökfelt, T., 1994.
receptor type 1 deficiency enhances hippocampal serotonergic Distribution of ␣1 adrenoceptors in rat brain revealed by in
neurotransmission: an in vivo microdialysis study in mutant mice. situ hybridization experiments utilizing subtype-specific probes. J.
Neuroscience 109, 253–266. Neurosci. 14, 4252–4268.
Penn, D.L., Hope, D.A., Spaulding, W., Kucera, J., 1994. Social anxiety Pierrefiche, G., Zerbib, R., Laborit, H., 1993. Anxiolytic activity of
in schizophrenia. Schizophr. Res. 11, 277–284. melatonin in mice: involvement of benzodiazepine receptors. Res.
Perez, S.E., Wynick, D., Steiner, R.A., Mufson, E.J., 2001. Distribution Commun. Chem. Pathol. Pharmacol. 82, 131–141.
of galaninergic immunoreactivity in the brain of the mouse. J. Piggins, H.D., Stamp, J.A., Burns, J., Rusak, B., Semba, K., 1996.
Comp. Neurol. 434, 158–185. Distribution of pituitary adenylate cyclase activating polypeptide
Perez De La Mora, M., Mendez-Franco, J., Salceda, R., Aguirre, K.A., (PACAP) immunoreactivity in the hypothalamus and extended
Fuxe, K., 1991. Neurochemical effects of nicotine on glutamate and amygdala. J. Comp. Neurol. 376, 278–294.
GABA mechanisms in the rat brain. Acta Physiol. Scand. 141, 241–250. Piggott, M., Owens, J., O’Brien, J., Paling, S., Wyper, D., Fenwick, J.,
Pérez-Garcia, C., Morales, L., Cano, M.V., Sancho, I., Alguacil, L.F., Johnson, M., Perry, R., Perry, E., 2002. Comparative distribution of
1999. Effects of histamine H3 receptor ligands in experimental models binding of the muscarinic receptor ligands pirenzepine, AF-DX 384,
of anxiety and depression. Psychopharmaocology 142, 215–220. (R,R)-I-QNB and (R,S)-I-QNB to human brain. J. Chem. Neuroanat.
Peričić, D., Pivac, N., 1995. Sex differences in conflict behaviour and 24, 211–223.
in plasma corticosterone levels. J. Neural Transm. Gen. Sect. 101, Pikkarainen, M., Rönkkö, S., Savander, V., Insausti, R., Pitkänen, A.,
213–221. 1999. Projections from the lateral, basal, and accessory basal nuclei of
Peričić, D., Tvrdeić, A., 1993. Dihydroergosine: anticonflict effect in rat the amygdala to the hippocampal formation in rat. J. Comp. Neurol.
and enhancing effects on [3 H]muscimol binding in the human brain 403, 229–260.
post mortem. Eur. J. Pharmacol. 235, 267–274. Pilc, A., Klodzinska, A., Branski, P., Nowak, G., Palucha, A., Szewczyk,
Pernow, B., 1983. Substance P. Pharmacol. Rev. 35, 85–141. B., Tatarczynska, E., Chojnacka-Wojcik, E., Wieronska, J.M., 2002a.
Pertwee, R.G., 1997. Pharmacology of cannabinoid CB1 and CB2 Multiple MPEP administration evoke anxiolytic- and antidepressant-
receptors. Pharmacol. Ther. 74, 129–180. like effects in rats. Neuropharmacology 43, 181–187.
Pertwee, R.G., 2000. Cannabinoid receptor ligands: clinical and
Pilc, A., Klodzinska, A., Nowak, G., 2002b. A role of glutamate in the
neuropharmacological considerations, relevant to future drug discovery
treatment of anxiety and depression: focus on group I metabotropic
and development. Expert Opin. Investig. Drugs 9, 1553–1571.
glutamate (mGlu) receptors. Drugs Future 27, 753–763.
Petersen, E.N., Buus Lassen, J., 1981. A water lick conflict paradigm
Pillot, C., Heron, A., Cochois, V., Tardivel-Lacombe, J., Ligneau,
using drug experienced rats. Psychopharmacology 75, 236–239.
Petersen, E.N., Braestrup, C., Scheel-Kruger, J., 1985. Evidence that the X., Schwartz, J.-C., Arrang, J.-M., 2002. A detailed mapping of
anticonflict effect of midazolam in amygdala is mediated by specific the histamine H3 receptor and its gene transcripts in rat brain.
benzodiazepine receptors. Neurosci. Lett. 53, 285–288. Neuroscience 114, 173–193.
Petty, F., Trivedi, M.H., Fulton, M., Rush, J., 1995. Benzodiazepines as Pin, J.-P., Duvoisin, R., 1995. The metabotropic glutamate receptors:
antidepressants: does GABA play a role depression? Biol. Psychiatry structure and functions. Neuropharmacology 34, 1–26.
38, 578–591. Pin, J.-P., De Colle, C., Bessis, A.-S., Acher, F., 1999. New perspectives
Pezze, M.A., Heidbreder, C.A., Feldon, J., Murphy, C.A., 2001. Selective for the development of selective metabotropic glutamate receptor
responding of nucleus accumbens core and shell dopamine to aversively ligands. Eur. J. Pharmacol. 375, 277–294.
conditioned contextual and discrete stimuli. Neuroscience 108, 91–102. Pin, J.-P., Parmentier, M.-L., Prézeau, L., 2001. Positive allosteric
Pezze, M.A., Bast, T., Feldon, J., 2003. Significance of dopamine modulators for ␥-aminobutyric acidB receptors open new routes for the
transmission in the rat medial prefrontal cortex for conditioned fear. development of drugs targeting family 3 G-protein-coupled receptors.
Cereb. Cortex 13, 371–380. Mol. Pharmacol. 60, 881–884.
Phan, K., Wager, T., Taylor, S., Liberzon, I., 2002. Functional Pinnock, R.D., Woodruff, G.N., 1991. Bombesin excites a subpopulation
neuroanatomy of emotion: a meta-analysis of emotion activation of 5-hydroxytryptamine-sensitive neurones in the rat dorsal raphe
studies in PET and fMRI. Neuroimage 16, 331–348. nucleus in vitro. J. Physiol. 440, 55–65.
224 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Pinnock, R.D., Reynolds, T., Woodruff, G.N., 1994. Different types of Porcu, P., Sogliano, C., Cinus, M., Purdy, R.H., Biggio, G., Concas, A.,
bombesin receptor on neurons in the dorsal raphe nucleus and the rostral 2003. Nicotine-induced changes in cerebrocortical neuroactive steroids
hypothalamus in rat brain slices in vitro. Brain Res. 653, 119–124. and plasma corticosterone concentrations in the rat. Pharmacol.
Piomelli, D., Giuffrida, A., Calignano, A., Rodriguez de Fonseca, F., Biochem. Behav. 74, 683–690.
2000. The endocannabinoid system as a target for therapeutic drugs. Porter, A.C., Felder, C.C., 2001. The endocannabinoid nervous system:
Trends Pharmacol. Sci. 21, 218–224. unique opportunities for therapeutic intervention. Pharmacol. Ther.
Piper, D., Upton, N., Thomas, D., Nicholas, J., 1988. The effects of 90, 45–60.
the 5-HT3 receptor antagonists BRL43694 and GR38032F in animal Porter, A.C., Bymaster, F.P., DeLapp, N.W., Yamada, M., Wess, J.,
behavioural models of anxiety. Br. J. Pharmacol. 94, 314P. Hamilton, S.E., Nathanson, N.M., Felder, C.C., 2002. M1 muscarinic
Pirker, S., Schwarzer, C., Wieselthaler, A., Wieghart, W., Sperk, G., 2000. receptor signaling in mouse hippocampus and cortex. Brain Res. 944,
GABAA receptors: immunocytochemical distribution of 13 subunits 82–89.
in the adult rat brain. Neuroscience 101, 815–850. Potokar, J., Nutt, D.J., 1994. Anxiolytic potential of benzodiazepine
Pistis, M., Ferraro, L., Pira, L., Flore, G., Tanganelli, S., Gessa, G.L., receptor partial agonists. CNS Drugs 1, 305–315.
Devoto, P., 2002. 9 -Tetrahydrocannabinol decreases extracellular Powell, K.R., Barrett, J.E., 1991. Evaluation of the effects of PD 134308
GABA and increases extracellular glutamate and dopamine levels in (Cl-988), a CCK-B antagonist, on the punished responding of squirrel
the rat prefrontal cortex: an in vivo microdialysis study. Brain Res. monkeys. Neuropeptides 19, 75–78.
948, 155–158. Power, A.E., McGaugh, J.L., 2002. Cholinergic activation of the
Pitchot, W., Ansseau, M., Gonzalez Moreno, A., Hansenne, M., Von basolateral amygdala regulates unlearned freezing behavior in rats.
Frenckell, R., 1992. Dopaminergic function in panic disorder: Behav. Brain Res. 134, 307–315.
comparison with major and minor depression. Biol. Psychiatry 32, Pozzi, L., Acconcia, S., Ceglia, I., Invernizzi, R.W., Samanin, R.,
1004–1011. 2002. Stimulation of 5-hydroxytryptamine (5-HT2C ) receptors in the
Pitkänen, A., Savander, V., LeDoux, J.E., 1997. Organization of ventrotegmental area inhibits stress-induced but not basal dopamine
intra-amygdaloid circuitries in the rat: an emerging framework for release in the rat prefrontal cortex. J. Neurochem. 82, 93–100.
understanding functions of the amygdala. Trends Neurosci. 20, 517– Pozzo-Miller, L.D., Inoue, T., Dieuliis Murphy, D., 1999. Estradiol
523. increases spine density and NMDA-dependent Ca2+ transients in
Pittaluga, A., Feligioni, M., Ghersi, C., Gemignani, A., Raiteri, M., spines of CA1 pyramidal neurons from hippocampal slices. J. Neuro-
2001. Potentiation of NMDA receptor function through somatostatin physiol. 81, 1404–1411.
release: a possible mechanism for the cognition-enhancing activity of Pradhan, A.A.A., Cumming, P., Clarke, P.B.S., 2002. [125 I]Epibatidine-
GABAB receptor antagonists. Neuropharmacology 41, 301–310. labelled nicotinic receptors in the extended striatum and cerebral
cortex: lack of association with serotonergic afferents. Brain Res. 954,
Platt, D.M., Rowlett, J.K., Spealman, R.D., Cook, J., Ma, C., 2002.
227–236.
Selective antagonism of the ataxic effects of zolpidem and tirazolam
Pralong, E., Magistretti, P., Stoop, R., 2002. Cellular perspectives on the
by the GABAA /␣1 -preferring antagonist ␤-CCt in squirrel monkeys.
glutamate–monoamine interactions in limbic lobe structures and their
Psychopharmarmacology 164, 151–159.
relevance for some psychiatric disorders. Prog. Neurobiol. 67, 173–202.
Plaznik, A., Jessa, M., Bidzinski, A., Nazar, M., 1994a. The effect of
Prasad, P.D., Li, H.W., Fei, Y.J., Ganapathy, M.E., Fujita, T., Plumley,
serotonin depletion and intra-hippocampal midazolam on rat behavior
L.H., Yang-Feng, T.L., Leibach, F.H., Ganapathy, V., 1998. Exon–intron
in the Vogel Conflict Test. Eur. J. Pharmacol. 257, 293–296.
structure, analysis of promotor region, and chromosomal localization
Plaznik, A., Nazar, M., Jessa, M., 1994b. The limbic location of some
of the human type 1 sigma receptor gene. J. Neurochem. 70, 443–451.
central effects of competitive and noncompetitive NMDA receptor
Pratt, J.A., 1992. The neuroanatomical basis of anxiety. Pharmacol. Ther.
antagonists: the role in emotional control. Eur. Neuropsychopharmacol.
55, 149–181.
4, 335.
Pribilla, I., Neuhaus, R., Huba, R., Hillmann, M., Turner, J.D., Stephens,
Plaznik, A., Palejko, W., Nazar, M., Jessa, M., 1994c. Effects of
D.N., Schneider, H.H., 1993. Abecarnil is a full agonist at some and
antagonists at the NMDA receptor complex in two models of anxiety.
a partial agonist at other recombinant ␥-aminobutyric acid type A
Eur. Neuropsychopharmacol. 4, 503–512.
receptor subtypes. In: Stephens, N.D. (Ed.), Anxiolytic ␤-Carbolines:
Plihal, W., Krug, R., Pietrowski, R., Fehm, H.L., Born, J., 1996. From Molecular Biology to the Clinic. Springer, Berlin, pp. 50–61.
Corticosteroid receptor mediated effects on mood in humans. Privette, T.H., Terrian, D.M., 1995. Kappa opioid agonists produce
Psychoneuroendocrinology 21, 515–523. anxiolytic-like behavior on the elevated plus-maze. Psychopharmaco-
Poelchen, W., Sieler, D., Wirkner, K., Illes, P., 2001. Co-transmitter logy 118, 444–450.
function of ATP in central catecholaminergic neurons of the rat. Privou, C., Knoche, A., Hasenöhrl, R.U., Huston, J.P., 1998. The H1 -
Neuroscience 102, 593–602. and H2 -histamine blockers chlorpheniramine and ranitidine applied
Pohorecky, L.A., 1990. Interaction of ethanol and stress: research with to the nucleus basalis magnocellularis region modulate anxiety and
experimental animals—an update. Alcohol 25, 263–276. reinforcement related processes. Neuropharmacology 37, 1019–1032.
Pollack, S.J., Harper, S.J., 2002. Small molecule Trk receptor agonists Prut, L., Belzung, C., 2003. The open field as a paradigm to measure the
and other neurotrophic factor mimetics. Curr. Drug. Targets 1, 59–80. effects of drugs on anxiety-like behaviors: a review. Eur. J. Pharmacol.
Pollard, G.T., Howard, J.C.L., 1989. Effects of drugs on punished 463, 3–33.
behaviour: preclinical test for anxiolytics. Pharmacol. Ther. 45, Przegaliński, E., Chojnacka-Wojcik, E., Filip, M., 1992. Stimulation of
403–424. postsynaptic 5-HT1A receptors is responsible for the anticonflict effect
Pontieri, F.E., Tanda, G., Orzi, F., Di Chiara, G., 1996. Effects of nicotine of ipsapirone in rats. J. Pharm. Pharmacol. 44, 780–782.
on the nucleus accumbens and similarity to those of addictive drugs. Przegaliński, E., Malgorzata, F., Chojnacka-Wójcik, E., Tatarczynska,
Nature 382, 255–257. E., 1994a. The role of 5-hydroxytryptamine1A (5-HT1A ) receptors
Popik, P., Van Ree, J.M., 1998. Neurohypophyseal peptides and social in the anticonflict activity of ␤-adrenoceptor antagonists. Pharmacol.
recognition in rats. Prog. Brain Res. 119, 415–436. Biochem. Behav. 47, 873–878.
Popoli, P., Betto, P., Reggio, R., Ricciarrello, G., 1995. Adenosine A2A Przegaliński, E., Tatarczyńska, E., Klodzińska, A., Chojnacka-Wójcik, E.,
receptor stimulation enhances striatal extracellular glutamate levels in 1994b. The role of postsynaptic 5-HT1A receptors in the anticonflict
rats. Eur. J. Pharmacol. 287, 215–217. effect of ipsapirone. Neuropharmacology 33, 1109–1115.
Popoli, M., Gennarelli, M., Racagni, G., 2002. Modulation of synaptic Przegaliński, E., Tatarczyńska, E., Chojnacka-Wójcik, E., 1995. The role
plasticity by stress and antidepressants. Bipolar Disord. 4, 166–182. of hippocampal 5-hydroxytryptamine1A (5-HT1A ) receptors in the
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 225

anticonflict activity of ␤-adrenoceptor antagonists. Neuropharmacology Raffa, R.B., 1998. Possible role(s) of neurokinins in CNS development
34, 1211–1217. and neurodegenerative or other disorders. Neurosci. Biobehav. Rev.
Przegaliński, E., Tatarczyńska, E., Chojnacka-Wójcik, E., 1996. 22, 789–813.
Anticonflict effect of a competitive NMDA receptor antagonist and a Ragnauth, A., Schuller, A., Morgan, M., Chan, J., Ogawa, S., Pintar, J.,
partial agonist at strychnine-insensitive glycine receptors. Pharmacol. Bodnar, R.J., Pfaff, D.W., 2001. Female preproenkephalin-knock-out
Biochem. Behav. 54, 73–77. mice display altered emotional responses. Proc. Natl. Acad. Sci.
Przegaliński, E., Tatarczyńska, E., Klodzińska, A., Chojnacka-Wójcik, U.S.A. 98, 1958–1963.
E., 1999. Tolerance to anxiolytic- and antidepressant-like effects of Rägo, L., Kiivet, R.-A., Harro, J., Pold, M., 1988. Behavioral differences
a partial agonist of glycineB receptors. Pharmacol. Biochem. Behav. in an elevated plus-maze: correlation between anxiety and decreased
64, 461–466. number of GABA and benzodiazepine receptors in mouse cerebral
Przegaliński, E., Tatarczyńska, E., Chojnacka-Wójcik, E., 2000. The cortex. Naunyn Schmiedebergs Arch. Pharmacol. 337, 675–678.
influence of the benzodiazepine receptor antagonist flumazenil Ragsdale, D.S., Avoli, M., 1998. Sodium channels as molecular targets
on the anxiolytic-like effects of CGP37849 and ACPC in rats. for antiepileptic drugs. Brain Res. Rev. 26, 16–28.
Neuropharmacology 39, 1858–1864. Rainbow, T.C., Parsons, B., Wolfe, B.B., 1984. Quantitative
Pudovkina, O.L., Cremers, T.I.F.H., Westerink, B.H.C., 2002. The autoradiography of ␤1 - and ␤2 -adrenergic receptors in rat brain. Proc.
interaction between the locus coeruleus and dorsal raphe nucleus Natl. Acad. Sci. U.S.A. 81, 1585–1589.
studied with dual-probe microdialysis. Eur. J. Pharmacol. 445, 37–42. Rainnie, D.G., 1999. Serotonergic modulation of neurotransmission in
Puia, G., Vicini, S., Seeburg, P.H., Costa, E., 1991. Influence of the rat basolateral amygdala. J. Neurophysiol. 82, 69–85.
recombinant ␥-aminobutyric acid-A receptor subunit composition on Rainville, P., 2002. Brain mechanisms of pain affect and pain modulation.
the action of allosteric modulators of ␥-aminobutyric acid-gated Cl− Curr. Opin. Neurobiol. 12, 195–204.
currents. Mol. Pharmacol. 39, 691–696. Raiteri, M., Paudice, P., Vallebuona, F., 1993. Inhibition by 5-HT3 receptor
Purdy, R.H., Morrow, A.L., Moore, P.H., Paul, S.M., 1991. Stress-induced antagonists of release of cholecystokinin-like immunoreactivity from
elevations of ␦-aminobutyricacid type A receptor-active steroids in the frontal cortex of freely moving rats. Naunyn Schmiedebergs Arch.
the rat brain. Proc. Natl. Acad. Sci. U.S.A. 88, 4553–4557. Pharmacol. 347, 111–114.
Purdy, R.H., Moore, P.H., Morrow, A.L., Paul, S.M., 1992. Neurosteroids Ralevic, V., Burnstock, G., 1998. Receptors for purines and pyrimidines.
and GABAA receptor function. Adv. Biochem. Psychopharmacol. 47, Pharmacol. Rev. 50, 413–492.
87–92. Ramboz, S., Saudou, F., Amara, D.A., Belzung, C., Segu, L., Misslin, R.,
Qiu, J., Lou, L.-G., Huang, X.-Y., Lou, S.-J., Pei, G., Chen, Y.-Z., 1998. Buhot, M.C., Hen, R., 1996. 5-HT1B receptor knock out—behavioral
Nongenomic mechanisms of glucocorticoid inhibition of nicotine- consequences. Behav. Brain Res. 73, 305–312.
induced calcium influx in PC12 cells: involvement of protein kinase Rao, T.S., Cler, J.A., Mick, S.J., Dilworth, V.M., Contreras, P.C.,
C. Endocrinology 139, 5103–5108. Iyengar, S., Wood, P.L., 1990. Neurochemical characterization of
Quartara, L., Maggi, C.A., 1997. The tachykinin NK1 receptor. Part I.
dopaminergic effects of opipramol, a potent sigma receptor ligand, in
Ligands and mechanisms of cellular activation. Neuropeptides 31,
vivo. Neuropsychopharmacology 29, 1191–1197.
537–563.
Rasmussen, K., Kendrick, W.T., Kogan, J.H., Aghajanian, G.K., 1996.
Quinlan, J.J., Firestone, L.L., Homanics, G.E., 2000. Mice lacking the
A selective AMPA antagonist, LY293558, antagonizes morphine
long splice variant of the ␥2 subunit of the GABAA receptor are
withdrawal-induced activation of locus coeruleus neurons and
more sensitive to benzodiazepines. Pharmacol. Biochem. Behav. 66,
behavioral signs of morphine withdrawal. Neuropsychopharmacology
371–374.
15, 497–505.
Quirion, R., Bowen, W.D., Itzhak, Y., Junien, J.L., Musacchio, J.M.,
Rasmusson, A.M., Goldstein, L.E., Deutch, A.Y., Bunney, B.S., Roth,
Rothman, R.B., Su, T.P., Tam, S.W., Taylor, D.P., 1992. A proposal
R.H., 1994. 5-HT1A agonist 8-OH-DPAT modulates basal and stress-
for the classification of sigma binding sites. Trends Pharmacol. Sci.
induced changes in medial prefrontal cortical dopamine. Synapse 18,
13, 85–86.
218–224.
Quirk, J.C., Nisenbaum, E.S., 2002. LY404187: a novel positive allosteric
Ravard, S., Dourish, C.T., 1990. Cholecystokinin and anxiety. Trends
modulator of AMPA receptors. CNS Drug Rev. 8, 255–282.
Quitkin, F.M., Taylor, B.P., Kremer, C., 2001. Does mirtazapine have a Pharmacol. Sci. 11, 271–273.
more rapid onset than SSRIs? J. Clin. Psychiatry 62, 358–361. Ravard, S., Hervé, D., Thiébot, M.-H., Soubrié, P., Tassin, J.-P., 1989.
Quock, R.M., Nguyen, E., 1992. Possible involvement of nitric oxide in Anticonflict-like effect of a prefrontal dopaminergic lesion in rats:
chlordiazepoxide-induced anxiolysis in mice. Life Sci. 51, 255–260. permissive role of noradrenergic neurons. Behav. Pharmacol. 1,
Raber, J., 1998. Detrimental effects of chronic hypothalamic–pituitary– 255–259.
adrenal axis activation. Mol. Neurobiol. 18, 1–22. Ravard, S., Carnoy, P., Hervé, D., Tassin, J.P., Thiébot, M.-H., Soubrié,
Rabiner, E.A., Bhagwagar, Z., Gunn, R.N., Sargent, P.A., Bench, C.J., P., 1990. Involvement of prefrontal dopamine neurones in behavioural
Cowen, P.J., Grasby, P.M., 2001. Pindolol augmentation of selective blockade induced by controllable vs. uncontrollable negative events
serotonin reuptake inhibitors: PET evidence that the dose used in in rats. Behav. Brain Res. 37, 9–18.
clinical trials in too low. Am. J. Psychiatry 158, 2080–2082. Ravard, S., Betschart, J., Fardin, V., Flamand, O., Blanchard, J.C., 1994.
Rademacher, D.J., Schuyler, A.L., Druschel, C.K., Steinpreis, R.E., 2002. Differential ability of tachykinin NK1 and NK2 agonists to produce
Effects of cocaine and putative atypical antipsychotics on rat social scratching and grooming behaviours in mice. Brain Res. 651, 199–208.
behavior: an ethopharmacological study. Pharmacol. Biochem. Behav. Raymond, M.J., Lucan, C.J., Beesley, M.L., O’Connell, B.A., Roberts,
73, 769–778. J.A.F., 1957. A trial of five tranquilizing drugs in psychoneurosis. Br.
Radu, D., Ahlin, A., Svanborg, P., Lindefors, N., 2003. Pentagastrin test Med. J. 2, 63–66.
for anxiety—psychophysiology and personality. Psychopharmacology Raymond, J.R., Mukhin, Y.V., Gettys, T.W., Garnovsakaya, M.N., 1999.
166, 139–145. The recombinant 5-HT1A receptor: G-protein coupling and signalling
Radulovic, J., Rühmann, A., Liepold, T., Spiess, J., 1999. Modulation of pathways. Br. J. Pharmacol. 127, 1751–1764.
learning and anxiety by corticotropin-releasing factor (CRF) and stress: Razandi, M., Pedram, A., Greene, G.L., Levin, E.R., 1999. Cell membrane
differential roles of CRF receptors 1 and 2. J. Neurosci. 19, 5016–5025. and nuclear estrogen receptors (ERs) originate from a single transcript:
Radulovic, J., Fischer, A., Sananbenesi, F., Radulovic, M., Schrick, C., studies of ER␣ and ER␤ expressed in chinese hamster ovary cells.
Kishimoto, T., Rosenfeld, M.G., Spiess, J., 2002. Corticotrophin- Mol. Endocrinol. 13, 307–319.
releasing factor receptor 2 deletion increases anxiety—but decreases Redfern, W.S., Williams, A., 1995. A re-evaluation of the role of
neophobic-like behavior. Soc. Neurosci. Abstr. 370.19. ␣2 -adrenoceptors in the anxiogenic effects of yohimbine, using the
226 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

