<<

AND BRAIN DEVELOPMENT

Monsheel S. K. Sodhi and Elaine Sanders-Bush Departments of Pharmacology and Psychiatry Vanderbilt University Nashville, Tennessee 37232

I. Introduction II. The Discovery of Serotonin and Classification of Serotonin Receptors A. Distribution and Projections of the System III. The Role of Serotonin in Developmental Plasticity A. Serotonergic Projections during Brain Development B. Growth Factors Influencing the Development of Serotonergic Neurons C. The Role of Serotonin as a Growth Factor D. Serotonin Receptors and Developmental Plasticity IV. Manipulation of the Serotonergic System Alters Synaptic Plasticity A. and Serotonin Depletion Studies B. Experimental Models of Synaptic Plasticity V. Does Dysfunction of Serotonergic Signaling Result in Impaired Brain Development? A. The Role of Serotonin in Learning and Memory B. Autism and Serotonin C. The Role of Serotonin in Stress and D. Serotonergic Influences on Synaptic Plasticity in AVective Disorders E. Altered Synaptic Plasticity in Schizophrenia F. Down’s Syndrome, Mental Retardation, and Serotonin VI. Conclusions References

The role of the serotonergic system in the neuroplastic events that create, repair, and degenerate the brain has been explored. Synaptic plasticity occurs throughout life and is critical during brain development. Evidence from bio- chemical, pharmacological, and clinical studies demonstrates the huge import- ance of an intact serotonergic system for normal central nervous system (CNS) function. Serotonin acts as a growth factor during embryogenesis, and serotonin activity forms a crucial part of the cascade of events leading to changes in brain structure. The serotonergic system interacts with brain-derived neuro- trophic factor (BDNF), S100 , and other chemical messengers, in addition to its cross talk with the GABAergic, glutamatergic, and dopaminergic neurotrans- mitter systems. Disruption of these processes may contribute to CNS disorders that have been associated with impaired development. Furthermore, many psy- chiatric alter serotonergic activity and have been shown to create changes in brain structure with long-term treatment. However, the mechanisms for their

INTERNATIONAL REVIEW OF 111 Copyright 2004, Elsevier Inc. NEUROBIOLOGY, VOL. 59 All rights reserved. 0074-7742/04 $35.00 112 SODHI AND SANDERS-BUSH therapeutic eYcacy are still unclear. Treatments for psychiatric illness are usually chronic and alleviate psychiatric symptoms, rather than cure these diseases. Therefore, greater exploration of the serotonin system during brain development and growth could lead to real progress in the discovery of treatments for mental disorders.

I. Introduction

Serotonin (5-hydroxytryptamine, 5-HT), the ‘‘happy hormone,’’ has a phylo- genetically ancient role in neural transmission (Turlejski, 1996). Because the serotonergic system has a widespread distribution in the CNS, it influences almost every sphere of mammalian physiology, from cardiovascular regulation (Miyata et al., 2000; Nebigil et al., 2000; Thorin et al., 1990), respiration, the gastrointestinal system (Kato et al., 1999), pain sensitivity, and thermoregulation to more centrally controlled functions. The latter include the maintenance of circadian rhythm, appetite, aggression, sensorimotor activity, sexual behavior, mood, cognition, learning, and memory. Hence drugs with serotonergic activity are used to treat the aVective disorders schizophrenia (AbiDargham et al., 1996; Breier, 1995; Kapur and Remington, 1996; Meltzer, 2002; Ohuoha et al., 1993; Sodhi and Murray, 1997), anxiety (Gross et al., 2002), stress, eating disorders (Bray, 2000; GuyGrand, 1995; Halford, 2001; Heal et al., 1998; Heisler et al., 1998b; Hesselink and Sambunaris, 1995; Jallon and Picard, 2001; Koponen et al., 2002; Luque and Rey, 1999; McNeely and Goa, 1998; Prasad, 1998; Weissman, 2001), and deliberate self-harm (Holden, 1995). In addition, personality dysfunctions such as addictive behaviors, aggression, psychopathic and sociopathic behavior, attention-deficit hyperactivity, and autism are also as- sociated with altered serotonergic transmission. Indeed, new serotonin receptor ligands are being explored as possible treatments for Alzheimer’s disease, as they appear to improve memory (Sumiyoshi et al., 2001), obesity (Bray, 2000; Rothman and Baumann, 2002; Stunkard and Allison, 2003; Wechsler, 1998), and epilepsy (Chadwick et al., 1977; Chugani and Chugani, 2003; Dailey et al., 1992; Deahl and Trimble, 1991; Fromm et al., 1977; Heisler et al., 1998b; Lunardi et al., 1995; Monaco et al., 1995; Savic et al., 2001; Statnick et al., 1996; Yan et al., 1994). Although increasing knowledge of serotonergic function is propelling many advances in the therapeutics of psychiatric and behavioral disorders, drugs in clinical use often treat the disease symptoms instead of relieving or preventing the causes. Moreover, treatment regimes are often lengthy or lifelong, sometimes with severe side eVects. As yet the causes of psychiatric disease are unknown, therefore the role of serotonin in the etiology or progression of these disorders SEROTONIN AND BRAIN DEVELOPMENT 113 requires exploration in order to facilitate improvements in medication and prog- noses. There is increasing support for the hypothesis that impaired development and synaptic plasticity contribute to the etiologies of many central nervous system (CNS) diseases. Plasticity is defined as functionally relevant structural adaptations performed by the CNS following genetic or environmental challenges. Neuronal plasticity is essential for the survival of an individual in a constantly changing environment. It is a dynamic process based on the ability of neuronal systems, brain nuclei synapses, single nerve cells, and receptors to adapt to challenges. Plasticity reveals itself in a number of ways, which range from altered gene expression or changes in neurotransmitter release to changes in behavior or phenotype. Synaptic plasti- city is constant throughout life and is especially important during development. Connections between neurons of the CNS are capable of being dismantled and reconstructed in response to changes in the physiological environment, therefore stress, malnutrition, sleep, hormones, and drugs can all produce changes in brain structure. Accumulating research suggests that serotonin plays an important role in synaptic plasticity and brain development. In this review we attempt to explore this evidence and its implications for impaired brain development and psychiatric illness.

II. The Discovery of Serotonin and Classification of Serotonin Receptors

The chemical 5-hydroxytryptamine (5-HT) was first isolated in serum and, because of its powerful vasoconstrictive eVects, was dubbed ‘‘serotonin’’ (Rapport, 1948). Serotonin was later detected in the brain (Twarog and Page, 1953). In 1957, Gaddum and Picarelli reported the existence of multiple sero- tonin receptor subtypes, which they called 5-HT-M and 5-HT-D, after their an- tagonists, morphine and dibenzyline, respectively. Peroutka and Snyder (1979) reclassified these receptors based on radioligand-binding studies in brain hom- 3 ogenates. The 5-HT1 receptor was labeled by [ H]5-HT, whereas the 5-HT2 receptor (corresponding to 5-HT-D) was sensitive to the receptor 3 [ H]. By 1986, the M receptors were renamed 5-HT3 receptors (Bradley,1986),werefoundtobetheonlyionotropicsubtypeofthe5-HTreceptor, and were detected to be at low density in limbic and striatal areas (Abi-Dargham et al., 1993). [3H]Lysergic acid diethylamide (LSD), a psychotomimetic com- pound with a structure similar to serotonin, was found to have high aYnity for serotonin receptors. Subsequently, heterogeneity was revealed in the 5-HT1 re- ceptor class; 5-HT1A receptors could be distinguished from 5-HT1B receptors (equivalent to the human 5-HT1D receptor) by the high aYnity of the former for spiperone (Pedigo et al., 1981). The use of receptor autoradiographic 114 SODHI AND SANDERS-BUSH techniques demonstrated the existence of a third 5-HT1 receptor in the porcine choroid plexus, the 5-HT1C subtype (later renamed 5-HT2C), through its high aYnity for [3H] and [3H]5-HT (Pazos et al., 1984). The application of molecular cloning techniques in the late 1980s revolutionized protein discov- ery and to date, 14 distinct subtypes of mammalian serotonin receptors have been cloned. Both molecular structure and pharmacological properties determine their classification, and under a revised system, the serotonin receptors are now allocated to seven distinct families (Hoyer et al., 1994). The characteristics and distributions of the diVerent subtypes of serotonin receptors are summarized in Table I. A detailed review of their pharmacology has been compiled by Barnes and Sharp (1999).

A. Distribution and Projections of the Serotonergic System

Serotonergic neurons are part of one of the most widely distributed neuronal systems in the mammalian brain; this neuronal network is also one of the earliest to develop in the embryo. Serotonin-containing neurons projecting to the forebrain originate in four brain stem nuclei, the principal of these being median and dorsal raphe nuclei. The dorsal raphe nucleus projects thin serotonin fibers, which are more abundant in the cortex, whereas the median raphe nucleus provides thick serotonin fibers with large varicosities that are relatively sparse and more abun- dant in the hippocampus (Kosofsky and Molliver, 1987). The hippocampus re- ceives fibers following a dorsomedial course through the cingulate cortex (reviewed by Rubenstein, 1998). During development of the cortex, the cortical regions have diVerent profiles for the maturation of serotonin axon terminals (D’Amato et al., 1987; DeFelipe et al., 1991). The thick and thin serotonin fiber systems innervate the cortex in diVerent time frames (Vu and Tork, 1992). Because the fibers also change after injury and according to the target cell innervated, clas- sification according to the thickness of the fibers is not consistent (Azmitia, 1999). Serotonin neurons innervate almost all areas of the brain ( Jacobs and Azmitia, 1992), and their projections and targets are summarized in Fig. 1.

III. The Role of Serotonin in Developmental Plasticity

A central tenet of neuroscience is that synaptic plasticity underlies behavioral plasticity and that information is coded by alterations in synaptic strength and connectivity in networks of neurons (Kandel and Spencer, 1968; Martin et al., 2000). Prior to its vital role as a neurotransmitter in adult brain, serotonin acts as a regulator of brain development. The latter is inextricably linked to the TABLE I The Serotonin Receptorsa

Structure Second Principal Effect of Clinical effects of Receptor (genetic locus) messenger distribution receptor activation receptor activation

The 5-HT1 family 5-HT1A 7 TMD, G-protein-coupled Hippocampus, lateral septum, # 5-HT release and , intronless # adenylate cingulate and entorrhinal " ACh release in e.g., , , (5q11.2-q13) cyclase cortices, dorsal and median septohippocampal neurons 5-HT syndrome, raphe nuclei " NE release in hypothalamus, hypothermia, hyperphagia, and hippocampus, frontal altered sexual behavior cortex VTA # Glutamate release " Prolactin release,

115 " growth hormone, " ACTH release 5-HT1B 7 TMD, G-protein-coupled Basal ganglia, especially # 5-HT release Hypophagia, hypothermia, (previously intronless # adenylate substantia nigra, globus # ACh release myoclonic jerks, 5-HT1D ) (6q13) cyclase pallidus, ventral pallidum, antiaggressive properties, and entopeduncular nucleus penile erection 5-HT1D 7 TMD, G-protein-coupled # Basal ganglia, especially " GABA release? (previously intronless adenylate cyclase substantia nigra, globus # 5-HT release 5-HT1D´ ) (1p34.3-36.3) pallidus, caudate putamen, # ACh release? periaquaductal grey, hippocampus, cortex, olfactory cortex, dorsal raphe nucleus, locus coeruleus, and spinal cord 5-HT1E 7 TMD, G-protein-coupled Cortex including entorrhinal intronless # adenylate cortex, basal ganglia, (6q14-15) cyclase claustrum, hippocampus (subiculum), amygdale, hypothalamus

(Continued ) TABLE I (Continued ) Structure Second Principal Effect of Clinical effects of Receptor (genetic locus) messenger distribution receptor activation receptor activation

5-HT1F 7 TMD, G-protein-coupled Cortex, especially cingulate and Relief of migraine? intronless # adenylate entorrhinal cortices, hippocampus (3q1) cyclase (CA1-CA3 layers), basal ganglia, claustrum, caudate nucleus, dorsal raphe nucleus The 5-HT2 family 5-HT2A 7 TMD, G-protein-coupled Olfactory bulb, hippocampus, Excitation of 5-HT neurons Antipsychotic (antagonist), 2 introns within " phospholipase C cortex (neocortex, claustrum, Inhibition of NE neurons anxiolytic (antagonist), relief

116 ORF (13q14-21) pyriform and entorhinal cortices), in locus coeruleus of sleep and eating disorders, caudate nucleus, and nucleus # NE release in hippocampus relief of migraine, accumbens " BDNF expression? " cortisol, hallucinogenic, hyperthermia " ACTH, " renin, " prolactin 5-HT2B 7 TMD, G-protein-coupled Cerebellum, lateral septum, Mediation of mitogenic effects Anxiolysis (antagonist)? 2 introns within " phospholipase C dorsal hippocampus and of 5-HT during neural ORF (2q36.3-37.1) medial amygdala, stomach development? fundus, heart 5-HT2C 7 TMD, 3 introns G-protein-coupled Choroid plexusolfactory nucleus, Excitation of 5-HT neurons Antipsychotic (antagonist), (previously within ORF " phospholipase C pyriform, cingulate and # NE and DA release in anxiolytic (antagonist), 5-HT1C ) (Xq24) retrospinal cortices, nucleus mesocortical/mesolimbic NE and antidepressant, relief of accumbens, hippocampus, DA projections " cortisol, migraine and sleep amygdala, subiculum, " ACTH, " prolactin disorders, hypolocomotion, entorhinal cortex, caudate hypophagia, penile nucleus, substantia nigra pars erection, anticonvulsant compacta, striatum, thalamus, hypothalamus, frontal cortex 5-HT3 Ligand-gated cation " intracellular Hippocampus, amygdale, Fast synaptic Vomiting, relief of migraine channel, several Na+:K+ ratio superficial layers of cerebral transmission in brain (antagonist), anxiolysis introns cortex, limbic and " 5-HT release (antagonist), " cognition (11q23.1-23.2) basal ganglia structures, dorsal # ACh release (antagonist), # locomotion vagal complex in brain stem, " GABA release (antagonist), # reward gastrointestinal tract " CCK release (antagonist), # LTP, " DA release analgesia (antagonist)? Na+:K+ ratio 5-HT4 7 TMD, >5 G-protein-coupled Hippocampus, basal ganglia and " 5-HT release " cognition, anxiolysis and introns " adenylate substantia nigra, atrium, and anxiogenesis (antagonist) (5q31-33) cyclase gastrointestinal tract " ACh release " DA release 5-HT5A 7 TMD, G-protein-coupled Cortex, hippocampus, thalamus, " locomotor activity 1 intron # adenylate cyclase olfactory bulb, hypothalamus, (7q36.1) cerebellum, spinal cord 5-HT5B 7 TMD, G-protein-coupled Hippocampus, medial and lateral 117 1 intron # adenylate cyclase habenula, dorsal raphe nucleus, (2q11-13) olfactory bulb, entorrhinal and pyriform cortices 5-HT6 7 TMD, G-protein-coupled Caudate nucleus, olfactory tubercles, Antipsychotic (antagonist), 2 introns " adenylate cyclase nucleus accumbens, hippocampus, antidepressant (antagonist), (1p36-35) cerebral cortex, striatum, stomach, anxiolysis (antagonist) adrenal glands 5-HT7 7 TMD, G-protein-coupled Thalamus, hypothalamus, Antipsychotic (antagonist), 2 introns " adenylate cyclase hippocampus, cerebral cortex, antidepressant (antagonist), (10q21-24) amygdala anticonvulsant (antagonist)

aReviewed by Barnes and Sharpe (1999) and Crossland (2000). TMD, transmembrane domain; ORF, open reading frame; VTA, ventral tegmental area; NE, norepinepherine/noradrenaline; ACh, acetylcholine; GABA, -aminobutyric acid; CCK, cholescystokinin; DA, dopamine. 118 SODHI AND SANDERS-BUSH

Fig. 1. Serotonergic projections in the human brain, arising from the raphe nuclei. The serotonergic projections innervate sympathetic preganglionic neurons, the sensory glomeruli in olfactory bulb, the intermediate lobe of pituitary, the epithelial cells of choroid plexus, the lateral ventricles, the motor neurons of brainstem, the spinal cord, visual cortex and all regions of cerebral cortex. Other transmitter systems make specialized contacts with serotonergic targets: the dopaminergic neurons in substantia nigra, the noradrenergic neurons in locus coeruleus, pacemaker neurons of the suprachiasmatic nucleus, specialized calbindin GABA interneurons in hippocampus, and pyramidal cortical neurons. Cell types in close proximity to the serotonin fibers include glia, endothelial cells, eppendymal cells in addition to the pineal gland and subcommissural organ. processes of long-term potentiation (LTP, described in Section V,A) and synaptic plasticity. The immature brain is thought to overdevelop, producing excess con- nections and cells. The redundant inputs are pruned according to their activity levels by apoptotic mechanisms guided by existing chemical systems. Therefore the ‘‘use it or lose it’’ principle states that unused connections are removed. Be- cause serotonin is likely to be present earlier in development than other monoa- mine transmitter systems and because the turnover rate of serotonin is higher in the immature mammalian brain than at any other period (Hamon and Bourgoin, 1979), serotonin probably plays a key role in this developmental process (reviewed by Whitaker-Azmitia, 2001).