selective antagonist, delequamine, in the rat. Br. J. Pharmacol. 116, Rickels, K., Schweizer, E., DeMartinis, N., Mandos, L., Mercer, C.,
2081–2089. 1997. Gepirone and diazepam in generalized anxiety disorder: a
Redrobe, J.P., Dumont, Y., Chabot, J.G., Herzog, H., Quirion, R., 2002a. placebo-controlled trial. J. Clin. Psychopharmacol. 17, 272–277.
Neuropeptide Y Y2 receptors mediate behaviour in two animal models Rickels, K., Pollack, M.H., Lydiard, R.B., Bielski, R.J., Fletner, D.E.,
of anxiety evidence from Y2 receptor knock-out mice. Soc. Neurosci. Pande, A.C., Kavoussi, R.J., 2002. Comparison of the efficacy and
Abstr. 398.2. safety of pregabalin and alprazolam in generalized anxiety disorder.
Redrobe, J.P., Dumont, Y., Quirion, R., 2002b. Neuropeptide Y (NPY) Int. J. Neuropsychopharmacol. 5, 14–15.
and depression: from animal studies to the human condition. Life Sci. Riedel, G., Wetzel, W., Reymann, K.G., 1996. Comparing the role of
71, 2921–2937. metabotropic glutamate receptors in long-term potentiation and in
Redrobe, J.P., Dumont, Y., Herzog, H., Quirion, R., 2003. Neuropeptide learning and memory. Prog. Neuropsychopharmacol. Biol. Psychiatry
Y (NPY) Y2 receptors mediate behaviour in two animal models of 20, 761–789.
anxiety: evidence from Y2 receptor knock-out mice. Behav. Brain Riedel, G., Sandager-Nielsen, K., Macphail, E.M., 2002. Impairment of
Res. 141, 251–255. contextual fear conditioning in rats by group I mGluRs: reversal by
Rehfeld, J.H., Nielsen, F.C., 1995. Molecular forms and regional the nootropic nefiracetam. Pharmacol. Biochem. Behav. 73, 391–399.
distribution of cholecystokinin in the central cervous system. In: Risold, P.Y., Swanson, L.W., 1997. Connections of the rat lateral septal
Bradwejn, J., Vasar, E. (Eds.), Cholecystokinin and Anxiety. RG complex. Brain Res. Rev. 24, 115–195.
Landes Company, Georgetown, pp. 33–56. Rivet, J.-M., Melon, C., Millan, M.J., 2001. Blockade of DOI-induced
Reiman, E.M., Fusselman, M.J., Fox, P.T., Raichle, M.E., 1989. Neuro- corticosterone secretion in rats by diverse antidepressant agents reflects
anatomical correlates of anticipatory anxiety. Science 243, 1071–1074. antagonist properties at 5-HT2A receptors. In: O’Connor, W.T., Lowry,
Reimann, F., Ashcroft, F.M., 1999. Inwardly rectifying potassium J.P., O’Connor, J.J., O’Neill, R.D. (Eds.), Monitoring Molecules in
channels. Curr. Opin. Cell Biol. 11, 503–508. Neuroscience. University College Dublin, Dublin, pp. 407–408.
Reinscheid, R.K., Civelli, O., 2002. The orphaninFQ/nociceptin knock- Rivkees, S.A., Price, S.L., Zhou, F.C., 1995. Immunohistochemical
out mouse: a behavioral model for stress responses. Neuropeptides detection of A1 adenosine receptors in rat brain with emphasis on
36, 72–76. localization in the hippocampal formation, cerebral cortex, cerebellum,
Reinscheid, R.K., Nothacker, H.P., Civelli, O., 2000. The orphaninFQ/ and basal ganglia. Brain Res. 677, 193–203.
nociceptin gene: structure, tissue distribution of expression and Rivkees, S.A., Thevananther, S., Hao, H., 2000. Are A3 adenosine
functional implications obtained from knock-out mice. Peptides 21, receptors expressed in the brain? NeuroReport 11, 1025–1030.
Robe, D., Alonso, G., Duchamp, F., Bockaert, J., Manzoni, O.J., 2001.
901–906.
Localization and mechanisms of action of cannabinoid receptors at
Reiriz, J., Holm, P.C., Alberch, J., Arenas, E., 2002. BMP2 and CAMP
glutamatergic synapses of the mouse nucleus accumbens. J. Neurosci.
elevation confer locus coeruleus neurons responsiveness to multiple
21, 109–116.
neurotrophic factors. J. Neurobiol. 50, 291–304.
Robello, M., Amico, C., Bucossi, G., Cupello, A., Rapallino, M.V.,
René, F., Muller, A., Jover, E., Kieffer, B., Koch, B., Loeffler, J.-P.,
Thellung, S., 1996. Nitric oxide and GABAA function in the rat
1998. Melanocortin receptors and ␦-opioidreceptor mediate opposite
cerebral cortex and cerebellar granule cells. Neuroscience 74, 99–106.
signalling actions of POMC-derived peptides in CATH.a cells. Eur. J.
Roberts, A.J., Cole, M., Koob, G.F., 1996. Intra-amygdala muscimol
Neurosci. 10, 1885–1894.
decreases operant ethanol self-administration in dependent rats.
Reppert, S.M., Godson, C., Mahle, C.D., Weaver, D.R., Slaugenhaupt,
Alcohol. Clin. Exp. Res. 20, 1289–1298.
S.A., Gusella, J.F., 1995. Molecular characterization of a second
Roberts, C., Allen, L., Langmead, C.J., Hagan, J.J., Middlemiss, D.N.,
melatonin receptor expressed in human retina and brain: the Mel1b
Price, G.W., 2001. The effect of SB-269970, a 5-HT7 receptor
melatonin receptor. Proc. Natl. Acad. Sci. U.S.A. 92, 8734–8738.
antagonist, on 5-HT release from serotonergic terminals and cell
Reppert, S.M., Weaver, D.R., Godson, C., 1996. Melatonin receptors
bodies. Br. J. Pharmacol. 132, 1574–1580.
step into the light: cloning and classification of subtypes. Trends Roberts, C., Thomas, D.R., Kew, J.N.C., 2002a. GABAergic modulation
Pharmacol. Sci. 17, 100–102. of 5-HT7 receptor mediated effects on 5-HT efflux: an in vitro fast
Reul, J.M.H.M., Holsboer, F., 2002. On the role of corticotropin- cyclic voltametry study. Soc. Neurosci. Abstr. 398.17.
releasing hormone receptors in anxiety and depression. Dialogues Roberts, J.C., Reavill, C., East, S.Z., Harrison, P.J., Patel, S., Routledge,
Clin. Neurosci. 4, 31–46. C., Leslie, R.A., 2002b. The distribution of 5-HT6 receptors in rat
Reyes, T.M., Lewis, K., Perrin, M.H., 2001. Urocortin II: a member of brain: an autoradiographic binding study using the radiolabelled
the corticotrophin-releasing factor (CRF) neuropeptide family that is 5-HT6 receptor antagonist [125 I]SB-258585. Brain Res. 934, 49–57.
selectively bound by type 2 CRF receptors. Proc. Natl. Acad. Sci. Robinson, R.T., Drafts, B.C., Fisher, J.L., 2003. Fluoxetine increases
U.S.A. 98, 2843–2848. GABAA receptor activity through a novel modulatory site. J.
Reynolds, D.S., McKernan, R.P., Dawson, G.R., 2001. Anxiolytic-like Pharmacol. Exp. Ther. 304, 978–984.
action of diazepam: which GABAA receptor subtype is involved? Robson, P., 2001. Therapeutic aspects of cannabis and cannabinoids. Br.
Trends Pharmacol. Sci. 22, 402. J. Psychiatry 178, 107–115.
Ribeiro, R.L., De Lima, T.C.M., 2002. Participation of GABAA receptors Roca, C.A., Schmidt, P.J., Smith, M.J., Danaceau, M.A., Murphy,
in the modulation of experimental anxiety by tachykinin agonists and D.L., Rubinow, D.R., 2002. Effects of metergoline on symptoms in
antagonists in mice. Prog. Neuropsychopharmacol. Biol. Psychiatry women with premenstrual dysphoric disorder. Am. J. Psychiatry 159,
26, 861–869. 1876–1881.
Ribeiro, E.B., Bettiker, R.L., Bogdanov, M., Wurtman, R.J., 1993. Effects Rocha, B., Rigo, M., Di Scala, G., Sandner, G., Hoyer, D., 1994.
of systemic nicotine on serotonin release in rat brain. Brain Res. 621, Chronic mianserin or eltoprazine treatment in rats: effects on the
311–318. elevated plus-maze test and on limbic 5-HT2C receptor levels. Eur. J.
Ribeiro, S.J., Teixeira, R.M., Calixto, J.B., De Lima, T.C.M., 1999. Pharmacol. 262, 125–131.
Tachykinin NK3 receptor involvement in anxiety. Neuropeptides 33, Rocha, B.A., Goulding, E.H., O’Dell, L.E., Mead, A.N., Coufal, N.G.,
181–188. Parsons, L.H., Tecott, L.H., 2002. Enhanced locomotor, reinforcing, and
Ribeiro-da-Silva, A., Hökfelt, T., 2000. Neuroanatomical localisation neurochemical effects of cocaine in serotonin 5-hydroxytryptamine2C
of substance P in the CNS and sensory neurons. Neuropeptides 34, receptor mutant mice. J. Neurosci. 22, 10039–10045.
256–271. Roche, M., Commons, K.G., Peoples, A., Valentino, R.J., 2003. Circuitry
Richter-Levin, G., Akirav, I., 2000. Amygdala–hippocampus dynamic underlying regulation of the serotonergic system by swim stress. J.
interaction in relation to memory. Mol. Neurobiol. 22, 11–20. Neurosci. 23, 970–977.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 227

Rocher, C., Jacquot, C., Gardier, A.M., 1999. Simultaneous effects Rominger, A., Förster, S., Zentner, J., Dooley, D.J., McKnight, A.T.,
of local dexfenfluramine application on extracellular glutamate and Feuerstein, T.J., Jackisch, R., Vlaskovska, M., 2002. Comparison of
serotonin levels in rat frontal cortex: a reverse microdialysis study. the ORL1 receptor-mediated inhibition of noradrenaline release in
Neuropharmacology 38, 513–523. human and rat neocortical slices. Br. J. Pharmacol. 135, 800–806.
Rochford, J., Beaulieu, S., Rousse, I., Glowa, J.-R., Barden, N., 1997. Rondi-Reig, L., Lemaigre Dubreuil, Y., Martinou, J.-C., Delhaye-
Behavioral reactivity to aversive stimuli in a transgenic mouse model Bouchaud, N., Caston, J., Mariani, J., 1997. Fear decrease in transgenic
of impaired glucocorticoid (type II) receptor function: effects of mice over-expressing blc-2 in neurons. NeuroReport 8, 2429–2432.
diazepam and FG-7142. Psychopharmacology 132, 145–152. Roozendaal, B., Bohus, B., McGaugh, J.L., 1996. Dose-dependent
Rodgers, R.J., 1997. Animal models of “anxiety”: where next? Behav. suppression of adrenocortical activity with metyrapone: effects on
Pharmacol. 8, 477–796. emotion and memory. Psychoneuroendocrinology 21, 681–693.
Rodgers, R.J., Cole, J.C., 1995. Effects of scopolamine and its quaternary Roozendaal, B., Nguyen, B.T., Power, A.E., McGaugh, J.L., 1999.
analogue in the murine elevated plus-maze test of anxiety. Behav. Basolateral amygdala noradrenergic influence enables enhancement
Pharmacol. 6, 283–289. of memory consolidation induced by hippocampal glucocorticoid
Rodgers, R.J., Johnson, N.J.T., 1995. Factor analysis of spatiotemporal receptor activation. Proc. Natl. Acad. Sci. U.S.A. 96, 11642–11647.
and ethological measures in the murine elevated plus-maze test of Roozendaal, B., Brunson, K.L., Holloway, B.L., McGaugh, J.L.,
anxiety. Pharmacol. Biochem. Behav. 52, 297–303. Baram, T.Z., 2002. Involvement of stress-released corticotrophin-
Rodgers, R.J., Nikulina, E.M., Cole, J.C., 1994. Dopamine D1 and releasing hormone in the basolateral amygdala in regulating memory
D2 receptor ligands modulate the behaviour of mice in the elevated consolidation. Proc. Natl. Acad. Sci. U.S.A. 99, 13908–13913.
plus-maze. Pharmacol. Biochem. Behav. 49, 985–995. Ropert, N., Guy, N., 1991. Serotonin facilitates GABAergic transmission
Rodgers, R.J., Cao, B.-J., Dalvi, A., Holmes, A., 1997. Animal models in the CA1 region of rat hippocampus in vitro. J. Physiol. 441, 121–136.
of anxiety: an ethological perspective. Braz. J. Med. Biol. Res. 30, Rosa, M.L., Guimaraes, F.S., Pearson, R.C., Del Bel, E.A., 2002. Effects
289–304. of single or repeated restraint stress on GluR1 and GluR2 flip and
Rodgers, D.C., Costall, B., Domeney, A.M., Gerrard, P.A., Greener, flop mRNA expression in the hippocampal formation. Brain Res.
M., Kelly, M.E., Hagan, J.J., Hunter, A.J., 2000. Anxiolytic Bull. 59, 117–124.
profile of ropinirole in the rat, mouse and common marmoset. Roselli-Rehfuss, L., Mountjoy, K.G., Robbins, L.S., Mortrud, M.T.,
Psychopharmacology 151, 91–97. Löw, M.J., Tatro, J.B., Entwistle, M.L., Simerly, R.B., Cone, R.D.,
Rodgers, R.J., Boullier, E., Chatzimichalaki, P., Cooper, G.D., Shorten, 1993. Identification of a receptor for ␥ melanotropin and other
A., 2002a. Contrasting phenotypes of C57BL/6JolaHsd, 129S2/SvHsd proopiomelanocortin peptides in the hypothalamus and limbic system.
and 129/SvEv mice in two exploration-based tests of anxiety-related Proc. Natl. Acad. Sci. U.S.A. 90, 8856–8860.
Rosen, J.B., Schulkin, J., 1998. Froom normal fear to pathological
behaviour. Physiol. Behav. 77, 301–310.
anxiety. Psychopharmacol. Rev. 10, 325–350.
Rodgers, R.J., Davies, B., Shore, R., 2002b. Absence of anxiolytic
Rosenkranz, J.A., Grace, A.A., 2001. Dopamine attenuates prefrontal
response to chlordiazepoxide in two common background strains
cortical suppression of sensory inputs to the basolateral amygdala of
exposed to the elevated plus-maze: importance and implications of
rats. J. Neurosci. 21, 4090–4103.
behavioural baseline. Genes Brain Behav. 1, 241–251.
Rosenkranz, J.A., Grace, A.A., 2002. Cellular mechanisms of infralimbic
Rodriguez, M., Beauverger, P., Naime, I., Rique, H., Ouvry, C., Souchaud,
and prelimbic prefrontal cortical inhibition and dopaminergic
S., Dromaint, S., Nagel, N., Suply, T., Audinot, V., Boutin, J.A.,
modulation of basolateral amygdala neurons in vivo. J. Neurosci. 22,
Galizzi, J.P., 2001. Cloning and molecular characterization of the
324–337.
novel human melanin-concentrating hormone receptor MCH2 . Mol.
Rosin, D.L., Talley, E.M., Lee, A., Stornetta, R.L., Gaylinn, B.D.,
Pharmacol. 60, 632–639.
Guyenet, P.G., Lynch, K.R., 1996. Distribution of ␣2C -adrenergic
Rodrigues, S.M., Bauer, E.P., Farb, C.R., Schafe, G.E., LeDoux, J.E.,
receptor-like immunoreactivity in the rat central nervous system. J.
2002. The group I metabotropic, glutamate receptor mGluR is required
Comp. Neurol. 372, 135–165.
for fear memory formation and long-term potentiation in the lateral Rosin, D.L., Robeva, A., Woodard, R.L., Guyenet, P.G., Linden, J., 1998.
amygdala. J. Neurosci. 22, 5219–5229. Immunohistochemical localization of adenosine A2A receptors in the
Rodriguez-Arias, M., Felip, C.M., Broseta, I., Minarro, J., 1999. The rat central nervous system. J. Comp. Neurol. 401, 163–186.
dopamine D3 antagonist U-99194A maleate increases social behaviors Ross, S.A., Wong, J.Y.E., Clifford, J.J., Kinsella, A., Massalas, J.S.,
of isolation-induced aggressive male mice. Psychopharmacology 144, Horne, M.K., Scheffer, I.E., Kola, I., Waddington, J.L., Berkovic,
90–94. S.F., Drago, J., 2000. Phenotype characterization of an ␣4 neuronal
Rodriguez de Fonseca, F., Rubio, P., Menzahi, F., Merlo-Pich, E., nicotinic acetylcholine receptor subunit knock-out mouse. J. Neurosci.
Rivier, J., Koob, G.F., Navarro, M., 1996. Corticotrophin-releasing 20, 6431–6441.
factor (CRF) antagonist (d-Phe12, Nle21,38,C alpha MeLeu37) Rossetti, Z.L., Carboni, S., 1995. Ethanol withdrawal is associated with
CRF attenuates the acute actions of the highly potent cannabinoid increased extracellular glutamate in the rat striatum. Eur. J. Pharmacol.
receptor agonist HU-210 on defensive-withdrawal behavior in rats. J. 283, 177–183.
Pharmacol. Exp. Ther. 276, 56–64. Rossetti, Z.L., Pani, L., Portas, C., Gessa, G., 1989. Brain dialysis provides
Rodriguez de Fonseca, F., Carrera, M.R., Navarro, M., Koob, G.F., Weiss, evidence for D2 -dopamine receptors modulating noradrenaline release
F., 1997. Activation of corticotrophin-releasing factor in the limbic in the rat frontal cortex. Eur. J. Pharmacol. 163, 393–395.
system during cannabinoid withdrawal. Science 276, 2050–2054. Rostène, W.H., Alexander, M.J., 1997. Neurotensin and neuroendocrine
Rogoz, Z., Klodzinska, A., Maj, J., 2000. The anxiolytic-like effect regulation. Front. Neuroendocrinol. 18, 115–173.
of nafadotride, a dopamine D3 receptor antagonist, in rats. Eur. Roth, T., 2001. The relationship between psychiatric diseases and
Neuropsychopharmacol. 10, S349. insomnia. Int. J. Clin. Pract. 116, 3–8.
Rollema, H., Lu, Y., Schmidt, A.W., Sprouse, J.S., Zorn, S.H., 2000. 5- Rougé-Pont, F., Deroche, V., Le Moal, M., Piazza, P.V., 1998. Individual
HT1A receptor activation contributes to ziprasidone-induced dopamine differences in stress-induced dopamine release in the nucleus
release in the rat prefrontal cortex. Biol. Psychiatry 48, 229–237. accumbens are influenced by corticosterone. Eur. J. Neurosci. 10,
Romero, G., Pérez, M.P., Carceller, A., Monroy, X., Farré, A.J., Guitart, 3903–3907.
X., 2000. Changes in phosphoinositide signalling activity and levels Rowe, W., Viau, V., Meaney, M.J., Quirion, R., 1995. Stimulation
of the alpha subunit of Gq/11 protein in rat brain induced by E-5842, a of CRH-mediated ACTH secretion by central administration of
sigma1 receptor ligand and potential atypical antipsychotic. Neurosci. neurotensin: evidence for the participation of the paraventricular
Lett. 290, 189–192. nucleus. J. Neuroendocrinol. 7, 109–117.
228 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Rowlett, J.K., Tornatzky, W., Cook, J.M., Ma, C., Miczek, K.A., 2001. Ryan, R.R., Katsuno, T., Mantey, S.A., Pradhan, T.K., Weber, H.C., Coy,
Zolpidem, triazolam, and diazepam decrease distress vocalizations in D.H., Battey, J.F., Jensen, R.T., 1999. Comparative pharmacology
mouse pups: differential antagonism by flumazenil and ␤-carbonline- of the nonpeptide neuromedinB receptor antagonist PD 168368. J.
3-carboxylate-t-butyl ester (␤-CCt). J. Pharmacol. Exp. Ther. 297, Pharmacol. Exp. Ther. 290, 1202–1211.
247–253. Saarelainen, T., Hendolin, P., Lucas, G., Koponen, E., Sairanen, M.,
Rowlett, J.K., Lelas, S., Spealman, R.D., 2002. Anti-conflict effects of MacDonald, E., Agerman, K., Haapasalo, A., Nawa, H., Aloyz, R.,
benzodiazepines in monkeys: GABAA receptor mechanisms. Soc. Ernfors, P., Castrén, E., 2003. Activation of the TrkB neurotrophin
Neurosci. Abstr. 396.11. receptor is induced by antidepressant drugs and is required for
Rowlett, J.K., Spealman, R.D., Lelas, S., Cook, J.M., Yin, W., 2003. antidepressant-induced behavioral effects. J. Neurosci. 23, 349–357.
Discriminative stimulus effects of zolpidem in squirrel monkeys: role Sadja, R., Alagem, N., Reuveny, E., 2002. Graded contribution of the
of GABAA/␣1 receptors. Psychopharmacology 165, 209–215. G␤␥ binding domains to GIRK channel activation. Proc. Natl. Acad.
Roy-Byrne, P.P., Uhde, T.W., Sack, D.A., Linnoila, M., Post, R.M., Sci. U.S.A. 99, 10783–10788.
1986. Plasma HVA and anxiety in patients with panic disorder. Biol. Saffroy, M., Torrens, Y., Glowinski, J., Beaujouan, J.-C., 2001. Presence
Psychiatry 21, 849–853. of NK2 binding sites in the rat brain. J. Neurochem. 79, 985–996.
Saffroy, M., Torrens, Y., Glowinski, J., Beaujouan, J.-C., 2003.
Rubinow, D.R., Schmidt, P.J., Roca, C.A., 1998. Estrogen–serotonin
Autoradiographic distribution of tachykinin NK2 binding sites in the
interactions: implications for affective regulation. Biol. Psychiatry 44,
rat brain: comparison with NK1 and NK3 binding sites. Neuroscience
839–850.
116, 761–773.
Rubinstein, M., Phillips, T.J., Nunzow, J.R., Falzone, T.L.,
Saito, Y., Cheng, M., Leslie, F.M., Civelli, O., 2001. Expression of the
Dziewczapolski, G., Zhang, G., Fang, Y., Larson, J.L., McDougall,
melanin-concentrating hormone (MCH) receptor mRNA in the rat
J.A., Chestor, J.A., Saez, C., Pugsley, T.A., Gershanik, O., Löw,
brain. J. Comp. Neurol. 435, 26–40.
M.J., Grandy, D.K., 1997. Mice lacking dopamine D4 receptoros are Sajdyk, T.J., Gehlert, D.R., 2000. Astressin, a corticotropin releasing
supersensitive to ethanol, cocaine, and methamphetamine. Cell 90, factor antagonist, reverses the anxiogenic effects of urocortin when
991–1001. administered into the basolateral amygdala. Brain Res. 877, 226–234.
Rudolph, U., Crestani, F., Benke, D., Brünig, I., Benson, J.A., Fritschy, Sajdyk, T.J., Shekhar, A., 1997. Excitatory amino acid receptors in
J.M., Martin, J.A., Bluethmann, H., Möhler, H., 1999. Benzodiazepine the basolateral amygdala regulate anxiety responses in the social
actions mediated by specific ␥-aminobutyric acidA receptor subtypes. interaction test. Brain Res. 764, 262–264.
Nature 401, 796–800. Sajdyk, T.J., Schober, D.A., Gelhert, D.R., Shekhar, A., 1999a. Role
Ruel, J., Guitton, M.J., Puel, J.-L., 2002. Negative allosteric modulation of corticotropin-releasing factor and urocortin within the basolateral
of AMPA-preferring receptors by the selective isomer GYKI 53784 amygdala of rats in anxiety and panic responses. Behav. Brain Res.
(LY303070), a specific non-competitive AMPA antagonist. CNS Drug 100, 207–215.
Rev. 8, 235–254. Sajdyk, T.J., Vandergriff, M.G., Gelhert, D.R., 1999b. Amygdalar
Rueter, L.E., Fornal, C.A., Jacobs, B.L., 1997. A critical review of 5-HT neuropeptide Y Y1 receptors mediate the anxiolytic-like actions of
brain microdialysis and behavior. Rev. Neurosci. 8, 117–137. neuropeptide Y in the social interaction test. Eur. J. Pharmacol. 368,
Ruggiero, D.A., Underwood, M.D., Rice, P.M., Mann, J.J., Arango, 143–147.
V., 1999. Corticotropic-releasing hormones and serotonin interact Sajdyk, T.J., Schober, D.A., Gelhert, D.R., 2002a. Neuropeptide Y
in the human brainstem: behavioral implications. Neuroscience 91, receptor subtypes in the basolateral nucleus of the amygdala modulate
1343–1354. anxiogenic responses in rats. Neuropharmacology 43, 1165–1172.
Rupniak, N.M.J., Kramer, M.S., 1999. Discovery of the anti-depressant Sajdyk, T.J., Schober, D.A., Smiley, D.L., Gehlert, D.R., 2002b.
and anti-emetic efficacy of substance P receptor (NK1 ) antagonists. Neuropeptide Y Y2 receptors mediate anxiety in the amygdala.
Trends Pharmacol. Sci. 20, 485–490. Pharmacol. Biochem. Behav. 71, 419–423.
Rupniak, N.M.J., Carlson, E.-C., Harrison, T., Oates, B., Seward, Sajdyk, T.J., Zink, C., Gackenheimer, S.L., Fitz, S.D., Shekhar, A.,
E., Owen, S., De Felipe, C., Hunt, S., Wheeldon, A., 2000. Gehlert, D.R., 2002c. The nociceptin/ORL1 system modulates anxiety-
Pharmacological blockade or genetic deletion of substance P NK(1) like behavior in a modified social interaction test. Soc. Neurosci.
receptors attenuates neonatal vocalisation in guinea-pigs and mice. Abstr. 396.5.
Neuropharmacology 39, 1413–1421. Sakamoto, H., Matsumoto, K., Ohno, Y., Nakamura, M., 1998. Anxiolytic-
like effects of perospirone, a novel serotonin2 and dopamine2
Rupniak, N.M.J., Carlson, E.-C., Webb, J.K., Harrison, T., Porsolt,
antagonist (SDA)-type antipsychotic agent. Pharmacol. Biochem.
R.D., Roux, S., De Felipe, C., Hunt, S.P., Oates, B., Wheeldon,
Behav. 60, 873–878.
A., 2001. Comparison of the phenotype of NK1R −/− mice with
Sakaue, M., Hoffman, B.B., 1991. Glucocorticoids induce transcription
pharmacological blockade of the substance P (NK1 ) receptor in
and expression of the alpha-1B adrenergic receptor gene in DTT1
assays for antidepressant and anxiolytic drugs. Behav. Pharmacol. 12,
MF-2 smooth muscle cells. J. Clin. Invest. 88, 385–389.
497–508.
Sakaue, M., Ago, Y., Murakami, C., Sowa, C., Sakamoto, Y., Koyama,
Rupprecht, R., Holsboer, F., 1999. Neuroactive steroids: mechanisms of
Y., Baba, A., Matsuda, T., 2001. Involvement of benzodiazepine
action and neuropsychopharmacological perspectives. Trends Neurosci.
binding sites in an antiaggressive effect by 5-HT1A receptor activation
22, 410–416.
in isolated mice. Eur. J. Pharmacol. 432, 163–166.
Rupprecht, R., Di Michele, F., Hermann, B., Ströhle, A., Lancel, M., Sakic, B., Szechtman, H., Talangbayan, H., Denburg, S.D., Carbotte,
Romeo, E., Holsboer, F., 2001. Neuroactive steroids: molecular R.M., Denburg, J.A., 1994. Disturbed emotionality in autoimmune
mechanisms of action and implications for neuropsychopharmacology. MRL-1pr mice. Physiol. Behav. 56, 609–617.
Brain Res. Rev. 37, 59–67. Salas, R., Pieri, F., Fung, B., Dani, J.A., De Biasi, M., 2002. Altered
Rush, C.R., 1998. Behavioral pharmacology of zolpidem relative to anxiety-related responses in mutant mice lacking the ␤4 subunit of
benzodiazepines: a review. Pharmacol. Biochem. Behav. 61, 253–269. the nicotinic receptor. Soc. Neurosci. Abstr. 283.6.
Rush, C.R., Frey, J.M., Griffiths, R.R., 1999. Zaleplon and triazolam Saldivar-Gonzalez, J.A., Posadas-Andrews, A., Rodriguez, R., Gomez,
in humans: acute behavioural effects and abuse potential. C., Hernandez-Manjarrez, M.E., Ortiz-Leon, S., Martinez-Pineda, A.,
Psychopharmacology 145, 39–51. Gomez-Laguna, D., Salgado, V., Manjarrez, J., Alvarado, R., 2003.
Russel, A., Banes, A., Berlin, H., Fink, G.D., Watts, S.W., 2002. 5- Effect of electrical stimulation of the baso-lateral amygdala nucleus
Hydroxytryptamine2B receptor function is enhanced in the N ␻ -nitro- on defensive burying shock probe test and elevated plus maze in rats.
l-arginine hypertensive rat. J. Pharmacol. Exp. Ther. 303, 179–187. Life Sci. 72, 819–829.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 229