A. Serotonergic Projections during Brain Development

The highest levels of serotonergic activity are detected early in development (Lidov and Molliver, 1982). Serotonergic neurons form a superior and inferior group of immature cells, which have distinct maturational and migration patterns SEROTONIN AND BRAIN DEVELOPMENT 119

(Lidov and Molliver, 1982; Wallace and Lauder, 1983). In both groups, seroto- nergic neurons form di Verent subsets of cells, and the raphe nuclei from which serotonergic fibers project in the brain all have di Verent origins (Azmitia and Gannon, 1986). In the human, serotonergic neurons can be detected when the embryo is just 5 weeks old (Sundstrom et al., 1993), with rapid growth and multi- plication until the 10th week of gestation (Kontur et al., 1993; Levallois et al., 1997; Shen et al., 1989). After 15 weeks, clustering of the serotonin cell bodies in the raphe nuclei is observed. The synaptic density of biogenic amine systems in the human cerebral cortex doubles from birth to 1 year old, when it reaches a peak value and decreases thereafter to adult levels by the age of 10 (Huttenlocher and Dabholkar, 1997). Similarly, levels of serotonin increase during the first 2 years after birth and then decline to adult levels after the age of 5 (Hedner et al., 1986; Toth and Fekete, 1986). An equivalent time course for the development of the serotonergic system has been observed in the chicken (Kojima et al., 1988) and rodents (Rubenstein, 1998) indicating that these changes are conserved in evolution. The early arrival of the serotonergic system before other monoamines indicates that it may be required to guide the development of other neuro- transmitter systems (Benes et al., 2000; Whitaker-Azmitia et al., 1996). The role of serotonin as a developmental signal is discussed in Section III,C. In the rat, axonal projections from the rostral raphe nuclei ascend to the mid- brain and forebrain, whereas those of the caudal nuclei descend to the spinal cord (Wallace and Lauder, 1983). The descending fibers enter the spinal cord by em- bryonic day 14 (E14) and innervate the preganglionic sympathetic neurons and somatic motor neurons. Here they start to form synapses at E17, with innervation of dorsal horn neurons occurring later. The rostral projections are visible soon after serotonin can be detected in the brain stem. The unbranched fibers grow in the marginal zone as a fascicle in the median forebrain bundle, and by E15 they reach the diencephalon, where they branch out. By E17, medial fibers from the medial forebrain bundle project to the frontal pole of the telencephalon, while lateral fibers project to the hypothalamus and arrive at the rostral end of the brain, some crossing the midline in the supraoptic commissure. There is a simultaneous entry by serotonin fibers into the telencephalon, the majority pass- ing through the diagonal bands of Broca, the septal areas, and then projecting into the cerebral cortex. A small number of serotonin fibers enter through the ganglionic eminences (Rubenstein, 1998). Interestingly, after birth there are non- raphe thalamocortical fibers containing serotonin in the sensory neocortex. Be- cause there is no serotonin synthesis within these neurons, serotonin must be transported to neighboring neurons from adjacent synapses (Lebrand et al., 1996), suggesting that all cells found to contain serotonin may not necessarily manufacture it. This may indicate a mechanism by which serotonin can act as a developmental signal in nonserotonergic cells (discussed in Section III,C). 120 SODHI AND SANDERS-BUSH

B. Growth Factors Influencing the Development of Serotonergic Neurons

Several growth factors that influence serotonergic development are also im- portant in synaptic plasticity events. These include the astroglial growth factor, S100 , levels of which are increased by serotonin, indicating that serotonergic nerves can stimulate their own growth. Inhibition occurs via serotonergic recep- tors at the nerve terminals. Serotonin not only regulates the growth and develop- ment of its targets, but it is also autoregulatory (Whitaker-Azmitia, 2001). S100 is described in greater detail in the review by Rothermundt, Ponath, and Arolt. The serotonin system regulates its own diVerentiation by sequential activation of the 5-HT1A receptor, brain-derived neurotrophic factor and its receptor trkB (see the review by Guillin and colleagues), CREB, CREM, and ATF-1, among many of the signaling molecules under serotonergic influence (Herdegen and Leah, 1998). In addition, Petl, an ETS domain transcription factor, is closely as- sociated with developing raphe serotonergic neurons. Other neurotrophins and growth factors control the diVerentiation of serotonergic neurons, including bone morphogenetic protein, and ciliary neurotrophic factor. Furthermore, consensus binding motifs for Petl have been detected in the 50-untranslated regions of the human and mouse genes for the 5-HT1A receptor, , aro- matic l-amino acid decarboxylase (dopa decarboxylase), and tryptophan hydro- xylase. These play important roles in the biosynthesis and degradation of serotonin from dietary tryptophan, as illustrated in Fig. 2. It follows that Petl may also be critical for the regulation of serotonin levels (Hendricks et al., 1999), which have profound eVects on brain development.

C. The Role of Serotonin as a Growth Factor

The dietary precursor for serotonin, tryptophan, is found across the phylo- genetic spectrum from lower plants to higher mammals. It is therefore an evolu- tionarily ancient chemical. Its chemical structure is shown in Fig. 2. Because tryptophan possesses an indole ring, it is capable of absorbing light and is vitally important for energy production during plant photosynthesis. Tryptophan can also be converted to auxin, which is a plant tropic factor, guiding plant growth and cell diVerentiation (reviewed by Azmitia, 2001). In animals, tryptophan is also the precursor for , a light-sensitive amine vital for the control of circadian rhythm. It follows that if auxin is a plant growth factor, then derivatives of tryptophan in animals could have similar functions. Serotonin alters the morphology of many mammalian cell types, including skeletal muscle (O’Steen et al., 1967), platelets (Leven et al., 1983), and fibroblasts (Boswell et al., 1992). Once serotonergic terminals have developed in a target SEROTONIN AND BRAIN DEVELOPMENT 121

Fig. 2. A diagram summarizing the proposed signal transduction pathways of 5-HT2 receptor subtypes. These receptors belong to the superfamily of G-protein-coupled receptors and are specifically linked to the Gq protein, which activates phopholipase C (PLC) when stimulated by the receptor- complex. This initiates the phosphoinositol second messenger cascade, producing inositol triphosphate (IP3) and diacylglycerol (DAG), which stimulate the release of calcium from intracellular stores and the activation of protein kinase C (PKC). Receptors with similar activity include the cholinergic muscarinic, metabotropic glutamate, and 2- receptors. The

5-HT2C and M1 receptors may play a major role in the regulation of synaptic plasticity, since both receptors increase the mobilization of calcium ions (Ca2þ) from intracellular stores via PLC activation, which then activates Ca2þ/calmodulin, which opens L-type Ca2þ channels, leading to plastic events. region, serotonin release may influence neurogenesis (Lauder and Krebs, 1976, 1978), neuronal removal through apoptotic mechanisms, dendritic refinement, cell migration, and synaptic plasticity (Chubakov et al., 1986, 1993; Lauder, 1990; Lauder et al., 1981). When these events are combined, they could produce the highly sophisticated organization of the hippocampus or the somatosensory maps called ‘‘barrel fields’’ (discussed in Section III,E,2). At a later developmental stage, serotonin directs dendritic growth. Serotonin activity alters the overall length of dendrites, the formation of dendritic spines, and branches into the hippocampus and cortex (Faber and Haring, 1999; Mazer et al., 1997; Okado et al., 1993; Wilson et al., 1998; Yan et al., 1997a). As the animal matures, 122 SODHI AND SANDERS-BUSH

Fig. 3. L-tryptophan is a dietary precursor of brain serotonin. Serotonin synthesis occurs as summarized above in the postganglionic serotonergic neurons. increased serotonin levels inhibit normal spine formation, perhaps by a 5-HT1A receptor-induced mechanism indicated in Fig. 3 (discussed in Section III,D,1). Depletion of serotonin levels during the developmental years may cause a loss of synapses, which can be corrected by adulthood. However, in the aged animal this may lead to an increase in the number of synapses, a ‘‘reactive synaptogen- esis,’’ or perhaps a retention of synapses normally lost (Whitaker-Azmitia, 2001). Therefore, early developmental disturbance has the potential to alter brain structure and repair much later in life. SEROTONIN AND BRAIN DEVELOPMENT 123

Studies of neurotoxin-induced serotonergic lesions of dorsal and medial raphe nuclei have detected changes in cell growth in basal ganglia regions and hypothalamic nuclei of adult rats. Newly generated granule cells can be identified by immunostaining for bromo-2 0-deoxyuridine (BrdU) and the polysialylated form of neural cell adhesion molecule (PSA-NCAM). Decreases in PSA-NCAM staining have been observed after the inhibition of serotonin synthesis induced by parachlorophenylalanine (PCPA) administration, suggesting that serotonin may reduce adhesion by acting on PSA-NCAM expression in its environment and thus facilitate plasticity in adult brain. A normalization of PSA-NCAM staining two months after neurotoxin lesions has been associated with a partial restoration in serotonin fiber density in the nucleus accumbens and the supraoptic nucleus, indicating that PSA-NCAM may facilitate the sprouting of serotonin fibers. As no changes in striatal PSA-NCAM staining have been observed after selective lesions of the dopaminergic pathway, serotonin appears to have a selective and critical role in adult brain plasticity (Brezun and Daszuta, 1999). Prenatal depletion of serotonin delays the onset of neurogenesis in serotoner- gic target regions. In the fetus evidence suggests that serotonin promotes the di- Verentiation of cortical and hippocampal neurons. In the adult brain, studies indicate that serotonin may play a role in neuronal plasticity by maintaining the synaptic connections in the cortex and hippocampus (Azmitia et al., 1995; Chen et al., 1994; Mazer et al., 1997). Neuronal precursor cells persist in adult- hood in two discrete regions, the subventricular zone and the hippocampal sub- granular zone, as demonstrated in primates (Gould et al., 1998). Both inhibition of serotonin synthesis and selective lesions of serotonergic neurons are associated with decreases in the number of newly generated cells in the dentate gyrus, as well as in the subventricular zone (Brezun and Daszuta, 1999). Serotonin depletion studies are discussed in more detail in Section IV,A. In the periphery, serotonin has been shown to potentiate the mitogenic eVects of platelet-derived growth factor BB (Eddahibi et al., 1999). The mechanism for these eVects on morphology and cell proliferation could involve the phosphoryl- ation of guanidine triphosphatase-activating protein, which is required for signal in smooth muscle cell mitogenesis induced by serotonin (Lee et al., 1997). Seroto- nergic activity also targets the cytoskeleton by inducing actin polymerization in addition to the regulation and maintenance of microtubules and microfilaments. The mechanism by which serotonin influences this growth is thought to involve the expression of S100 , a neurotrophic factor derived from astroglial cells. S100 is released from glial cells after the activation of glial 5-HT1A receptors. S100 activity increases synaptic stability and promotes neuronal development (reviewed in the chapter by Rothermundt and colleagues). Interestingly, expres- sion of the microtubule-associated protein tau is decreased in undiVerentiated neuroblastoma cells by high levels of serotonin, whereas low serotonin levels increased tau expression ( John et al., 1991). Alterations in tau phosphorylation 124 SODHI AND SANDERS-BUSH induced by serotonin have been linked with the neuropathology of Alzheimer’s disease and other forms of neurodegeneration (Gudelsky and Yamamoto, 2003; Doraiswamy, 2003; Kovacs et al., 2003; Satoh et al., 1992; Yang and Schmitt, 2001).

D. S erotonin Receptorsand D evelopmental P lasticity

Several studies have demonstrated fluctuations in the expression of serotonin receptor subtypes during brain development. In early developmental stages of the neocortex, the transient overexpression of these receptor subtypes has been dem- onstrated by autoradiography (Bar-Peled et al., 1991a; Dyck and Cynader, 1993a; Leslie et al., 1992; Lidow and Rakic, 1992). As outlined previously in Section III,C, accumulating data suggest that serotonin acts as a developmental signal for brain stem serotonin neurons and their target cells through the activation of serotonin receptors. Serotonin may act presynaptically by generating cAMP, which modulates hyperpolarization-activated cation channels (Ih channels) in axons. Compounds increasing intracellular cyclic adenosine monophosphate (cAMP) mimic and potentiate serotonin action (Dixon and Atwood, 1989; Enyeart, 1981), whereas adenyl cyclase inhibition reduces the serotonergic en- hancement of synapse transmission (Dixon and Atwood, 1989). This modulaton increases synaptic strength and could be a mechanism by which serotonin regulates synaptic plasticity (Beaumont and Zucker, 2000). Table I summarizes the distribution and activity of the serotonin receptors, whereas Fig. 4 outlines the adaptive eVects of environmental challenges mediated by serotonin. The involvement of specific serotonin receptors in synaptic plasticity events is outlined next.

1. The Influence of the 5-HT1A Receptor on Synaptic Plasticity

Evidence suggests that the 5-HT1A receptor is present early in development and is involved in the regulation of serotonergic system development (Bar-Peled et al., 1991a; Hillion et al., 1993). Experiments on pregnant rats have demon- strated that in utero exposure to PCPA, which depletes serotonin in rat brain (Lauder et al., 1985; Sanders-Bush et al., 1972a,b, 1974), reduces the postnatal ex- pression of 5-HT1A receptors, suggesting that the system is autoregulatory (Lauder et al., 2000). 5-HT1A receptors are likely to be functional before birth because they can be up- or downregulated in fetal brain (Lauder et al., 2000; Whitaker-Azmitia et al., 1990). In human fetal brain, 5-HT1A mRNA expression peaks between 16 and 22 weeks (Bar-Peled et al., 1991a). Investigations in several brain regions demonstrate that the highest level of 5-HT1A mRNA expression occurs during the development of that region and declines after maturation has occurred. This phenomenon has been reported in the brain stem (Hillion et al., SEROTONIN AND BRAIN DEVELOPMENT 125

1993), cerebellum (Daval et al., 1987), and visual cortex (Dyck and Cynader, 1993a). The 5-HT1A receptor is expressed at high density in the dentate gyrus. Evi- dence indicates that this receptor is involved in the development of dentate gran- ule cells, and it is probable that it also plays a role in maintaining synaptic integrity in the adult. The 5-HT1A receptor has also been detected on glial cells, and its activation reduces the immunoreactivity of S100 in the dentate area of adult rats, indicating that loss of synapses is due to less neurotrophic activity. In a study performed by Wilson et al. (1998), decreased serotonin in the hippocampus produced a reduction of synaptic density in the dentate molecular layer, which was attributed to lower levels of S100 reductions by astrocytic 5-HT1A receptors (Wilson et al., 1998). In addition to eVects in the hippocampus, amygdala, and cortex, the seroto- nergic raphe neurons are thought to provide important modulatory eVects on motor output systems. Developmental increases in serotonergic innervation of the rat hypoglossal nucleus, which coincide with decreased 5-HT1A receptor ex- pression by hypoglossal motor neurons (HMs), have been detected. After-spike hyperpolarization inhibition by serotonin on neonatal HMs is lost in juvenile HMs, probably due to 5-HT1A receptor activity. Therefore, 5-HT1A receptor regulation of neonatal HM function is lost in the adult (Bayliss et al., 1997). HMs contribute to innervation of the tongue muscles and to the maintenance of upper airway patency during respiration, which is important in the prevention of obstructive apneas during REM sleep (Remmers, 1990).

2. The Role of 5-HT2 Receptors in the Modulation of Plasticity In addition to the strong evidence supporting a developmental role for 5-HT1A receptors, similar data have been unearthed in investigations of 5-HT2 receptor subtypes. As with 5-HT1A receptor, the activity of 5-HT2 receptors is in- creased at critical stages of brain development. When 5-HT2R-mediated inositol phosphate levels have been measured, they are approximately 10-fold higher in developing brain compared with mature adult brain (Claustre et al., 1988). An- other similarity with 5-HT1A receptor is that pre- or postnatal environmental stressors acting via the glutamatergic system during development can alter the number and function of 5-HT2 receptors permanently (Aghajanian and Marek, 1999; Meaney et al., 1994; Peters, 1988). Furthermore, pharmacological studies by Niitsu and colleagues (1995) suggest that 5-HT2A receptor activity regulates synaptogenesis in the embryo. In the mouse, serotonin modulates embryogenesis in a dose-dependent manner by the activation of 5-HT2 receptors (Choi et al., 1997, 1998; Moiseiwitsch and Lauder, 1995, 1997; Moiseiwitsch et al., 1998; Shuey et al., 1992, 1993; Yavarone et al., 1993). Lauder and colleagues (2000) investigated expression patterns of 5-HT2 receptors during mouse embryogenesis. DiVerential 126 SODHI AND SANDERS-BUSH and overlapping spatiotemporal patterns of 5-HT2A, 5-HT2B, and 5-HT2C re- ceptor immunoreactivity were observed during active phases of morphogenesis in a variety of embryonic tissues, including neuroepithelia of brain and spinal cord, notochord, somites, cranial neural crest, craniofacial mesenchyme and epi- thelia, heart myocardium and endocardial cushions, tooth germs, whisker fol- licles, cartilage, and striated muscle. Exposure of mouse embryos at the head fold stage to selective 5-HT2 receptor antagonists revealed potent developmental eVects. The most pronounced eVect was observed after the administration of , which has high aYnity for all the 5-HT2 receptors, especially the 5-HT2B receptor subtype, and produced 100% malformed embryos. The 5-HT2A/ was 10-fold less potent, whereas ketan- serin, which primarily targets 5-HT2A receptor, did not cause a significant number of malformed embroys at any dose tested (Lambert and Lauder, 1999; Lauder, 2000). These data support previous evidence that serotonin acts as an important morphoregulatory signal during embryogenesis. As indicated in Table I and Fig. 4, 5-HT2A receptor and 5-HT2C receptor couple to the phosphoinositide (PI) hydrolysis signal transduction pathway. In 1995, Ike and colleagues demonstrated that during development a switch occurs in the functional 5-HT receptor in the rat hippocampus, from 5-HT2A to 5-HT2C, between the first and third weeks of life. 5-HT2A receptor antagonists blocked serotonin-induced PI hydrolysis in the hippocampus of 7-day-old rats

Fig. 4. Schematic diagram illustrating the eVects of stress and synaptic plasticity on behavior. This scheme focuses on the role of serotonin. Excitatory amino acids, glucocorticoids, and brain- derived neurotrophic factor (BDNF) are the topics of other articles in this issue. SEROTONIN AND BRAIN DEVELOPMENT 127 but not in 21-day-old rats. In contrast, nonselective 5-HT2A/2C receptor antagon- ists blocked serotonin-mediated PI hydrolysis in both 7- and 21-day-old rats. These data support the idea that the serotonin-induced PI hydrolysis signalling in the hippocampus of 7-day-old rats is mediated predominantly by 5-HT2A re- ceptor, but in 21-day-old rats the PI hydrolysis signal is mediated mainly by 5-HT2C receptor.Becausetherewasnoconcurrentchangeinthegeneexpression, developmental changes in receptor density were ruled out as a mechanism for the observed di Verences, but instead the explanation probably lies in altered receptor activation. Therefore, 5-HT2A receptor predominates in neonatal hippocampus, whereas 5-HT2C receptor prevails as development progresses (Ike et al., 1995). 5-HT2A receptor promotes neuronal firing by potentiating the activity of AMPA receptors and also by increasing intracellular calcium levels (Fig. 4), which increase the metabolic activity of the cell. There is enhanced glucogenesis, pro- moting cell proliferation. During development, 5-HT2A receptor has been local- ized to the neural folds, where intense cell growth occurs. Increased intracellular calcium concentration due to 5-HT2A receptor activation also stimulates protein kinase C (PKC) and increases the expression of several transcription factors. These include the immediate early gene, c-fos, and it is also coupled to the Jak/STAT pathway. The latter regulates the expression of myogenic genes and the glucose transporter GLUT3 in cultured skeletal myoblasts (Broydell et al., 1997). 5-HT2A receptors and 5-HT2C receptors also mediate the mitogenic e Vects for serotonin when expressed at high density in fibroblasts ( Julius et al., 1989, 1990). Therefore, the activation of 5-HT2 receptors influences apoptosis and cell growth and perhaps apoptotic mechanisms. 5-HT2A receptor has also been shown to influence postnatal development in the prefrontal cortex (PFC). This region of the brain comprises a large portion of the frontal lobe and undergoes progressive growth in mammals, reaching its greatest development in humans (Rakic and Goldman-Rakic, 1982). Functional studies have demonstrated that PFC plays a key role in cognitive functions (Fuster, 1991; Goldman-Rakic, 1987, 1995) and that PFC damage causes deficits in memory (Kolb, 1984) and in the organization of future events (Fuster, 1985). Dysfunction of the PFC is thought to influence many CNS disorders, particularly in schizophrenia (discussed in Section V,E), as deficiencies in cognition and working memory are prominent symptoms in the disease (Goldman-Rakic, 1994; Weinberger and Berman, 1996). A critical developmental period for the neocortex in rodents occurs during the first 2 postnatal weeks, which is also a period of intense synaptogenesis, as synaptic density increases fivefold between P10 and P15 when it is almost at the level of adult brain (Micheva and Beaulieu, 1996). Neuronal activity plays a critical role during this time, and disruption during development produces enduring changes in cortical circuitry. The strong excitatory eVect produced by serotonin during this period is modulated mainly by 5-HT2A receptors and perhaps 5-HT7 receptors (Zhang, 2003). 128 SODHI AND SANDERS-BUSH

Cytoarchitecture is also altered in the presence of 5-HT2A receptor antagon- ists by modulating expression of the activity-regulated, cytoskeleton-associated protein (ARC). ARC is an e Vector immediate early gene localized mainly in neuronal dendrites. Elevation of brain serotonin has been shown to increase the abundance of ARC mRNA abundance in the cingulate, orbital, frontal, and parietal cortices, as well as in the striatum, but conversely, produces a reduction in the CA1 region of the hippocampus. Pretreatment with a 5-HT2A receptor antagonist significantly attenuates this e Vect in the cortex. However, this attenuation is only partial and is absent in the CA1 region (Pei et al., 2000). Furthermore, administration of a single dose of lysergic acid diethylamide (LSD), which is a hallucinogen and 5-HT2 receptor agonist, has been shown to increase ARC expression in the cortex (Nichols and Sanders-Bush, 2002). The 5-HT2 receptor agonist DOI also stimulate a dose-dependent increase in ARC mRNA abundance in cortical areas, but has much weaker e Vects on ARC mRNA in the striatum and no significant e Vects in the CA1, CA3, and the dentate gyrus (DG) of the hippocampus. This DOI e Vect was completely blocked by , indicating that it is mediated by 5-HT2A receptor (Pei et al., 2000). Astroglial cell growth is also modulated by 5-HT2 receptor activity. Several glial and glioma cell lines have been shown to express these receptors (Brismar, 1995; Ding et al., 1993; Hirst et al., 1998; Merzak et al., 1996; Wu et al., 1999). The serotonin-induced cell turnover in primary cultures of astroglia from the cortex, striatum, hippocampus, and brain stem is blocked by 5-HT2A receptor antagonists. Moreover, glioma cell migration and invasion can be modulated by serotonin (Merzak et al., 1996). Therefore, the role of serotonin as a growth factor in neurogenesis and synap- tic plasticity is mediated in part by the activity of 5-HT2 receptors, and evidence that these processes may be impaired in psychiatric disorders is discussed in Sections V,D and V,E.