Sallinen, J., Link, R.E., Haapalinna, A., Viitamma, T., Kulatunga, M., Sanger, D.J., Morel, E., Perrault, G., 1996. Comparison of the
Sjöholm, B., MacDonald, E., Pelto-Huikko, M., Leino, T., Barsh, pharmacological profiles of the hypnotic drugs, zaleplon and zolpidem.
G.S., Kobilka, B.K., Scheinin, M., 1997. Genetic alteraton of alpha2C - Eur. J. Pharmacol. 313, 35–42.
adrenoceptor expression in mice: influence on locomotor, hypothermic, Sanna, E., Motzo, C., Usala, M., Serra, M., Dazzi, L., Maciocco, E.,
and neurochemical effects of dexmedetomidine, a subtype-nonselective Trapani, G., Latrofa, A., Liso, G., Biggio, G., 1999. Characterization
alpha2C -adrenoceptor agonists. Mol. Pharmacol. 51, 36–46. of the electrophysiological and pharmacological effects of 4-iodo-2,6-
Sallinen, J., Haapalinna, A., Viitamaa, T., Kobilka, B.K., Scheinin, M., diisopropylphelol, a propofol analogue devoid of sedative–anaesthetic
1998. Adrenergic ␣2C -receptors modulate the acoustic startle reflex, properties. Br. J. Pharmacol. 126, 1444–1454.
prepulse inhibition, and aggression in mice. J. Neurosci. 18, 3035–3042. Sanna, E., Busonero, F., Talani, G., Carta, M., Massa, F., Peis, M.,
Sallinen, J., Haapalinna, A., MacDonald, E., Viitamaa, T., Lähdesmäki, Maciocco, E., Biggio, G., 2002. Comparison of the effects of zaleplon,
J., Rybnikova, E., Pelto-Huikko, M., Kobilka, B., Scheinin, M., 1999. zolpidem, and triazolam at various GABAA receptor subtypes. Eur. J.
Genetic alteration of the alpha2 -adrenoceptor subtype C in mice affects Pharmacol. 451, 103–110.
the development of behavioral despair and stress-induced increases in Sansum, A.J., Chessel, I.P., Hicks, G.A., Trezise, D.J., Humphrey,
plasma corticosterone levels. Mol. Psychiatry 4, 443–452. P.P.A., 1998. Evidence that P2X purinoceptors mediate the excitatory
Salonen, M., Onaivi, E.S., Maze, M., 1992. Dexmedetomidine effects of ␣␤methylene-ADP in rat locus coeruleus neurones.
synergism with midazolam in the elevated plus-maze test in rats. Neuropharmacology 37, 875–885.
Psychopharmacology 108, 229–234. Santarelli, L., Gobbi, G., Debs, P.C., Sibille, E.L., Blier, P., Hen, R., Heath,
Sanacora, G., Mason, G., Rothman, D., Krystal, J., 2002. Increased M.J.S., 2001. Genetic and pharmacological disruption of neurokinin1
occipital cortex GABA concentrations in depressed patients after receptor function decreases anxiety-related behaviors and increases
therapy with selective serotonin reuptake inhibitors. Am. J. Psychiatry serotonergic function. Proc. Natl. Acad. Sci. U.S.A. 98, 1912–1917.
159, 663–665. Santarelli, L., Gobbi, G., Blier, P., Hen, R., 2002. Behavioral and
Sanchez, C., 2003. Stress-induced vocalisation in adult animals. A valid physiologic effects of genetic or pharmacologic inactivation of the
model of anxiety? Eur. J. Pharmacol. 463, 133–143. substance P receptor (NK1 ). J. Clin. Psychiatry 63, 11–17.
Sanchez, C., Mork, A., 1999. N-Ethoxycarbonyl-2-ethoxy-1,2-dihydro- Sante, A.B., Nobre, M.J., Brandao, M.L., 2000. Place aversion induced
quinoline studies on the role of 5-HT1A and 5-HT2 receptors in by blockade of mu or activation of kappa opioid receptors in the
mediating foot-shock-induced ultrasonic vocalisation in adult rats. dorsal periaqueductal gray matter. Behav. Pharmacol. 11, 583–589.
Eur. Neuropsychopharmacol. 9, 287–294. Saphier, D., Welch, J.E., 1994. Central stimulation of adrenocortical
Sanchez, C., Arnt, J., Costall, B., Domeney, A.M., Kelly, E., Naylor, R.J., secretion by 5-hydroxytryptamine1A agonists is mediated by
1995. Sertindole: a limbic selective neuroleptic with potent anxiolytic sympathomedullary activation. J. Pharmacol. Exp. Ther. 270, 905–917.
effects. Drug Dev. Res. 34, 19–29.
Sapolsky, R.M., 2000a. Glucocorticoids and hippocampal atrophy in
Sanchez, C., Arnt, J., Moltzen, E., 1996. Assessment of relative efficacies
neuropsychiatric disorders. Arch. Gen. Psychiatry 57, 925–935.
of 5-HT1A receptor ligands by means of in vivo animal models. Eur.
Sapolsky, R.M., 2000b. Stress hormones: good and bad. Neurobiol. Dis.
J. Pharmacol. 315, 245–254.
7, 540–542.
Sanchez, C., Arnt, J., Costall, B., Kelly, M.E., Meier, E., Naylor, R.J.,
Sapolsky, R.M., Romero, L.M., Munck, A.U., 2000. How do
Perregaard, J., 1997. The selective ␴2 -ligand Lu 28–179 has potent
glucocorticoids influence stress responses? Integrating permissive,
anxiolytic-like effects in rodents. J. Pharmacol. Exp. Ther. 283, 1323–
suppressive, stimulatory, and preparative actions. Endocr. Rev. 21,
1332.
55–89.
Sanchez, M.M., Young, L.J., Plotsky, P.M., Insel, T.R., 1999.
Saria, A., 1999. The tachykinin NK1 receptor in the brain: pharmacology
Autoradiographic and in situ hybridization localization of corticotropin-
and putative functions. Eur. J. Pharmacol. 375, 51–60.
releasing factor 1 and 2 receptors in nonhuman primate brain. J.
Comp. Neurol. 408, 365–377. Sarret, P., Beaudet, A., Vincent, J.-P., Mazella, J., 1998. Regional and
Sanders-Bush, E., 1990. Adaptive regulation of central serotonin receptors cellular distribution of low affinity neurotensin receptor mRNA in
linked to phosphoinositide hydrolysis. Neuropsychopharmacology 3, adult and developing mouse brain. J. Comp. Neurol. 394, 344–356.
411–416. Sarter, M., Bruno, J.P., 1999. Abnormal regulation of corticopetal
Sanders, S.K., Shekhar, A., 1995. Regulation of anxiety by GABAA cholinergic neurons and impaired information processing in
receptors in the rat amygdala. Pharmacol. Biochem. Behav. 52, neuropsychiatric disorders. Trends Neurosci. 22, 67–74.
701–706. Sarter, M., Bruno, J.P., Berntson, G.G., 2001. Psychotogenic properties
Sandford, J.J., Forshall, S., Bell, C., Argyropoulos, S., Rich, A., of benzodiazepine receptor inverse agonists. Psychopharmacology
Orlando, K.J., Gammans, R.E., Nutt, D.J., 2001. Crossover trial of 156, 1–13.
pagoclone and placebo in patients with DSM-IV panic disorder. J. Sasson, Y., Iancu, I., Fux, M., Taub, M., Dannon, P.N., Zohar, J., 1999. A
Psychopharmacol. 15, 205–208. double-blind crossover comparison of clomipramine and desipramine
Sands, S.A., Strong, R., Corbitt, J., Morilak, D.A., 2000. Effects of acute in the treatment of panic disorder. Eur. Neuropsychopharmacol. 9,
restraint stress on tyrosine hydroxylase mRNA expression in locus 191–196.
coeruleus of Wistar and Wistar–Kyoto rats. Mol. Brain Res. 75, 1–7. Satoh, M., Minami, M., 1995. Molecular pharmacology of the opioid
Sands, S.A., Reisman, S.A., Enna, S.J., 2003. Effects of stress and receptors. Pharmacol. Ther. 68, 343–364.
tranylcypromine on amphetamine-induced locomotor activity and Saulskaya, N., Marsden, C.A., 1995. Extracellular glutamate in the
GABAB receptor function in rat brain. Life Sci. 72, 1085–1092. nucleus accumbens during a conditioned emotional response in the
Sanger, D.J., 1985. GABA and the behavioural effects of anxiolytic rat. Brain Res. 698, 114–120.
drugs. Life Sci. 36, 1503–1513. Sauvage, M., Steckler, T., 2001. Detection of corticotropin-relasing
Sanger, D.J., 1995. Behavioural effects of novel benzodiazepine () hormone receptor I. Immunoreactivity in cholinergic, dopaminergic
receptor agonists and partial agonists: increases in punished responding and noradrenergic neurons of the murine basal forebrain and brainstem
and antagonism of the pentylenetetrazole cue. Behav. Pharmacol. 6, nuclei—potential implication for arousal and attention. Neuroscience
116–126. 104, 643–652.
Sanger, D.J., Cohen, C., 1995. Fear and anxiety induced by benzodiazepine Sauvage, M., Brabet, P., Holsboer, F., Bockaert, J., Steckler, T., 2000.
receptor inverse agonists. In: Sarter M., Nutt D.J., Lister, R.G. (Eds.), Mild deficits in mice lacking pituitary adenylate cyclase-activating
Benzodiazepine Receptor Inverse Agonists. Wiley–Liss, New York, polypeptide receptor type 1 (PAC1) performing on memory tasks.
pp. 185–212. Mol. Brain Res. 84, 79–89.
230 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Sbrenna, S., Marti, M., Calo, G., Guerrini, R., Beani, L., Bianchi, C., Schmitz, D., Empson, R.M., Heinemann, U., 1995. Serotonin reduces
2000. Modulation of 5-hydroxytryptamine efflux from rat cortical inhibition via 5-HT1A receptors in area CA1 of rat hippocampal slices
synaptosomes by opioids and nociceptin. Br. J. Pharmacol. 130, in vitro. J. Neurosci. 15, 7217–7225.
425–433. Schneier, F.R., 2001. Treatment of social phobia with antidepressants. J.
Scarselli, M., Novi, F., Schallmach, E., Lin, R., Baragli, A., Colzi, A., Clin. Psychiatry 62, 43–48.
Griffon, N., Corsini, G.U., Sokoloff, P., Levenson, R., Vogel, Z., Schneier, F.R., Liebowitz, M.R., Abi-Dargham, A., Zea-Ponce, Y., Lin,
Maggio, R., 2001. D2 /D3 dopamine receptor heterodimers exhibit S.H., Laruelle, M., 2000. Low dopamine D2 receptor binding potential
unique functional properties. J. Biol. Chem. 276, 30308–30314. in social phobia. Am. J. Psychiatry 157, 457–459.
Schaaf, M.J.M., Hoetelmans, R.W.M., De Kloet, E.R., Vreugdenhil, E., Schoepp, D.D., 2001. Unveiling the functions of presynaptic metabotropic
1997. Corticosterone regulates expression of BDNF and trkB but not glutamate receptos in the central nervous system. J. Pharmacol.
NT-3 and trkC mRNA in the rat hippocampus. J. Neurosci. Res. 48, Exp. Ther. 299, 12–20.
334–341. Schoepp, D.D., Marek, G.J., 2002. Preclinical pharmacology of mGlu2/3
Schaefer, M., Engelbrecht, A., Gut, O., Fiebich, B.L., Bauer, J., Schmidt, receptor agonists: novel agents for schizophrenia? Curr. Drug Targets
F., Grunze, H., Lieb, K., 2002. Interferon alpha (IFN␣) and psychiatric 1, 215–225.
syndromes: a review. Prog. Neuropsychopharmacol. Biol. Psychiatry Schoepp, D.D., Jane, D.E., Monn, J.A., 1999. Pharmacological
26, 731–746. agents acting at subtypes of metabotropic glutamate receptors.
Schatzberg, A.F., Schildkraut, J.J., 1995. Recent studies on norepinephrine Neuropharmacology 38, 1431–1476.
systems in mood disorders. In: Bloom, F.E., Kupper, D.J. (Eds.), Schooler, N.R., Siu, C., 2000. Ziprasidone’s effect on anxiety in a group
Psychopharmacology: The Fourth Generation of Progress. Raven, of outpatients with stable schizophrenia. CINP Abstr. (Brussels) S288.
New York, pp. 911–920. Schramm, N.L., McDonald, M.P., Limbird, L.E., 2001. The ␣2A -
Scheel-Krüger, J., Petersen, E.N., 1982. Anticonflict effect of the adrenergic receptor plays a protective role in mouse behavioural
benzodiazepines mediated by a GABAergic mechanism in the models of depression and anxiety. J. Neurosci. 21, 4875–4882.
amygdala. Eur. J. Pharmacol. 82, 115–116. Schreiber, R., De Vry, J., 1993a. 5-HT1A receptor ligands in animal
Schefke, D.M., Fontana, D.J., Commissaris, R.L., 1989. Anti-conflict models of anxiety, impulsivity and depression: multiple mechanisms
efficacy of buspirone following acute versus chronic treatment. of action? Prog. Neuropsychopharmacol. Biol. Psychiatry 17, 87–104.
Psychopharmacology 99, 427–429. Schreiber, R., De Vry, J., 1993b. Neuronal circuits involved in the
Scheibner, J., Trendelenburg, A.-U., Hein, L., Starke, K., 2001. ␣2 - anxiolytic effects of the 5-HT1A receptor agonists 8-OH-DPAT,
Adrenoceptors modulating neuronal serotonin release: a study in ␣2 - ipsapirone and buspirone in the rat. Eur. J. Pharmacol. 249, 341–351.
adrenoceptor subtype-deficient mice. Br. J. Pharmacol. 132, 925–933. Schreiber, R., Melon, C., De Vry, J., 1998. The role of 5-HT receptor
Schenberg, L.C., Rustamante Capucho, L., Vatanabe, R.O., Vargas, subtypes in the anxiolytic effects of selective serotonin reuptake
L.C., 2002. Acute effects of clomipramine and fluoxetine on dorsal inhibitors in the rat ultrasonic vocalization test. Psychopharmacology
periaqueductal grey-evoked unconditioned defensive behaviours of the 135, 383–391.
rat. Psychopharmacology 159, 138–144. Schroeder, B.E., Schiltz, C.A., Kelley, A.E., 2003. Neural activation profile
Scherrer, J.F., True, W.R., Xian, H., Lyons, M.J., Eisen, S.A., Goldberg, elicited by cues associated with the anxiogenic drug yohimbine differs
J., Lin, N., Tsuang, M.T., 2000. Evidence for genetic influences from that observed for reward-paired cues. Neuropsychopharmacology
common and specific to symptoms of generalized anxiety and panic. 28, 14–21.
J. Affect. Disord. 57, 25–35. Schrott, L.M., Crnic, L.S., 1996. Increased anxiety behaviors in
Schilström, B., Svensson, H.M., Svensson, T.H., Nomikos, G.G., 1998. autoimmune mice. Behav. Neurosci. 110, 492–502.
Nicotine and food induced dopamine release in the nucleus accumbens Schuckit, M.A., Hesselbrock, V., 1994. Alcohol dependence and anxiety
of the rat: putative role of ␣7 nicotinic receptors in the ventral disorders: what is the relationship? Am. J. Psychiatry 141, 1723–1734.
tegmental area. Neuroscience 85, 1005–1009. Schuler, V., Lüscher, C., Blanchet, C., Klix, N., Sansig, G., Klebs, K.,
Schilström, B., Rawal, N., Mameli-Engvall, M., Nomikos, G.G., Svensson, Schmutz, M., Heid, J., Gentry, C., Urban, L., Fox, A., Spooren, W.,
T.H., 2003. Dual effects of nicotine on dopamine neurons mediated Jaton, A.L., Vigouret, J.M., Pozza, M., Kelly, P.H., Mosbacher, J.,
by different nicotinic recepor subtypes. Int. J. Neuropsychopharmacol. Froestl, W., Käslin, E., Korn, R., Bischoff, S., Kaupmann, K., Van
6, 3–11. Der Putten, H., Bettler, B., 2001. Epilepsy, hyperalgesia, impaired
Schlicker, E., Kathman, M., 2001. Modulation of transmitter release memory, and loss of pre- and postsynaptic GABAB responses in mice
via presynaptic cannabinoid receptors. Trends Pharmacol. Sci. 22, lacking GABAB(1) . Neuron 31, 47–58.
565–572. Schulkin, J., Gold, P.W., McEwen, B.S., 1998. Induction of corticotropin-
Schlicker, E., Morari, M., 2000. Nociceptin/orphaninFQ and neurotra- releasing hormone gene expression by glucocorticoids: implication
nsmitter release in the central nervous system. Peptides 21, 1023–1029. for understanding the states of fear and anxiety and allostatic load.
Schlicker, E., Timm, J., Zentner, J., Göthert, M., 1997. Cannabinoid CB1 Psychoneuroendocrinology 23, 219–243.
receptor-mediated inhibition of noradrenaline release in the human and Schulte, D., Callado, L.F., Davidson, C., Phillips, P.E., Roewer, N., Schulte
guinea-pig hippocampus. Naunyn Schmiedebergs Arch. Pharmacol. am Esch, J., Stamford, J.A., 2000. Propofol decreases stimulated
356, 583–589. dopamine release in the rat nucleus accumbens by a mechanism
Schmidt, M.E., Oshinsky, R.J., Kim, H.-G., Schouten, J.L., Folley, independent of dopamine D2 , GABAA and NMDA receptors. Br. J.
B.S., Potter, W.Z., 1999. Cerebral glucose metabolic and plasma Anaesth. 84, 250–253.
catecholamine responses to the ␣2 -adrenoceptor antagonist ethoxyi- Schulz, B., Fendt, M., Gasparini, F., Lingenhohl, K., Kuhn, R., Koch,
dazoxan given to healthy volunteers. Psychopharmacology 146, 119– M., 2001. The metabotropic glutamate receptor antagonist 2-methyl-
127. 6(phenylethynyl)-pyridine (MPEP) blocks fear conditioning in rats.
Schmitt, M.L., Coelho, W., Lopes-de-Souza, A.S., Guimarães, F.S., Neuropharmacology 41, 1–7.
Carobrez, A.P., 1995. Anxiogenic-like effect of glycine and d-serine Schulz, B., Fendt, M., Schnitzler, H.-U., 2002. Clonidine injections into
microinjected into dorsal periaqueductal gray matter of rats. Neurosci. the lateral nucleus of the amygdala block acquisition and expression
Lett. 189, 93–96. of fear-potentiated startle. Eur. J. Neurosci. 15, 151–157.
Schmitt, U., Lüddens, H., Hiemke, C., 2002. Anxiolytic-like effects of Schwarcz, R., Pellicciari, R., 2002. Manipulation of brain kynurenines:
acute and chronic GABA transporter inhibition in rats. J. Neural glial targets, neuronal effects, and clinical opportunities. J. Pharmacol.
Transm. 109, 871–880. Exp. Ther. 303, 2–10.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 231