IV. Manipulation of the Serotonergic System Alters Synaptic Plasticity

The role of the serotonergic system has been tested in a variety of ways. The creation of lesions within serotonergic circuits alters transmission, as does reduc- ing the production of the neurotransmitter and its precursors, by dietary trypto- phan depletion. Conversely, inhibition of sensory or motor activity and investigations into the subsequent alterations in brain structure have provided insight into the serotonergic mechanisms involved in synaptic plasticity. This section discusses data from studies involving the manipulation of serotonergic activity. SEROTONIN AND BRAIN DEVELOPMENT 129

A. Tryptophanand S erotonin D epletion S tudies

Studies of serotonin depletion in rats using PCPA, a tryptophan hydroxylase inhibitor (Koe and Weissman, 1966), found that there was an increase in indu- cible nitric oxide synthase in the corpus callosum 2–5 weeks after treatment. In addition, neuronal nitric oxide synthase was increased in the striatum and hippocampus, demonstrating a close interaction between nitrergic and serotoner- gic systems during development (Ramos et al., 2002). PCPA depletion studies have alsoshownthatthedensityofCA1spinesinthehippocampusandthedorsalraphe´ nucleus are maintained by serotonergic projections to these regions (Alves et al., 2002). The importance of serotonin in synaptic plasticity during development and in the maintenance of synapses in the adult has also been demonstrated in similar studies of chicken spinal cord (Okado et al., 1993). The role of serotonin in synaptic plasticity is underscored by the strong relationship between serotonin, S100 , growth-associated protein-43 (GAP-43), and nitric oxide signaling. The latter increases cyclic guanidine monophosphate (cGMP) through guanyl cyclase activation(SouthamandGarthwaite,1993).Nitricoxideactivityisthoughttoplay acriticalroleinsynaptic plasticity andmammalian brain development(Dinerman et al., 1994; Kandel and O’Dell, 1992; Roskams et al., 1994). Long-term depletion of serotonin was modeled in adult rats by the administra- tion of PCPA for 1 week and was shown to increase binding at (S)-[3H] -amino-3- hydroxy-5-methylisoxazol-4-proprionic acid receptors (AMPAr) in the cerebral cortex. In contrast, N-methyl- d-aspartate receptor (NMDAr) binding remained constant (Shutoh et al., 2000). Moreover, the maturation of motor neurons has been investigated in rats injected with PCPA in the lumbar region after birth. Pos- tural dysfunctions were observed, including reduced flexion of the knee and ankle and reduced extension of the hip, probably due to arrested development of these neurons(Pfliegeretal.,2002).Insimilarexperiments,adultratshavebeenshownto produce an elevated glial expression of S100 , lowered serotonin levels, reduced densities of serotonin transporter (SERT), neurofilament-200 and -68 fibers (NF- 200 and NF-68, respectively), and altered cytoskeletal morphology. Even when serotonin levels normalize after treatment, cytoskeletal changes are still present in the striatum, whereas NF-200, NF-68, and SERT levels increased gradually (Ramos et al., 2000). These studies using PCPA demonstrate that serotonin appears to modulate synaptic plasticity by regulating AMPAr density and S100 release. An alternative method of depleting serotonergic transmission is through the control of tryptophan intake, which has been performed in rodents and in clinical studies (Section V,D). Tryptophan-restricted diets in rats produce increased glu- tamic acid decarboxylase (GAD) activity in the hippocampus and cerebral cortex between age 14 and 60 days. In contrast, -aminobutyric acid (GABA) activity decreases in these regions postnatally, but then increases in rats between 30 and 60 days of age. Reduced GABAergic inhibition by depletion in GABAergic 130 SODHI AND SANDERS-BUSH cells at this stage of development would result in enhanced serotonergic transmis- sion and glutamic acid decarboxylase (GAD) activity (Del Angel-Meza et al., 2002). In the pyramidal neurons of the prefrontal cortex, long-term plastic changes are observed in response to tryptophan depletion during development. At 40 days of age, rats have less profuse dendritic aborization and dendritic processes are enlarged, with an increased number of dendritic spines at 60 days (Gonzalez-Burgos et al., 1996). Additionally, reduced serotonin release is observed in prenatally malnourished anesthetized rats compared with well-nourished rats after electrical stimulation of the median raphe´ nucleus. However, the basal release of serotonin is higher before electrical stimulation in the malnourished group, perhaps due to the reduced density of serotonergic neurons leading to a decreased control of serotonin release. Electrical stimulation of the median raphe´ nucleus in the malnourished group would therefore enhance the inhibition of serotonin release through the negative feedback mechanisms controlled by 5-HT1A and 5-HT1D autoreceptors (Mokler et al., 1999). These changes in neurotransmitter function have important implications for cognitive processes, such as learning and memory (Section V,A), as tryptophan depletion has been shown to impair spatial learning, which is attributed to hippo- campal function. In contrast, short-term memory is improved due to serotonergic activity in the prefrontal cortex (Gonzalez-Burgos et al., 1996). These findings are consistent with reports that for 5-HT receptors and GABA receptors impair memory retention, whereas antagonists improve it (Farr et al., 2000). The clinical consequences of depletions in tryptophan intake or reduced serotonin levels in the brain are outlined in Section V,D.

B. Experimental Models of Synaptic Plasticity

1. The Visual Cortex The mammalian visual cortex is comparatively straightforward to manipulate and therefore has been used as a model of postnatal synaptic plasticity. If one eye and not the other is deprived of vision, then permanent impairment to the vision of the deprived eye results, with the reassignment of cortical territory to the other eye. In kittens, this process begins at postnatal day 21 (P21), peaks after 4 to 6 weeks, and decreases gradually in the subsequent months (Cynader et al., 1980; Hubel and Wiesel, 1970). During this critical developmental period there are transient, regional, laminar, and columnal changes in the distribution of seroto- nergic aVerents and receptors; increases in 5-HT1A receptor and 5-HT2C receptor densities have been noted, which peak at P30 and P40, respectively, followed by a gradual reduction to adult levels (Dyck and Cynader, 1993b). Inter- estingly, 5-HT2C receptor transient expression is concentrated in columns within the striate cortex during this period. If the eye was removed shortly after birth, SEROTONIN AND BRAIN DEVELOPMENT 131 the 5-HT2C receptor columns were absent, and if the 5-HT2C receptor antagon- ist mesulergine was administered, visual cortex plasticity was lowered during this critical period (Wang et al., 1997). The importance of this receptor was under- lined when electrophysiological studies of brain slices demonstrated that 5-HT activation of 5-HT2C receptor facilitated LTP or LTD in P60–P80 kittens (Kojic et al., 1997, 2000). It has been demonstrated further that 5-HT2C receptor activity provides a major contribution to LTP in layer 4 (Kojic et al., 2001). These recep- tors could play a significant role in the regulation of synaptic plasticity, as they increase the mobilization of calcium ions (Ca2+) from intracellular stores via phospholipase C activation (Conn and Sanders-Bush, 1984, 1986a,b, 1987; Conn et al., 1986; Sanders-Bush and Conn, 1986), which is illustrated in Fig. 2. Evidence demonstrates that synaptic plasticity is induced after intracellular Ca2+ concentration rises, perhaps via the activation of NMDA receptors or voltage- dependent Ca2+ channels (Perrier et al., 2002). 2. The Somatosensory Cortex Synaptogenesis, dendritic remodeling, and neurogenesis may all contribute to cortical columnarization. The primary somatosensory cortex (S1) of the rat has been studied extensively because of the segmented ‘‘barrel field’’-checkered patterns on layer IV of S1 when immunostained for serotonin receptors. Each barrel is produced by axons extending to the individual whiskers of the rat, making this an ideal model to test the impact of serotonergic manipulations during brain development and synaptic plasticity. A transient production of serotonergic neurons occurs in the embryo, which may regulate the develop- ment of this area. Furthermore, several serotonin receptor subtypes, including 5-HT1B and 5-HT2A, transiently appear in a vibrissae-related pattern in S1 during early postnatal development (Chiaia et al., 1994; D’Amato et al., 1987; Lebrand et al., 1996; Rhoades et al., 1990). Raising serotonin levels in rats by the administration of specific monoamine oxidase A (MAOA) inhibitors leads to an increase in the size of the barrels, with apparent fusing of the adjacent barrels (Kesterson et al., 2002; Vitalis et al., 1998). In addition, MAOA gene knockout mice are devoid of the barrel pattern, which reappears when serotonin synthesis is inhibited (Cases et al., 1996). It has also been postulated that the elevation of cortical serotonin promotes GAP-43 expression before P8 in layer IV of S1. Increasing cortical serotonin leads to accelerated development of GAP-43 axons targeting the septal regions (Kesterson et al., 2002). Conversely, if serotonergic input is reduced, altered barrel patterns are observed and delayed maturation occurs (Bennett-Clarke et al., 1994; Blue et al., 1991; Osterheld-Haas et al., 1994; Persico et al., 2000). Moreover, serotonin transporter gene knockout mice display significantly thinner barrels in the posteromedial barrel subfield layer IV and an absence of barrel patterns in other subfields of S1 (Persico et al., 2001). 132 SODHI AND SANDERS-BUSH

V. Does Dysfunction of Serotonergic Signaling Result in Impaired Brain Development?

We have already described the vastness of the serotonergic system within the brain and its widespread influence in almost every sphere of mammalian physi- ology (Section II). Furthermore, we have outlined the role of serotonin in the development of many neurotransmitter pathways and several regions of the brain (Section III). It therefore follows that a malfunctioning serotonergic system could be a contributory factor, if not a primary cause, of some developmental disorders. Synapses are probably the cellular substrate for learning and memory, therefore disruption of the serotonergic processes in brain development would lead to cog- nitive impairment. There are several behavioral and psychiatric disorders with cognitive dysfunction in which deviations from normal serotonergic activity have been discovered, such as schizophrenia. Addictive substances such as cocaine (Whitaker-Azmitia, 1998), (Kim et al., 1997; Sari et al., 2001; Tajuddin and Druse, 1993; Zhou et al., 2001), and (Muneoka et al., 1997) alter serotonin levels. These and similar substances are strongly contraindicated in pregnant women, as many have been shown to impair brain development after prenatal exposure. The serotonergic hypotheses of many psychiatric disorders are underpinned by pharmacological evidence. Many diseases are treated with serotonergic drugs, including those disorders thought to have developmental origins, such as schizo- phrenia, the aVective disorders, anxiety, and disorders of learning, such as autism and mental retardation. Behavioral problems, such as eating disorders, addiction, and stress-related disorders are also thought to be influenced by impaired seroto- nergic acitvity. Drugs are often administered for several weeks before a clinical response is observed, hence it is possible that structural changes are needed to al- leviate psychiatric symptoms. The relationship of these disorders with serotonin and synaptic plasticity is discussed in the following sections.

A. The Role of Serotonin in Learning and Memory

Long-term potentiation (LTP) is thought to be the cellular mechanism through which memories are stored in the hippocampus (Teyler and Discenna, 1984). LTP is characterized by a long-term elevation in synaptic eYcacy after a short, high-frequency electrical stimulation of aVerent fibers. Changes in synaptic density and the structure of neurons have been associated with LTP, i.e., LTP in- duces synaptic plasticity, which has a central role in nearly all models of learning and memory (reviewed by Silva, 2003). The phenomenon was first reported in the perforant path–dentate gyrus synapse (Bliss and Gardner-Medwin, 1973) and has subsequently been detected in many other hippocampal pathways. The formation of new memories is thought to require the hippocampus and SEROTONIN AND BRAIN DEVELOPMENT 133 adjacent medial temporal lobe (including the archicortical hippocampus, en- torrhinal, perirhinal, and parahippocampal cortices), but the final storage of memories is probably in a widely distributed neocortical network. Lesion studies have suggested that there is a wide distribution of neocortical memory traces en- coded in the strength of synaptic connections among neurons across large areas of the neocortex (Fries et al., 2003). Although glutamate and GABA receptors play a more major role in learning and memory (Zilles et al., 2000), serotonergic e Vects have also been detected (Beaumont and Zucker, 2000; Klancnik et al., 1991; Mazer et al., 1997; Sarihi et al., 2000, 2003), and the regions implicated in memory storage are richly innervated by the serotonergic system (as described in Section II,A, Fig. 1,andTable I). Serotonin depletion inhibits LTP in the dentate gyrus, indicating a serotoner- gic modulation of LTP in this region. Also, stimulation of the median raphe´ nu- cleus has been shown to induce LTP in the dentate gyrus (Klancnik et al., 1991). Furthermore, reversible inactivation of serotonergic projections from the median raphe´ nucleus has been observed to enhance the maintenance of LTP in the den- tate gyrus so that consolidation and retrieval but not acquisition of learned behav- ior occurs (Sarihi et al., 1999). This serotonergic lesion also enhances working memory tested by the Morris water maze (Sarihi et al., 2000). Therefore the median raphe´ nucleus has an inhibitory role in memory consolidation and re- trieval in classic conditioning experiments and spatial memory (Sarihi et al., 2003). Developing serotonergic neurons are thought to regulate growth factors for dopamine neurons, a potential mechanism for serotonin and dopamine inter- actions in schizophrenia (Whitaker-Azmitia et al., 1995). Serotonin is required during synaptogenesis: if serotonin is depleted during the critical period for sy- naptogenesis in developing rats, the result is a decreased density of dendritic and synaptic markers in the brains of the adult animals. These rats show deficits in learning and memory (Matsukawa et al., 1997; Mazer et al., 1997). Serotonin also plays a role in maintenance of the adult brain, as depletion of serotonin results in a loss of synapses (Azmitia, 1999; Okado et al., 2001). Behavioral experiments on 5-HT1A receptor gene knockout mice have demon- strated a deficit in hippocampal-dependent spatial learning and memory tasks, such as the hidden platform version of the Morris water maze, but not in nonspa- tial tasks, such as the visible platform version of the Morris water maze. Absence of paired-pulse inhibition in the CA1 region of the hippocampus has also been reported, resulting in an abnormality in short-term neuroplasticity, although LTP was not impaired (Sarnyai et al., 2000). Similar dissociations between short- term and long-term plasticity have been observed in -calmodulin kinase II ( -CaMKII) knockout mice (Frankland et al., 2001; Glazewski et al., 2000; Waxham et al., 1996) and ataxin-1 knockout mice (Matilla et al., 1998), and it has been suggested that short-term plasticity enables storage of information about the timing of events (Buonomano and Merzenich, 1995). In the basolateral amygdala, 134 SODHI AND SANDERS-BUSH

-adrenergicreceptors( AR)arecolocalizedwith5-HT1A receptorsonexcitatory nerve endings (Cheng et al., 1998; Huang et al., 1996), and when ARs are acti- vated, they induce long-term enhancement of synaptic transmission in these neurons (Huang et al., 1996). Application of serotonin lowers this synaptic trans- mission. This e Vect was mimicked by a 5-HT1A receptor agonist and blocked by a selective 5-HT1A receptor antagonist. Hence, LTP in the basal amygdala seems to be modulated by 5-HT1A receptor cross talk with ARs (Wang et al., 1999). 5-HT1A receptors also interact with glutamatergic receptors in prefrontal cortical (PFC) neurons. Postsynaptic PFC glutamatergic transmission is altered by serotonin via 5-HT1A receptor activation. This leads to a reduction in AMPA-evoked currents in PFC pyramidal neurons through a CaMKII-mediated mechanism, operated by the activation of protein phosphatase 1 and inhibition of protein kinase A. The 5-HT1A receptor/CaMKII mechanism may regulate the synaptic plasticity of PFC neurons (Cai et al., 2002). These events are crucial for learning and memory and for several forms of synaptic plasticity (Mayford and Kandel, 1999).

B. Autismand S erotonin

Autism is a behavioral disorder of unknown etiology that is four times more common in boys than in girls. Symptoms include repetitive movements, lack of imaginative play, ritualistic behavior, and poor communication, leading to re- duced social interaction. Bonding emotionally is limited and there is greater sen- sitivity to tactile (Ayres and Tickle, 1980) and auditory stimulation (Kientz and Dunn, 1997). The deficit in communication only becomes apparent at age 2, when diagnoses are usually confirmed. It was first recognized that serotonin may be connected with the etiology of autism, when elevated platelet serotonin levels were reported in autistic patients (Boullin et al., 1970, 1971; Ritvo et al., 1970). Since then, neurobiological, pharmacological, and genetic data have added to the hypothesis that dysfunction of the serotonergic system plays a role in the development of autism (Betancur et al., 2002; Cook and Leventhal, 1996; Marazziti et al., 2000). The hypothesis is strengthened by increasing evidence for serotonergic involvement in brain development and the especially rich seroto- nergic innervation of limbic areas critical for emotional expression and social behavior. Depletion of the serotonin precursor tryptophan (discussed in Section IV, A) has been reported to cause a significant deterioration in autistic patients (Cook and Leventhal, 1996; McDougle et al., 1993), and epidemiological studies detected a high frequency of prenatal exposure to cocaine (Davis et al., 1992)or alcohol (Nanson, 1992), which are both known to alter serotonergic activity. A positron emission tomography (PET) study of male autistic patients detected decreased serotonin synthesis in the cortex and thalamus, although there was an SEROTONIN AND BRAIN DEVELOPMENT 135 increase in the dentate nucleus (Chugani et al., 1997), which could lead to develop- mental abnormalities. In fact, postmortem studies have found reduced branching of dendrites in CA1 and CA3 hippocampal regions (Raymond et al., 1996), along with reduced hippocampal volumes, suggesting that the cells failed to mature and connections with the cortex may not have fully developed (Aylward et al., 1999). It is possible that the high levels of serotonin detected in autism would create a loss of serotonergic nerve terminals during brain development. This may ex- plain why eVicacious treatments for autism enhance serotonergic transmission (Buitelaar and Willemsen-Swinkels, 2000; McDougle et al., 2000; Posey and McDougle, 2000). Drugs targeting the serotonin transporter (SERT), including the selective serotonin reuptake inhibitors (SSRIs) fluoxetine, , fluvox- amine, and , are now widely used to treat autism (Fatemi et al., 1998; Hollander et al., 2000a; Namerow et al., 2003). SSRIs appear to alter many autistic symptoms, including social relatedness. , an atypical antipsy- chotic with strong aYnity for many serotonin receptors, particularly 5-HT2A re- ceptor, is also used frequently to treat autism. Neuroendocrine and behavioral challenge paradigms have found abnormal responses to the increased serotonin levels after the administration of fenfluramine or the serotonin precursor 5-hy- droxytryptophan or the direct 5-HT1B/D receptor agonist . It has been suggested that 5-HT1D receptor supersensitivity may be responsible for the repetitive behaviors exhibited by autistic children (Hollander et al., 2000b). Evidence suggests that autism is a disorder of impaired brain development. The limited pharmacological data available provide circumstantial evidence that impaired serotoninergic activity may contribute to the etiology of this disorder. It is possible that the development of brain regions involved in communication and emotion are arrested by prenatal processes altering synaptic plasticity via a sero- tonergic mechanism that is triggered by either genetic or environmental factors.

C. The Role of Serotonin in Stress and Anxiety

Serotonin is a chemical mediator of inflammation. Its secretion and physio- logical actions mediate stress and pain, aVecting both immune and nervous system functions through the hypothalamic–pituitary–adrenal (HPA) axis. Sero- tonin receptor dysfunction is well characterized in mental disturbances such as depression and anxiety. Considerable evidence supports the idea that the early postnatal period is a critical time for the establishment of lifelong anxiety-related behavior, which is a component of many psychiatric disorders. Moreover, clinical studies demon- strate that early life stressors, such as divorce or bereavement, increase suscepti- bility to anxiety and mood disorders in adulthood (Breier, 1989; Bulik et al., 2001; Kendler et al., 1992, 1993, 1996, 2000, 2002a,b; Kessler et al., 1997). 136 SODHI AND SANDERS-BUSH

The mechanism linking stress with abnormal plastic events during brain de- velopment is thought to involve 5-HT1A receptor activity (discussed in Section III,D). Chronic stress has been shown to produce specific downregulation of 5-HT1A receptors in the hippocampus ( Lopez et al., 1999; Pare and Tejani-Butt, 1996). This hippocampal deficit may partially explain the cognitive dysfunction observed in patients with aVective disorders. Mice lacking the 5-HT1A receptor gene display increased anxiety in behavioral models such as the open field test or elevated plus maze (Heisler et al., 1998a; Parks et al., 1998; Sarnyai et al., 2000; Sibille et al., 2000), and the anxious phenotype is associated with changes in GABAAR neurotransmission (Olivier et al., 2001). This behavior can be re- versed by the selective reinstatement of 5-HT1A expression in the hippocampus and cortex during early postnatal development, but not by reinstatement in the adult or in raphe nuclei at any age (Gross et al., 2002). These data indicate that interrupted postnatal 5-HT1A receptor developmental processes contribute to anxiety behavior in adulthood. Active and passive stress responses have been compared in short (SAL) and long attack latency (LAL) mice, which are genetically selected mouse lines re- sponding aggressively to an opponent in the intermale resident–intruder experi- ment. LAL mice have a characteristic chronic elevation in corticosterone levels and a decreased level of 5-HT1A receptors in the dentate gyrus, hippocampal CA1 region, lateral septum, and frontal cortex, but not in the dorsal raphe nu- cleus (Korte et al., 1996; van Riel et al., 2002). These findings are consistent with the reduced 5-HT1AR expression in mice exposed to chronic stress and the resultant elevation in glucocorticoid levels (Fernandes et al., 1997; Flugge, 1995; Lopez et al., 1998; McKittrick et al., 1995; Meijer et al., 1997a,b; Watanabe et al., 1993; Wissink et al., 2000). In addition to corticosteroids, sex steroids have been shown to alter behavior and synaptic plasticity. Stress-induced reductions in 5-HT1A receptors in the hippocampus can be renormalized by the administration of androgens (Flugge et al., 1997, 1998). Furthermore, hormone levels are known to alter mood, and there are strong links between estrogens and aVective disorders. Fluctuations in these hormones can cause mood swings, anxiety, or postnatal depression. The aVective symptoms, irritability, and anxiety associated with low estradiol levels in postmenopausal women are alleviated by estrogen replacement ( JoVe and Cohen, 1998; Soares et al., 2001). Furthermore, high levels of estrogens have been shown to increase dendrite growth and synaptic plasticity in the rat hippo- campus (Foy et al., 1999; Good et al., 1999; Leranth et al., 2000). As with 5-HT1A receptor activity, estrogens reduce neuronal excitability in the basolateral amyg- dala (Edwards et al., 1999). Female mice lacking the estrogen receptor (ER ) were shown to be more anxious and to have a lower threshold for the induction of synaptic plasticity in the basolateral amygdala, which coincided with increased 5-HT1A receptor expression in the medial amygdala (Krezel et al., 2001). Thus, SEROTONIN AND BRAIN DEVELOPMENT 137 the serotonin system may influence the steroid hormone control of synaptic plasticity. As described in Section III,D,1, 5-HT1A receptor activity inhibits synaptic plasticity in the amygdala, which is considered to be the sensorimotor interface for conditioned fear. Classic fear conditioning, a paradigm for the study of aver- sive learning and memory, is thought to depend on the integrity of the amygdala (Fendt and Fanselow, 1999; Rogan and LeDoux, 1996), although there is debate regarding the role of the hippocampus in this response (Maren et al., 1997; Radulovic et al., 1999). Studies either depleting or increasing serotonin have pro- vided evidence for the role of serotonin in fear conditioning (Archer et al., 1982, 1984; Hashimoto et al., 1996; Inoue et al., 1996). A specific role for 5-HT1A recep- tor activity in this response has been indicated by studies of 5-HT1A receptor an- tagonists in combination with serotonin reuptake inhibition (Hashimoto et al., 1997). The 5-HT1A receptor agonist 8-OH-DPAT caused a deficit in contextual fear due to postsynaptic 5-HT1A receptor activation in rats. Both responses were blocked by the 5-HT1A receptor antagonist WAY100635. Therefore, postsynap- tic 5-HT1A receptor activation has been shown to interfere with learning pro- cesses in the acquisition of fear (Stiedl et al., 2000). In addition, PET studies of patients with major depression show decreased 5-HT1A receptor binding in the temporal lobe and in the limbic system (Drevets et al., 1999; Sargent et al., 2000), and 5-HT1A receptor agonists have anxiolytic properties in clinical and animal models (Feighner and Boyer, 1989; Menard and Treit, 1999). Although the 5-HT1A receptor plays an important role in anxiogenic pro- cesses, the 5-HT2B/2C receptor agonist m-chlorophenylpiperazine (mCPP) also induces anxiety. mCPP has been shown to reduce novelty-seeking behavior. In a two-box light/dark choice situation, mCPP has been found to decrease the time spent by mice in the lit box and the number of transitions between the light and dark boxes. This behavioral test has been validated for the assessment of novel compounds with anxiolytic or anxiogenic properties. Therefore, activation of 5-HT2C receptor enhances anxiety responses toward novel and aversive places (Griebel et al., 1991; Meert et al., 1997). Immobilization of animals is another stress-inducing behavioral paradigm. Im- mobilization decreases the expression of BDNF mRNA in the rat hippocampus, and this eVect could contribute to the atrophy of hippocampal neurons. Pretreatment with a selective 5-HT2A receptor antagonist significantly blocks the influence of stress on the expression of BDNF mRNA. In contrast, pretreatment with either a selective 5-HT2C or a 5-HT1A receptor antagonist did not influence the stress-induced de- crease in levels of BDNF mRNA levels (Vaidya et al., 1999). Furthermore, the inhibi- tory eVects of stress on the activity of periventricular hypophysial dopaminergic (PHDA) neurons have been shown to be mediated by serotonergic neurons, acting via 5HT2Rs, and this activity leads to an increase in the secretion of -melanocyte- stimulating hormone ( -MSH) (Goudreau et al., 1993). Therefore, behavioral 138 SODHI AND SANDERS-BUSH pharmacological evidence indicates an important role for the serotonergic system, together with steroid hormones, in the etiology for anxiety disorders.

D. Serotonergic Influences on Synaptic Plasticity in Affective Disorders

The catecholamine hypothesis of the aVective disorders was proposed in 1965 based on the observation that the antihypertensive agent reserpine was found to lower the mood of patients (Schildkraut, 1965). This depletes catecholamines, therefore it was suggested that depression was caused by a dysfunction in the cat- echolamine system. Coppen (1967) modified this theory after the eVects of mono- amine oxidase inhibitors (MAOI) on serotonin were observed by suggesting that depression was caused by reduced serotonin in the synapse (Fig. 2)(Coppen et al., 1967). However, the delay in patient response to after administra- tion (which is also seen in antipsychotic therapy) could not be correlated with the rapid eVects of antidepressant drugs on serotonin levels, and their eYcacy is thought to be related to more long-term changes caused by plastic events under the control of the monoamine transmitter systems (Charney et al., 1981). The monoamines, noradrenaline, serotonin, and dopamine, and their metab- olites, 3-methoxy-4-hydroxyphenylglycol (MOPEG), 5-HIAA, and homovanillic acid (HVA), respectively, have been studied extensively in the blood, cerebro- spinal fluid (CSF), and postmortem brain of depressed patients. There is a general consensus after many clinical studies that MOPEG in the urine and CSF of de- pressed patients is reduced by about 25% in comparison with controls and that this shows a cyclic change when manic and depressed states are experienced by bipolar patients ( Johnstone, 1982). In 1976, Asberg and colleagues measured 5-HIAA in the CSF as an index of brain serotonin turnover (Fig. 2) and found a bimodal dis- tribution in depressed patients. It was also observed that the severity of depression increased with decreased 5-HIAA levels in a subgroup of patients who were more prone to violent suicide attempts. Several studies have also measured decreases in HVA levels in depressed patients, but once methodological details were stream- lined, no diVerences in CSF HVA, the dopamine metabolite, were observed be- tween patients and controls. By contrast, more than 20 studies have detected a consistent association between high 5-HIAA levels and suicidal behavior, schizo- phrenia, personality disorder, and impulse control but not bipolar aVective dis- order. A low concentration of 5-HIAA in patients is associated with an increased short-term risk for suicide in patients (reviewed by Asberg, 1997). It has been suggested that abnormalities in receptor sensitivity are present in de- pressed patients (Charney et al., 1981) and that they fail to make adaptive responses to stress or adverse stimuli due to dysfunctional neural plasticity. Therefore the mechanism of antidepressant action and perhaps the etiology of aVective illnesses may involve the induction of specific plastic changes by serotonergic activation. SEROTONIN AND BRAIN DEVELOPMENT 139

Alterations in levels of serotonin during the postnatal period produce long- lasting changes in brain plasticity, including disruption of the maps of the sensory and visual cortices (Section IV,B), decreased density of dendritic spines (Yan et al., 1997b), and changes in the dentate gyrus (Section IV). One hypothesis has sug- gested that depression develops as a consequence of the poor availability of diet- ary tryptophan to the brain, which would reduce the availability of serotonin (Fig. 3). A study of plasma tryptophan in aggressive males found that high trypto- phan levels correlated with anxiety (Cleare and Bond, 1995). Other investigations have associated tryptophan with feelings of well-being (Charney et al., 1982; Smith et al., 1997). Eighty to 90% acute depletion of dietary tryptophan has been associated with a lowering of mood in normal females but not in males in one study (Ellenbogen et al., 1996), and depletion of tryptophan has been shown to induce relapse of depression in more than 50% of patients during remission of the disease. This relapse was reversed when tryptophan levels were restored; free plasma tryptophan levels were found to be correlated inversely with depression scores during acute tryptophan depletion (Delgado et al., 1990; Smith et al., 1997). Changes in synaptic plasticity after tryptophan depletion are discussed in Section IV,A, and it is possible that these changes are involved in the etiology and treatment of the aVective disorders. Estrogens and aVective disorders are strongly linked, as described in the previ- ous section, since fluctuations in these hormones may cause mood swings, anxiety, or postnatal depression. High levels of estrogens have been shown to increase den- drite growth and synaptic plasticity in the rat hippocampus (Foy et al., 1999; Good et al., 1999; Leranth et al., 2000). Furthermore, as with 5-HT1A receptor stimula- tion, estrogens reduce neuronal excitability in the basolateral amygdala (Edwards et al., 1999). Investigations into the mechanisms by which endogenous factors stimulate neurogenesis have determined that serotonin and estradiol act through a common pathway to increase cell proliferation in the adult dentate gyrus. Neurogenesis has also been studied in animal models of depression. Ovariec- tomy is a method of depleting estrogens, creating depressive behavior in rodents. Combining ovariectomy with inhibition of serotonin synthesis using PCPA treat- ment produced approximately the same decreases in the number of BrdU and PSA-NCAM-immunolabeled cells (indicating a reduction in newly generated cells) in the subgranular layer as ovariectomy alone. Administration of 5-hydro- xytryptophan (5-HTP), a precursor of serotonin (Fig. 3), has been shown to restore cell proliferation decreased primarily by ovariectomy, whereas estradiol does not reverse this change. Estrogen may regulate structural plasticity by stimu- lating PSA-NCAM expression independently of neurogenesis, as shown by the increases in the number of PSA-NCAM-labeled cells in pregnant rats. These data indicate that positive regulation of cell proliferation and neuroplasticity by serotonin and estrogen may contribute to the reduced hippocampal connectivity observed in depressed patients (Banasr et al., 2001). 140 SODHI AND SANDERS-BUSH

The aVective symptoms of irritability and anxiety, which are associated with low estradiol levels in postmenopausal women, are alleviated by estrogen replace- ment ( Jo Ve and Cohen, 1998; Soares and Cohen, 2001). Female mice lacking the estrogen receptor (ER ) were shown to be more anxious and to have a lower threshold for the induction of synaptic plasticity in the basolateral amygdala, which coincided with increased 5-HT1A receptor expression in the medial amyg- dala (Krezel et al., 2001). Together, these studies demonstrate the link between the serotoninergic system and the steroid hormone control of synaptic plasticity in the aVective disorders.