Schwarz, S.K.W., Puil, E., 1998. Analgesic and sedative concentrations Seth, P., Cheeta, S., Tucci, S., File, S.E., 2002. Nicotinic-serotonergic
of lognocaine shunt tonic and burst firing in thalamocortical neurones. interactions in brain and behaviour. Pharmacol. Biochem. Behav. 71,
Br. J. Pharmacol. 124, 1633–1642. 795–805.
Schweiler, L., Delbro, D.S., Engberg, G., Erhardt, S., 2003. The Seyfried, C.A., Greiner, H.E., Haase, A.F., 1989. Biochemical and
anaesthetic agent propofol interacts with GABAB -receptors: an functional studies on EMD 49980: a potent, selectively presynaptic
electrophysiological study in rat. Life Sci. 72, 2793–2801. D2 dopamine agonist with actions on serotonin systems. Eur. J.
Schwendt, M., Jezová, D., 2000. Gene expression of two glutamate Pharmacol. 160, 31–41.
receptor subunits in response to repeated stress exposure in rat Shafer, R.A., Levant, B., 1998. The D3 dopamine receptor in cellular
hippocampus. Cell. Mol. Neurobiol. 20, 319–329. and organismal functions. Psychopharmacology 135, 1–16.
Seamans, J.K., Gorelova, N., Durstewitz, D., Yang, C.R., 2001. Shannon, H.E., Lutz, E.A., 2000. Yohimbine produces antinociception
Bidirectional dopamine modulation of GABAergic inhibition in in the formalin test in rats: involvement of serotonin1A receptors.
prefrontal cortical pyramidal neurons. J. Neurosci. 21, 3628–3638. Psychopharmacology 149, 93–97.
Seasholtz, A.F., Burrows, H.L., Karolyi, I.J., Camper, S.A., 2001. Mouse Sharma, A.C., Kulkarni, S.K., 1993. Evidence for benzodiazepine
models of altered CRH-binding protein expression. Peptides. 22, receptor interaction with MK 801 in anxiety related behaviour in rats.
743–751. Indian J. Exp. Biol. 31, 191–193.
Segars, J.H., Driggers, P.H., 2002. Estrogen action and cytoplasmic Sharma, A.C., Punhani, T., Fone, K.C.F., 1997. Distribution of the 5-
signalling cascades. Part I. Membrane-associated signalling complexes. hydroxytryptamine2C receptor protein in adult rat brain and spinal
Trends Endocrinol. Metab. 13, 349–354. cord determined using a receptor-directed antibody: effect of 5,7-
Seggie, J., Campbell, L., Brown, G., Grota, L., 1985. Melatonin and dihydroxytryptamine. Synapse 27, 45–56.
N-acetyl-serotonin stress responses: effects of type of stimulation and Shekhar, A., Keim, S.R., 2002. LY354740, a potent group II metabotropic
housing conditions. J. Pineal. Res. 2, 39–49. receptor agonist, prevents lactate-induced panic-like response in
Segieth, J., Pearce, B., Fowler, L., Whitton, P.S., 2001. Regulatory panic-prone rats. Neuropharmacology 39, 1139–1146.
role of nitric oxide over hippocampal 5-HT release in vivo. Naunyn Shekhar, A., McCann, U.D., Meaney, M.J., Blanchard, D.C., Davis,
Schmiedebergs Arch. Pharmacol. 363, 302–306. M., Frey, K.A., Liberzon, I., Overall, K.L., Shear, M.K., Tecott,
Seifert, R., Wenzel-Seifert, D., 2002. Constitutive activity of G-protein- L.H., Winsky, L., 2001. Summary of a National Institute of Mental
coupled receptors: cause of disease and common property of wild-type Health Workshop: developing animal models of anxiety disorders.
receptors. Naunyn Schmiedebergs Arch. Pharmacol. 366, 381–416. Psychopharmacology 157, 327–339.
Selmer, I.-S., Schindler, M., Allen, J.P., Humphrey, P.P.A., Emson, P.C., Shekhar, A., Katner, J.S., Sajdyk, T.J., Kohl, R.R., 2002. Role of
2000. Advances in understanding neuronal somatostatin receptors. norepinephrine in the dorsomedial hypothalamic panic response: an in
Regul. Pept. 90, 1–18. vivo microdialysis study. Pharmacol. Biochem. Behav. 71, 493–500.
Shen, R.Y., Andrade, R., 1998. 5-Hydroxytryptamine2 receptor facilitates
Semeniuk, T., Jhangri, G.S., Le Melledo, J.M., 2001. Neuroactive steroid
GABAergic neurotransmission in rat hippocampus. J. Pharmacol.
levels in patients with generalized anxiety disorder. J. Neuropsychiatry
Exp. Ther. 285, 805–812.
Clin. Neurosci. 13, 396–398.
Shen, M., Piser, T.M., Seybold, V.S., Thayer, S.A., 1996. Cannabinoid
Sepinwall, J., Cook, L., 1980. Mechanism of action of the
receptor agonists inhibit glutamatergic synaptic transmission in rat
benzodiazepines: behavioral aspect. Fed. Proc. 39, 3024–3031.
hippocampal cultures. J. Neurosci. 16, 4322–4334.
Sergeyev, V., Hökfelt, T., Hurd, Y., 1999. Serotonin and substance P
Shen, C., Li, H., Meller, E., 2002. Repeated treatment with antidepressants
co-exist in dorsal raphe neurons of the human brain. NeuroReport 10,
differentially alters 5-HT1A agonist-stimulated [35 S]GTP␥S binding
3967–3970.
in rat brain regions. Neuropharmacology 42, 1031–1038.
Serra, M., Pisu, M.G., Muggironi, M., Parodo, V., Papi, G., Sari,
Shepard, J.D., Barron, K.W., Myers, D.A., 2000. Corticosterone delivery
R., Dazzi, L., Spiga, F., Purdy, R.H., Biggio, G., 2001. Opposite
to the amygdala increases corticotropin-releasing factor mRNA in the
effects of short- versus long-term administration of fluoxetine on
central amygdaloid nucleus and anxiety-like behavior. Brain Res. 861,
the concentrations of neuroactive steroids in rat plasma and brain.
288–295.
Psychopharmacology 158, 48–54. Shephard, R.A., 1986. Neurotransmitters, anxiety and benzodiazepines:
Serradeil-Le Gal, C., Wagnon, J., Simiand, J., Griebel, G., Lacour, C., a behavioural review. Neurosci. Biobehav. Rev. 10, 449–461.
Guillon, G., Barberis, C., Brossard, G., Soubrié, P., Nisato, D., Pascal, Shephard, R.A., 1987. Behavioral effects of GABA agonists in relation
M., Pruss, R., Scatton, B., Maffrand, J.-P., Le Fur, G., 2002. Characteri- to anxiety and benzodiazepine action. Life Sci. 40, 2429–2436.
zation of (2S,4R)-1-[5-chloro-1-[(2,4-dimethoxyphenyl)sulfonyl]-3-(2- Shephard, R.A., Wedlock, P., Wilson, N.E., 1992. Direct evidence for the
methoxy-phenyl)-2-oxo-2,3-dihydro-1H-indol-3-yl]-4-hydroxy-N,Ndi- mediation of an anticonflict effect of baclofen by GABAB receptors.
methyl-2-pyrrolidine carboxamide (SSR149415), a selective and orally Pharmacol. Biochem. Behav. 41, 651–653.
active vasopressin V1b receptor antagonist. J. Pharmacol. Exp. Ther. Shepherd, J., Bill, D.J., Dourish, C.T., Grewal, S.S., McLenachan, A.,
300, 1122–1130. Stanhope, K.J., 1996. Effects of the selective angiotensin II receptor
Serrano, A., D’Angio, M., Scatton, B., 1989. NMDA antagonists block antagonists iosartan and PD123177 in animal models of anxiety and
restraint-induced increase in extracellular DOPAC in rat nucleus memory. Psychopharmacology 126, 206–218.
accumbens. Eur. J. Pharmacol. 162, 157–166. Sher, L., 2003. Alcoholism, anxiety, and opioid–dopaminergic interactions.
Serrats, J., Artigas, F., Mengod, G., Cortés, R., 2003. GABAB receptor Psychopharmacology 165, 202–203.
mRNA in the raphe nuclei: co-expression with serotonin transporter Sherif, F.M., Ahmad, S.S., 1995. Basic aspects of GABA-transaminase
and glutamic acid decarboxylase. J. Neurochem. 84, 743–752. in neuropsychiatric disorders. Clin. Biochem. 28, 145–154.
Seta, K.A., Jansen, H.T., Kreitel, K.D., Lehman, M., Behbehani, M.M., Sherif, F.M., Oreland, L., 1995. Effect of the GABA-transaminase
2001. Cold water swim stress increases the expression of neurotensin inhibitor vigabatrin on exploratory behaviour in socially isolated rats.
mRNA in the lateral hypothalamus and medial preoptic regions of Behav. Brain Res. 72, 135–140.
the rat brain. Mol. Brain Res. 86, 145–152. Sherwood-Brown, E., Rush, A.J., McEwen, B.S., 1999. Hippocampal
Setem, J., Pinheiro, A.P., Motta, V.A., Morato, S., Cruz, A.P.M., 1999. remodelling and damage by corticosteroids: implications for mood
Ethopharmacological analysis of 5-HT ligands on the rat elevated disorders. Neuropsychopharmacology 21, 474–484.
plus-maze. Pharmacol. Biochem. Behav. 62, 515–521. Shiba, M., Bower, J.H., Maraganore, D.M., McDonnell, S.K., Peterson,
Seth, P., Fei, Y.J., Li, H.W., Huang, W., Leibach, F.H., Ganapathy, V., B.J., Ahlskog, J.E., Schaid, D.J., Rocca, W.A., 2000. Anxiety disorders
1998. Cloning and functional characterization of a sigma receptor and depressive disorders preceding Parkinson’s disease: a case–control
from rat brain. J. Neurochem. 70, 922–931. study. Mov. Disorder Soc. 15, 669–677.
232 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Shibata, S., Yamashita, K., Yamamoto, E., Ozaki, T., Ueki, S., 1989. Siemiatkowski, M., Rokicki, D., Członkowska, A.I., Sienkiewicz-Jarosz,
Effects of benzodiazepine and GABA antagonists on anticonflict H., Bidziński, A., Płaźnik, A., 2000a. Locomotor activity and
effects of antianxiety drugs injected into the rat amygdala in a conditioned fear response: correlation with cortical and subcortical
water-lick suppression test. Psychopharmacology 98, 38–44. binding of the dopamine D1 receptor antagonist. NeuroReport 11,
Shih, J.C., Chen, K., Ridd, M.J., 1999. Monoamine oxidase: from genes 3953–3956.
to behavior. Annu. Rev. Neurosci. 22, 197–217. Siemiatkowski, M., Sienkiewicz-Jarosz, H., Członkowska, A.I., Szyndler,
Shigemoto, R., Kinoshita, A., Normura, S., Ohishi, H., Takada, M., Flor, J., Bidziński, A., Płaźnik, A., 2000b. The effects of dopamine D2
P.J., Neki, A., Abe, T., Nakanishi, S., Mizuno, N., 1997. Differential receptor ligands on novelty-induced behavior in the rat open field
presynaptic localization of metabotropic glutamate receptor subtypes test. Neurosci. Res. Commun. 27, 155–163.
in the rat hippocampus. J. Neurosci. 17, 7503–7522. Siemiatkowski, M., Maciejak, P., Sienkiewicz-Jarosz, H., Czlonkowska,
Shimada, M., Tritos, N.A., Lowell, B.B., Flier, J.S., Maratos-Fler, E., A.I., Szyndler, J., Gryczynska, A., Plaznik, A., 2001. Opposite effects
1998. Mice lacking melanin-concentrating hormone are hypophagic of olanzapine and haloperidol in rat ultrasonic vocalization test. Pol.
and lean. Nature 380, 243–247. J. Pharmacol. 53, 669–673.
Shimizu, H., Hirose, A., Tatsuno, T., Nakamura, M., Katsube, J., Sigel, E., Baur, R., Furtmueller, R., Razet, R., Dodd, R.H., Sieghart,
1987. Pharmacological properties of SM-3997: a new anxioselective W., 2001. Differential cross talk of ROD compounds with the
anxiolytic candidate. Jpn. J. Pharmacol. 45, 493–500. benzodiazepine binding site. Mol. Pharmacol. 59, 1470–1477.
Shimizu, H., Tatsuno, T., Tanaka, H., Hirose, A., Araki, Y., Nakamura, M., Sillaber, I., Montkowski, A., Landgraf, R., Barden, N., Holsboer, F.,
1992. Serotonergic mechanisms in anxiolytic effect of tandospirone Spanagel, R., 1998. Enhanced morphine-induced behavioural effects
in the Vogel Conflict Test. Jpn. J. Pharmacol. 59, 105–112. and dopamine release in the nucleus accumbens in a transgenic
Shimuzu, K., Matsubara, K., Uezono, T., Kimura, K., Shiono, H., 1998. mouse model of impaired glucocorticoid (type II) receptor function:
Reduced dorsal hippocampal glutamate release significantly correlates influence of long-term treatment with the antidepressant moclobemide.
with the spatial memory deficits produced by benzodiazepines and Neuroscience 85, 415–425.
ethanol. Neuroscience 83, 701–706. Silva, R.C.B., Brandao, M.L., 2000. Acute and chronic effects of gepirone
Shipley, M.T., McLean, J.H., Behbehani, M.M., 1987. Heterogeneous and fluoxetine in rats tested in the elevated plus-maze: an ethological
distribution of neurotensin-like immunoreactive neurons and fibers in analysis. Pharmacol. Biochem. Behav. 65, 209–216.
the midbrain periaqueductal gray of the rat. J. Neurosci. 7, 2025–2034. Silva, R.H., Frussa-Filho, R., 2002. Naltrexone potentiates both amnestic
Shirayama, Y., Mitsuhio, H., Takashima, M., Ichikawa, H., Takahashi, K., and anxiolytic effects of chlordiazepoxide in mice. Life Sci. 72,
1996. Reduction of substance P after chronic antidepressant treatment 721–730.
in the striatum, substantia nigra and amygdala of the rat. Brain Res.
Silva, M., Hetem, L.A.B., Guimaraes, F.S., Graeff, F.G., 2001.
739, 70–78.
Opposite effects of nefazodone in two human models of anxiety.
Shirayama, Y., Chen, A.C., Nakagawa, S., Russell, D.S., Duman, R.S.,
Psychopharmacology 156, 454–460.
2002. Brain-derived neurotrophic factor produces antidepressant effects
Silva, S.M., Paula-Barbosa, M.M., Dulce Madeira, M., 2002. Prolonged
in behavioral models of depression. J. Neurosci. 22, 3251–3261.
alcohol intake leads to reversible depression of corticotropin-releasing
Shirazi-Southall, S., Rodriguez, D.E., Nomikos, G.G., 2002. Effects
hormone and vasopressin immunoreactivity and mRNA levels in the
of typical and atypical antipsychotics and receptor selective
parvocellular neurons of the paraventricular nucleus. Brain Res. 954,
compounds on acetylcholine efflux in the hippocampus of the rat.
82–93.
Neuropsychopharmacology 26, 583–594.
Silvestre, J.S., Fernandez, A.G., Palacios, J.M., 1996. Effects of 5-HT4
Shors, F., Oreland, L., 1995. Effect of the GABA-transaminase inhibitor
receptor antagonists on rat behaviour in the elevated plus-maze test.
vigabatrin on exploratory behaviour in socially isolated rats. Behav.
Eur. J. Pharmacol. 309, 219–222.
Brain Res. 72, 135–140.
Silvestre, J.S., Pallarés, M., Nadal, R., Ferré, M., 2002. Opposite effects
Shors, T.J., Elkabes, S., Selcher, J.C., Black, I.B., 1997. Stress persistently
of ethanol and ketamine in the elevated plus-maze test in Wistar
increases NMDA receptor-mediated binding of [3 H] PDBu (a marker
rats undergoing a chronic oral voluntary consumption procedure. J.
for protein kinase C) in the amygdala, and re-exposure to the stressful
Psychopharmacol. 16, 305–312.
context reactivates the increase. Brain Res. 750, 293–300.
Shughrue, P.J., Lane, M.V., Merchenthaler, I., 1997. Comparative Simon, P., Panissaud, C., Costentin, J., 1993. Anxiogenic-like effects
distribution of estrogen receptor-␣ and ␤ mRNA in the rat central induced by stimulation of dopamine receptors. Pharmacol. Biochem.
nervous system. J. Comp. Neurol. 388, 507–525. Behav. 45, 685–690.
Shumyatsky, G.P., Tavetkov, E., Malleret, G., Vronskaya, S., Hatton, Simon, P., Panissaud, C., Costentin, J., 1994. The stimulant effect of
M., Hampton, L., Battey, J.F., Dulac, C., Kandel, E.R., Bolshakov, modafinil on wakefulness is not associated with an increase in anxiety
V.Y., 2002. Identification of a signaling network in lateral nucleus in mice. Psychopharmacology 114, 597–600.
of amygdala important for inhibiting memory specifically related to Singewald, N., Sharp, T., 2000. Neuroanatomical targets of anxiogenic
learned fear. Cell 111, 905–918. drugs in the hindbrain as revealed by Fos immunocytochemistry.
Sibille, E., Hen, R., 2001. Serotonin1A receptors in mood disorders: Neuroscience 98, 759–770.
a combined genetic and genomic approach. Behav. Pharmacol. 12, Singh, L., Field, M.J., Hughes, J., Menzies, R., Oles, R.J., Vass, C.A.,
429–438. Woodruff, G.N., 1991. The behavioral properties of CI-988, a selective
Sibille, E., Pavlides, C., Benke, D., Toth, M., 2000. Genetic inactivation cholecystokininB receptor antagonist. Br. J. Pharmacol. 104, 239–245.
of the serotonin1A receptor in mice results in downregulation of major Singh, L., Field, M.J., Ferris, P., Hunter, J.C., Oles, R.J., Williams,
GABAA receptor ␣ subunits, reduction of GABAA receptor binding, R.G., Woodruff, G.N., 1996. The antiepileptic agent GABApentin
and benzodiazepine-resistant anxiety. J. Neurosci. 20, 2758–2765. (neurontin) possesses anxiolytic-like and antinociceptive actions that
Sibley, D.R., 1999. New insights into dopaminergic receptor function are reversed by d-serine. Psychopharmacology 127, 1–9.
using antisense and genetically altered animals. Annu. Rev. Pharmacol. Siniscalchi, A., Rodi, D., Cavallini, S., Marino, S., Ferraro, L., Beani,
Toxicol. 39, 313–341. L., Bianchi, C., 2003. Effects of cholecystokinin tetrapeptide (CCK4 )
Sibug, R.M., Compaan, J.C., Meijer, O.C., Van Der Gugten, J., Olivier, B., and of anxiolytic drugs on GABA outflow from the cerebral cortex
De Kloet, E.R., 2000. Effects of flesinoxan on corticosteroid receptor of freely moving rats. Neurochem. Int. 42, 87–92.
expression in the rat hippocampus. Eur. J. Pharmacol. 404, 111–119. Sinnott, R.S., Mark, G.P., Finn, D.A., 2002. Reinforcing effects of the
Sieghart, W., 1995. Structure and pharmacology of ␥-aminobutyric acidA neurosteroid allopregnanolone in rats. Pharmacol. Biochem. Behav.
receptor subtypes. Pharmacol. Rev. 47, 181–234. 72, 923–929.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 233

Siuciak, J.A., Wong, V., Pearsall, D., Wiegand, S.J., Lindsay, R.M., Smith, M.T., Perlis, M.L., Park, A., Smith, M.S., Pennington, J.,
1995. BDNF produces analgesia in the formalin test and modifies Giles, D.E., Buysse, D.J., 2002a. Comparative meta-analysis of
neuropeptide levels in rat brain and spinal cord areas associated with pharmacotherapy and behavior therapy for persistent insomnia. Am.
nociception. Eur. J. Neurosci. 7, 663–670. J. Psychiatry 159, 5–11.
Siuciak, J.A., Boylan, C., Fritsche, M., Altar, C.A., Lindsay, R.M., Smith, W., Feltner, D., Kavoussi, R., 2002b. Pregabalin in generalized
1996. BDNF increases monoaminergic activity in rat brain following anxiety disorder: long term efficacy and relapse prevention. Eur.
intracerebroventricular or intraparenchymal administration. Brain Res. Neuropsychopharmacol. 12, S350.
710, 11–20. Smythe, J.W., Bhatnagar, S., Murphy, D., Timothy, C., Costall, B., 1998.
Siuciak, J.A., Lewis, D., Wiegand, S.L., Lindsay, R.M., 1997. The effects of intrahippocampal scopolamine infusions on anxiety in
Antidepressant-like effect of brain-derived neurotrophic factor (BDNF). rats as measured by the black-white box test. Brain Res. Bull. 45,
Pharmacol. Biochem. Behav. 56, 131–137. 89–93.
Siuciak, J.A., Clark, M.S., Rind, H.B., Whittemore, S.R., Russo, A.F., Soby, K., Mikkelsen, J.D., Meier, E., Thomsen, C., 2002. Lu 28–179 labels
1998. BDNF induction of tryptophan hydroxylase mRNA levels in a ␴2 -site in rat and human brain. Neuropharmacology 43, 95–100.
the rat brain. J. Neurosci. Res. 52, 149–158. Söderpalm, B., 1989. The SHR exhibits less “anxiety” but increased
Skelton, K.H., Nemeroff, C.B., Knight, D.L., Owens, M.J., 2000. Chronic sensitivity to the anticonflict effect of clonidine compared to
administration of the triazolobenzodiazepine alprazolam produces normotensive controls. Pharmacol. Toxicol. 65, 381–386.
opposite effects on corticotropin-releasing factor and urocortin Söderpalm, B., Engel, J.A., 1988. Biphasic effects of clonidine on conflict
neuronal systems. J. Neurosci. 20, 1240–1248. behavior: involvement of different alpha-adrenoceptors. Pharmacol.
Skofitsch, G., Jacobowitz, D.M., 1985. Immunohistochemical mapping Biochem. Behav. 30, 471–477.
of galanin-like neurons in the rat central nervous system. Peptides 6, Söderpalm, B., Engel, J.A., 1989. ␣2 -Adrenoceptor antagonists potentiate
509–546. the anticonflict and the rotarod impairing effects of benzodiazepines.
Skolnick, P., Miller, R., Young, A., Boje, K., Trullas, R., 1992. Chronic J. Neural Transm. 76, 191–204.
treatment with l-aminocyclopropanecarboxylic acid desensitizes Söderpalm, B., Engel, J.A., 1990. ␣1 - and ␤-adrenoceptor stimulation
behavioral responses to compounds acting at the N-methyl-d-aspartate potentiate the anticonflict effect of a benzodiazepine. J. Neural
receptor complex. Psychopharmacology 107, 489–499. Transm. 79, 155–167.
Skrebuhhova, T., Allikmets, L., Matto, V., 1999. [3 H]-Ketanserin binding
Söderpalm, B., Engel, J.A., 1991. Involvement of the GABAA /benzo-
and elevated plus-maze behavior after chronic antidepressant treatment
diazepine chloride ionophore receptor complex in the 5,7-DHT
in DSP-4 and PCPA pretreated rats: evidence for partial involvement of
induced anticonflict effect. Life Sci. 49, 139–153.
5-HT2A receptors. Methods Find Exp. Clin. Pharmacol. 21, 483–490.
Söderpalm, B., Eriksson, E., Engel, J.A., 1989. Anticonflict and rotarod
Slassi, A., Isaac, M., O’Brien, A., 2002. Recent progress in 5-HT6
impairing effects of alprazolam and diazepam in rat after acute
receptor antagonists for the treatment of CNS diseases. Expert Opin.
and subchronic administration. Prog. Neuropsychopharmacol. Biol.
Ther. Patents 12, 513–527.
Psychiatry 13, 269–283.
Sloviter, R.S., Ali-Akbarian, L., Horvath, K.D., Menkens, K.A., 2001.
Söderpalm, B., Andersson, G., Johannessen, K., Engel, J.A., 1992.
Substance P receptor expression by inhibitory interneurons of the
Intracerebroventricular 5,7-DHT alters the in vitro function of
rat hippocampus: enhanced detection using improved immuno-
rat cortical GABAA /benzodiazepine chloride ionophore receptor
cytochemical methods for the preservation and colocalization of
complexes. Life Sci. 51, 327–335.
GABA and other neuronal markers. J. Comp. Neurol. 430, 283–305.
Söderpalm, A., Blomqvist, O., Söderpalm, B., 1995a. The yohimbine-
Slowe, S.J., Clarke, S., Lena, I., Goody, R.J., Lattanzi, R., Negri,
induced anticonflict effect in the rat. Part I. Involvement of
L., Simonin, F., Matthes, H.W., Filliol, D., Kieffer, B.L., Kitchen,
noradrenergic, serotonergic and endozepinergic mechanisms. J. Neural
J., 2001. Autoradiographic mapping of the opioid receptor-like 1
Transm. 100, 175–189.
(ORL1) receptor in the brains of mu-, delta-, or kappa-opioid receptor
Söderpalm, A., Ehrenström, F., Söderpalm, B., 1995b. The yohimbine-
knock-out mice. Neuroscience 106, 469–480.
Smagin, G.N., Heinrichs, S.C., Dunn, A.J., 2001. The role of CRH in induced anticonflict effect in the rat. Part II. Neurochemical findings.
behavioral responses to stress. Peptides 22, 713–724. J. Neural Transm. 100, 191–206.
Smith, B.N., Dudek, F.E., 1996. Amino acid-mediated regulation of Söderpalm, B., Andersson, G., Enerbäck, C., Engel, J.A., 1997. In
spontaneous synaptic activity patterns in the rat basolateral amygdala. vivo administration of the 5-HT1A receptor agonist 8-OH-DPAT
J. Neurophysiol. 76, 1958–1967. interferes with brain GABAA /benzodiazepine receptor complexes.
Smith, M.A., Makino, S., Altemus, M., Michelson, D., Hong, S.- Neuropharmacology 36, 1071–1077.
K., Kvetnansky, R., Post, R.M., 1995a. Stress and antidepressants Solinas, M., Ferré, S., You, Z.-B., Karcz-Kubicha, M., Popoli, P.,
differentially regulate neurotrophin 3 mRNA expression in the locus Goldberg, S.R., 2002. Caffeine induces dopamine and glutamate release
coeruleus. Proc. Natl. Acad. Sci. U.S.A. 92, 8788–8792. in the shell of the nucleus accumbens. J. Neurosci. 22, 6321–6324.
Smith, M.A., Makino, S., Kvetnansky, R., Post, R.M., 1995b. Stress and Sommer, B., Keinanen, K., Verdoorn, T.A., Wisden, W., Burnashev, N.,
glucocorticoids affect the expression of brain-derived neurotrophic Herb, A., Kohler, M., Takagi, T., Sakmann, B., Seeburg, P.H., 1990.
factor and neurotrophin-3 mRNAs in the hippocampus. J. Neurosci. Flip and flop: a cell-specific functional switch in glutamate-operated
15, 1768–1777. channels of the CNS. Science 249, 1580–1585.
Smith, G.W., Aubry, J.-M., Dellu, F., Contarino, A., Bilezikjian, L.M., Somogyi, J., Llewellyn-Smith, I.J., 2001. Patterns of colocalization of
Gold, L.H., Chen, R., Marchuk, Y., Hauser, C., Bentley, C.A., GABA, glutamate and glycine immunoreactivities in terminals that
Sawchenko, P.E., Koob, G.F., Vale, W., Lee, K.-F., 1998a. Corticotropin synapse on dendrites of noradrenergic neurons in rat locus coeruleus.
releasing factor 1-deficient mice display decreased anxiety, impaired Eur. J. Neurosci. 14, 219–228.
stress response, and aberrant neuroendocrine development. Neuron Son, L.Z., Yanai, K., Mobarakeh, J.I., Kuramasu, A., Li, Z.Y., Sakurai,
20, 1093–1102. E., Hashimoto, Y., Watanabe, T., Watanabe, T., 2001. Histamine
Smith, S.S., Gong, Q.H., Hsu, F.C., Markowitz, R.S., French-Mullen, J.M., H1 receptor-mediated inhibition of potassium-evoked release of 5-
Li, X., 1998b. GABAA receptor alpha4 subunit suppression prevents hydroxytryptamine from mouse forebrain. Behav. Brain Res. 124,
withdrawal properties of an endogenous steroid. Nature 392, 926–930. 113–120.
Smith, D.W., Hewson, L., Fuller, P., Williams, A.R., Wheeldon, A., Sorbera, L.A., Leeson, P.A., Silvestre, J., Castaner, J., 2001. Pagoclone.
Rupniak, N.M.J., 1999. The substance P antagonist L-760,735 Drugs Future 26, 651–657.
inhibits stress-induced NK1 receptor internalisation in the basolateral Sousa, N., Lukoyano, N.V., Madeira, M.D., Almeida, O.F., Paula-Barbosi,
amygdala. Brain Res. 848, 90–95. M.M., 2000. Reorganization of the morphology of hippocampal
234 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