1. Serotonin Receptors and Depression The regulatory actions of serotonin are mediated by serotonin receptors, therefore the distinct temporal expression patterns of serotonin receptors may re- flect their changing roles during development. Serotonin is thought to regulate the production of neurotrophic factors in the CSF. 5-HT2C receptors are present at high density and probably play a role in CSF production. 5-HT2c receptors may mediate these regulatory functions of serotonin (Esterle and Sanders-Bush, 1992). Psychiatric disorders comprise an array of overlapping symptoms, and pa- tients diagnosed with a variety of CNS disorders often su Ver with sleep disturb- ances. Staner and colleagues (1992) demonstrated that 5-HT2 receptor antagonists produced increased slow wave sleep in healthy individuals, whereas depressed patients exhibited a smaller increase in slow wave sleep at the same dose of drug (Staner et al., 1992). 5-HT2 receptor antagonists have antidepressant potential (Eison, 1990; Marek et al., 1989), which may be related to an interaction with 5-HT1A receptors (Eison, 1990; Yocca et al., 1990). Alterations in 5-HT2A receptor expression and activity have been demon- strated in patients with a range of CNS disorders, and the changes could play a role in the pathogenesis and treatment of psychiatric disease (Burnet et al., 1996, 1999; Eastwood et al., 2001; Harrison and Geddes, 1996; Harrison and Burnet, 1997). An increase in postsynaptic 5-HT2A receptor binding in postmortem brains of patients may reveal a pathophysiological mechanism in aVective dis- orders (McKeith et al., 1987; Yates et al., 1990). The deficiency of serotonergic ac- tivity in depressed patients is thought to increase their vulnerability to the disease. Studies have shown low levels of serotonin in the brains of depressed patients who have committed suicide (Coppen and Doogan, 1988) and that there are altered densities of 5-HT2A receptors in the frontal cortex of suicide victims (Mann et al., 1986a,b,c; Stanley et al., 1986a,b), in depressed patients (McKeith et al., 1987), and in individuals at risk for suicidal behavior (Arango et al., 1997; Audenaert et al., 2001; Pandey et al., 1995, 1997; Stockmeier et al., 1997). Drug treatments for the aVective disorders, for example, antidepressants such as , mianserin, and fluoxetine, are thought to act through the serotonergic system and usually raise serotonin levels. Therefore, it has been proposed that SEROTONIN AND BRAIN DEVELOPMENT 141 there is a decrease in serotonergic activity in the brains of depressed patients, leading to an upregulation of 5-HT2A receptors. Evidence supporting this hy- pothesis includes the finding that there is increased 5-HT2A receptor-mediated phosphoinositide turnover in the platelets of patients with aVective disorders (Mikuni et al., 1992). Moreover, downregulation of brain 5-HT2 receptors has been demonstrated 2–4 weeks after the administration of many antidepressants (Peroutka and Snyder, 1980), such as lithium (Treiser and Kellar, 1980; Wajda et al., 1986), monoamine oxidase inhibitors (Kellar et al., 1981), some selective serotonin reuptake inhibitors (SSRIs) (Eison et al., 1991; Nelson et al., 1989; Stolz et al., 1983), 5-HT1A receptor partial agonists (Lafaille et al., 1991; Yocca et al., 1991), and (Hietala et al., 1992). Conversely, electroconvulsive therapy upregulates 5-HT2A receptor-binding sites (Burnet et al., 1996; Kellar et al., 1981). Alterations in 5-HT2A receptor binding also appear to be related to the severity of depressive symptoms, as clinically improved patients have similar densities of 5-HT2A receptors to controls (Yates et al., 1990). This evidence has been con- firmed by studies in platelets when patients were treated successfully with antidepressants (Biegon et al., 1990). Raising the level of serotonin in the synapse by the administration of anti- depressants produces a cascade of events mediated by 5-HT2, 5-HT1A, 5-HT4, and 5-HT7 receptor subtypes. 5-HT2 receptor stimulation raises intracellular Ca2+ levels and activates CaM kinases, which lead to CREB phosphorylation at ser133. This increases expression of the BDNF gene, which promotes neuronal plasticity and survival. Similarly, 5-HT1A receptor and 5-HT7 receptor stimulation increases Raf, MEK, Erk 1, Erk 2, and Rsk 2 (critical for cell survival), which leads to phosphorylation of the proapoptotic protein BAD and inactivates it. Rsk also activates CREB phosphorylation, conferring cell survival by increasing expression of the antiapoptotic protein Bcl2 (reviewed by D’Sa, 2002; Duman, 2000). Ritanserin, a 5-HT2A receptor antagonist, has been reported to have anxio- lytic or antidepressant properties in several studies (Bakish et al., 1993; Bersani et al., 1991; Ceulemans et al., 1985). blocks both noradrenaline and serotonin reuptake in addition to 5-HT2 receptors and has demonstrated anti- depressant eYcacy in several controlled trials and was well tolerated by patients (Feighner et al., 1998). Another 5-HT2 receptor antagonist, , has been launched as a treatment for depression. Mirtazapine also acts as an antagonist at 2 and 5-HT3 receptors and this combined activity increases noradrenergic and serotonergic transmission, contributing to its therapeutic eYcacy (de Boer, 1996; de Boer et al., 1996; Dinan, 1996). The drug clozapine has also demonstrated eYcacy in the treatment of depressive symptoms of bipolar aVective disorder, but may exacerbate mania (Frye et al., 1998). Although 5-HT2A receptors have been investigated extensively in studies of mood disorders, there is also evidence implicating altered 5-HT2C receptor func- tion in these illnesses. Molecular genetic studies show that a single nucleotide 142 SODHI AND SANDERS-BUSH polymorphism in the 5-HT2C receptor gene is associated with bipolar aVective disorder (Lerer et al., 2001) and that the locus of the 5-HT2C receptor gene, Xq24, is linked to this disease (Craddock and Jones, 1999, 2001). SSRIs have been shown to have high aYnity for 5-HT2C receptor, and antipsychotic drugs with high 5-HT2C receptor aYnity, such as clozapine, are also indicated for the treatment of bipolar disorder (Calabrese et al., 1991, 1996; Kimmel et al., 1994). Increased 5-HT2C receptor responsiveness accompanies the isolation-rearing model of anxiety/depression and may contribute to the enhanced response to stress and the increased neophobia observed. In isolation-reared rats, rapid down- regulation of supersensitive 5-HT2C receptors may occur in the hippocampus following a serotonergic agonist challenge (Fone et al., 1996). 2. The Serotonin Transporter and Antidepressants A major site of action for many antidepressants is the serotonin transporter (SERT). A growing body of evidence indicates that variability in SERT gene ex- pression influences temperamental traits, which could be determined by genetic factors, and may lead to psychopathology. Anxiety, depression, and aggressive behavior are all thought to be alleviated by altered SERT activity, but the mech- anism of therapeutic eYcacy of SSRIs is still unclear and animal models provide a valuable tool for investigation. Animal models of depression have been varied, including procedures such as olfactory bulbectomy, maternal deprivation, and clomipramine administration. Dendritic spines could represent an anatomical marker for the enduring changes accompanying depression. Dendritic spines are the postsynaptic sites of most ex- citatory synapses, and their density increases during the first 2 postnatal months in rat hippocampus. Synaptic development is altered significantly in the hippo- campus if there is variation in the levels of serotonin and norepinepherine during this time period. Norrholm and Ouimet (2000) have examined dendritic spine density in the CA1 region of the hippocampus and dentate gyrus of juvenile rats after acute and chronic exposure to antidepressant drugs. Acute antidepressant treatment has been reported to increase dendritic length and spine density, whereas chronic treatment with fluoxetine, a selective serotonin reuptake inhibi- tor, arrests spine development into young adulthood. Further investigations have revealed that olfactory bulbectomy reduces spine density in CA1, CA3, and den- tate gyrus compared to sham-operated controls. Chronic treatment with a non- specific antidepressant reverses the bulbectomy-induced reduction in dendritic spine density in CA1, CA3, and dentate gyrus. However, the mianserin, with 5-HT2A/2C receptor antagonist properties, only reversed this reduction in dentate gyrus. By contrast, other models of depression do not demonstrate this eVect. Hence enduring changes in hippocampal den- dritic spine density could contribute to a neural mechanism specific to this model of depression (Norrholm and Quimet, 2001). SEROTONIN AND BRAIN DEVELOPMENT 143

It has been reported that the projections arising from the hippocampal struc- tures to the hippocampo-medial prefrontal cortex (mPFC) are involved in the exe- cution of higher cognitive functions in rats. The eVects of single and repeated antidepressant treatment on the synaptic eYcacy and synaptic plasticity in the rat medical (mPFC) pathway have been examined using the drug fluvoxamine, a select- ive serotonin similar to fluoxetine. has been reported to enhance synaptic eYcacy in the hippocampo-mPFC pathway in a dose-dependent manner, and repeated treatments enhance synaptic plasticity, so that the establishment of long-term potentiation in the hippocampo-mPFC pathway improves significantly. These findings may indicate the mechanism by which SSRIs produce their therapeutic eVects in depressive disorders (Ohashi et al., 2002). Investigations in mice lacking the SERT gene have demonstrated adaptive changes in 5-HT2A receptor function. Autoradiographic labeling of these recep- tors by the selective antagonist [3H]MDL 100,907 and saturation experiments with cortical membranes have revealed a new localization of 5-HT2A receptors in the external field of striatum and regional variations in adaptive changes in the density of 5-HT2A receptors in SERT (-/-) mutants: a reduction of 30–40% in the claustrum, cerebral cortex, and lateral striatum when compared to wild-type mice (Rioux et al., 1999). Furthermore, immunohistochemistry of the cerebral cortex of SERT knockout mice has revealed a nearly complete absence of sero- tonin and of barrels, both at P7 and adulthood. VMAT2 knockout mice, which completely lack an activity-dependent vesicular release of monoamines, including serotonin, also display an absence of serotonin in the cortex but have almost normal barrel fields, albeit with some reduced postnatal growth. These data sup- port the idea that transient SERT expression is required for barrel pattern forma- tion, whereas activity-dependent vesicular serotonin release is not essential for this process (Persico et al., 2001). Although the significance of these observations has yet to be fully explained, the importance of serotonin receptors and SERT in the etiology and treatment of the aVective disorders has been established. The development of depressive illness is reviewed more comprehensively by Lesch (2000, 2001).

E. Altered Synaptic Plasticity in Schizophrenia

1. Is Schizophrenia Caused by Impaired Brain Development? The neurodevelopmental and neurodegeneration hypotheses of schizophre- nia have been opposing etiological theories for many decades. The major glitch in the neurodegenerative hypothesis has been the failure to detect gliosis in postmortem schizophrenic brain (reviewed by Harrison, 1995). The alternative, neurodevelopmental hypothesis, proposes dysfunctional brain development and growth during embryogenesis and childhood before the emergence of 144 SODHI AND SANDERS-BUSH schizophrenia during or after adolescence (Weinberger, 1995). Although the neu- rodevelopmental theory predominates, concurrent neurodegenerative processes that do not involve sustained gliosis have not been excluded (Woods, 1998). The neurodevelopmental hypothesis is supported by findings of delayed de- velopmental milestones in preschizophrenic children, including learning and be- havioral abnormalities, indicating abnormal brain function prior to the diagnosis of schizophrenia (Cannon et al., 1997; Chua et al., 1996; Hanson et al., 1976; Jones et al., 1994; Lewis and Levitt, 2002; Marenco et al., 2000; van Os et al., 1995). Murray and colleagues (1992) suggested that the age of disease onset could be used to categorize three forms of schizophrenia: congenital, adult onset, and late onset (Murray et al., 1992). The former can be traced to brain abnormal- ities during pre- or perinatal stages of development, with gradual increases in behavioral disturbances until the disease can be diagnosed in adolescence or early adulthood. The congenital form is thought to be more common in males (Pilowsky et al., 1993; Stober et al., 1998), whereas the late-onset disorder is more prevalent in females (Hafner, 2003; Palmer et al., 2001). Several neurodevelopmental mechanisms have been proposed. In addition to theories of brain lesions that occur before infancy and lie dormant until adoles- cence (Weinberger, 1987), it is also possible that a lesion occurs during ad- olescence before the onset of disease symptoms. Alternatively, computer simulations have proposed a model of schizophrenia in which reduced synaptic connectivity occurs after disturbed neural development during both perinatal and adolescent periods (McGlashan and HoVman, 2000). Alterations in brain structure have been observed in postmortem studies and neuroimaging of schizophrenia subjects. These changes have not been conclu- sive, as there is considerable overlap with the normal range measured in controls. In addition, there are several problems with these studies, including small sample sizes and patient drug history, because antipsychotic drugs have been shown to produce structural changes in the brain (Crow et al., 1986; Meredith et al., 2000). Even so, a consensus has been reached for some neurobiological changes that have been confirmed by several independent studies summarized later. Initially, anatomical research of schizophrenic brain identified enlargements of the lateral and third ventricles in conjunction with a decrease in cortical volume. The latter was particularly prominent in the temporal lobe and medial temporal lobe structures, especially the hippocampal formation and amygdala (Lawrie and Abukmeil, 1998; Wright et al., 2000). Subcortical structures appear to be reduced in size, including some thalamic nuclei (Pakkenberg and Gundersen, 1989; Popken et al., 2000; Young et al., 2000) and the striatum (Keshavan et al., 1998). Reduced normal brain asymmetry is another consistent finding with increased structural vari- ability in the left hemisphere of schizophrenia brains (reviewed by Harrison, 1999b). Considering the early anatomical evidence, Feinberg (1982) proposed that the central pathogenetic process in schizophrenia included altered synaptic pruning SEROTONIN AND BRAIN DEVELOPMENT 145 during adolescence (Feinberg, 1982). A smaller brain volume in schizophrenia could be explained by increased cell death, which would account for the decreased volume of the thalamus (Young et al., 2000), but not the changes in the cortex and hippocampus. These structures do not display cell loss (Pakkenberg, 1993) but in- stead there are smaller neurons and more densely packed cortical cells (Selemon et al., 1995, 1998; Weinberger, 1999). Altered synaptic plasticity has been inferred by these data, as neuronal cell body size correlates with the extent of dendritic ar- borization (Selemon and Goldman-Rakic, 1999). Moreover, fewer dendritic spines have been observed on cortical and hippocampal pyramidal neurons in schizo- phrenia subjects (Garey et al., 1998; Glantz et al., 2000; Rosoklija et al., 2000), and a decrease in the number of synapses in the prefrontal cortex and hippocam- pus in schizophrenia is suggested by the lower abundance of synaptic proteins in these areas (reviewed in the chapter by Eastwood). Altered synaptic plasticity would probably release a cascade of events altering the expression of genes control- ling synaptic function, especially during critical developmental periods. Although the jury is still out on the underlying mechanisms surrounding these developmental disturbances, because the serotoninergic system plays such an important role both in neurodevelopment and in the pharmacotherapy of schizophrenia, it is possible that an aberrant serotonin-dependent developmental process contributes to the etiology of the disease. 2. The Serotonergic Hypothesis of Schizophrenia The idea that serotonergic activity contributes to the etiology of schizophre- nia evolved from the observation that lysergic acid diethylamide, a drug structur- ally similar to serotonin, was hallucinogenic (Gaddum and Picarelli 1954; Jacobs et al., 1979; Pieri et al., 1978). Initially there were doubts about the relevance of these findings, because the psychosis induced by LSD involved primarily percep- tual disturbances, i.e., there was a prevalence of visual rather than auditory hal- lucinations, an absence of thought disorder, and the preservation of aVect and insight, which is only found in a small proportion of patients with schizophrenic psychosis (Szara, 1967). However, these diVerences were minimized when LSD psychosis was compared with the early stages of schizophrenia rather than the chronic illness (Freedman and Chapman, 1973). The LSD psychosis has been found to be a close model for the reality distortion syndrome in schizophrenia but not for the negative symptoms prevalent in a subset of patients (Slade, 1976). Evidence suggests that LSD produces its psychotomimetic eVects through the stimulation of 5-HT2 receptors (Sanders-Bush et al., 1988; Sanders-Bush and Breeding, 1991), and because atypical antipsychotic drugs have high aYnities for these receptors, there is circumstantial pharmacological evidence for the role of serotonin receptors in the development of schizophrenia. The behaviors observed after the administration of serotonin receptor ligands and the distribution of sero- tonin receptors in brain regions involved in cognition and mood also support this 146 SODHI AND SANDERS-BUSH hypothesis and are summarized in Table I. Also the importance of serotonin in development and neural plasticity has been established in previous sections of this review. Further evidence connecting serotonin with the development of schizophrenia includes the fact that serotonin levels can be altered during viral infections in development (Pletnikov et al., 2000; Popenenkova et al., 1977; Yamashita et al., 1989), malnutrition (Blatt et al., 1994; Kaye et al., 1991; Manjarrez et al., 1996), social isolation (Whitaker-Azmitia et al., 2000), hypoxia (Kim et al., 1994), and stress (Lapiz et al., 2001; Read et al., 2001; Robinson and Becker, 1986) (discussed in Section V,C). Most of these are epidemiological risk factors for schizophrenia later in life (reviewed by Lewis and Levitt, 2002) and are associated with altered brain development in the embryo and early childhood (Miller and Azmitia, 1999). a. Serotonin Receptors and Pharmacological Studies of Schizophrenia. The core symp- toms distinguishing schizophrenia from other mental disorders according to DSM IV diagnostic criteria are delusions and hallucinations, which contribute to the loss of insight, which is characteristic of the disease. It is recognized that hallucinations occur after activation of the 5-HT2 receptors because, LSD acts potently at 5-HT2 receptor sites, as do the substituted hallucino- gens such as (Aghajanian, 1994). Moreover, the aYnity of these com- pounds for 5-HT2 receptor sites has been found to be closely related to their as hallucinogens in humans (Glennon et al., 1984; Titeler et al., 1988). Humanstudies ofthepsilocybinindicatethat 5-HT2 receptors areinvolved inhal- lucinogenesis (Vollenweider and Geyer, 2001; Vollenweider et al., 1998). Evidence fromantipsychotic treatment studies also supports theserotonergic hypothesis.Al- though the discovery of antidopaminergic antipsychotic drugs meant that this hy- pothesis was eclipsed temporarily by the dopamine theory of schizophrenia, interest in serotonin reemerged with the discovery of atypical antipsychotic treat- ments.Thearchetypeatypicaldrugclozapineisanantipsychoticwithhighe Ycacy in patients exhibiting a poor response to conventional antidopaminergic antipsy- chotics. Clozapine has been found to have a higher a Ynity for serotonin receptors thanfordopaminereceptors(Meltzeretal.,1989a,b,c).Itisstillundecidedwhether thesepharmacologicalindicationsaresymptomaticorcausalinschizophreniaand therefore biochemical and postmortem studies have been conducted to address this question. The proposed dysfunction in synaptic plasticity in schizophrenia could un- leash a cascade of changes in the expression of genes controlling synaptic func- tion, especially during critical developmental periods. Because serotonin plays a vital role during development (discussed in Sections III and IV), investigations of gene expression in schizophrenia may provide etiological clues. In fact, postmor- tem studies have found alterations in the abundance of several 5-HTRs. Binding sites for 5-HT1A receptors are increased in the prefrontal cortex, cingulate cortex, and temporal cortex in schizophrenia (Burnet et al., 1996; Gurevich et al., 1997; SEROTONIN AND BRAIN DEVELOPMENT 147

Hashimoto et al., 1991, 1993; Simpson et al., 1996; Sumiyoshi et al., 1996), with no changes in the encoding mRNA (Burnet et al., 1996). Reductions in the abundance of 5-HT2A receptor mRNA have been reported in the prefrontal, cin- gulate, temporal, and occipital cortices (Burnet et al., 1996; Hernandez et al., 2000), and although there have been some failures to replicate these data, the majority of studies have demonstrated a decrease in 5-HT2A-binding sites in the prefrontal cortex in schizophrenia (reviewed by Harrison and Burnet, 1997; Harrison and Geddes, 1996). Postmortem studies have also demonstrated that the abundance of 5-HT2A receptor is reduced in schizophrenia in brain regions shown to have important neuropathological alterations specific to the disease (reviewed by Harrison, 1999a). Molecular genetic studies have shown associ- ations between single nucleotide polymorphisms (SNPs) in the 5-HT2A receptor gene and schizophrenia (Inayama et al., 1996; Sodhi et al., 1999a; Spurlock et al., 1998; Williams et al., 1996, 1997) and antipsychotic drugs are being increasingly developed to target this receptor (reviewed by Sodhi and Murray, 1997). Because some behavioral responses elicited by acute doses of LSD resemble symptoms of schizophrenia, characterization of gene expression profiles after LSD may provide clues about the etiology of the disease. A small group of genes within the rat prefrontal cortex were found to have altered expression: ania3, ARC, c-fos, I- , krox-20 (erg2), neuron-derived 1(Nor1), and serum gluco- corticoid kinase. These findings lend weight to the developmental hypothesis of schizophrenia because many of these proteins alter synaptic plasticity (Nichols and Sanders-Bush, 2002) and the prefrontal cortex is a region associated with neuro- chemical and structural alteractionsinschizophrenia(Goldman-Rakic,1994;Wein- berger and Berman, 1996), these findings provide support for the developmental hypothesisofschizophrenia(SectionV,E,1).TheresponseofthesegenestoLSDad- ministration was reported to be dynamic with diVering rates of expression changes. ARCwasthe mosthighly expressed gene andwasalsothegenewiththegreatestin- crease in expression after LSD treatment (Nichols et al., 2003). The majority of expression increases were due to activation of 5-HT2A receptor. Furthermore, Pei and colleagues (2000) observed that 5-HT2A receptor antagonists produced cy- toarchitectural changes by modulating the expression of ARC. In another study, LSD administration produced a five- to eightfold increase in fos-like immunoreactiv- ity in the medial prefrontal cortex, anterior cingulate cortex, and central nucleus of amygdala. LSD activation of the medial prefrontal cortex and anterior cingulate cortex was found to be mediated by 5-HT2A receptor, whereas 5-HT2A receptor activation in the amygdala was just a component of the response (Gresch et al., 2002). Although attention has focused on the 5-HT2C receptor subtype, the 5-HT2C receptor subtype may also be relevant, as lysergic acid diethylamide is a high- aYnity agonist at 5-HT2CR (Burris et al., 1991). Furthermore, several genetic studies have detected associations between 5-HT2C receptor cys23ser poly- morphism and long-term hospitalization in schizophrenia (Segman et al., 1997), 148 SODHI AND SANDERS-BUSH and between both 5-HT2A receptor and 5-HT2C receptor genes and hallucinations in dementia (Holmes et al., 1998). Variation in both these genes has also been associated with a therapeutic response to clozapine (Arranz et al., 1995, 1996, 1998; Masellis et al., 1995, 1998; Sodhi et al., 1995, 1999b). The importance of 5-HT2C receptor in CNS disorders is also emphasized by the potentially fatal neurological deficits observed in 5-HT2C receptor knockout mice (Tecott et al., 1995). In addition, interesting changes have been detected in 5-HT2C receptor RNA editing, a posttranscriptional process by which 5-HT2C receptor mRNA undergoes editing to produce several receptor variants, some with pharmacological di Verences (Burns et al., 1997). Elevated expression of 5-HT2C receptor mRNA unedited and partially edited isoforms has been detected in the dorsolateral prefrontal cortex of schizophrenia subjects treated with antido- paminergic drugs (Sodhi et al., 2001). Because the unedited 5-HT2C receptor ex- hibits greater G-protein coupling and constitutive activity (Herrick-Davis et al., 1999; Niswender et al., 1999), reduced RNA editing of the receptor may increase 5-HT2C receptor activity in the dorsolateral prefrontal cortex in schizophrenia. Although the association between 5-HT2c RNA editing and schizophrenia has not been replicated (Dracheva et al., 2003), 5-HT2c RNA editing has been asso- ciated with suicide in several studies (Gurevich et al., 2002; Iwamoto and kato, 2003; Niswender et al., 2001). Improved methodology and increased sample numbers will facilitate investigations into this exciting phenomenon. Examination of other serotonin receptors has revealed reductions in hippo- campal levels of both 5-HT6 receptor mRNA and 5-HT7 receptor mRNA in the prefrontal cortex in schizophrenia (East et al., 2002a, b). Serotonin also binds to the serotonin transporter, which has also been examined in schizophrenia research because it is well established that ligands for SERT alter mood and SERT antagonists are widely used to treat depressive disorders (Section V, D). Some studies have reported a decreased density of transporter-binding sites in the schizophrenic frontal cortex (Joyce et al., 1993; Laruelle et al., 1993; Ohuoha et al., 1993), although others have reported no change (Dean et al., 1995, 1996; Gurevich and Joyce, 1997). In contrast, SERT mRNA levels appear to be in- creased in this brain region. However, the drug histories of the psychiatric case and control groups di Ver in postmortem studies and could confound these data (Hernandez et al., 2000). Furthermore, reduced SERT-binding site densities have been observed in the cingulate cortex in schizophrenia, but are increased in stria- tum ( Joyce et al., 1993). In the hippocampus, no alterations in the density of transporter-binding sites have been found, but a lower aYnity of the transporter for [3H] in schizophrenia has been reported (Dean et al., 1995). Therefore, changes in serotonin receptor and SERT gene expression, altered serotonin levels, and behavioral studies provide support for the hypothesis that a dysfunctional serotonergic system alters synaptic plasticity suViciently to cause developmental changes leading to schizophrenia. SEROTONIN AND BRAIN DEVELOPMENT 149

F. Down’s Syndrome,Mental Retardation, and Serotonin

In contrast with disorders such as schizophrenia and autism, Down’s syndrome is associated with reduced levels of serotonin, detected first in blood (Boullin and O’Brien, 1971; Tu and Zellweger, 1965), then in cerebrospinal fluid (Scott et al., 1983), and finally in studies of postmortem brain (Mann et al., 1985). Neurotransmit- ter deficits in Down’s Syndrome are comparable to those found in Alzheimer’sdis- ease (Godridge et al., 1987). In Down’s syndrome, the expected peak of 5-HT1A receptor density during development is greater than in control postmortem brain, but this decreases below controls after birth (Bar-Peled et al., 1991b). This may ex- plain the findings of initial dendritic overdevelopment, followed by hypertrophy (Becker et al., 1986; Takashima et al., 1994). It has been proposed that similar to autism, the serotonergic system fails to mature and form appropriate connections in the brains of Down’s syndrome children (Okado et al., 2001). In Down’ssyndrome adults there is a region-specific increase in serotonin. These increases occur in the frontal and occipital cortices (Gulesserian et al., 2000), while there are reductions in the thalamus, caudate, cerebellum, and temporal cortex (Seidl et al., 1999). Pharmacological evidence supports a serotonergic role in the development of the symptoms of Down’s syndrome, as serotonergic drugs are especially useful in the treatment of the aggressive symptoms, self-harm, cognitive impairment, and depres- sion exhibited by patients (Gedye, 1990, 1991). Interestingly, the gene for S100 is located on chromosome 21, which is in trisomy in Down’ssyndrome(Ueda et al., 1994a,b). Overexpression of the S100 protein has been detected in postmortem brain and lymphocytes of Down’s syndrome patients, prenatally and in adults (Kato et al., 1990). S100 produced by astrocytes is used as a trophic factor by serotonergic neurons, and a positive correlation occurs among serotonergic neuron density, spine density of pyramidal cell dendrites, 5HT1AR-binding sites, and S100 expression (Ueda et al., 1995, 1996). The level of S100 also correlates with the degree of mental impairment in the patients (Kato et al., 1990). Moreover, a transgenic mouse de- veloped to overexpress S100 possesses neuropathology and symptoms resembling Down’s syndrome (reviewed by Whitaker-Azmitia, 2001). Therefore, as with the other psychiatric disorders discussed, a component of the developmental pathology observed in Down’s syndrome can be attributed to dysfunctions of synaptic plasticity modulated by the serotonergic system.

VI. Conclusions

This review has considered the role of serotonin and serotonergic receptors in the neuroplastic events that create, repair, and degenerate the brain. Research spanning more than five decades has shown that serotonergic projections in the 150 SODHI AND SANDERS-BUSH brain have a widespread distribution and that these projections interact with other neurotransmitter systems, thereby influencing many, if not all, physiological functions. Evidence from biochemical, pharmacological, and clinical studies demonstrates the huge importance of an intact serotonergic system to support brain development and neurogenesis in the maintenance of normal CNS func- tion. Serotonin acts as a growth factor and influences other growth factors during development. The high level of serotonin function during these crucial stages of brain development and the pharmacological evidence for impaired serotonergic function in several disparate brain disorders underpins the importance of under- standing this highly complex system. Increased insight may facilitate real progress in drug development, with the ultimate goal of preventing and even curing these debilitating illnesses.

References

AbiDargham, A., Laruelle, M., Charney, D., and Krystal, J. (1996). Serotonin and Schizophrenia: A review. Drugs Today 32, 171–185. Abi-Dargham, A., Laruelle, M., Lipska, B., Jaskiw, G. E., Wong, D. T., Robertson, D. W., Weinberger, D. R., and Kleinman, J. E. (1993). Serotonin 5-HT3 receptors in schizophrenia: A postmortem study of the amygdala. Brain Res. 616, 53–57. Aghajanian, G. K. (1994). Electrophysiological studies on the actions of hallucinogenic drugs at 5-HT2 receptors in rat brain. NIDA Res. Monogr. 146, 183–202. Aghajanian, G. K., and Marek, G. J. (1999). Serotonin, via 5-HT2A receptors, increases EPSCs in layer V pyramidal cells of prefrontal cortex by an asynchronous mode of glutamate release. Brain Res. 825, 161–171. Alves, S. E., Hoskin, E., Lee, S. J., Brake, W. G., Ferguson, D., Luine, V., Allen, P. B., Greengard, P., and McEwen, B. S. (2002). Serotonin mediates CA1 spine density but is not crucial for ovarian steroid regulation of synaptic plasticity in the adult rat dorsal hippocampus. Synapse 45, 143–151. Arango, V., Underwood, M. D., and Mann, J. J. (1997). Postmortem findings in suicide victims: Implications for in vivo imaging studies. Ann. N. Y. Acad. Sci. 836, 269–287. Archer, T., Ogren, S. O., and Ross, S. B. (1982). Serotonin involvement in aversive conditioning: Reversal of the fear retention deficit by long-term p-chloroamphetamine but not p-chlorophenylalanine. Neurosci. Lett. 34, 75–82. Archer, T., Ogren, S. O., Ross, S. B., and Magnusson, O. (1984). Retention deficits induced by acute p-chloroamphetamine following fear conditioning in the rat. Psychopharmacology (Berl.) 82, 14–19. Arranz, M., Collier, D., Sodhi, M., Ball, D., Roberts, G., Price, J., Sham, P., and Kerwin, R. (1995). Association between clozapine response and allelic variation in 5-HT2A receptor gene. Lancet 346, 281–282. Arranz, M. J., Collier, D. A., Munro, J., Sham, P., Kirov, G., Sodhi, M., Roberts, G., Price, J., and Kerwin, R. W. (1996). Analysis of a structural polymorphism in the 5-HT2A receptor and clinical response to clozapine. Neurosci. Lett. 217, 177–178. Arranz, M. J., Munro, J., Owen, M. J., Spurlock, G., Sham, P. C., Zhao, J., Kirov, G., Collier, D. A., and Kerwin, R. W. (1998). Evidence for association between polymorphisms in the promoter and coding regions of the 5-HT2A receptor gene and response to clozapine. Mol. Psychiat. 3, 61–66. Asberg, M. (1997). Neurotransmitters and suicidal behavior: The evidence from cerebrospinal fluid studies. Ann. N. Y. Acad. Sci. 836, 158–181. SEROTONIN AND BRAIN DEVELOPMENT 151

Audenaert, K., Van Laere, K., Dumont, F., Slegers, G., Mertens, J., van Heeringen, C., and Dierckx, R. A. (2001). Decreased frontal serotonin 5-HT 2a receptor binding index in deliberate self-harm patients. Eur. J. Nuclear Med. 28, 175–182. Aylward, E. H., Minshew, N. J., Goldstein, G., Honeycutt, N. A., Augustine, A. M., Yates, K. O., Barta, P. E., and Pearlson, G. D. (1999). MRI volumes of amygdala and hippocampus in non- mentally retarded autistic adolescents and adults. Neurology 53, 2145–2150. Ayres, A. J., and Tickle, L. S. (1980). Hyper-responsivity to touch and vestibular stimuli as a predictor of positive response to sensory integration procedures by autistic children. Am. J. Occup. Ther. 34, 375–381. Azmitia, E. C. (1999). Serotonin neurons, neuroplasticity, and homeostasis of neural tissue. Neuropsychopharmacology 21, 33S–45S. Azmitia, E. C. (2001). Modern views on an ancient chemical: Serotonin effects on cell proliferation, maturation, and apoptosis. Brain Res. Bull. 56, 413–424. Azmitia, E. C., and Gannon, P. J. (1986). The primate serotonergic system: A review of human and animal studies and a report on Macaca fascicularis. Adv. Neurol. 43, 407–468. Azmitia, E. C., Rubinstein, V. J., Strafaci, J. A., Rios, J. C., and Whitaker-Azmitia, P. M. (1995). 5-HT1A agonist and dexamethasone reversal of para-chloroamphetamine induced loss of MAP- 2 and synaptophysin immunoreactivity in adult rat brain. Brain Res. 677, 181–192. Bakish, D., Lapierre, Y. D., Weinstein, R., Klein, J., Wiens, A., Jones, B., Horn, E., Browne, M., Bourget, D., Blanchard, A., et al. (1993). Ritanserin, , and placebo in the treatment of dysthymic disorder. J. Clin. Psychopharmacol. 13, 409–414. Banasr, M., Hery, M., Brezun, J. M., and Daszuta, A. (2001). Serotonin mediates oestrogen stimulation of cell proliferation in the adult dentate gyrus. Eur. J. Neurosci. 14, 1417–1424. Barnes, N. M., and Sharp, T. (1999). A review of central 5-HT receptors and their function. Neuropharmacology 38, 1083–1152. Bar-Peled, O., Gross-Isseroff, R., Ben-Hur, H., Hoskins, I., Groner, Y., and Biegon, A. (1991a). Fetal human brain exhibits a prenatal peak in the density of serotonin 5-HT1A receptors. Neurosci. Lett. 127, 173–176. Bar-Peled, O., Israeli, M., Ben-Hur, H., Hoskins, I., Groner, Y., and Biegon, A. (1991b). Developmental pattern of muscarinic receptors in normal and Down’s syndrome fetal brain: An autoradiographic study. Neurosci. Lett. 133, 154–158. Bayliss, D. A., Viana, F., Talley, E. M., and Berger, A. J. (1997). Neuromodulation of hypoglossal motoneurons: Cellular and developmental mechanisms. Respir. Physiol. 110, 139–150. Beaumont, V., and Zucker, R. S. (2000). Enhancement of synaptic transmission by cyclic AMP modulation of presynaptic lh channels. Nature Neurosci. 3, 133–141. Becker, L. E., Armstrong, D. L., and Chan, F. (1986). Dendritic atrophy in children with Down’s syndrome. Ann. Neurol. 20, 520–526. Benes, F. M., Taylor, J. B., and Cunningham, M. C. (2000). Convergence and plasticity of systems in the medial prefrontal cortex during the postnatal period: Implications for the development of psychopathology. Cereb. Cortex 10, 1014–1027. Bennett-Clarke, C. A., Hankin, M. H., Leslie, M. J., Chiaia, N. L., and Rhoades, R. W. (1994). Patterning of the neocortical projections from the raphe nuclei in perinatal rats: Investigation of potential organizational mechanisms. J. Comp. Neurol. 348, 277–290. Bersani, G., Pozzi, F., Marini, S., Grispini, A., Pasini, A., and Ciani, N. (1991). 5-HT2 receptor antagonism in dysthymic disorder: A double-blind placebo-controlled study with ritanserin. Acta Psychiatr. Scand. 83, 244–248. Betancur, C., Corbex, M., Spielewoy, C., Philippe, A., Laplanche, J. L., Launay, J. M., Gillberg, C., Mouren-Simeoni, M. C., Hamon, M., Giros, B., Nosten-Bertrand, M., and Leboyer, M. (2002). Serotonin transporter gene polymorphisms and hyperserotonemia in autistic disorder. Mol. Psychiat. 7, 67–71. 152 SODHI AND SANDERS-BUSH

Biegon, A., Essar, N., Israeli, M., Elizur, A., Bruch, S., and Bar-Nathan, A. A. (1990). Serotonin 5-HT2 receptor binding on blood platelets as a state dependent marker in major affective disorder. Psychopharmacology (Berl.) 102, 73–75. Blatt, G. J., Chen, J. C., Rosene, D. L., Volicer, L., and Galler, J. R. (1994). Prenatal protein malnutrition effects on the serotonergic system in the hippocampal formation: An immunocytochemical, ligand binding, and neurochemical study. Brain Res. Bull. 34, 507–518. Bliss, T. V., and Gardner-Medwin, A. R. (1973). Long-lasting potentiation of synaptic transmission in the dentate area of the unanaestetized rabbit following stimulation of the perforant path. J. Physiol. 232, 357–374. Blue, M. E., Erzurumlu, R. S., and Jhaveri, S. (1991). A comparison of pattern formation by thalamocortical and serotonergic afferents in the rat barrel field cortex. Cereb. Cortex 1, 380–389. Boswell, C. A., Majno, G., Joris, I., and Ostrom, K. A. (1992). Acute endothelial cell contraction in vitro: A comparison with vascular smooth muscle cells and fibroblasts. Microvasc. Res. 43, 178–191. Boullin,D.J.,Coleman,M.,andO’Brien, R. A. (1970). Abnormalities in platelet 5-hydroxytryptamine efflux in patients with infantile autism. Nature 226, 371–372. Boullin, D. J., Coleman, M., O’Brien, R. A., and Rimland, B. (1971). Laboratory predictions of infantile autism based on 5-hydroxytryptamine efflux from blood platelets and their correlation with the Rimland E-2 score. J. Autism Child. Schizophr. 1, 63–71. Boullin, D. J., and O’Brien, R. A. (1971). Abnormalities of 5-hydroxytryptamine uptake and binding by blood platelets from children with Down’s syndrome. J. Physiol. 212, 287–297. Bradley, P. B., Feniuk, W., Fozard, J. R., Humphrey, P. P., Middlemiss, D. N., Mylecharane, E. J., Richardson, B. P., and Saxena, P. R. (1986). Proposals for the classification and nomenclature of functional receptors for 5-hydroxytryptamine. Neuropharmacology 25, 563–576. Bray, G. A. (2000). A concise review on the therapeutics of obesity. Nutrition 16, 953–960. Breier, A. (1989). A. E. Bennett award paper. Experimental approaches to human stress research: Assessment of neurobiological mechanisms of stress in volunteers and psychiatric patients. Biol. Psychiat. 26, 438–462. Breier, A. (1995). Serotonin, schizophrenia and antipsychotic drug-action. Schizophr. Res. 14, 187–202. Brezun, J. M., and Daszuta, A. (1999). Serotonin depletion in the adult rat produces differential changes in highly polysialylated form of neural cell adhesion molecule and tenascin-C immunoreactivity. J. Neurosci. Res. 55, 54–70. Brismar, T. (1995). Physiology of transformed glial cells. Glia 15, 231–243. Broydell, M., Mazzuca, D. M., Abidi, F., Kudo, P. A., and Lo, T. C. (1997). Involvement of the GLUT 3 transporter in myogenic regulation. Biochem. Mol. Biol. Int. 43, 847–866. Buitelaar, J. K., and Willemsen-Swinkels, S. H. (2000). Autism: Current theories regarding its pathogenesis and implications for rational pharmacotherapy. Paediatr. Drugs. 2, 67–81. Bulik, C. M., Prescott, C. A., and Kendler, K. S. (2001). Features of childhood sexual abuse and the development of psychiatric and substance use disorders. Br. J. Psychiat. 179, 444–449. Buonomano, D. V., and Merzenich, M. M. (1995). Temporal information transformed into a spatial code by a neural network with realistic properties. Science 267, 1028–1030. Burnet, P. W., Eastwood, S. L., and Harrison, P. J. (1996). 5-HT1A and 5-HT2A receptor mRNAs and binding site densities are differentially altered in schizophrenia. Neuropsychopharmacology 15, 442–455. Burnet, P. W., Sharp, T., LeCorre, S. M., and Harrison, P. J. (1999). Expression of 5-HT receptors and the 5-HT transporter in rat brain after electroconvulsive shock. Neurosci. Lett. 277, 79–82. Burns, C. M., Chu, H., Rueter, S. M., Hutchinson, L. K., Canton, H., Sanders-Bush, E., and Emeson, R. B. (1997). Regulation of serotonin-2C receptor G-protein coupling by RNA editing. Nature 387, 303–308. SEROTONIN AND BRAIN DEVELOPMENT 153