neurites and synapses after stress-induced damage correlates with Steckler, T., Holsboer, F., 1999. Corticotropin-releasing hormone receptor
behavioural improvement. Neuroscience 97, 253–266. subtypes and emotion. Biol. Psychiatry 46, 1480–1508.
Spadoni, F., Hainsworth, A.H., Mercuri, N.B., Caputi, L., Martella, G., Stefani, A., Spadoni, F., Bernardi, G., 1998a. GABApentin inhibits calcium
Lavaroni, F., Bernardi, G., Stefani, A., 2002. Lamotrigine derivatives currents in isolated rat brain neurons. Neuropharmacology 37, 83–91.
and riluzole inhibit INa,P in cortical neurons. NeuroReport 13, Stefani, A., Spadoni, F., Bernardi, G., 1998b. Group III metabotropic
1167–1170. glutamate receptor agonists modulate high voltage-activated Ca2+
Spanagel, R., Montkowski, A., Allingham, K., Stohr, T., Shoaib, M., currents in pyramidal neurons of the adult rat. Exp. Brain Res. 119,
Holsboer, F., Landgraf, R., 1995. Anxiety—a potential predictor of 237–244.
vulnerability to the initiation of ethanol self administration in rats. Stefański, R., Palejko, W., Kostowski, W., Plaźnik, A., 1992. The
Psychopharmacology 122, 369–373. comparison of benzodiazepine derivatives and serotonergic agonists
Spealman, R.D., Kelleher, R.T., Goldberg, S.R., DeWeese, J., Goldberg, and antagonists in two animal models of anxiety. Neuropharmacology
D.M., 1983. Behavioral effects of clozapine: comparison with 31, 1251–1258.
thioridazine, chlorpromazine, haloperidol and chlordiazepoxide in Stefański, R., Palejko, W., Bidzinski, A., Kostowski, W., Plaźnik, A.,
squirrel monkeys. J. Pharmacol. Exp. Ther. 224, 127–134. 1993a. Serotonergic innervation of the hippocampus and nucleus
Spencer, R.L., McEwen, B.S., 1990. Adaptation of the hypothalamic– accumbens septi and the anxiolytic-like action of midazolam and
pituitary–adrenal axis to chronic ethanol stress. Neuroendocrinology 5-HT1A receptor agonists. Neuropharmacology 32, 977–985.
52, 481–489. Stefański, R., Palejko, W., Bidzinski, A., Kostowski, W., Plaźnik, A.,
Sperling, W.L., Demling, J., 1997. New tetracyclic antidepressants. Drugs 1993b. Serotonergic innervation of the hippocampus and nucleus
Today 33, 95–102. accumbens septi and the anxiolytic-like action of the 5-HT3 receptor
Spooren, W.P.J.M., Vassout, A., Neijt, H.C., Kuhn, R., Gasparini, F., antagonists. Neuropharmacology 32, 987–993.
Roux, S., Porsolt, R.D., Gentsch, C., 2000. Anxiolytic-like effects Steimer, T., Driscoll, P., Schulz, P.E., 1997. Brain metabolism of
of the prototypical metabotropic glutamate receptor 5 antagonist 2- progesterone, coping behaviour and emotional reactivity in male rats
methyl-6-(phenylethynyl)pyridine in rodents. J. Pharmacol. Exp. Ther. from two psychogenetically selected lines. J. Neuroendocrinol. 9,
295, 1267–1275. 169–175.
Stein, M.B., Chartier, M.J., Lizak, K.L., 2001a. Familial aggregation of
Spooren, W.P.J.M., Gazparini, F., Salt, T.E., Kuhn, R., 2001. Novel
anxiety-related quantitative traits in generalized social phobia: clues to
allosteric antagonists shed light on mglu5 receptors and CNS disorders.
understanding “disorder” heritability? Am. J. Med. Genet. 105, 79–83.
Trends Pharmacol. Sci. 22, 331–337.
Stein, M.B., Sareen, J., Hami, S., Chao, J., 2001b. Pindolol potentiation
Sprouse, J.S., Reynolds, L.S., Braselton, J.P., Rollema, H., Zorn, S.H.,
of paroxetine for generalized social phobia: a double-blind, placebo-
1999. Comparison of the novel antipsychotic ziprasidone with clozapine
controlled, crossover study. Am. J. Psychiatry 158, 1725–1727.
and olanzapine: inhibition of dorsal raphe cell firing and the role of
Stein, D.J., Westenberg, H.G., Liebowitz, M.R., 2002. Social anxiety
5-HT1A receptor activation. Neuropsychopharmacology 21, 622–631.
disorder and generalized anxiety disorder: serotonergic and
Sprouse, J.S., Braselton, J., Reynolds, L., 2000. 5-HT1A agonist potential
dopaminergic neurocircuity. J. Clin. Psychiatry 63, 12–19.
of pindolol: electrophysiologic studies in the dorsal raphe nucleus and
Steinberg, V., Marco, N., Voutsinos, B., Bensaid, M., Rodier, D.,
hippocampus. Biol. Psychiatry 47, 1050–1055.
Souilhac, J., Alonso, R., Oury-Donat, F., Le Fur, G., Soubrié, P., 1998.
Spruijt, B.M., Van Hooff, J.A., Gispen, W.H., 1992. Ethology and
Expression and presence of septal neurokinin2 receptors controlling
neurobiology of grooming behaviour. Physiol. Rev. 72, 825–852.
hippocampal acetylcholine release during sensory stimulation in rat.
Sramek, J.J., Costa, J.F., Adams, J.B., 1995. Single-site findings in a study
Eur. J. Neurosci. 10, 2337–2345.
of the safety and efficacy of a CCKB receptor antagonist, CI-988, in Steinberg, R., Alonso, R., Griebel, G., Bert, L., Jung, M., Oury-Donat, F.,
the treatment of generalized anxiety disorder. Anxiety 1, 242–243. Poncelet, M., Gueudet, C., Desvignes, C., Le Fur, G., Soubrié, P., 2001.
Sramek, J.J., Zarotsky, V., Cutler, N.R., 2002. Generalised anxiety Selective blockade of neurokinin2 receptors produces antidepressant-
disorder. Drugs 62, 1635–1648. like effects associated with reduced corticotropin-releasing factor
Srisurapanon, M., Boonyanaruthee, V., 1997. Alprazolam and standard function. J. Pharmacol. Exp. Ther. 299, 449–458.
antidepressants in the treatment of depression: a meta-analysis of the Steinberg, R., Alonso, R., Rouquier, L., Desvignes, C., Michaud, J.-
antidepressant effect. J. Med. Assoc. Thai. 80, 183–188. C., Cudennec, A., Jung, M., Simiand, J., Griebel, G., Emonds-Alt,
Stacey, A.E., Woodhall, G.L., Jones, R.S.G., 2002a. Activation of X., Le Fur, G., Soubrié, P., 2002. SSR240600 [(R)-2-(1-{2-[4-
neurokinin-1 receptors promotes GABA release at synapses in the rat {2-[35-bis(trifluoromethyl)phenyl]acetyl}- 2 - (3,4 - dichlorophenyl) -2-
entorhinal cortex. Neuroscience 115, 575–586. morpholinyl]ethyl}-4-piperidinyl)-2-methypropanamide], a centrally-
Stacey, A.E., Woodhall, G.L., Jones, R.S.G., 2002b. Neurokinin-receptor- active, nonpeptide antagonist of the tachykinin neurokinin 1 receptor.
mediated depolarization of cortical neurons elicits an increase in II. Neurochemical and behavioral characterization. J. Pharmacol.
glutamate release at excitatory synapses. Eur. J. Neuroscience 16, Exp. Ther. 303, 1180–1188.
1896–1906. Steiner, H., Fuchs, S., Accili, D., 1998. D3 dopamine receptor-deficient
Stahl, S.M., Mendels, J., Schwartz, G.E., 2002. Effects of reboxetine mouse: evidence for reduced anxiety. Physiol. Behav. 63, 137–141.
on anxiety, agitation, and insomnia: results of a pooled evaluation of Stell, B.M., Mody, I., 2002. Receptors with different affinities mediate
randomized clinical trials. J. Clin. Psychopharmacol. 22, 388–392. phasic and tonic GABAA conductances in hippocampal neurons. J.
Stanek, L., Walker, D.L., Davis, M., 2000. Amygdala infusion of Neurosci. 22, 1–5.
LY354,740, a group II metabotropic glutamate receptor agonist, blocks Stella, N., Piomelli, D., 2001. Receptor-dependent formation of
fear potentiated startle in rats. Soc. Neurosci. Abstr. 26, 2020. endogenous cannabinoids in cortical neurons. Eur. J. Pharmacol. 425,
Stanhope, K.J., Dourish, C.T., 1996. Effects of 5-HT1A receptor agonists, 189–196.
partial agonists and a silent antagonist on the performance of the Stenzel-Poore, M.P., Heinrichs, S.C., Rivest, S., Koob, G.F., Vale, W.W.,
conditioned emotional response test in the rat. Psychopharmacology 1994. Overproduction of corticotropin-releasing factor in transgenic
128, 293–303. mice: a genetic model of anxiogenic behavior. J. Neurosci. 14,
Stanley, S.A., Small, C.J., Murphy, K.G., Rayes, E., Abbott, C.R., Seal, 2579–2584.
L.J., Morgan, D.G., Sunter, D., Dakin, C.L., Kim, M.S., Hunter, R., Stephens, D.N., Schneider, H.H., Kehr, W., Andrews, J.S., Rettig, K.-J.,
Kuhar, M., Ghatei, M.A., Bloom, S.R., 2001. Actions of cocaine- and Turski, L., Schmiechen, R., Turner, J.D., Jensen, L.H., Petersen,
amphetamine-regulated transcript (CART) peptide on regulation of E.N., Honore, T., Hansen, B.J., 1990. Abecarnil, a metabolically
appetite and hypothalamo-pituitary axes in vitro and in vivo in male stable, anxioselective ␤-carboline acting at benzodiazepine receptors.
rats. Brain Res. 893, 186–194. J. Pharmacol. Exp. Ther. 253, 334–343.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 235

Stephens, D.N., Mead, A.N., Ripley, T.L., 2002. Studying the neurobiology Ströhle, A., Poettig, M., Barden, N., Holsboer, F., Montkowski, A., 1998.
of stimulant and alcohol abuse and dependence in genetically Age- and stimulus-dependent changes in anxiety-related behaviour of
manipulated mice. Behav. Pharmacol. 13, 327–345. transgenic mice with GR dysfunction. NeuroReport 9, 2099–2102.
Steppuhn, K.G., Turski, L., 1993. Diazepam dependence prevented by Ströhle, A., Romeo, E., Hermann, B., Pasini, A., Spalletta, G., Di
glutamate antagonists. Proc. Natl. Acad. Sci. U.S.A. 90, 6889–6893. Michele, F., Holsboer, F., Rupprecht, R., 1999. Concentrations of 3␣-
Stewart, R.B., Gatto, G.J., Lumeng, L., Li, T.K., Murphy, J.M., 1993. reduced neuroactive steroids and their precursors in plasma of patients
Comparison of alcohol-preferring (P) and nonpreferring (NP) rats on with major depression and after clinical recovery. Biol. Psychiatry 45,
tests of anxiety and for the anxiolytic effects of ethanol. Alcohol 10, 274–277.
1–10. Ströhle, A., Romeo, E., Di Michele, F., Pasini, A., Yassouridis, A.,
Stine, S.M., Southwick, S.M., Petrakis, I.L., Kosten, T.R., Charney, Holsboer, F., Rupprecht, R., 2002. GABAA receptor-modulating
D.S., Krystal, J.H., 2002. Yohimbine-induced withdrawal and anxiety neuroactive steroid composition in patients with panic disorder before
symptoms in opioid-dependent patients. Biol. Psychiatry 51, 642–651. and during paroxetine treatment. Am. J. Psychiatry 159, 145–147.
Stoehr, J.D., Cramer, C.P., North, W.G., 1992. Oxytocin and vasopressin Ströhle, A., Romeo, E., Di Michele, F., Pasini, A., Hermann, B.,
hexapeptide fragments have opposing influences on conditioned Gajewsky, G., Hosboer, F., Rupprecht, R., 2003. Induced panic attacks
freezing behaviour. Psychoneuroendocrinology 17, 267–271. shift ␥-aminobutyric acid type A receptor modulatory neuroactive
Stone, T.W., 2001. Kynurenic acid antagonists and kynurenine pathway steroid composition in patients with panic disorder. Arch. Gen.
inhibitors. Exp. Opin. Invest. Drugs 10, 633–645. Psychiatry 60, 161–168.
Stone, T.W., Darlington, G.L., 2002. Endogenous kynurenines as targets Stroud, L.R., Salovey, P., Epel, E.S., 2002. Sex differences in stress
for drug discovery and development. Nat. Rev. 1, 609–620. responses: social rejection versus achievement stress. Biol. Psychiatry
Stone, E.A., Quartermain, D., 1999. Alpha1 -noradrenergic neurotran- 52, 318–327.
smission, corticosterone, and behavioral depression. Biol. Psychiatry Stutzmann, G.E., LeDoux, J.E., 1999. GABAergic antagonists block the
46, 1287–1300. inhibitory effects of serotonin in the lateral amygdala: a mechanism
for modulation of sensory inputs related to fear conditioning. J.
Stone, E.A., Platt, J.E., Herrera, A.S., Kirk, K.L., 1986. The effect of
Neurosci. 19 (RC8), 1–4.
repeated restraint stress, desmethylimipramine or adrenocorticotropin
Stutzmann, J., Cintrat, P., Laduron, P., Blanchard, J., 1989. Riluzole
on the ␣ and ␤ adrenergic components of the cyclic AMP response
antagonizes the anxiogenic properties of the beta-carboline FG 7142
to norepinephrine in rat brain slices. J. Pharmacol. Exp. Ther. 230,
in rats. Psychopharmacology 99, 515–519.
702–707.
Stutzmann, J.M., Eon, B., Darche, F., Lucas, M., Rataud, J., Piot, O.,
Stone, E., Zhang, Y., Rosengarten, H., Yeretsian, J., Quartermain,
Blanchard, J.C., Laduron, P.M., 1991. Are 5-HT2 antagonists endowed
D., 1999. Brain ␣1-adrenergic neurotransmission in necessary for
with anxiolytic properties in rodents? Neurosci. Lett. 128, 4–8.
behavioral activation to environmental change in mice. Neuroscience
Stutzmann, G.E., McEwen, B.S., LeDoux, J.E., 1998. Serotonin
94, 1245–1252.
modulation of sensory inputs to the lateral amygdala: dependency on
Stone, E.A., Lin, Y., Itteera, A., Quatermain, D., 2001. Pharmacological
corticosterone. J. Neurosci. 18, 9529–9538.
evidence for the role of brain alpha1B -adrenoceptors in the motor
Su, T.P., Pagliaro, M., Schmidt, P.J., Pickar, D., Wolkowtiz, O., Rubinow,
activity and spontaneous movement of mice. Neuropharmacology 40,
D.R., 1993. Neuropsychiatric effects of anabolic steroids in male
254–261.
normal volunteers. J. Am. Med. Assoc. 269, 2760–2764.
Stone, E.A., Lin, Y., Suckow, R.F., Quartermain, D., 2002. Stress-induced
Sudakov, S.K., Medvedeva, O.F., Rusakova, I.V., Terebilina, N., Goldberg,
subsensitivity to modafinil and its prevention by corticosteroids.
S.R., 2001. Differences in genetic predisposition to high-anxiety in
Pharmacol. Biochem. Behav. 73, 971–978.
two inbred rat stains: role of substance P, diazepam binding inhibitor
Stork, O., Ji, F.-Y., Stork, S., Yoshinobu, Y., Moriya, T., Shibata, S., fragment and neuropeptide Y. Psychopharmacology 154, 327–335.
Obata, K., 2000a. Decreased GABA levels and disturbance of neural Sugita, S., North, R.A., 1993. Opioid actions on neurons of rat lateral
functions in mice lacking the 65 kDa isoform of glutamic acid amygdala in vitro. Brain Res. 612, 151–155.
decarboxylase. Brain Res. 865, 45–58. Sullivan, G.M., Coplan, J.D., Kent, J.M., Gorman, J.M., 1999. The
Stork, O., Welzl, H., Wolfer, D., Schuster, T., Mantei, N., Stork, S., noradrenergic system in pathological anxiety: a focus on panic with
Hoyer, D., Lipp, H.-P., Obata, K., Schachner, M., 2000b. Recovery relevance to generalized anxiety and phobias. Biol. Psychiatry 46,
of emotional behaviour in neural cell adhesion molecule (NCAM) 1205–1218.
null mutant mice through transgenic expression of NCAM180. Eur. Sulzer, D., Joyce, M.P., Lin, L., Geldwert, D., Haber, S.N., Hattori, T.,
J. Neurosci. 12, 3291–3306. Rayport, S., 1998. Dopamine neurons make glutamatergic synapses
Stork, O., Ji, F.-Y., Obata, K., 2002a. Reduction of extracellular GABA in vitro. J. Neurosci. 18, 4588–4602.
in the mouse amygdala during and following confrontation with a Sumner, B.E., Fink, G., 1995. Estrogen increases the density of 5-HT2A
conditioned fear stimulus. Neurosci. Lett. 327, 138–142. receptors in cerebral cortex and nucleus accumbens in the female rat.
Stork, O., Kojima, N., Stork, S., Kume, N., Obata, K., 2002b. Resistance J. Steroid Biochem. Biol. 54, 15–20.
to alcohol withdrawal-induced behaviour in Fyn transgenic mice and Sundstrom-Poromaa, I., Smith, D.H., Gong, Q.H., Sabado, T.N., Li, X.,
its reversal by ifenprodil. Mol. Brain Res. 105, 126–135. Light, A., Wiedmann, M., Williams, K., Smith, S.S., 2002. Homonally
Stork, O., Yamanaka, H., Stork, S., Kume, N., Obata, K., 2003. Altered regulated ␣4 ␤2 ␦GABAA receptors are a target for alcohol. Nat.
conditioned fear behavior in glutamate decarboxylase 65 null mutant Neurosci. 5, 721–722.
mice. Genes Brain Behav. 2, 65–70. Sur, C., Wafford, K.A., Reynolds, D.S., Hadingham, K.L., Bromidge, F.,
Stout, B.D., Clarke, W.P., Berg, K.A., 2002. Rapid desensitization of Macaulay, A., Collinson, N., O’Meara, G., Howell, O., Newman, R.,
serotonin2C receptor system: effector pathway and agonist dependence. Myers, J., Atack, J.R., Dawson, G.R., McKernan, R.M., Whiting, P.J.,
J. Pharmacol. Exp. Ther. 302, 957–962. Rosahl, T.W., 2001. Loss of the major GABAA receptor subtype in
Stringer, J.L., Lorenzo, N., 1999. The reduction in paired-pulse inhibition the brain is not lethal in mice. J. Neurosci. 21, 3409–3418.
in the rat hippocampus by GABApentin is independent of GABAB Suzuki, E., Nakaki, T., Shintani, F., Kanba, S., Miyaoka, H., 2002a.
receptor activation. Epilepsy Res. 33, 169–176. Antipsychotic, antidepressant, anxiolytic, and anticonvulsant drugs
Ströhle, A., Jahn, H., Montkowski, A., Liebsch, G., Boll, E., Landgraf, induce type II nitric oxide synthase mRN in rat brain. Neurosci. Lett.
R., Holsboer, F., Wiedemann, K., 1997. Central and peripheral 333, 217–219.
administration of atriopeptin is anxiolytic in rats. Neuroendocrinology Suzuki, T., Ishigooka, J., Watanabe, S., Miyaoka, H., 2002b. Enhancement
65, 210–215. of delayed release of dopamine in the amygdala induced by conditioned
236 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

fear stress in methamphetamine-sensitized rats. Eur. J. Pharmacol. Kanatani, A., Van Der Ploeg, L.H.T., Howard, A.D., 2002b. Melanin-
435, 59–65. concentrating hormone receptor subtypes 1 and 2: species-specific
Swanson, L.W., 1982. The projections of the ventral tegmental area gene expression. Genomics 79, 785–792.
and adjacent regions: a combined fluorescent retrograde tracer and Tanaka, E., North, R.A., 1993. Actions of 5-hydroxytryptamine on
immunofluorescence study in the rat. Brain Res. Bull. 9, 321–353. neurons of the rat cingulate cortex. J. Neurophysiol. 69, 1749–1757.
Swanson, L.W., Petrovich, G.D., 1998. What is the amygdala? Trends Tanaka, M., Yoshida, H., Emoto, H., Ishii, H., 2000. Noradrenaline systems
Neurosci. 21, 323–331. in the hypothalamus, amygdala and locus coeruleus are involved in the
Swanson, T.H., Drazba, J.A., Rivkees, S.A., 1995. Adenosine A1 provocation of anxiety: basic studies. Eur. J. Pharmacol. 405, 397–406.
receptors are located predominantly on axons in the rat hippocampal Tanaka, S., Tsuchida, A., Kiuchi, Y., Oguchi, K., Numazawa, S., Yoshida,
formation. J. Comp. Neurol. 363, 517–531. T., 2003. GABAergic modulation of hippocampal glutamatergic
Swanson, G.T., Green, T., Sakai, R., Contractor, A., Che, W., Kamiya, H., neurons: an in vivo microdialysis study. Eur. J. Pharmacol. 465, 61–67.
Heinemann, S.F., 2002. Differential activation of individual subunits Tanay, VA.M.I., Greenshaw, A.J., Baker, G.B., Bateson, A.N., 2001.
in heteromeric kainate receptors. Neuron 34, 589–598. Common effects of chronically administered antipanic drugs on
Szabo, C., 1996. Physiological and pathophysiological roles of nitric brainstem GABAA receptor subunit gene expression. Mol. Psychiatry
oxide in the central nervous system. Brain Res. Bull. 41, 131–141. 6, 404–412.
Szabo, S.T., Blier, P., 2001. Serotonin1A receptor ligands act on Tang, Y.P., Shimuzu, E., Dube, G.R., Rampon, C., Kerchner, G.A., Zhuo,
norepinephrine neuron firing through excitatory amino acid and M., Liu, G., Tsien, J.Z., 1999. Genetic enhancement of learning and
GABAA receptors: a microiontophoretic study in the rat locus memory in mice. Nature 401, 25–27.
coeruleus. Synapse 42, 203–212. Tang, Y.P., Wang, H., Feng, R., Kyin, M., Tsien, J.Z., 2001. Differential
Szabo, S.T., De Montigny, C., Blier, P., 2000. Progressive attenuation effects of enrichment on learning and memory function in NR2B
of the firing activity of locus coeruleus noradrenergic neurons by transgenic mice. Neuropharmacology 41, 779–790.
sustained administration of selective serotonin reuptake inhibitors. Int. Tanganelli, S., Fuxe, K., Antonelli, T., O’Connor, W.T., Ferraro, L.,
J. Neuropsychopharmacol. 3, 1–11. 2001. Cholecystokinin/dopamine/GABA interaction in the nucleus
Sze, P.Y., Iqbal, Z., 1994. Glucocorticoid action on depolarization- accumbens: biochemical and functional correlates. Peptides 22, 1229–
dependent calcium influx in brain synaptosomes. Neuroendocrinology 1234.
59, 457–465. Tanoue, A., Koshimizu, T.-A., Tsujimoto, G., 2002. Transgenic studies
Szewczak, M.R., Corbett, R., Rush, D.K., Wilmot, C.A., Conway, P.G., of ␣1 -adrenergic receptor subtype function. Life Sci. 71, 2207–2215.
Strupczewski, J.T., Cornfeldt, M., 1995. The pharmacological profile Tao, R., Auerbach, S.B., 2002a. GABAergic and glutamatergic afferents
of iloperidone, a novel atypical antipsychotic agent. J. Pharmacol. in the dorsal raphe nucleus mediate morphine-induced increases in
Exp. Ther. 274, 1404–1413. serotonin efflux in the rat central nervous system. J. Pharmacol.
Szumlinski, K.K., Szechtman, H., 2002. D2 receptor blockade in Exp. Ther. 303, 704–710.
the dorsal raphe increases quinpirole-induced locomotor excitation. Tao, R., Auerbach, S.B., 2002b. Opioid receptor subtypes differentially
NeuroReport 13, 563–566. modulate serotonin efflux in the rat central nervous system. J.
Szyndler, J., Jarosz, H., Maciejak, P., Siemiatkowski, M., Rokicki, D., Pharmacol. Exp. Ther. 303, 549–556.
Czlonkowska, A.I., Plaznik, A., 2001. The anxiolytic-like effect of Tao, R.W., Poo, M.M., 2001. Retrograde signalling at central synapses.
nicotine undergoes rapid tolerance in a model of contextual fear Proc. Natl. Acad. Sci. U.S.A. 98, 11009–11015.
conditioning in rats. Pharmacol. Biochem. Behav. 69, 511–518. Tao, R., Ma, Z., Auerbach, S.B., 1996. Differential regulation of 5-
Takahashi, L.K., 2001. Role of CRF1 and CRF2 receptors in fear and hydroxytryptamine release by GABAA and GABAB receptors in
anxiety. Neurosci. Biobehav. Rev. 25, 627–636. midbrain raphe nuclei and forebrain of rats. Br. J. Pharmacol. 119,
Takahashi, L.K., Ho, S.P., Livanov, V., Graciani, N., Arneric, S.P., 2001a. 1375–1384.
Antagonism of CRF(2) receptors produces anxiolytic behavior in Tao, R., Ma, Z., Auerbach, S.B., 1997. Influence of AMPA/kainate
animal models of anxiety. Brain Res. 902, 135–142. receptors on extracellular 5-hydroxytryptamine in rat midbrain raphe
Takahashi, S., Horikomi, K., Kato, T., 2001b. MS-377, a novel selective and forebrain. Br. J. Pharmacol. 121, 1707–1715.
␴1 receptor ligand, reverses phencyclidine-induced release of dopamine Tarantino, L.M., Bucan, M., 2000. Dissection of behavior and psychatric
and serotonin in rat brain. Eur. J. Pharmacol. 427, 211–219. disorders using the mouse as a model. Hum. Mol. Gen. 9, 953–965.
Takahata, R., Moghaddam, B., 1998. Glutamatergic regulation of basal Tatarczyńska, E., Klodzińska, A., Kroczka, B., Chojnacka-Wójcik, E.,
and stimulus-activated dopamine release in the prefrontal cortex. J. Palucha, A., Gasparini, F., Kuhn, R., Pilc, A., 2001a. Potential
Neurochem. 71, 1443–1449. anxiolytic- and antidepressant-like effects of MPEP, a potent, selective
Takamori, S., Rhee, J.S., Rosenmund, C., Jahn, R., 2001. Identification and systematically active mGlu5 receptor antagonist. Br. J. Pharmacol.
of differentiation-associated brain-specific phosphate transporter as 132, 1423–1430.
a second vesicular glutamate transporter (VGLUT2). J. Neurosci. Tatarczyńska, E., Klodzińska, A., Kroczka, B., Chojnacka-Wójcik, E.,
21 (RC182), 1–6. Pilc, A., 2001b. The antianxiety-like effects of antagonists of group I
Takekawa, S., Asami, A., Ishihara, Y., Terauchi, J., Kato, K., Shimomura, and agonists of group II and III metabotropic glutamate receptors after
Y., Mori, M., Murakoshi, H., Kato, K., Suzuki, N., Nishimura, O., intrahippocampal administration. Psychopharmacology 158, 94–99.
Fujino, M., 2002. T-226296: a novel, orally active and selective Tauscher, J., Bagby, R.M., Javanmard, M., Christensen, B.K., Kasper,
melanin-concentrating hormone receptor antagonist. Eur. J. Pharmacol. S., Kapur, S., 2001. Inverse relationship between serotonin 5-HT1A
438, 129–135. receptor binding and anxiety: a [11 C]WAY-100635 PET investigation
Talley, E.M., Rosin, D.L., Lee, A., Guyenet, P.G., Lynch, K.R., 1996. in healthy volunteers. Am. J. Psychiatry 158, 1326–1328.
Distribution of ␣2A -adrenergic receptor-like immunoreactivity in the Taylor, C.P., Narasimhan, L.S., 1997. Sodium channels and therapy of
rat central nervous system. J. Comp. Neurol. 372, 111–134. central nervous system diseases. Adv. Pharmacol. 39, 47–98.
Tan, C.M., Wilson, M.H., MacMillan, L.B., Kobilka, B.K., Limbird, Taylor, C.P., Gee, N.S., Su, T.Z., Kocsis, J.D., Welty, D.F., Brown, J.P.,
L.E., 2002a. Heterozygous alpha2A -adrenergic receptor mice unveil Dooley, D.J., Boden, P., Singh, L., 1998. A summary of mechanistic
unique therapeutic benefits of partial agonists. Proc. Natl. Acad. Sci. hypotheses of GABApentin pharmacology. Epilepsy Res. 29, 233–249.
U.S.A. 99, 12471–12476. Tecott, L.H., Chu, H.M., Brennan, T.J., 1998. Neurobehavioral analysis
Tan, C.P., Sano, H., Iwaasa, H., Pan, J., Sailer, A.W., Hrniuk, D.L., of 5-HT6 receptor null mutant mice. In: Proceedings of the Fourth
feighner, S.D., Palyha, O.C., Pong, S.-S., Figueroa, D.J., Austin, C.P., IUPHAR Satellite Meeting on Serotonin, Rotterdam, July 1998, p. 23
Jiang, M.M., Yu, H., Ito, J., Ito, M., Guan, X.M., MacNeil, D.J., (Abstract).
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 237

Teixeira, K.V., Carobrez, A.P., 1999. Effects of glycine or (±)-3-amino- Thorsell, A., Carlsson, K., Ekman, R., Heilig, M., 1999. Behavioral
1-hydroxy-2-pyrrolidone microinjections along the rostrocaudal axis and endocrine adaptation and up-regulation of NPY expression in
of the dorsal periaqueductal gray matter on rat performance in the rat amygdala following repeated restraint stress. NeuroReport 10,
elevated plus-maze task. Behav. Neurosci. 113, 196–203. 3003–3007.
Teixeira, R.M., Santos, A.R.S., Ribeiro, S.J., Calixto, J.B., Rae, G.A., De Thorsell, A., Michalkiewicz, M., Dumont, Y., Quirion, R., Caberlotto, L.,
Lima, T.C.M., 1996. Effects of central administration of tachykinin Rimondini, R., Mathé, A.A., Heilig, M., 2000. Behavioral insensitivity
receptor agonists and antagonists on plus-maze behavior in mice. Eur. to restraint stress, absent fear suppression of behavior and impaired
J. Pharmacol. 311, 7–14. spatial learning in transgenic rats with hippocampal neuropeptide Y
Tejani-Butt, S.M., Pare, W.P., Yang, J., 1994. Effect of repeated novel overexpression. Proc. Natl. Acad. Sci. U.S.A. 97, 12852–12853.
stressors on depressive behavior and brain norepinephrine receptor Tiihonen, J., Kuikka, J., Bergström, K., Lepola, U., Koponen, H.,
system in Sprague–Dawley and Wistar Kyoto (WKY) rats. Brain Res. Leinonen, E., 1997a. Dopamine reuptake site densities in patients
649, 27–35. with social phobia. Am. J. Psychiatry 154, 239–242.
Telch, M.J., Silverman, A., Schmidt, N.B., 1996. Effects of anxiety Tiihonen, J., Kuikka, J., Rasanen, P., Lepola, U., Koponen, H., Liuska,
sensitivity and perceived control on emotional responding to caffeine A., Lehmusvaara, A., Vainio, P., Kononen, M., Bergstrom, K., Yu, M.,
challenge. J. Anxiety Disord. 10, 21–35. Kinnunen, I., Akerman, K., Karhu, J., 1997b. Cerebral benzodiazepine
Tellez, S., Colpaert, F., Marien, M., 1997. Acetylcholine release in the receptor binding and distribution in generalized anxiety disorder: a
rat prefrontal cortex in vivo: modulation by ␣2 -adrenoceptor agonists fractal analysis. Mol. Psychiatry 2, 463–471.
and antagonists. J. Neurochem. 68, 778–785. Timmerman, W., Cisci, G., Nap, A., De Vries, J.B., Westerink, B.H.C.,
Temel, Y., Helmy, A., Pinnock, S., Herbert, J., 2003. Effect of serotonin 1999. Effects of handling on extracellular levels of glutamate and other
depletion on the neuronal, endocrine and behavioural responses to amino acids in various areas of the brain measured by microdialysis.
corticotropin-releasing factor in the rat. Neurosci. Lett. 338, 139–142. Brain Res. 833, 150–160.
Terry, P., Salmon, P., 1991. Anxiolytic-like action of beta-blockers: effects Timothy, C., Costall, B., Smythe, J.W., 1999. Effects of SCH23390 and
of stimulus salience. Pharmacol. Biochem. Behav. 39, 597–603. raclopride on anxiety-like behaviour in rats tested in the black–white
Tessitore, A., Hariri, A.R., Fera, F., Smith, W.G., Chase, T.N., Hyde, box. Pharmacol. Biochem. Behav. 62, 323–327.
T.M., Weinberger, D.R., Mattay, V.S., 2002. Dopamine modulates the Timpl, P., Spanagel, R., Sillaber, I., Kresse, A., Reul, J.M.H.M., Stalla,
response of the human amygdala: a study in Parkinson’s disease. J. G.K., Blanquet, V., Steckler, T., Holsboer, F., Wurst, W., 1998. Impaired
Neurosci. 22, 9099–9103. stress response and reduced anxiety in mice lacking a functional
Thanky, N.R., Son, J.H., Herbison, A.E., 2002. Sex differences in the corticotropin-releasing hormone receptor. Nat. Genet. 19, 162–166.
regulation of tyrosine hydroxylase gene transcription by estrogen in Tizzano, J.P., 2002. Behavioral characterization of the selective AMPA
the locus coeruleus of TH9-LacZ transgenic mice. Mol. Brain Res. antagonist (GYKI 52466) in the fear potentiated startle model in rats.
104, 220–226. Soc. Neurosci. Abstr. 396.1.
Thiébot, M.-H., Hamon, M., Soubrié, P., 1983. The involvement of Tizzano, J.P., Griffey, K.I., Schoepp, D.D., 2002. The anxiolytic action
nigral serotonin innervation in the control of punishment-induced of mGlu2/3 receptor agonist LY354740, in the fear-potentiated startle
behavioural inhibition in rats. Biochem. Behav. 19, 225–229. model in rats is mechanistically distinct from diazepam. Pharmacol.
Thiébot, M.-H., Soubrié, P., Sanger, D., 1988. Anxiogenic properties Biochem. Behav. 73, 367–374.
of beta-CCE and FG 7142: a review of promises and pitfalls. To, C.T., Bagdy, G., 1999. Anxiogenic effect of central CCK
Psychopharmacology 94, 452–463. administration is attenuated by chronic fluoxetine or ipsapirone
Thiébot, M.-H., Dangoumau, L., Richard, G., Puech, A.J., 1991. treatment. Neuropharmacology 38, 279–282.
Safety signal withdrawal: a behavioural paradigm sensitive to both To, C.T., Anheuer, Z.E., Bagdy, G., 1999. Effects of acute and chronic
“anxiolytic” and “anxiogenic” drugs under identical experimental fluoxetine treatment on CRH-induced anxiety. NeuroReport. 10,
conditions. Psychopharmacology 103, 415–424. 553–555.
Thiel, C.M., Schwarting, R.K., 2001. Dopaminergic lateralization in the Todorovic, C., Radulovic, J., Spiess, J., 2002. Interrelationship between
forebrain: relations to behavioural asymmetries and anxiety in male anxiety-like behavior and fear conditioning—the role of corticotropin
Wistar rats. Neuropsychobiology 43, 192–199. releasing receptor 2 . Soc. Neurosci. Abstr. 370.15.
Thomas, L.S., Jane, D.E., Harris, J.R., Croucher, M.J., 2000. Tokunaga, S., McDaniel, J.R., Morrow, A.L., Matthews, D.B., 2003.
Metabotropic glutamate autoreceptors of the mGlu(5) subtype Effect of acute ethanol administration and acute allopregnanolone
positively modulate neuronal glutamate release in the rat forebrain in administration on spontaneous hippocampal pyramidal cell neural
vitro. Neuropharmacology 39, 1554–1566. activity. Brain Res. 967, 273–280.
Thomas, L.S., Jane, D.E., Gasparini, F., Croucher, M.J., 2001. Glutamate Tokuyama, S., Hirata, K., Yoshida, A., Maruo, J., Matsumo, K., Mita,
release inhibiting properties of the novel mGlu(5) receptor antagonist S., Ueda, H., 1999. Selective coupling of mouse brain metabotropic
2-methyl-6-(phenylethynyl)-pyridine (MPEP): complementary in vitro sigma (␴) receptor with recombinant Gi1 . Neurosci. Lett. 268, 85–88.
and in vivo evidence. Neuropharmacology 41, 523–527. Tollefson, G.D., Sanger, T.M., Beasley, C.M., Tran, P.V., 1998. A double-
Thomas, E., Rana, H.Q., Hoeldtke, K.I., 2002. Anxiolytic effects of a blind, controlled comparison of the novel antipsychotic olanzapine
corticotropin releasing factor antagonist administered to the lateral versus haloperidol or placebo on anxious and depressive symptoms
septum of the rat. Soc. Neurosci. Abstr. 571.2. accompanying schizophrenia. Biol. Psychiatry 43, 803–810.
Thomas, E., Pernar, L., Lucki, I., Valentino, R.J., 2003. Corticotropin- Tolwany, R.J., Buckmaster, P.S., Varma, S., Cosgaya, J.M., Wu, Y., Suri,
releasing factor in the dorsal raphe nucleus regulates activity of lateral C., Shooter, E.M., 2002. BDNF overexpression increases dendrite
septal neurons. Brain Res. 960, 201–208. complexity in hippocampal dentate gyrus. Neuroscience 114, 795–805.
Thomson, I.R., Peterson, M.D., Hudson, R.J., 1998. A comparison of Tomasini, M.C., Ferraro, L., Bebe, B.W., Tanganelli, S., Cassano, T.,
clonidine with conventional pre-anesthetic medication in patients Cuomo, V., Antonelli, T., 2002. 9 -Tetrahydrocannabinol increases
undergoing coronary artery bypass grafting. Anesth. Analg. 87, endogenous extracellular glutamate levels in primary cultures of rat
292–299. cerebral cortex neurons: involvement of CB1 receptors. J. Neurosci.
Thompson, S.A., Bonnert, T.P., Cagetti, E., Whiting, P.J., Wafford, K.A., Res. 68, 449–453.
2002. Overexpression of the GABAA receptor ε subunit results in Tordera, R.M., Monge, A., Del Rio, J., Lasheras, B., 2002. Antidepressant-
insensitivity to anaesthetics. Neuropharmacology 43, 662–668. like actions of VN2222, a serotonin reuptake inhibitor with high
Thorsell, A., Heilig, M., 2002. Diverse functions of neuropeptide Y affinity at 5-HT1A receptors. Eur. J. Pharmacol. 442, 63–71.
revealed using genetivoltage-dependentified animals. Neuropeptides Toth, M., 2003. 5-HT1A receptor knock-out mouse as a genetic model
36, 182–193. of anxiety. Eur. J. Pharmacol. 463, 177–184.
238 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Treit, D., 1985. Animal models for the study of anti-anxiety agents: a Umezu, T., 1999. Effects of psychoactive drugs in the Vogel Conflict
review. Neurosci. Biobehav. Rev. 9, 203–222. Test in mice. Jpn. J. Pharmacol. 80, 111–118.
Treit, D., 1994. Animal models of anxiety and anxiolytic drug action. Urani, A., Romieu, P., Portales-Casamar, E., Roman, F.J., Maurice,
In: Den Boer, J.A., Sutsen, J.M.A. (Eds.), Handbook of Depression T., 2002. The antidepressant-like effect induced by the sigma1 (␴1 )
and Anxiety. Marcel Dekker, New York, pp. 201–224. receptor agonist igmesine involves modulation of intracellular calcium
Treit, D., Degroot, A., Kashluba, S., Bartoszyk, G.D., 2001. Systemic mobilization. Psychopharmacology 163, 26–35.
EMD 68843 injections reduce anxiety in the shock-probe, but not the Urwyler, S., Mosbacher, J., Lingenhoehl, K., Heid, J., Hofstetter, K.,
plus-maze test. Eur. J. Pharmacol. 414, 245–248. Froestl, W., Bettler, B., Kaupmann, K., 2001. Positive allosteric
Tronche, F., Kellendonk, C., Kretz, G., Gass, P., Anlag, K., Orban, P.C., modulation of native and recombinant ␥-aminobutyric acidB
Bock, R., Klein, R., Schutz, G., 1999. Disruption of the glucocorticoid receptors by 2,6-di-tert-butyl-4-(3-hydroxy-2,2-dimethyl-propyl)-
receptor gene in the nervous system results in reduced anxiety. Nat. phenol (CGP7930) and its aldehyde analog CGP13501. Am. Soc.
Genet. 23, 99–103. Pharmacol. Ther. 60, 963–971.
Trottier, S., Chotard, C., Traiffort, E., unmehopa, U., Fisser, B., Swaab, Uvnäs-Moberg, K., 1998. Antistress pattern induced by oxytocin. News
D.F., Schwartz, J.-C., 2002. Co-localization of histamine with GABA Physiol. Sci. 13, 22–25.
but not with galanin in the human tuberomamillary nucleus. Brain Uvnäs-Moberg, K., Ahlenius, S., Hillegaart, V., Alster, P., 1994. High
Res. 939, 52–64. doses of oxytocin cause sedation and low doses cause an anxiolytic-
Trullas, R., Folio, T., Young, A., Miller, R., Boje, K., Skolnick, P., like effect in male rats. Pharmacol. Biochem. Behav. 49, 101–106.
1991. 1-Aminocyclopropanecarboxylates exhibit antidepressant and Uzunov, D.P., Cooper, T.B., Costa, E., Guidotti, A., 1996. Fluoxetine-
anxiolytic actions in animal models. Eur. J. Pharmacol. 203, 379–385. elicited changes in brain neurosterol content measured by negative ion
Tsou, K., Brown, S., Sanudo-Pena, M.C., Mackie, K., Walker, J.M., mass fragmentography. Proc. Natl. Acad. Sci. U.S.A. 93, 12599–13604.
1998. Immunohistochemical distribution of cannabinoid CB1 receptors Uzunova, V., Sheline, Y., Davis, J.M., Rasmusson, A., Uzunov, D.P.,
in the rat central nervous system. Neuroscience 83, 393–411. Costa, E., Guidotti, A., 1998. Increase in the cerebrospinal fluid
Tsuda, A., Tanaka, M., Georgiev, V., Emoto, H., 1992. Effects of content of neurosteroids in patients with unipolar major depression
angiotensin II on behavioural responses of defensive burying paradigm who are receivieng fluoxetine of fluvoxamine. Proc. Natl. Acad. Sci.
in rats. Pharmacol. Biochem. Behav. 43, 729–732. U.S.A. 95, 3239–3244.
Tsuda, M., Suzuki, T., Misawa, M., Nagase, H., 1996. Involvement of Vaccari, C., Lolait, S.J., Ostrowski, N.L., 1998. Comparative distribution
the opioid system in the anxiolytic effect of diazepam in mice. Eur. of vasopressin V1b and oxytocin receptor messenger ribonucleic acids
J. Pharmacol. 307, 7–14. in brain. Endocrinology 139, 5015–5033.
Tsuiki, K., Blier, P., Diksic, M., 2000. Effect of the ␤-adrenoceptor
Vaidya, V.A., Terwilliger, R.Z., Duman, R.S., 1999. Role of 5-HT2A
agonist flerobuterol on serotonin synthesis in the rat brain. Biochem.
receptors in the stress-induced down-regulation of brain-derived
Pharmacol. 59, 673–679.
neurotrophic factor expression in rat hippocampus. Neurosci. Lett.
Tsutsumi, T., Akiyoshi, J., Isogawa, K., Kohno, Y., Hikichi, T., Nagayama,
262, 1–4.
H., 1999. Suppression of conditioned fear by administration of CCKB
Vaidya, V.A., Castro, M.E., Pei, Q., Sprakes, M.E., Grahame-Smith, D.G.,
receptor antagonist PD135158. Neuropeptides 33, 483–486.
2001. Influence of thyroid hormone on 5-HT1A and 5-HT2A receptor-
Tuinstra, T., Cools, A.R., 2000. High and low responders to novelty: effects
mediated regulation of hippocampal BDNF mRNA expression.
of adrenergic agents on the regulation of accumbal dopamine under
Neuropharmacology 40, 48–56.
challenged and non-challenged conditions. Neuroscience 99, 55–64.
Vaidya, V.A., Rosenthal, D., Crooke, J., Reitz, A.B., 2002. Anxiolytic
Turner, A.J., Hooper, N.M., 2002. The angiotensin-converting enzyme
activity of orally administered buspirone in the elevated plus-maze and
gene family: genomics and pharmacology. Trends Pharmacol. Sci. 23,
Vogel Conflict Tests in Long–Evans rats. Soc. Neurosci. Abstr. 398.16.
177–183.
Turski, L., Jacobsen, P., Honoré, T., Syephens, N., 1992. Relief of Valdez, G.R., Inoue, K., Koob, G.F., rivier, J., Vale, W., Zorrilla, E.P.,
experimental spasticity and anxiolytic/anticonvulsant actions of the 2002a. Human urocortin II: mild locomotor suppressive and delayed
alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionate antagonist anxiolytic-like effects of a novel corticotropin-releasing factor related
2,3-dihydroxy-6-nitro-7-sulfamoyl-benzo(F)quinoxaline. J. Pharmacol. peptide. Brain Res. 943, 142–150.
Exp. Ther. 260, 742–747. Valdez, G.R., Zorrilla, E.P., Rivier, J., Vale, W.W., Koob, G.F., 2002b.
Turton, M.D., O’Shea, D., Gunn, I., Beak, S.A., Edwards, C.M., Meeran, Behavioral effects of urocortin III, a novel, highly selective CRF2
K., Choi, S.J., Taylor, G.M., Heath, M.M., Lambert, P.D., Wilding, receptor agonist, in animal models of anxiety. Soc. Neurosci. Abstr.
J.P., Smith, D.M., Ghatei, M.A., Herbert, J., Bloom, S.R., 1996. A 683.8.
role for glucagon-like peptide-1 in the central regulation of feeding. Vale, A.L., Green, S., 1986. Effects of chlordiazepoxide, nicotine and
Nature 379, 69–72. d-amphetamine in the rat potentiated startle model of anxiety. Behav.
Twist, E.C., Mitchell, S., Brazell, C., Stahl, S.M., Campbell, I.C., 1990. Pharmacol. 7, 138–143.
5-HT2 receptor changes in rat cortex and platelets following chronic Vale, A.L., Green, S., Montgomery, A.M., Shafi, S., 1998. The nitric
ritanserin and clorgyline administration. Biochem. Pharmacol. 39, oxide synthesis inhibitor l-NAME produces anxiogenic-like effects in
161–166. the rat elevated plus-maze test, but not in the social interaction test.
Tyrer, P., 1988. Current status of ␤-blocking drugs in the treatment of J. Psychopharmacol. 12, 268–272.
anxiety disorders. Drugs 36, 773–783. Valentino, R.J., Aston-Jones, G., 1995. Physiological and anatomical
Tyrer, P., 1992. Anxiolytics not acting at the benzodiazepine receptor: beta determinants of locus coeruleus discharge. In: Bloom, F., Kupfer,
blockers. Prog. Neuropsychopharmacol. Biol. Psychiatry 16, 17–26. D. (Eds.), Psychopharmacology: The Fourth Generation of Progress.
Ueki, S., Watanabe, S., Yamamoto, T., Kataoka, Y., Shibata, S., Raven, New York, pp. 373–385.
Shibata, K., Ohta, H., Shimazoe, T., Kawamoto, H., 1987. Behavioral Valentino, R.J., Van Bockstaele, E., 2001. Opposing regulation of
and electroencephalographic effects of zopiclone, a cyclopyrrolone the locus coeruleus by corticotropin-releasing factor and opioids.
derivative. Jpn. J. Pharmacol. 43, 309–326. Psychopharmacology 158, 331–342.
Uhde, T.W., Boulenger, J.P., Jimerson, D.C., Post, R.M., 1984a. Valentino, R.J., Curtis, A.L., Parris, D.G., Wehby, R.G., 1990.
Caffeine and behaviour: relation to psychopathology and underlying Antidepressant actions on brain noradrenergic neurons. J. Pharmacol.
mechanisms. Psychopharmacol. Bull. 20, 426–430. Exp. Ther. 253, 833–840.
Uhde, T.W., Boulenger, J.P., Post, R., Siever, L.J., Vittone, B.J., Jimerson, Valentino, R.J., Page, M., Van Bockstaele, E., Aston-Jones, G., 1992.
D.C., Roy-Birne, P.P., 1984b. Fear and anxiety: relationship to Corticotropin-releasing factor innervation of the locus coeruleus region:
noradrenergic function. Psychopharmacology 17, 8–23. distribution of fibers and sources of input. Neuroscience 48, 689–705.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 239