Cai, X., Gu, Z., Zhong, P., Ren, Y., and Yan, Z. (2002). Serotonin 5-HT1A receptors regulate AMPA receptor channels through inhibiting Ca2þ/calmodulin-dependent kinase II in prefrontal cortical pyramidal neurons. J. Biol. Chem. 277, 36553–36562. Calabrese, J. R., Kimmel, S. E., Woyshville, M. J., Rapport, D. J., Faust, C. J., Thompson, P. A., and Meltzer, H. Y. (1996). Clozapine for treatment-refractory mania. Am. J. Psychiat. 153, 759–764. Calabrese, J. R., Meltzer, H. Y., and Markovitz, P. J. (1991). Clozapine prophylaxis in rapid cycling bipolar disorder. J. Clin. Psychopharmacol. 11, 396–397. Cannon, M., Jones, P., Gilvarry, C., Rifkin, L., McKenzie, K., Foerster, A., and Murray, R. M. (1997). Premorbid social functioning in schizophrenia and bipolar disorder: Similarities and differences. Am. J. Psychiat. 154, 1544–1550. Cases, O., Vitalis, T., Seif, I., De Maeyer, E., Sotelo, C., and Gaspar, P. (1996). Lack of barrels in the somatosensory cortex of monoamine oxidase A-deficient mice: role of a serotonin excess during the critical period. Neuron 16, 297–307. Ceulemans, D. L., Hoppenbrouwers, M. L., Gelders, Y. G., and Reyntjens, A. J. (1985). The influence of ritanserin, a serotonin antagonist, in anxiety disorders: A double-blind placebo- controlled study versus . Pharmacopsychiatry 18, 303–305. Chadwick, D., Jenner, P., and Reynolds, E. H. (1977). Serotonin metabolism in human epilepsy: Influence of anticonvulsant drugs. Epilepsia 18, 288. Charney, D. S., Heninger, G. R., Reinhard, J. F., Jr., Sternberg, D. E., and Hafstead, K. M. (1982). The effect of intravenous L-tryptophan on prolactin and growth hormone and mood in healthy subjects. Psychopharmacology (Berl.) 77, 217–222. Charney, D. S., Menkes, D. B., and Heninger, G. R. (1981). Receptor sensitivity and the mechanism of action of antidepressant treatment: Implications for the etiology and therapy of depression. Arch. Gen. Psychiat. 38, 1160–1180. Chen, L., Hamaguchi, K., Ogawa, M., Hamada, S., and Okada, N. (1994). PCPA reduces both monoaminergic afferents and nonmonoaminergic synapses in the cerebral cortex. Neurosci. Res. 19, 111–115. Cheng, L. L., Wang, S. J., and Gean, P. W. (1998). Serotonin depresses excitatory synaptic transmission and depolarization-evoked Ca2þ influx in rat basolateral amygdala via 5-HT1A receptors. Eur. J. Neurosci. 10, 2163–2172. Chiaia, N. L., Fish, S. E., Bauer, W. R., Figley, B. A., Eck, M., Bennett-Clarke, C. A., and Rhoades, R. W. (1994). Effects of postnatal blockade of cortical activity with tetrodotoxin upon the development and plasticity of vibrissa-related patterns in the somatosensory cortex of hamsters. Somatosens. Motil. Res. 11, 219–228. Choi, D. S., Kellermann, O., Richard, S., Colas, J. F., Bolanos-Jimenez, F., Tournois, C., Launay, J. M., and Maroteaux, L. (1998). Mouse 5-HT2B receptor-mediated serotonin trophic functions. Ann. N. Y. Acad. Sci. 861, 67–73. Choi, D. S., Ward, S. J., Messaddeq, N., Launay, J. M., and Maroteaux, L. (1997). 5-HT2B receptor- mediated serotonin morphogenetic functions in mouse cranial neural crest and myocardiac cells. Development 124, 1745–1755. Chua, S. E., and Murray, R. M. (1996). The neurodevelopmental theory of schizophrenia: Evidence concerning structure and neuropsychology. Ann. Med. 28, 547–555. Chubakov, A. R., Gromova, E. A., Konovalov, G. V., Sarkisova, E. F., and Chumasov, E. I. (1986). The effects of serotonin on the morpho-functional development of rat cerebral neocortex in tissue culture. Brain Res. 369, 285–297. Chubakov, A. R., Tsyganova, V. G., and Sarkisova, E. F. (1993). The stimulating influence of the raphe nuclei on the morphofunctional development of the hippocampus during their combined cultivation. Neurosci. Behav. Physiol. 23, 271–276. Chugani, D. C., and Chugani, H. T. (2003). Does serotonin have trophic effects in temporal lobe epilepsy? Neurology 60, 736–737. 154 SODHI AND SANDERS-BUSH

Chugani, D. C., Muzik, O., Rothermel, R., Behen, M., Chakraborty, P., Mangner, T., da Silva, E. A., and Chugani, H. T. (1997). Altered serotonin synthesis in the dentatothalamocortical pathway in autistic boys. Ann. Neurol. 42, 666–669. Claustre, Y., Rouquier, L., and Scatton, B. (1988). Pharmacological characterization of serotonin- stimulated phosphoinositide turnover in brain regions of the immature rat. J. Pharmacol. Exp. Ther. 244, 1051–1056. Cleare, A. J., and Bond, A. J. (1995). The effect of tryptophan depletion and enhancement on subjective and behavioural aggression in normal male subjects. Psychopharmacology (Berl.) 118, 72–81. Conn, P. J., and Sanders-Bush, E. (1984). Selective 5HT-2 antagonists inhibit serotonin stimulated phosphatidylinositol metabolism in cerebral cortex. Neuropharmacology 23, 993–996. Conn, P. J., and Sanders-Bush, E. (1986a). Biochemical characterization of serotonin stimulated phosphoinositide turnover. Life Sci. 38, 663–669. Conn, P. J., and Sanders-Bush, E. (1986b). Regulation of serotonin-stimulated phosphoinositide hydrolysis: Relation to the serotonin 5-HT-2 binding site. J. Neurosci. 6, 3669–3675. Conn, P. J., and Sanders-Bush, E. (1987). Central serotonin receptors: Effector systems, physiological roles and regulation. Psychopharmacology (Berl.) 92, 267–277. Conn, P. J., Sanders-Bush, E., Hoffman, B. J., and Hartig, P. R. (1986). A unique serotonin receptor in choroid plexus is linked to phosphatidylinositol turnover. Proc. Natl. Acad. Sci. USA 83, 4086–4088. Cook, E. H., and Leventhal, B. L. (1996). The serotonin system in autism. Curr. Opin. Pediatr. 8, 348–354. Coppen, A., Shaw, D. M., Herzberg, B., and Maggs, R. (1967). Tryptophan in the treatment of depression. Lancet 2, 1178–1180. Coppen, A. J., and Doogan, D. P. (1988). Serotonin and its place in the pathogenesis of depression. J. Clin. Psychiat. 49(Suppl.), 4–11. Craddock, N., and Jones, I. (1999). Genetics of bipolar disorder. J. Med. Genet. 36, 585–594. Craddock, N., and Jones, I. (2001). Molecular genetics of bipolar disorder. Br. J. Psychiat. Suppl. 41, s128–s133. Crossland, N. (2001). Expression of the 5-HT2C receptor gene in schizophrenia. G6R DPhil Oxford, 51–164–111. Crow, T. J., Ferrier, I. N., and Johnstone, E. C. (1986). The two-syndrome concept and neuroendocrinology of schizophrenia. Psychiat. Clin. North Am. 9, 99–113. Cynader, M., Timney, B. N., and Mitchell, D. E. (1980). Period of susceptibility of kitten visual cortex to the effects of monocular deprivation extends beyond six months of age. Brain Res. 191, 545–550. Dailey, J. W., Yan, Q. S., Mishra, P. K., Burger, R. L., and Jobe, P. C. (1992). Effects of on convulsions and on brain-serotonin as detected by microdialysis in genetically epilepsy-prone rats. J. Pharmacol. Exp. Ther. 260, 533–540. D’Amato, R. J., Blue, M. E., Largent, B. L., Lynch, D. R., Ledbetter, D. J., Molliver, M. E., and Snyder, S. H. (1987). Ontogeny of the serotonergic projection to rat neocortex: Transient expression of a dense innervation to primary sensory areas. Proc. Natl. Acad. Sci. USA 84, 4322–4326. Daval, G., Verge, D., Becerril, A., Gozlan, H., Spampinato, U., and Hamon, M. (1987). Transient expression of 5-HT1A receptor binding sites in some areas of the rat CNS during postnatal development. Int. J. Dev. Neurosci. 5, 171–180. Davis, E., Fennoy, I., Laraque, D., Kanem, N., Brown, G., and Mitchell, J. (1992). Autism and developmental abnormalities in children with perinatal cocaine exposure. J. Natl. Med. Assoc. 84, 315–319. de Boer, T. (1996). The pharmacologic profile of mirtazapine. J. Clin. Psychiat. 57(Suppl. 4), 19–25. SEROTONIN AND BRAIN DEVELOPMENT 155 de Boer, T. H., Nefkens, F., van Helvoirt, A., and van Delft, A. M. (1996). Differences in modulation of noradrenergic and serotonergic transmission by the alpha-2 adrenoceptor antagonists, mirtazapine, mianserin and . J. Pharmacol. Exp. Ther. 277, 852–860. Deahl, M., and Trimble, M. (1991). Serotonin reuptake inhibitors, epilepsy and myoclonus. Br. J. Psychiat. 159, 433–435. Dean, B., and Hayes, W. (1996). Decreased frontal cortical serotonin(2A) receptors in schizophrenia. Schizophr. Res. 21, 133–139. Dean, B., Hayes, W., Opeskin, K., Naylor, L., Pavey, G., Hill, C., Keks, N., and Copolov, D. L. (1996). Serotonin2 receptors and the serotonin transporter in the schizophrenic brain. Behav. Brain Res. 73, 169–175. Dean, B., Opeskin, K., Pavey, G., Naylor, L., Hill, C., Keks, N., and Copolov, D. L. (1995). [3H]Paroxetine binding is altered in the hippocampus but not the frontal cortex or caudate nucleus from subjects with schizophrenia. J. Neurochem. 64, 1197–1202. DeFelipe, J., Hendry, S. H., Hashikawa, T., and Jones, E. G. (1991). Synaptic relationships of serotonin-immunoreactive terminal baskets on GABA neurons in the cat auditory cortex. Cereb. Cortex 1, 117–133. Del Angel-Meza, A. R., Ramirez-Cortes, L., Adame-Gonzalez, I. G., Gonzalez Burgos, I., and Beas- Zarate, C. (2002). Cerebral GABA release and GAD activity in protein- and tryptophan- restricted rats during development. Int. J. Dev. Neurosci. 20, 47–54. Delgado, P. L., Charney, D. S., Price, L. H., Aghajanian, G. K., Landis, H., and Heninger, G. R. (1990). Serotonin function and the mechanism of antidepressant action: Reversal of antidepressant-induced remission by rapid depletion of plasma tryptophan. Arch. Gen. Psychiat. 47, 411–418. Dinan, T. G. (1996). Noradrenergic and serotonergic abnormalities in depression: Stress-induced dysfunction? J. Clin. Psychiat. 57(Suppl. 4), 14–18. Dinerman, J. L., Dawson, T. M., Schell, M. J., Snowman, A., and Snyder, S. H. (1994). Endothelial nitric oxide synthase localized to hippocampal pyramidal cells: Implications for synaptic plasticity. Proc. Natl. Acad. Sci. USA 91, 4214–4218. Ding, D., Toth, M., Zhou, Y., Parks, C., Hoffman, B. J., and Shenk, T. (1993). Glial cell-specific expression of the serotonin 2 receptor gene: Selective reactivation of a repressed promoter. Brain Res. Mol. Brain Res. 20, 181–191. Dixon, D., and Atwood, H. L. (1989). Conjoint action of phosphatidylinositol and adenylate cyclase systems in serotonin-induced facilitation at the crayfish neuromuscular junction. J. Neurophysiol. 62, 1251–1259. Doraiswamy, P. M. (2003). The role of the N-methyl-d-aspartate receptor in Alzheimer’s disease: Therapeutic potential. Curr. Neurol. Neurosci. Rep. 3, 373–378. Dracheva, S., Elhakem, S. L. et al. (2003). RNA editing and alternative splicing of human serotonin 2C receptor in schizophrenia. J. Neurochem. 87, 1402–1412. Drevets, W. C., Frank, E., Price, J. C., Kupfer, D. J., Holt, D., Greer, P. J., Huang, Y., Gautier, C., and Mathis, C. (1999). PET imaging of serotonin 1A receptor binding in depression. Biol. Psychiat. 46, 1375–1387. Dyck, R. H., and Cynader, M. S. (1993a). Autoradiographic localization of serotonin receptor subtypes in cat visual cortex: Transient regional, laminar, and columnar distributions during postnatal development. J. Neurosci. 13, 4316–4338. Dyck, R. H., and Cynader, M. S. (1993b). An interdigitated columnar mosaic of cytochrome oxidase, zinc, and neurotransmitter-related molecules in cat and monkey visual cortex. Proc. Natl. Acad. Sci. USA 90, 9066–9069. East, S. Z., Burnet, P. W., Kerwin, R. W., and Harrison, P. J. (2002a). An RT-PCR study of 5-HT(6) and 5-HT(7) receptor mRNAs in the hippocampal formation and prefrontal cortex in schizophrenia. Schizophr. Res. 57, 15–26. 156 SODHI AND SANDERS-BUSH

East, S. Z., Burnet, P. W., Leslie, R. A., Roberts, J. C., and Harrison, P. J. (2002b). 5-HT6 receptor binding sites in schizophrenia and following antipsychotic drug administration: Autoradio- graphic studies with [125I]SB-258585. Synapse 45, 191–199. Eastwood, S. L., Burnet, P. W., Gittins, R., Baker, K., and Harrison, P. J. (2001). Expression of serotonin 5-HT(2A) receptors in the human cerebellum and alterations in schizophrenia. Synapse 42, 104–114. Eddahibi, S., Fabre, V., Boni, C., Martres, M. P., Raffestin, B., Hamon, M., and Adnot, S. (1999). Induction of serotonin transporter by hypoxia in pulmonary vascular smooth muscle cells: Relationship with the mitogenic action of serotonin. Circ. Res. 84, 329–336. Edwards, H. E., Burnham, W. M., and MacLusky, N. J. (1999). Testosterone and its metabolites affect afterdischarge thresholds and the development of amygdala kindled seizures. Brain Res. 838, 151–157. Eison, A. S., Yocca, F. D., and Gianutsos, G. (1991). Effect of chronic administration of antidepressant drugs on 5-HT2-mediated behavior in the rat following noradrenergic or serotonergic denervation. J. Neural Transm. Gen. Sect. 84, 19–32. Eison, M. S. (1990). Serotonin: A common neurobiologic substrate in anxiety and depression. J. Clin. Psychopharmacol. 10, 26S–30S. Ellenbogen, M. A., Young, S. N., Dean, P., Palmour, R. M., and Benkelfat, C. (1996). Mood response to acute tryptophan depletion in healthy volunteers: Sex differences and temporal stability. Neuropsychopharmacology 15, 465–474. Enyeart, J. (1981). Cyclic AMP, 5-HT, and the modulation of transmitter release at the crayfish neuromuscular junction. J. Neurobiol. 12, 505–513. Esterle, T. M., and Sanders-Bush, E. (1992). Serotonin agonists increase transferrin levels via activation of 5-HT1C receptors in choroid plexus epithelium. J. Neurosci. 12, 4775–4782. Faber, K. M., and Haring, J. H. (1999). Synaptogenesis in the postnatal rat fascia dentata is influenced by 5-HT1a receptor activation. Brain Res. Dev. Brain Res. 114, 245–252. Farr, S. A., Uezu, K., Creonte, T. A., Flood, J. F., and Morley, J. E. (2000). Modulation of memory processing in the cingulate cortex of mice. Pharmacol. Biochem. Behav. 65, 363–368. Fatemi, S. H., Realmuto, G. M., Khan, L., and Thuras, P. (1998). Fluoxetine in treatment of adolescent patients with autism: A longitudinal open trial. J. Autism Dev. Disord. 28, 303–307. Feighner, J., Targum, S. D., Bennett, M. E., Roberts, D. L., Kensler, T. T., D’Amico, M. F., and Hardy, S. A. (1998). A double-blind, placebo-controlled trial of nefazodone in the treatment of patients hospitalized for major depression. J. Clin. Psychiat. 59, 246–253. Feighner, J. P., and Boyer, W. F. (1989). Serotonin-1A : An overview. Psychopathology 22(Suppl. 1), 21–26. Feinberg, I. (1982). Schizophrenia: Caused by a fault in programmed synaptic elimination during adolescence? J. Psychiat. Res. 17, 319–334. Fendt, M., and Fanselow, M. S. (1999). The neuroanatomical and neurochemical basis of conditioned fear. Neurosci. Biobehav. Rev. 23, 743–760. Fernandes, C., McKittrick, C. R., File, S. E., and McEwen, B. S. (1997). Decreased 5-HT1A and increased 5-HT2A receptor binding after chronic corticosterone associated with a behavioural indication of depression but not anxiety. Psychoneuroendocrinology 22, 477–491. Flugge, G. (1995). Dynamics of central nervous 5-HT1A-receptors under psychosocial stress. J. Neurosci. 15, 7132–7140. Flugge, G., Ahrens, O., and Fuchs, E. (1997). Monoamine receptors in the prefrontal cortex of Tupaja belangeri during chronic psychosocial stress. Cell Tissue Res. 288, 1–10. Flugge, G., Kramer, M., Rensing, S., and Fuchs, E. (1998). 5HT1A-receptors and behaviour under chronic stress: Selective counteraction by testosterone. Eur. J. Neurosci. 10, 2685–2693. Fone, K. C., Shalders, K., Fox, Z. D., Arthur, R., and Marsden, C. A. (1996). Increased 5-HT2C receptor responsiveness occurs on rearing rats in social isolation. Psychopharmacology (Berl.) 123, 346–352. SEROTONIN AND BRAIN DEVELOPMENT 157