Vallee, M., Rivera, J.D., Koob, G.F., Purdy, R.H., Fitzgerald, R.L., 2000. Characterization of the anxiolytic properties of a novel neuroactive
Quantification of neurosteroids in rat plasma and brain following swim steroid, Co 2-6749 (GMA-839; WAY-141839; 3␣,21-dihydroxy-3␤-
stress and allopregnanolone administration using negative chemical trifluoromethyl-19-nor-5␤-pregnan-20-one), a selective modulator of ␥-
ionization gas chromatography/mass spectrometry. Anal. Biochem. aminobutyric acidA receptors. J. Pharmacol. Exp. Ther. 295, 337–345.
287, 153–166. Van Pett, K., Viau, V., Bittencourt, J.C., Chan, R.K.W., Li, H.Y.,
Vallone, D., Picetti, R., Borrelli, E., 2000. Structure and function of Arias, C., Prins, G.S., Perrin, M., Vale, W., Sawchenko, P.E., 2000.
dopamine receptors. Neurosci. Biobehav. Rev. 24, 125–132. Distribution of mRNAs encoding CRF receptors in brain and pituitary
Vallone, D., Pignatelli, M., Grammatikopoulos, G., Ruocco, L., Bozzi, Y., of rat and mouse. J. Comp. Neurol. 428, 191–212.
Westphal, H., Borrelli, E., Sadile, A.G., 2002. Activity, non-selective Van Sickle, B.J., Tietz, E.I., 2002. Selective enhancement of AMPA
attention and emotionality in dopamine D2 /D3 receptor knock-out receptor-mediated function in hippocampal CA1 neurons from chronic
mice. Behav. Brain Res. 130, 141–148. benzodiazepine-treated rats. Neuropharmacology 43, 11–27.
Van Ameringen, M., Mancini, C., Oakman, J.M., 1999. Nefazodone in Van Sickle, B.J., Cox, A.S., Achak, K., Greenfield, L.J., Tietz, E.I.,
social phobia. J. Clin. Psychiatry 60, 96–100. 2002. Chronic benzodiazepine administration alters hippocampal
Van Amelsvoort, T., Compton, J., Murphy, D., 2001. In vivo assessment CA1 neuron excitability: NMDA receptor function and expression.
of the effects of estrogen on human brain. Trends Endocrinol. Metab. Neuropharmacology 43, 595–606.
12, 273–276. Van Veen, J.F., Van Vliet, I.M., Westenberg, H.G.M., 2002. Mirtazapine
Van Bockstaele, E.J., 2000. Multiple substrates for serotonergic in social anxiety disorder: a pilot study. Int. Clin. Psychopharmacol.
modulation of rat locus coeruleus neurons and relationships with 17, 315–317.
kainate receptors. Brain Res. Bull. 51, 433–442. Varga, V., Sik, A., Freund, T.F., Kocsis, B., 2002. GABAB receptors in
Van Broekhoven, F., Verkes, R.J., 2003. Neurosteroids in depression: a the median raphe nucleus: distribution and role in the serotonergic
review. Psychopharmacology 165, 97–110. control of hippocampal activity. Neuroscience 109, 119–132.
Van De Kar, L.D., Blair, M.L., 1999. Forebrain pathways mediating Varia, I., Rauscher, F., 2002. Treatment of generalized anxiety disorder
stress-induced hormone sercretion. Front. Neuroendocrinol. 20, 1–48. with citalopram. Int. Clin. Psychopharmacol. 17, 103–107.
Vandergriff, J., Rasmussen, K., 1999. The selective mGlu2/3 receptor Varoqui, H., Schafer, M.K., Zhu, H., Weihe, E., Erickson, J.D., 2002.
agonist, LY354740, attenuates morphine withdrawal-induced activation Identification of the differentiation-associated Na+ /PI transporter as
of locus coeruleus neurons and behavioral signs of morphine a novel vesicular glutamate transporter expressed in a distinct set of
withdrawal. Neuropharmacology 38, 217–222. glutamatergic synapses. J. Neurosci. 22, 142–155.
Van Der Linden, G.J.H., Stein, D.J., Van Balkom, A.J.L.M., 2000. The Varty, G.B., Cohen-Williams, M.E., Morgan, C.A., Pylak, U., Duffy,
efficacy of the selective serotonin reuptake inhibitors for social anxiety R.A., Lachowicz, J.E., Carey, G.J., Coffin, V.L., 2002a. The gerbil
disorder (social phobia): a meta-analysis of randomised controlled elevated plus-maze II: anxiolytic-like effects of selective neurokinin
trials. Int. Clin. Psychopharmacol. 15, S15–S23. NK1 receptor antagonists. Neuropsychopharmacology 27, 371–379.
Vanderschuren, L.J.M.J., Warden, G., De Vries, T.J., Mulder, A.H., Varty, G.B., Morgan, C.A., Cohen-Williams, M.E., Coffin, V.L.,
Schoffelmeer, A.N.M., 1999. Opposing role of dopamine D1 and D2 Carey, G.J., 2002b. The gerbil elevated plus-maze I: behavioural
receptors in the modulation of rat nucleus accumbens noradrenaline characterization and pharmacological validation. Neuropsycho-
release. J. Neurosci. 19, 4123–4131. pharmacology 27, 357–370.
Vanecek, J., 1998. Cellular mechanisms of melatonin action. Pharmacol. Vassout, A., Veenstra, S., Hauser, K., Ofner, S., Brugger, F., Schilling, W.,
Rev. 78, 687–721. Gentsch, C., 2000. NKP608: a selective NK1 receptor antagonist with
Van Erp, A.M.M., Tachi, N., Miczek, K.A., 2001. Short or continuous anxiolytic-like effects in the social interaction and social exploration
social stress: suppression of continuously available ethanol intake in test in rats. Regul. Pept. 96, 7–16.
subordinate rats. Behav. Pharmacol. 12, 335–342. Vasudevan, N., Ogawa, S., Pfaff, D., 2002. Estrogen and thyroid hormone
Van Gaalen, M.M., Steckler, T., 2000. Behavioural analysis of four receptor interactions: physiological flexibility by molecular specificity.
mouse strains in an anxiety test battery. Behav. Res. 115, 95–106. Physiol. Rev. 82, 923–944.
Van Gaalen, M.M., Stenzel-Poore, M.P., Holsboer, F., Steckler, T., Vatta, M., Travaglianti, M., Bianciotti, L., Coll, C., Perazzo, J., Fernandez,
2002. Effects of transgenic overproduction of CRH on anxiety-like B., 1994. Atrial natriuretic factor effects on norepinephrine uptake in
behaviour. Eur. J. Pharmacol. 15, 2007–2015. discrete telencephalic and diencephalic nuclei of the rat. Brain Res.
Van Hensbeek, A.J., Schutte, A., Reimitz, P., 2000. Onset of action 646, 324–326.
of mirtazapine on anxiety symptoms related to depression. Int. J. Vaudry, D., Gonzalez, B.J., Basille, M., Yvon, L., Fournier, A., Vaudry,
Neuropsychopharmacol. 3, S227. H., 2000. Pituitary adenylate cyclase-activating polypeptide and its
Vanhoenacker, P., Haegeman, G., Leysen, J.E., 2000. 5-HT7 receptors: receptors: from structure to functions. Pharmacol. Rev. 52, 269–324.
current knowledge and future prospects. Trends Pharmacol. Sci. 21, Venault, P., Jacquot, F., Save, E., Sara, S., Chapouthier, G., 1993.
70–77. Anxiogenic-like effects of yohimbine and idazoxan in two behavioural
Van Hooft, J.A., Dougherty, J.J., Endeman, D., Nichols, R.A., Wadman, situations in mice. Life Sci. 52, 639–645.
W.J., 2002. GABApentin inhibits presynaptic Ca2+ -influx and synaptic Venero, C., Tilling, T., Hermans-Borgmeyer, I., Schmidt, R., Schachner,
transmission in rat hippocampus and neocortex. Eur. J. Pharmacol. M., Sandi, C., 2002. Chronic stress induces opposite changes in the
449, 221–228. mRNA expression of the cell adhesion molecules NCAM and L1.
Van Megen, H.J.G.M., Westenberg, H.G.M., Den Boer, J.A., Kahn, R.S., Neuroscience 115, 1211–1219.
1996. Cholecystokinin in anxiety. Eur. Neuropsychopharmacol. 6, Versiani, M., Cassano, G., Perugi, G., Benedetti, A., Mastalli, L., Nardi,
263–280. A., Savino, M., 2002. Reboxetine, a selective norepinephrine reuptake
Van Oekelen, D., Luyten, W.H.M.L., Leysen, J.E., 2003. 5-HT2A and inhibitor, is an effective and well-tolerated treatment for panic disorder.
5-HT2C receptors and their atypical regulation properties. Life Sci. J. Clin. Psychiatry 63, 31–37.
72, 2429–2449. Vertes, R.P., 1992. Analysis of projections from the supramammillary
Vanover, K.E., Robledo, S., Huber, M., Carter, R.B., 1999. nucleus in the rat. J. Comp. Neurol. 326, 595–622.
Pharmacological evaluation of a modified conflict procedure: punished Vertes, R.P., Fortin, W.J., Crane, A.M., 1999. Projections of the median
drinking in non-water-deprived rats. Psychopharmacology 145, 333– raphe nucleus in the rat. J. Comp. Neurol. 407, 555–582.
341. Vertongen, P., Schiffmann, S.N., Gourlet, P., Robberecht, P., 1997.
Vanover, K.E., Rosenzweig-Lipson, S., Hawkinson, J.E., Lan, N.C., Autoradiographic visualization of the receptor subclasses for vasoactive
Belluzzi, J.D., Stein, L., Barrett, J.E., Wood, P.L., Carter, R.B., 2000. intestinal polypeptide (VIP) in rat brain. Peptides 18, 1547–1554.
240 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Vetter, D.E., Li, C., Zhao, L., Contarino, A., Liberman, M.C., Smith, Wada, E., Way, J., Shapira, H., Kusano, K., Lebacq-Verheyden, A.-M.,
G.W., Marchuk, Y., Koob, G.F., Heinemann, S.F., Vale, W., Lee, Battey, J.F., 1990. NeuromedinB and gastrin-releasing peptide mRNAs
K.-F., 2002. Urocortin-deficient mice show hearing impairment and are differentially distributed in the rat nervous system. J. Neurosci.
increased anxiety-like behavior. Nat. Genet. 31, 363–369. 10, 2917–2930.
Vetulani, J., Nalepa, I., 2000. Antidepressants: past, present and future. Wada, E., Way, J., Shapira, H., Kusano, K., Lebacq-Verheyden, A.-M.,
Eur. J. Pharmacol. 405, 351–363. Coy, D., Jensen, R., Battey, J., 1991. cDNA cloning, characterization,
Vetulani, J., Antkiewicz-Michaluk, L., Rokosz-Pelc, A., 1984. Chronic and brain region-specific expression of a neuromedinB -preferring
administration of antidepressant drugs increases the density of cortical bombesin receptor. Neuron 6, 421–430.
[3 H]prazosin binding sites in the rat. Brain Res. 310, 360–362. Wada, E., Watase, K., Yamada, K., Ogura, H., Yamano, M., Inomata, Y.,
Vianna, M.R.M., Cammarota, M.P., Coitinho, A.S., Medina, J.H., Eguchi, J., Yamamoto, K., Sunday, M.E., Maeno, H., Mikoshiba, K.,
Izquierdo, I., 2003. Pharmacological studies of the molecular basis of Ohki-Hamazaki, H., Wada, K., 1997. Generation and characterization
memory extinction. Curr. Neuropharmacol. 1, 89–98. of mice lacking gastrin-releasing peptide receptor. Biochem. Biophys.
Vicini, S., Losi, G., Homanics, G.E., 2002. GABAA receptor ␦ subunit Res. Commun. 239, 28–33.
deletion prevents neurosteroid modulation of inhibitory synaptic Waddington, J.L., Clifford, J.J., McNamara, F.N., Tomiyama, K.,
currents in cerebellar neurons. Neuropharmacology 73, 646–650. Koshikawa, N., Croke, D.T., 2001. The psychopharmacology–
Vilner, B.J., Bowen, W.D., 2000. Modulation of cellular calcium by molecular biology interface: exploring the behavioural roles of
␴2 -receptors: release from intracellular stores in human SK-N-SH dopamine receptor subtypes using targeted gene deletion (‘knock-out’).
neuroblastoma cells. J. Pharmacol. Exp. Ther. 292, 900–911. Prog. Neuropsychopharmacol. Biol. Psychiatry 25, 925–964.
Vincent, J.-P., Mazella, J., Kitabgi, P., 1999. Neurotensin and neurotensin Waeber, C., Sebben, M., Nieoullon, A., Bockaert, J., Dumuis, A., 1994.
receptors. Trends Pharmacol. Sci. 20, 302–309. Regional distribution and ontogeny of 5-HT4 binding sites in rodent
Vincler, M.A., Eisenach, J.C., 2003. Immunocytochemical localization brain. Neuropharmacology 33, 527–541.
of the ␣3 , ␣4 , ␣5 , ␣7 , ␤2 , ␤3 and ␤4 nicotinic acetylcholine receptor Wafford, K.A., Whiting, P.J., 1992. Ethanol potentiation of GABAA
subunits in the locus coeruleus of the rat. Brain Res. 974, 25–36. receptors requires phosphorylation of the alternatively spiced variant
Viollet, C., Vaillend, C., Videau, C., Bluet-Pajot, M.T., Ungerer, A., of the ␥2 subutnit. FEBS Lett. 313, 133–137.
L’Héritier, A., Kopp, C., Potier, B., Billard, J.-M., Schaeffer, J., Wagstaff, A.J., Cheer, S.M., Matheson, A.J., Ormrod, D., Goa, K.L.,
Smith, R.G., Rohrer, S.P., Wilkinson, H., Zheng, H., Epelbaum, 2002. Paroxetine: an update of its use in psychiatric disorders in
J., 2000. Involvement of sst2 somatostatin receptor in locomotor, adults. Drugs 62, 655–703.
exploratory activity and emotional reactivity in mice. Eur. J. Neurosci.
Wahlestedt, C., Pich, E.M., Koob, G.F., Yee, F., Heilig, M., 1993.
12, 3761–3770.
Modulation of anxiety and neuropeptide Y-Y1 receptors by antisense
Visser, S.A.G., Gladdines, W.W.F.T., Van Der Graaf, P.H., Peletier, L.A.,
oligodeoxynucleotides. Science 259, 528–531.
Danhof, M., 2002. Neuroactive steroids differ in potency but not
Waldmeier, P.C., Wicki, P., Feldtrauer, J.J., Mickel, S.J., Bittiger, H.,
in intrinsic efficacy at the GABAA receptor in vivo. J. Pharmacol.
Baumann, P.A., 1994. GABA and glutamate release affected by
Exp. Ther. 303, 616–626.
GABAB receptor antagonists with similar potency: no evidence for
Vogel, J.R. Beer, Clody, D.E., 1971. A simple and reliable conflict
pharmacologically diferent presynaptic receptors. Br. J. Pharmacol.
procedure for testing anti-anxiety agents. Psychopharmacologia 21,
113, 1515–1521.
1–7.
Walker, D.L., Davis, M., 2000. Involvement of NMDA receptors within
Vogel, R.A., Frye, G.D., Wilson, J.H., Kuhn, C.M., Koepe, K.M.,
the amygdala in short- versus long-term memory for fear conditioning
Mailman, R.M., Mueller, R.A., Breese, G.R., 1980. Attenuation of the
as assessed with fear-potentiated startle. Behav. Neurosci. 114, 1019–
effects of punishment by ethanol: comparisons with chlordiazepoxide.
1033.
Psychopharmacology 71, 123–129.
Walker, N., Lepee-Lorgeoux, I., Fournier, J., Betaner, C., Rostene,
Volke, V., Soosaar, A., Kõks, S., Bourin, M., Mannsisto, P.T., Vasar, E.,
W., Ferrara, P., Caput, D., 1998. Tissue distribution and cellular
1997. 7-Nitroindazole, a nitric oxide synthase inhibitor, has anxiolytic-
localization of the levocabastine-sensitive neurotensin receptor mRNA
like properties in exploratory models of anxiety. Psychopharmacology
in adult rat brain. Mol. Brain Res. 57, 193–200.
131, 399–405.
Volke, V., Soosaar, A., Kõks, S., Vasar, E., Mannisto, P.T., 1998. l- Walker, D.L., Rattiner, L.M., Davis, M., 2002. Group II metabotropic
Arginine abolishes the anxiolytic effect of diazepam in the elevated glutamate receptors within the amygdala regulate fear as assessed
plus maze test in rats. Eur. J. Pharmacol. 351, 287–290. with potentiated startle in rats. Behav. Neurosci. 116, 1075–1083.
Volkow, N.D., Fowler, J.S., Wang, G.-L., 2002. Role of dopamine in Walker, D.L., Toufexis, D.J., Davis, M., 2003. Role of the bed nucleus
drug reinforcement and addiction in humans: results from imaging of the stria terminalis versus the amygdala in fear, stress, and anxiety.
studies. Behav. Pharmacol. 13, 355–366. Eur. J. Pharmacol. 463, 199–216.
Volonté, M., Ceci, A., Borsini, F., 1995. Effect of the 5- Wall, P.M., Messier, C., 2001. Methodological and conceptual issues in
hydroxytryptamine3 receptor antagonist itasertron (DAU 6215) on the use of the elevated plus-maze as a psychological measurement
(+)-N-allylnormetazocine-induced dopamine release in the nucleus instrument of animal anxiety-like behavior. Neurosci. Biobehav. Rev.
accumbens and in the corpus striatum of the rat: an in vivo 25, 275–286.
microdialysis study. J. Pharmacol. Exp. Ther. 275, 358–367. Wall, R.J., Messier, C., 2002. Infralimbic kappa opioid and muscarinic
Von Gall, C., Stehle, J.H., Weaver, D.R., 2002. Mammalian melatonin M1 receptor interactions in the concurrent modulation of anxiety and
receptors: molecular biology and signal transduction. Cell Tissue Res. memory. Psychopharmacology 160, 233–244.
309, 151–162. Wall, P.M., Flinn, J., Messier, C., 2001. Infralimbic muscarinic M1
Vrang, N., Tang-Christensen, M., Larsen, P.J., Kristensen, P., 1999. receptors modulate anxiety-like behaviour and spontaneous working
Recombinant CART peptide induces c-Fos expression in central areas memory in mice. Psychopharmacology 155, 58–68.
involved in control of feeding behaviour. Brain Res. 818, 499–509. Wallis, C.J., Lal, H., 1998. A discriminative stimulus produced by 1-
Vyas, A., Mitra, R., Rao, B.S., Chattarji, S., 2002. Chronic stress induces (3-chlorophenyl)-piperazine (mCPP) as a putative animal model of
contrasting patterns of dendritic remodelling in hippocampal and anxiety. Prog. Neuropsychopharmacol. Biol. Psychiatry 22, 547–565.
amygdaloid neurons. J. Neurosci. 22, 6810–6818. Walsh, D.M., Stratton, S.C., Harvey, F.J., Beresford, I.J.M., Hagan,
Wada, T., Fukuda, N., 1991. Effects of DN-2327, a new anxiolytic, R.M., 1995. The anxiolytic-like activity of GR159897, a non-peptide
diazepam and buspirone on exploratory activity of the rat in an NK2 receptor antagonist, in rodent and primate models of anxiety.
elevated plus-maze. Psychopharmacology 104, 444–450. Psychopharmacology 121, 186–191.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 241

Walters, R.J., Hadley, S.H., Morris, K.D.W., Amin, J., 2000. Wess, J., 1996. Molecular biology of muscarinic acetylcholine receptors.
Benzodiazepines act on GABAA receptors via two distinct and Crit. Rev. Neurobiol. 10, 69–99.
separable mechanisms. Nat. Neurosci. 3, 1274–1281. Westerink, B.H.C., Cremers, T.I.F.H., De Vries, J.B., Liefers, H., Tran,
Walther, T., Voigt, J.P., Fukamizu, A., Fink, H., Bader, M., 1999. Learning N., De Boer, P., 2002. Evidence for activation of histamine H3
and anxiety in angiotensin-deficient mice. Behav. Brain Res. 100, 1–4. autoreceptors during handling stress in the prefrontal cortex of the
Walther, D.J., Peter, J.U., Bashammakh, S., Hortnagl, H., Voits, M., Fink, rat. Synapse 43, 238–243.
H., Bader, M., 2003. Synthesis of serotonin by a second tryptophan Westphalen, R.I., Hemmings, H.C., 2003. Selective depression by general
hydroxylase isoform. Science 299, 76. anesthetics of glutamate versus GABA release from isolated nerve
Wan, Q., M, H.Y., Liu, F., Braunton, J., Niznik, H.R., Pang, S.F., Brown, terminals. J. Pharmacol. Exp. Ther. 304, 1188–1196.
G.M., Wang, Y.T., 1999. Differential modulation of GABAA receptor Wettstein, J.G., Buéno, L., Junien, J.L., 1994. CCK antagonists:
function by Mel1a and Mel1b receptors. Nat. Neurosci. 2, 401–403. pharmacology and therapeutic interest. Pharmacol. Ther. 62, 267–282.
Wang, Q.P., Ochiai, H., Nakai, Y., 1992. GABAergic innervation of Wettstein, J.G., Earley, B., Junien, J.L., 1995. Central nervous system
serotonergic neurons in the dorsal raphe nucleus of the rat studied pharmacology of neuropeptide Y. Pharmacol. Ther. 65, 397–414.
by electron microscopy double immunostaining. Brain Res. Bull. 29, Whitemore, E.R., Woodward, R.M., 1997. Antagonism of N-methyl-d-
943–948. aspartate receptors by ␴ site ligands: potency, subtype-selectivity and
Wang, M., Seippel, L., Purdy, R.H., Blackström, T., 1996a. Relationship mechanisms of inhibition. J. Pharmacol. Exp. Ther. 282, 326–338.
between symptom severity and steroid variation in women with Whiting, P.J., Bonnert, T.P., McKernan, R.M., Farrar, S., Le Bourdelles, B.,
premenstrual syndrome: study on serum pregnenolone, pregnenomone Heavens, R.P., Smith, D.W., Hewson, L., Rigby, M.R., Sirinathsinghji,
sulfate, 5-alpha-pregnane-3,20-dione and 3-alpha-hydroxy-5-alpha- D.J.S., Thompson, S.A., Wafford, K.A., 1999. Molecular and
pregnan-20-one. J. Clin. Endocrinol. Metab. 81, 1076–1082. functional diversity of the expanding GABAA receptor gene family.
Wang, R., MacMillan, L.B., Fremeau, R.T., Magnuson, M.A., Lindner, J., Ann. N. Y. Acad. Sci. 868, 645–653.
Limbird, L.E., 1996b. Expression of ␣2 -adrenergic receptor subtypes Wichems, C., Li, Q., Andrews, A., Lesch, K.P., Murphy, D.L., 2000.
in the mouse brain: evaluation of spatial and temporal information Serotonin transporter knock-out mice show a spontaneous behavioural
imparted by 3 kb of 5 regulatory sequence for the ␣2A AR-receptor phenotype of increased “anxiety” and stress responses. Int. J.
gene in transgenic animals. Neuroscience 74, 199–218. Neuropsychopharmacol. 3, S47.
Wang, S.-J., Coutinho, V., Sivra, T.S., 2002a. Presynaptic cross-talk of Wick, M.J., Radcliffe, R.A., Bowers, B.J., Mascia, M.P., Luscher, B.,
␤-adrenoceptor and 5-hydroxytryptamine receptor signalling in the Harris, R.A., Wehner, J.M., 2000. Behavioral changes produced by
modulation of glutamate release from cerebrocortical nerve terminals. transgenic overexpression of ␥2L and ␥2S subunits of the GABAA
Br. J. Pharmacol. 137, 1371–1379. receptor. Eur. J. Neurosci. 12, 2634–2638.
Wang, X., Zhong, P., Yan, Z., 2002b. Dopamine D4 receptors modulate
Widerlöv, E., Heiling, M., Ekman, R., Wahlestedt, C., 1989. Neuropeptide
GABAergic signaling in pyramidal neurons of prefrontal cortex. J.
Y—possible involvement in depression and anxiety. In: Mutt, V. (Ed.),
Neurosci. 22, 9185–9193.
Neuropeptide Y. Raven, New York, pp. 331–342.
Wang, F., Li, J., Wu, C., Yang, J., Xu, F., Zhao, Q., 2003. The
Wieland, S., Lan, N.C., Mirasedeghi, S., Gee, K.W., 1991. Anxiolytic
GABAA receptor mediates the hypnotic activity of melatonin in rats.
activity of the progesterone metabolite 5␣-pregnan-3␣-ol-20-one.
Pharmacol. Biochem. Behav. 74, 573–578.
Brain Res. 565, 263–268.
Waters, S.M., Krause, J.E., 2000. Distribution of galanin-1, -2, and
Wieland, S., Belluzzi, J., Stein, L., Lan, N.C., 1995. Comparative
-3 receptor messenger RNAs in central and peripheral rat tissues.
behavioral characterization of the neuroactive steroids 3␣-OH,5␣-
Neuroscience 95, 265–271.
pregnan20-one and 3␣-OH,5␤-pregnan-20-one in rodents. Psychophar-
Watson, C., Chen, G., Irving, P., Way, J., Chen, W.J., Kenakin, T., 2000.
macology 118, 65–71.
The use of stimulus-based assay systems to detect agonist-specific
Wieland, S., Belluzzi, J., Hawkinson, J.E., Hogenkamp, D., Upasani, R.,
receptor active states: implications for the trafficking of receptor
Stein, L., Wood, P.L., Gee, K.W., Lan, N.C., 1997. Anxiolytic and
stimulus by agonists. Mol. Pharmacol. 58, 1230–1238.
Waxman, S.G., Dib-Hajj, S., Cummins, T.R., Black, J.A., 1999. Sodium anticonvulsant activity of a synthetic neuroactive steroid Co 3-0593.
channels and pain. Brain Res. 96, 7635–7639. Psychopharmacology 134, 46–54.
Wedzony, K., Chocyk, A., Madkowiak, M., Fijal, K., Czyrak, A., 2000. Wieronska, J.M., Papp, M., Pilc, A., 2001. Effects of anxiolytic drugs on
Cortical localization of dopamine D4 receptors in the rat brain: some behavioral consequences in olfactory bulbectomized rats. Pol.
immunocytochemical study. J. Physiol. Pharmacol. 51, 205–221. J. Pharmacol. 53, 517–525.
Weeber, E.J., Atkins, C.M., Selcher, J.C., Varga, A.W., Mirnikjoo, B., Wiesinger, H., 2001. Arginine metabolism and the synthesis of nitric
Paylor, R., Leitges, M., Sweatt, J.D., 2000. A role of the ␤ isoform oxide in the nervous system. Prog. Neurobiol. 64, 365–391.
of protein kinase C in fear conditioning. J. Neurosci. 20, 5906–5914. Wightman, R.M., Robinson, D.L., 2002. Transient changes in mesolimbic
Wei, Q., Schafer, G.L., Hebda-Bauer, E., Shieh, K.R., Watson, S., dopamine and their association with “reward”. J. Neurochem. 82,
Seasholtz, A., Akil, H., 2001. Tissue-specific overexpression of the 721–735.
glucocorticoid receptor in the brain. Soc. Neurosci. Abstr. 27, 413.8. Wikberg, J.E.S., 1999. Melanocortin receptors: perspectives for novel
Weidenfeld, J., Newman, M.E., Itzik, A., Gur, E., Feldman, S., 2002. drugs. Eur. J. Pharmacol. 375, 295–310.
The amygdala regulates the pituitary–adrenocortical response and Wiley, J.L., Balster, R.L., 1992. Preclinical evaluation of N-methyl-d-
release of hypothalamic serotonin following electrical stimulation of aspartate antagonists for antianxiety effects: a review. In: Kamenka,
the dorsal raphe nucleus in the rat. Neuroendocrinology 76, 63–69. J.M., Domino, E.F. (Eds.), Multiple Sigma and PCP Receptor Ligand:
Weiss, S.M., Lightowler, S., Stanhope, K.J., Kennett, G.A., Dourish, C.T., Mechanisms for Neuromodulation and Neuroprotection? NPP Books,
2000. Measurement of anxiety in transgenic mice. Rev. Neurosci. 11, Ann Arbor, MI, pp. 801–815.
59–74. Wiley, J.L., Porter, J.H., Compton, A.D., Balster, R.L., 1992.
Welch, J.E., Saphier, D., 1994. Central and peripheral mechanisms in Antipunishment effects of acute and repeated administration of
the stimulation of adrenocortical secretion by 5-hydroxytryptamine2 phencyclidine and NPC 12626 in rats. Life Sci. 50, 1519–1528.
agonist (±)-1-(2,5-dimethoxy-4-iodophenyl)-2-aminopropane. J. Wiley, J.L., Compton, A.D., Porter, J.H., 1993. Effects of four
Pharmacol. Exp. Ther. 270, 918–928. antipsychotics on punished responding in rats. Pharmacol. Biochem.
Weninger, S.C., Dunn, A.J., Muglia, L.J., Dikkes, P., Miczek, K.A., Behav. 45, 263–267.
Swiergiel, A.H., Berridge, C.W., Majzoub, J.A., 1999. Stress-induced Wiley, J.L., Cristello, A.F., Balster, R.L., 1995. Effects of site-selective
behaviors require the corticotropin-releasing hormone (CRH) receptor, NMDA receptor antagonists in an elevated plus-maze model of anxiety
but not CRH. Proc. Natl. Acad. Sci. U.S.A. 96, 8283–8288. in mice. Eur. J. Pharmacol. 294, 101–107.
242 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Wiley, J.L., Compton, A.D., Holcomb, J.D., McCallum, S.E., Varvel, Wisden, W., Laurie, D.J., Monyer, H., Seeburg, P.H., 1992. The
S.A., Porter, J.H., Balster, R.L., 1998. Effects of modulation of distribution of 13 GABAA receptor subunit mRNAs in the rat brain.
NMDA neurotransmission on response rate and duration in a conflict I. Telencephalon, diencephalon, mesencephalon. J. Neurosci. 12,
procedure in rats. Neuropharmacology 37, 1527–1534. 1040–1062.
Wilke, R.A., Lupardus, P.J., Grandy, D.K., Rubinstein, M., Löw, Wise, P.M., Dubal, D.B., Wilson, M.E., Rau, S.W., Böttner, M., 2001a.
M.J., Jackson, M.B., 1999. K+ channel modulation in rodent Minireview: neuroprotective effects of estrogen—new insights into
neurohypophysial nerve terminals by sigma receptors and not by mechanisms of action. Endocrinology 142, 969–973.
dopamine receptors. J. Physiol. 517, 391–406. Wise, P.M., Dubal, D.B., Wilson, M.E., Rau, S.W., Liu, Y., 2001b.
Wilkinson, L.S., Humby, T., Killcross, A.S., Torres, E.M., Everitt, B.J., Estrogens: trophic and protective factors in the adult brain. Front.
Robbins, T.W., 1998. Dissociation in dopamine release in medial Neuroendocrinol. 22, 33–66.
prefrontal cortex and ventral striatum during the acquisition and Wissink, S., Meijer, O., Pearcet, D., Van Der Burg, B., Van Der
extinction of classical aversive conditioning in the rat. Eur. J. Neurosci. Saag, P., 2000. Regulation of the rat serotonin1A receptor gene by
10, 1019–1026. corticosteroids. J. Biol. Chem. 275, 1321–1326.
Williams, A.R., Carey, R.J., Miller, M., 1985. Altered emotionality of Witt, D.M., 1995. Oxytocin and rodent sociosexual responses: from
the vasopressin-deficient brattleboro rat. Peptides 6, 69–76. behavior to gene expression. Neurosci. Biobehav. Res. 19, 315–324.
Willetts, J., Balster, L., Leander, J.D., 1990. The behavioral pharmacology Wittchen, H.U., Kessler, R.C., Beesdo, K., Krause, P., Höfler, M.,
of NMDA receptor antagonists. Trends Pharmacol. Sci. 11, 423–428. Hoyer, J., 2002. Generalized anxiety and depression in primary care:
Willetts, J., Tokarz, M.E., Balster, R.L., 1991. Pentobarbital-like effects prevalence, recognition, and management. J. Clin. Psychiatry 8, 24–34.
of N-methyl-d-aspartate antagonists in mice. Life Sci. 48, 1795–1798. Witt-Enderby, P.A., Bennett, J., Jarzynka, M.J., Firestine, S., Melan,
Willetts, J., Clissold, D.B., Hartman, T.L., Brandsgaard, R.R., Hamilton, M.A., 2003. Melatonin receptors and their regulation: biochemical
G.S., Ferkany, J.W., 1993. Behavioral pharmacology of NPC 17742, a and structural mechanisms. Life Sci. 72, 2183–2198.
competitive N-methyl-d-aspartate (NMDA) antagonist. J. Pharmacol. Wohlfarth, K.M., Bianchi, M.T., MacDonald, R.L., 2002. Enhanced
Exp. Ther. 265, 1055–1062. neurosteroid potentiation of ternary GABAA receptors containing the
Williams, K., 1993. Ifenprodil discriminates subtypes of the N-methyl- ␦ subunit. J. Neurosci. 22, 1541–1549.
d-aspartate receptor: selectivity and mechanisms at recombinant Wolfer, D.P., Crusio, W.E., Lipp, H.-P., 2002. Knock-out mice: simple
heteromeric receptors. Mol. Pharmacol. 44, 851–859. solutions to the problems of genetic background and flanking genes.
Williams, K., 2001. Ifenprodil, a novel NMDA receptor antagonist: site Trends Neurosci. 25, 336–340.
and mechanism of action. Curr. Drug Targets 2, 285–298. Wonnacott, S., 1997. Presynaptic nicotinic ACh receptors. Trends
Williams, D.B., Akabas, M.H., 2002. Structural evidence that propofol
Neurosci. 20, 92–98.
stabilizes different GABAA receptor states at potentiating and
Woo, R.S., Park, E.Y., Shi, M.S., Jeong, M.S., Zhao, R.J., Shi, B.S., Kim,
activating concentrations. J. Neurosci. 22, 7417–7424.
C.J., Park, J.W., Kim, K.W., 2002. Mechanisms of nicotine-evoked
Williams, M., Jarvis, M.F., 2000. Purinergic and pyrimidinergic receptors
release of [3 H]noradrenaline in human cerebral cortex slices. Br. J.
as potential drug targets. Biochem. Pharmacol. 59, 1173–1185.
Pharmacol. 137, 1063–1070.
Willick, M.L., Kokkinidis, L., 1995. Cocaine enhances the expression of
Wood, S.J., Toth, M., 2001. Molecular pathways of anxiety revealed by
fear-potentiated startle: evaluation of state-dependent extinction and the
knock-out mice. Mol. Neurobiol. 23, 101–119.
shock-sensitization of acoustic startle. Behav. Neurosci. 109, 929–938.
Wood, M.D., Reavill, C., Trail, B., Wilson, A., Stean, T., Kennett,
Wilson, M.A., Mamounas, L.A., Fasman, K.H., Axt, K.J., Molliver, M.E.,
G.A., Lightowler, S., Blackburn, T.P., Thomas, D., Gager, T.L., Riley,
1993. Reactions of 5-HT neurons to drugs of abuse: neurotoxicity
G., Holland, V., Bromidge, S.M., Forbes, I.T., Middlemiss, D.N.,
and plasticity. NIDA Res. Monogr. 136, 155–178.
Wilson, R.I., Kunos, G., Nicoll, R.A., 2001. Presynaptic specificity of 2001. SB-243213, a selective 5-HT2C receptor inverse agonist with
endocannabinoid signaling in the hippocampus. Neuron 31, 453–462. improved anxiolytic profile: lack of tolerance and withdrawal anxiety.
Windle, R.J., Shanks, N., Lightman, S.L., Ingram, C.D., 1997. Central Neuropharmacology 41, 186–199.
oxytocin administration reduces stress-induced corticosterone release Woodruff, G.N., 1992. Cholecystokinin receptors in relation to brain
and anxiety behavior in rats. Endocrinology 138, 2829–2834. dopaminergic pathways. Neurochem. Int. 20, 339S–343S.
Winkler, J., Ramirrez, G.A., Thal, L.J., Waite, J.J., 2000. Nerve Woodruff, G.N., Hill, D.R., Boden, P., Pinnock, R., Singh, L., Hughes,
growth factor (NGF) augments cortical and hippocampal cholinergic J., 1991. Functional role of brain CCK receptors. Neuropeptides 19,
functioning after p75NGF receptor-mediated deafferentation but 45–56.
impairs inhibitory avoidance and induces fear-related behaviors. J. Woodward, J.J., 2000. Ethanol and NMDA receptor signaling. Crit. Rev.
Neurosci. 20, 834–844. Neurobiol. 14, 69–89.
Winslow, J.T., Insel, T.R., 2002. The social deficits of the oxytocin Wotjak, C.T., Ganster, J., Kohl, G., Holsboer, F., Landgraf, R., Engelmann,
knock-out mouse. Neuropeptides 36, 221–229. M., 1998. Dissociated central and peripheral release of vasopressin,
Winters, W.D., Yuwiler, A., Oxenkrug, G.F., 1991. The effects but not oxytocin, in response to repeated swim stress: new insights
of benzodiazepines on basal and isoproterenol-stimulated N- into the secretory capacities of peptidergic neurons. Neuroscience 85,
acetyltransferase activity by the rat pineal gland, in vivo and in vitro. 1209–1222.
J. Pineal Res. 10, 151–158. Wright, J.K., Harding, J.W., 1995. Brain angiotensin receptor subtypes,
Wirkner, K., Poelchen, W., Köles, L., Mühlberg, K., Scheibler, P., AT1 , AT2 and AT4 and their functions. Regul. Pept. 59, 269–295.
Allgaier, C., Illes, P., 1999. Ethanol-induced inhibition of NMDA Wright, J.K., Upton, N., Marsden, C.A., 1992. Effects of established
receptor channels. Neurochem. Int. 35, 153–162. and putative anxiolytics on the extracellular 5-HT and 5-HT1A levels
Wirtshafter, D., Sheppard, A.C., 2001. Localization of GABAB receptors in the ventral hippocampus of rats during behaviour on the elevated
in midbrain monoamine containing neurons in the rat. Brain Res. X-maze. Psychopharmacology 109, 338–346.
Bull. 56, 1–5. Wright, D.E., Seroogy, K.B., Lundgren, K.H., 1995. Comparative
Wirtshafter, D., Stratford, T.P., Pitzer, M.R., 1993. Studies on the localization of serotonin1A , IC, and 2 receptor subtype mRNAs in rat
behavioral activation produced by stimulation of GABAB receptors in brain. J. Comp. Neurol. 351, 357–373.
the median raphe nuleus. Behav. Brain Res. 59, 83–93. Wu, L.G., Saggau, P., 1997. Presynaptic inhibition of elicited
Wirtshafter, D., Stratford, T.P., Shim, I., 1998. Placement in a novel neurotransmitter release. Trends Neurosci. 20, 204–212.
environment induces Fos-like immunoreactivity in supramammillary Wu, M., Shanabrough, M., Leranth, C., Alreja, M., 2000. Cholinergic
cells projecting to the hippocampus and midbrain. Brain Res. 789, excitation of septohippocampal GABA but not cholinergic neurons:
331–334. implications for learning and memory. J. Neurosci. 20, 3900–3908.
M.J. Millan / Progress in Neurobiology 70 (2003) 83–244 243

Wu, W.R., Li, N., Sorg, B.A., 2002. Regulation of medial prefrontal Yamashita, S., Oishi, R., Gomita, Y., 1995. Anticonflict effects of acute
cortex dopamine by ␣-amino-3-hydroxy-5-methylisoxazole-4- and chronic treatments with buspirone and gepirone in rats. Pharmacol.
propionate/kainate receptors. Neuroscience 114, 507–516. Biochem. Behav. 50, 477–479.
Wunderlich, G.R., Raymond, R., DeSousa, N.J., Nobrega, J.N., Yamawaki, S., Kagaya, A., Tawaa, Y., Inagaki, M., 1998. Intracellular
Vaccarino, F.J., 2002. Decreased CCKB receptor binding in rat calcium signaling systems in the pathophysiology of affective
amygdala in animals demonstrating greater anxiety-like behavior. disorders. Life Sci. 62, 1665–1670.
Psychopharmacology 164, 193–199. Yan, Z., 2002. Regulation of GABAergic inhibition by serotonin signaling
Xi, Z.-X., Baker, D.A., Shen, H., Carson, D.S., Kalivas, P.W., 2002. Group in prefrontal cortex. Mol. Neurobiol. 26, 203–216.
II metabotropic glutamate receptors modulate extracellular glutamate Yan, Q., Radeke, C.R., Matheson, J., Talvenheimo, A.A., Welcher, S.C.,
in the nucleus accumbens. J. Pharmacol. Exp. Ther. 300, 162–171. 1997. Immunocytochemical localization of TrkB in the central nervous
Xie, Z.C., Buckner, E., Commissaris, R.L., 1995a. Anticonflict effect of system of the adult rat. J. Comp. Neurol. 378, 135–257.
MK-801 in rats: time course and chronic treatment studies. Pharmacol. Yan, Q.S., Reith, M.E., Yan, S.G., Jobe, P.C., 1998. Effect of systemic
Biochem. Behav. 51, 635–640. ethanol on basal and stimulated glutamate releases in the nucleus
Xie, X., Lancaster, B., Peakman, T., Garthwaite, J., 1995b. Interaction accumbens of freely moving Sprague–Dawley rats: a microdialysis
of the antiepileptic drug lamotrigine with recombinant rat brain type study. Neurosci. Lett. 258, 32–39.
IIA Na+ channels and with native Na+ channels in rat hippocampal Yanai, K., Son, L.Z., Endou, M., Sakurai, E., Watanabe, T., 1998.
neurones. Eur. J. Physiol. 430, 437–446. Targetted disruption of histamine H1 receptors in mice: behavioral
Xu, Z., Commissaris, R.L., 1992. Anxiolytic-like effects of the and neurochemical characterization. Life Sci. 62, 1607–1610.
noncompetitive NMDA antagonist MK 801. Pharmacol. Biochem. Yang, X.-M., Gorman, A.L., Dunn, A.J., 1990. The involvement of central
Behav. 43, 471–477. noradrenergic systems and corticotropin-releasing factor in defensive-
Xu, T., Pandey, S.C., 2000. Cellular localization of serotonin2A (5-HT2A ) withdrawal behavior in rats. J. Pharmacol. Exp. Ther. 255, 1064–1070.
receptors in the rat brain. Brain Res. Bull. 51, 499–505. Yang, X., Criswell, H.E., Breese, G.R., 1996. Nicotine-induced inhibition
Xu, M., Koeltzow, T.E., Santiago, G.T., Moratalla, R., Cooper, D.C., Hu, in medial septum involves activation of presynaptic nicotinic
X.-T., White, N.M., Graybiel, A.M., White, F.J., Tonegawa, S., 1997. cholinergic receptors on ␥-aminobutyric acid-containing neurons. J.
Dopamine D3 receptor mutant mice exhibit increased behavioural Pharmacol. Exp. Ther. 276, 482–489.
sensitivity to concurrent stimulation of D1 and D2 receptors. Neuron Yang, P., Jones, B.L., Henderson, L.P., 2002. Mechanisms of anabolic
19, 837–848. androgenic steroid modulation of ␣1 ␤3 ␥2L GABAA receptors.
Yadin, E., Thomas, E., Strickland, C.E., Grishkat, H.L., 1991. Neuropharmacology 43, 619–633.
Anxiolytic effects of benzodiazepines in amygdala-lesioned rats. Yang, H.S., Kim, S.Y., Choi, S.J., Kim, K.-J., Kim, O.N., Lee, S.B.,
Psychopharmacology 103, 473–479. Sung, K.-W., 2003. Effect of 5-hydroxyindole on ethanol potentiation
Yadin, E., Thomas, E., Grishkat, H., Strikland, C., 1993. The role of the of 5-hydroxytryptamine (5-HT)3 receptor-activated ion current in
lateral septum in anxiolysis. Physiol. Behav. 53, 1077–1083. NCB-20 neuroblastoma cells. Neurosci. Lett. 338, 72–76.
Yamada, K., Ohki-Hamazaki, H., Wada, K., 2000a. Differential effects Yasuda, K., Suemaru, K., Araki, H., Gomita, Y., 2002. Effect of nicotine
of social isolation upon body weight, food consumption, and cessation on the central serotonergic systems in mice: involvement
responsiveness to novel and social environment in bombesin receptor of 5-HT2 receptors. Naunyn Schmiedebergs Arch. Pharmacol. 366,
subtype-3 (BRS-3) deficient mice. Physiol. Behav. 68, 555–561. 276–281.
Yamada, K., Wada, E., Wada, K., 2000b. Male mice lacking the gastrin- Yau, J.L.W., Hibberd, C., Noble, J., Seckl, J.R., 2002. The effect of
releasing peptide receptor (GRP-R) display elevated preference for chronic fluoxetine treatment on brain corticosteroid receptor mRNA
conspecific odors and increased social investigatory behaviors. Brain expression and spatial memory in young and aged rats. Mol. Brain
Res. 870, 20–26. Res. 106, 113–117.
Yamada, K., Wada, E., Wada, K., 2001. Female gastrin-releasing peptide Yip, J., Chahl, L.A., 1997. Localization of Fos-like immuno-reactivity
receptor (GRP-R)-deficient mice exhibit altered social preference induced by the NK3 tachykinin receptor agonist, senktide, in the
for male conspecifics: implications for GRP/GRP-R modulation of ginea-pig brain. Br. J. Pharmacol. 122, 715–725.
GABAergic function. Brain Res. 894, 281–287. Yip, J., Chahl, L.A., 2001. Localization of NK1 and NK2 receptors in
Yamada, K., Santo-Yamada, Y., Wada, E., Wada, K., 2002a. Restraint guinea-pig brain. Regul. Pept. 98, 55–62.
stress impaired maternal behavior in female mice lacking the Yoshioka, M., Matsumoto, M., Tagashi, H., Mori, K., Saito, H., 1998.
neuromedinB receptor (NMB-R) gene. Neurosci. Lett. 330, 163–166. Central distribution and function of 5-HT6 receptor subtype in the rat
Yamada, K., Santo-Yamada, Y., Wada, E., Wada, K., 2002b. Role of brain. Life Sci. 62, 1473–1477.
bombesin (BN)-like peptides/receptors in emotional behavior by Yoshitake, T., Reenilä, I., Ögren, S.O., Hökfelt, T., Kehr, J., 2003a.
comparison of three strains of BN-like peptide receptor knock-out Galanin attenuates basal and antidepressant drug-induced increase of
mice. Mol. Psychiatry 7, 113–117. extracellular serotonin and noradrenaline levels in the rat hippocampus.
Yamada, K., Wada, E., Yamano, M., Sun, Y.-J., Ohara-Imaizumi, M., Neurosci. Lett. 339, 239–242.
Nagamastu, S., Wada, K., 2002c. Decreased marble burying behavior Yoshitake, T., Yoshitake, S., Yamaguchi, M., Ögren, S.O., Kehr, J., 2003b.
in female mice lacking neuromedinB receptor (NMB-R) implies the Activation of 5-HT1A autoreceptors enhances the inhibitory effect of
involvement of NMB/NMB-R in 5-HT neuron function. Brain Res. galanin on hippocampal 5-HT release in vivo. Neuropharmacology
942, 71–78. 44, 206–213.
Yamano, M., Ogura, H., Okuyama, S., Ohki-Hamazaki, H., 2002. Yuzurihara, M., Ikarashi, Y., Ishige, A., Sasaki, H., Kuribara, H.,
Modulation of 5-HT system in mice with a targeted disruption of Maruyama, Y., 2000. Effects of drugs acting as histamine releasers
neuromedinB receptor. J. Neurosci. Res. 68, 59–64. or histamine receptor blockers on an experimental anxiety model in
Yamashita, K., Kataoka, Y., Miyazaki, A., Shibata, K., Ozaki, T., mice. Pharmacol. Biochem. Behav. 67, 145–150.
Miyazaki, A., Kagoshima, M., 1989a. Neuroanatomical substrates Zacharko, R.M., Koszycki, D., Mendella, P.D., Bradwejn, J., 1995.
regulating rat conflict behavior evidenced by brain lesioning. Neurosci. Behavioral, neurochemical, anatomical and electrophysiological
Lett. 104, 195–200. correlates of panic disorder: multiple transmitter interaction and
Yamashita, K., Kataoka, Y., Miyazaki, A., Shibata, K., Tominaga, K., neuropeptide colocalization. Prog. Neurobiol. 47, 371–423.
Ueki, S., 1989b. A key role of the mammillary body in mediation of Zajecka, J.M., 1996. The effect of nefazodone on co-morbid anxiety
the anti-anxiety action of zopiclone, a cyclopyrrolone derivative. Jpn. symptoms associated with depression: experience in family practice
J. Pharmacol. 51, 438–442. and psychiatric outpatient setting. J. Clin. Psychiatry 57, 10–14.
244 M.J. Millan / Progress in Neurobiology 70 (2003) 83–244

Zangrossi, H., Viana, M.B., Graeff, F.G., 1999. Anxiolytic effect Zhou, F.M., Hablitz, J.J., 1999. Activation of serotonin receptors
of intra-amygdala injection of midazolam and 8-hydroxy-2-(di-n- modulates synaptic transmission in rat cerebral cortex. J. Neurophysiol.
propylamino)tetralin in the elevated T-maze. Eur. J. Pharmacol. 369, 82, 2989–2999.
267–270. Zhu, W.J., Wang, J.F., Krueger, K.E., Vicini, S., 1996. Delta subunit
Zangrossi, H., Viana, M.B., Zanoveli, J., Bueno, C., Nogueira, R.L., inhibits neurosteroid modulation of GABAA receptors. J. Neurosci.
Graeff, F.G., 2001. Serotonergic regulation of inhibitory avoidance 16, 6648–6656.
and one-way escape in the rat elevated T-maze. Neurosci. Biobehav. Zhuang, X., Gross, C., Santarelli, L., Compan, V., Trillat, A.C., Hen, R.,
Rev. 25, 637–645. 1999. Altered emotional states in knock-out mice lacking 5-HT1A or
Zapata, A., Witkin, J.M., Shippenberg, T.S., 2001. Selective D3 receptor 5-HT1B receptors. Neuropsychopharmacology 21, 52S–60S.
agonist effects of (+)-PD128,907 on dialysate dopamine at low doses. Ziegenbein, M., Steiger, A., Murck, H., 2000. Treatment with the
Neuropharmacology 41, 351–359. presynatptic 5-HT1A -antagonist pindolol in patients with panic
Zarrindast, M-R., Rostami, P., Sadeghi-Hariri, M., 2001. GABAA but disorder. Biol. Psychiatry 47, 81S.
not GABAB receptor stimulation induces antianxiety profile in rats. Zimanyi, I.A., Poindexter, G.S., 2000. NPY-ergic agents for the treatment
Pharmacol. Biochem. Behav. 69, 9–15. of obesity. Drug Dev. Res. 51, 94–111.
Zernig, G., Troger, J., Saria, A., 1993. Different behavioral profiles of Zimanyi, I.A., Fathi, Z., Poindexter, G.S., 1998. Central control of
the non-peptide substance P (NK1 ) antagonists CP-96,345 and RP feeding behavior by neuropeptide Y. Curr. Pharm. Des. 4, 349–366.
67580 in swiss albino mice in the black-and-white box. Neurosci. Zimmerman, M., McDermut, W., Mattia, J., 2000. Frequency of anxiety
Lett. 151, 64–66. disorders in psychiatric outpatients with major depressive disorder.
Zetterström, T.S., Pei, Q., Madhav, T.R., Coppell, A.L., Lewis, L., Am. J. Psychiatry 157, 1337–1340.
Grahame-Smith, D.G., 1999. Manipulations of brain 5-HT levels Zinder, O., Dar, D.E., 1999. Neuroactive steroids: their mechanism of
affect gene expression for BDNF in rat brain. Neuropharmacology action and their function in the stress response. Acta. Physiol. Scand.
38, 1063–1073. 167, 181–188.
Zhang, J., Snyder, S.H., 1995. Nitric oxide in the nervous system. Annu. Zohar, J., Insel, T.R., Zohar-Kadouch, R.C., Hill, J.L., Murphy, D.L., 1988.
Rev. Pharmacol. Toxicol. 35, 213–233. Serotonergic responsivity in obsessive–compulsive disorder: effects of
Zhang, J., Chiodo, L.A., Wettstein, J.G., Junien, J.-L., Freeman, A.S., chronic clomipramine treatment. Arch. Gen. Psychiatry 45, 167–172.
1993. Repeated administration of sigma ligands alters the population Zobel, A.W., Nickel, T., Kunzel, H.E., Ackl, N., Sonntag, A., Ising, M.,
activity of rat midbrain dopaminergic neurons. Synapse 13, 223–230. Holsboer, F., 2000. Effects of the high-affinity corticotropin-releasing
Zhang, D., Pan, Z.H., Awobuluyi, M., Lipton, S.A., 2001a. Structure hormone receptor 1 antagonist R121919 in major depression: the first
and function of GABAC receptors: a comparison of native versus 20 patients treated. J. Psychiatry Res. 34, 171–181.
recombinant receptors. Trends Pharmacol. Sci. 22, 121–132. Zorrilla, E.P., Redei, E., De Rueis, R.J., 1994. Reduced cytokine levels
Zhang, H-T., Frith, S.A., Wilkins, J., O’Donnell, J.M., 2001b. Comparison and T-cell function in healthy males: relation in individual differences
of the effects of isoproterenol administered into the hippocampus, in subclinical anxiety. Brain Behav. Immun. 8, 293–312.
frontal cortex or amygdala on behavior of rats maintained by differential Zorilla, E.P., Valdez, G.R., Nozulak, J., Koob, G.F., Markou, A., 2002.
reinforcement of low response rate. Psychopharmacology 159, 89–97. Effects of antalarmin, a CRF type 1 receptor antagonist, on anxiety-like
Zhang, W., Basile, A.S., Gomeza, J., Volpicelli, J., Volpicelli, L.A., behavior and motor activation in the rat. Brain Res. 952, 188–199.
Levey, A.I., Wess, J., 2002. Characterization of central inhibitory Zwanzger, P., Bahai, T., Boerner, R., Moller, H., Rupprecht, R.,
muscarinic autoreceptors by the use of muscarinic acetylcholine 2001a. Anxiolytic effects of vigabatrin in panic disorder. J. Clin.
receptor knock-out mice. J. Neurosci. 22, 1709–1717. Psychopharmacol. 21, 539–540.
Zhao, T.J., Chiu, T.H., Rosenberg, H.C., 1994. Reduced expression of ␥- Zwanzger, P., Bahai, T., Schuele, C., Strohle, A., Padberg, F., Kathmann,
aminobutyric acid type A/benzodiazepine receptor ␥2 and ␣5 subunit N., 2001b. Vigabatrin decreases cholecystokinin-tetrapeptide (CCK4 )
mRNAs in brain regions of flurazepam-treated rats. Mol. Pharmacol. induced panic in healthy volunteers. Neuropsychopharmacology 25,
45, 657–663. 699–703.

You might also like