Foy, M. R., Xu, J., Xie, X., Brinton, R. D., Thompson, R. F., and Berger, T. W. (1999). 17beta- estradiol enhances NMDA receptor-mediated EPSPs and long-term potentiation. J. Neurophysiol. 81, 925–929. Frankland, P. W., O’Brien, C., Ohno, M., Kirkwood, A., and Silva, A. J. (2001). Alpha-CaMKII- dependent plasticity in the cortex is required for permanent memory. Nature 411, 309–313. Freedman, B., and Chapman, L. J. (1973). Early subjective experience in schizophrenic episodes. J. Abnorm. Psychol. 82, 46–54. Fries, P., Fernandez, G., and Jensen, O. (2003). When neurons from memories. Trends Neurosci. 26, 123–124. Fromm, G. H., Glass, J. D., and Chattha, A. S. (1977). Role of serotonin in petit mal epilepsy. Electroen. Clin. Neuro. 43, 297. Frye, M. A., Ketter, T. A., Altshuler, L. L., Denicoff, K., Dunn, R. T., Kimbrell, T. A., Cora- Locatelli, G., and Post, R. M. (1998). Clozapine in bipolar disorder: Treatment implications for other atypical antipsychotics. J. Affect. Disord. 48, 91–104. Fuster, J. M. (1985). The prefrontal cortex, mediator of cross-temporal contingencies. Hum. Neurobiol. 4, 169–179. Fuster, J. M. (1991). The prefrontal cortex and its relation to behavior. Prog. Brain Res. 87, 201–211. Gaddum, J. H., and Picarelli, Z. P. (1957). Two kinds of receptor. Br. J. Pharmacol. 120, 134–139. Garey, L. J., Ong, W. Y., Patel, T. S., Kanani, M., Davis, A., Mortimer, A. M., Barnes, T. R., and Hirsch, S. R. (1998). Reduced dendritic spine density on cerebral cortical pyramidal neurons in schizophrenia. J. Neurol. Neurosurg. Psychiat. 65, 446–453. Gedye, A. (1990). Dietary increase in serotonin reduces self-injurious behaviour in a Down’s syndrome adult. J. Ment. Defic. Res. 34(Pt. 2), 195–203. Gedye, A. (1991). Serotonergic treatment for aggression in a Down’s syndrome adult showing signs of Alzheimer’s disease. J. Ment. Defic. Res. 35(Pt. 3), 247–258. Glantz, L. A., and Lewis, D. A. (2000). Decreased dendritic spine density on prefrontal cortical pyramidal neurons in schizophrenia. Arch. Gen. Psychiat. 57, 65–73. Glazewski, S., Giese, K. P., Silva, A., and Fox, K. (2000). The role of alpha-CaMKII autophosphorylation in neocortical experience-dependent plasticity. Nature Neurosci. 3, 911–918. Glennon, R. A., Titeler, M., and McKenney, J. D. (1984). Evidence for 5-HT2 involvement in the mechanism of action of hallucinogenic agents. Life Sci. 35, 2505–2511. Godridge, H., Reynolds, G. P., Czudek, C., Calcutt, N. A., and Benton, M. (1987). Alzheimer-like neurotransmitter deficits in adult Down’s syndrome brain tissue. J. Neurol. Neurosurg. Psychiat. 50, 775–778. Goldman-Rakic, P. S. (1987). Development of cortical circuitry and cognitive function. Child Dev. 58, 601–622. Goldman-Rakic, P. S. (1994). Working memory dysfunction in schizophrenia. J. Neuropsychiat. Clin. Neurosci. 6, 348–357. Goldman-Rakic, P. S. (1995). Architecture of the prefrontal cortex and the central executive. Ann. N. Y. Acad. Sci. 769, 71–83. Gonzalez-Burgos, I., del Angel-Meza, A. R., Barajas-Lopez, G., and Feria-Velasco, A. (1996). Tryptophan restriction causes long-term plastic changes in corticofrontal pyramidal neurons. Int. J. Dev. Neurosci. 14, 673–679. Good, M., Day, M., and Muir, J. L. (1999). Cyclical changes in endogenous levels of oestrogen modulate the induction of LTD and LTP in the hippocampal CA1 region. Eur. J. Neurosci. 11, 4476–4480. Goudreau, J. L., Manzanares, J., Lookingland, K. J., and Moore, K. E. (1993). 5HT2 receptors mediate the effects of stress on the activity of periventricular hypophysial dopaminergic neurons and the secretion of alpha-melanocyte-stimulating hormone. J. Pharmacol. Exp. Ther. 265, 303–307. 158 SODHI AND SANDERS-BUSH

Gould, E., Tanapat, P., McEwen, B. S., Flugge, G., and Fuchs, E. (1998). Proliferation of granule cell precursors in the dentate gyrus of adult monkeys is diminished by stress. Proc. Natl. Acad. Sci. USA 95, 3168–3171. Gresch, P. J., Strickland, L. V., and Sanders-Bush, E. (2002). Lysergic acid diethylamide-induced Fos expression in rat brain: Role of serotonin-2A receptors. Neuroscience 114, 707–713. Griebel, G., Misslin, R., Pawlowski, M., and Vogel, E. (1991). m-Chlorophenylpiperazine enhances neophobic and anxious behaviour in mice. Neuroreport. 2, 627–629. Gross, C., Zhuang, X., Stark, K., Ramboz, S., Oosting, R., Kirby, L., Santarelli, L., Beck, S., and Hen, R. (2002). Serotonin1A receptor acts during development to establish normal anxiety-like behaviour in the adult. Nature 416, 396–400. Gudelsky, G. A., and Yamamoto, B. K. (2003). Neuropharmacology and neurotoxicity of 3,4-methylenedioxymethamphetamine. Methods Mol. Med. 79, 55–73. Gulesserian, T., Engidawork, E., Cairns, N., and Lubec, G. (2000). Increased protein levels of serotonin transporter in frontal cortex of patients with Down syndrome. Neurosci. Lett. 296, 53–57. Gurevich, E. V., and Joyce, J. N. (1997). Alterations in the cortical serotonergic system in schizophrenia: A postmortem study. Biol. Psychiat. 42, 529–545. Gurevich, I., Tamir, H. et al. (2002). Altered editing of serotonin 2C receptor pre-mRNA in the prefrontal cortex of depressed suicide victims. Neuron 34, 349–356. GuyGrand, B. (1995). Clinical studies with : From past to future. Obes. Res. 3, S491–S496. Hafner, H. (2003). Gender differences in schizophrenia. Psychoneuroendocrinology 28(Suppl. 2), 17–54. Halford, J. C. G. (2001). Pharmacology of appetite suppression: Implication for the treatment of obesity. Curr. Drug Targets. 2, 353–370. Hamon, M., and Bourgoin, S. (1979). Ontogenesis of tryptophan transport in the rat brain. J. Neural Transm. Suppl. 93–105. Hanson, D. R., Gottesman, I. I., and Heston, L. L. (1976). Some possible childhood indicators of adult schizophrenia inferred from children of schizophrenics. Br. J. Psychiat. 129, 142–154. Harrison, P. J. (1995). On the neuropathology of schizophrenia and its dementia: Neurodevelop- mental, neurodegenerative, or both? Neurodegeneration 4, 1–12. Harrison, P. J. (1999a). Neurochemical alterations in schizophrenia affecting the putative receptor targets of atypical antipsychotics: Focus on dopamine (D1, D3, D4) and 5-HT2a receptors. Br. J. Psychiat. Suppl. 12–22. Harrison, P. J. (1999b). The neuropathology of schizophrenia: A critical review of the data and their interpretation. Brain 122(Pt. 4), 593–624. Harrison, P. J., and Burnet, P. W. (1997). The 5-HT2A (serotonin2A) receptor gene in the aetiology, pathophysiology and pharmacotherapy of schizophrenia. J. Psychopharmacol. 11, 18–20. Harrison, P. J., and Geddes, J. R. (1996). Schizophrenia and the 5-HT2A receptor gene. Lancet 347, 1274. Hashimoto, S., Inoue, T., and Koyama, T. (1996). Serotonin reuptake inhibitors reduce conditioned fear stress-induced freezing behavior in rats. Psychopharmacology (Berl.) 123, 182–186. Hashimoto, S., Inoue, T., and Koyama, T. (1997). Effects of the co-administration of 5-HT1A receptor antagonists with an SSRI in conditioned fear stress-induced freezing behavior. Pharmacol. Biochem. Behav. 58, 471–475. Hashimoto, T., Kitamura, N., Kajimoto, Y., Shirai, Y., Shirakawa, O., Mita, T., Nishino, N., and Tanaka, C. (1993). Differential changes in serotonin 5-HT1A and 5-HT2 receptor binding in patients with chronic schizophrenia. Psychopharmacology (Berl.) 112, S35–S39. Hashimoto, T., Nishino, N., Nakai, H., and Tanaka, C. (1991). Increase in serotonin 5-HT1A receptors in prefrontal and temporal cortices of brains from patients with chronic schizophrenia. Life Sci. 48, 355–363. SEROTONIN AND BRAIN DEVELOPMENT 159

Heal, D. J., Aspley, S., Prow, M. R., Jackson, H. C., Martin, K. F., and Cheetham, S. C. (1998). : A novel anti-obesity drug. A review of the pharmacological evidence to differentiate it from d- and d-. Int. J. Obesity 22, S18–S28. Hedner, J., Lundell, K. H., Breese, G. R., Mueller, R. A., and Hedner, T. (1986). Developmental variations in CSF monoamine metabolites during childhood. Biol. Neonate. 49, 190–197. Heisler, L. K., Chu, H. M., Brennan, T. J., Danao, J. A., Bajwa, P., Parsons, L. H., and Tecott, L. H. (1998a). Elevated anxiety and antidepressant-like responses in serotonin 5-HT1A receptor mutant mice. Proc. Natl. Acad. Sci. USA 95, 15049–15054. Heisler, L. K., Chu, H. M., and Tecott, L. H. (1998b). Epilepsy and obesity in serotonin 5-HT2C receptor mutant mice. Ann. N. Y. Acad. Sci. 861, 74–78. Hendricks, T., Francis, N., Fyodorov, D., and Deneris, E. S. (1999). The ETS domain factor Pet-1 is an early and precise marker of central serotonin neurons and interacts with a conserved element in serotonergic genes. J. Neurosci. 19, 10348–10356. Herdegen, T., and Leah, J. D. (1998). Inducible and constitutive transcription factors in the mammalian nervous system: Control of gene expression by Jun, Fos and Krox, and CREB/ATF proteins. Brain Res. Brain Res. Rev. 28, 370–490. Hernandez, I., and Sokolov, B. P. (2000). Abnormalities in 5-HT2A receptor mRNA expression in frontal cortex of chronic elderly schizophrenics with varying histories of neuroleptic treatment. J. Neurosci. Res. 59, 218–225. Herrick-Davis, K., Grinde, E., and Niswender, C. M. (1999). Serotonin 5-HT2C receptor RNA editing alters receptor basal activity: Implications for serotonergic signal transduction. J. Neurochem. 73, 1711–1717. Hesselink, J. M. K., and Sambunaris, A. (1995). Behavioral pharmacology of serotonin receptor subtypes: Hypotheses for clinical applications of selective serotonin ligands. Int. Rev. Psychiatr. 7, 41–53. Hietala, J., Koulu, M., Kuoppamaki, M., Lappalainen, J., and Syvalahti, E. (1992). Chronic clozapine treatment down-regulates serotonin 5-HT-1c receptors in rat brain. Prog. Neuropsycho- pharmacol. Biol. Psychiat. 16, 727–732. Hillion, J., Milne-Edwards, J. B., Catelon, J., de Vitry, F., Gros, F., and Hamon, M. (1993). Prenatal developmental expression of rat brain 5-HT1A receptor gene followed by PCR. Biochem. Biophys. Res. Commun. 191, 991–997. Hirst, W. D., Cheung, N. Y., Rattray, M., Price, G. W., and Wilkin, G. P. (1998). Cultured astrocytes express messenger RNA for multiple serotonin receptor subtypes, without functional coupling of 5-HT1 receptor subtypes to adenylyl cyclase. Brain Res. Mol. Brain Res. 61, 90–99. Holden, R. J. (1995). Schizophrenia, suicide and the serotonin story. Med. Hypoth. 44, 379–391. Hollander, E., Kaplan, A., Cartwright, C., and Reichman, D. (2000a). Venlafaxine in children, adolescents, and young adults with autism spectrum disorders: An open retrospective clinical report. J. Child Neurol. 15, 132–135. Hollander, E., Novotny, S., Allen, A., Aronowitz, B., Cartwright, C., and DeCaria, C. (2000b). The relationship between repetitive behaviors and growth hormone response to sumatriptan challenge in adult autistic disorder. Neuropsychopharmacology. 22, 163–167. Hoyer, D., Clarke, D. E., Fozard, J. R., Hartig, P. R., Martin, G. R., Mylecharane, E. J., Saxena, P. R., and Humphrey, P. P. (1994). International Union of Pharmacology classification of receptors for 5-hydroxytryptamine (serotonin). Pharmacol. Rev. 46, 157–203. Huang, C. C., Hsu, K. S., and Gean, P. W. (1996). Isoproterenol potentiates synaptic transmission primarily by enhancing presynaptic calcium influx via P- and/or Q-type calcium channels in the rat amygdala. J. Neurosci. 16, 1026–1033. Hubel, D. H., and Wiesel, T. N. (1970). The period of susceptibility to the physiological effects of unilateral eye closure in kittens. J. Physiol. 206, 419–436. Huttenlocher, P. R., and Dabholkar, A. S. (1997). Regional differences in synaptogenesis in human cerebral cortex. J. Comp. Neurol. 387, 167–178. 160 SODHI AND SANDERS-BUSH

Ike, J., Canton, H., and Sanders-Bush, E. (1995). Developmental switch in the hippocampal serotonin receptor linked to phosphoinositide hydrolysis. Brain Res. 678, 49–54. Inayama, Y., Yoneda, H., Sakai, T., Ishida, T., Nonomura, Y., Kono, Y., Takahata, R., Koh, J., Sakai, J., Takai, A., Inada, Y., and Asaba, H. (1996). Positive association between a DNA sequence variant in the serotonin 2A receptor gene and schizophrenia. Am. J. Med. Genet. 67, 103–105. Inoue, T., Hashimoto, S., Tsuchiya, K., Izumi, T., Ohmori, T., and Koyama, T. (1996). Effect of , a selective serotonin reuptake inhibitor, on the acquisition of conditioned freezing. Eur. J. Pharmacol. 311, 1–6. Iwamoto, K., and Kato, T. (2003). RNA editing of serotonin 2C receptors in human postmortem brains of major mental disorders. Neurosci. Lett. 346, 169–172. Jacobs, B. L., and Azmitia, E. C. (1992). Structure and function of the brain serotonin system. Physiol. Rev. 72, 165–229. Jacobs, B. L., and Trulson, M. E. (1979). Mechanisms of action of LSD. Am. Sci. 67, 396–404. Jallon, P., and Picard, F. (2001). Body weight gain and anticonvulsants: A comparative review. Drug Safety 24, 969–978. Joffe, H., and Cohen, L. S. (1998). Estrogen, serotonin, and mood disturbance: Where is the therapeutic bridge? Biol. Psychiat. 44, 798–811. John, N. J., Lew, G. M., Goya, L., and Timiras, P. S. (1991). Effects of serotonin on hydroxylase and tau protein in a human neuroblastoma cell line. Adv. Exp. Med. Biol. 296, 69–80. Jones, L. B. (2001). Recent cytoarchitechtonic changes in the prefrontal cortex of schizophrenics. Front. Biosci. 6, E148–E153. Jones, P. (1994). Schizophrenia after prenatal exposure to the Dutch hunger winter of 1944–1945. Arch. Gen. Psychiat. 51, 333–334. Jones, P., Rodgers, B., Murray, R., and Marmot, M. (1994). Child development risk factors for adult schizophrenia in the British 1946 birth cohort. Lancet 344, 1398–1402. Joyce, J. N., Shane, A., Lexow, N., Winokur, A., Casanova, M. F., and Kleinman, J. E. (1993). Serotonin uptake sites and serotonin receptors are altered in the limbic system of schizophrenics. Neuropsychopharmacology 8, 315–336. Julius, D., Huang, K. N., Livelli, T. J., Axel, R., and Jessell, T. M. (1990). The 5HT2 receptor defines a family of structurally distinct but functionally conserved serotonin receptors. Proc. Natl. Acad. Sci. USA 87, 928–932. Julius, D., Livelli, T. J., Jessell, T. M., and Axel, R. (1989). Ectopic expression of the serotonin 1c receptor and the triggering of malignant transformation. Science 244, 1057–1062. Kandel, E. R., and O’Dell, T. J. (1992). Are adult learning mechanisms also used for development? Science 258, 243–245. Kandel, E. R., and Spencer, W. A. (1968). Cellular neurophysiological approaches in the study of learning. Physiol. Rev. 48, 65–134. Kapur, S., and Remington, G. (1996). Serotonin-dopamine interaction and its relevance to schizophrenia. Am. J. Psychiat. 153, 466–476. Kato, K., Suzuki, F., Kurobe, N., Okajima, K., Ogasawara, N., Nagaya, M., and Yamanaka, T. (1990). Enhancement of S-100 beta protein in blood of patients with Down’s syndrome. J. Mol. Neurosci. 2, 109–113. Kato, S., Fujiwara, I., and Yoshida, N. (1999). Nitrogen-containing heteroalicycles with serotonin receptor binding affinity: Development of gastroprokinetic and antiemetic agents. Med. Res. Rev. 19, 25–73. Kaye, W. H., Gwirtsman, H. E., George, D. T., and Ebert, M. H. (1991). Altered serotonin activity in nervosa after long-term weight restoration: Does elevated cerebrospinal fluid 5-hydro- xyindoleacetic acid level correlate with rigid and obsessive behavior? Arch. Gen. Psychiat. 48, 556–562. Kellar, K. J., Cascio, C. S., Butler, J. A., and Kurtzke, R. N. (1981). Differential effects of electroconvulsive shock and antidepressant drugs on serotonin-2 receptors in rat brain. Eur. J. Pharmacol. 69, 515–518. SEROTONIN AND BRAIN DEVELOPMENT 161

Kendler, K. S., Bulik, C. M., Silberg, J., Hettema, J. M., Myers, J., and Prescott, C. A. (2000). Childhood sexual abuse and adult psychiatric and substance use disorders in women: An epidemiological and cotwin control analysis. Arch. Gen. Psychiat. 57, 953–959. Kendler, K. S., Gardner, C. O., and Prescott, C. A. (2002a). Toward a comprehensive developmental model for major depression in women. Am. J. Psychiatr. 159, 1133–1145. Kendler, K. S., Kessler, R. C., Neale, M. C., Heath, A. C., and Eaves, L. J. (1993). The prediction of major depression in women: Toward an integrated etiologic model. Am. J. Psychiat. 150, 1139–1148. Kendler, K. S., Neale, M. C., Kessler, R. C., Heath, A. C., and Eaves, L. J. (1992). Childhood parental loss and adult psychopathology in women: A twin study perspective. Arch. Gen. Psychiat. 49, 109–116. Kendler, K. S., Neale, M. C., Prescott, C. A., Kessler, R. C., Heath, A. C., Corey, L. A., and Eaves, L. J. (1996). Childhood parental loss and alcoholism in women: A causal analysis using a twin- family design. Psychol. Med. 26, 79–95. Kendler, K. S., Sheth, K., Gardner, C. O., and Prescott, C. A. (2002b). Childhood parental loss and risk for first-onset of major depression and alcohol dependence: The time-decay of risk and sex differences. Psychol. Med. 32, 1187–1194. Keshavan, M. S., Rosenberg, D., Sweeney, J. A., and Pettegrew, J. W. (1998). Decreased caudate volume in neuroleptic-naive psychotic patients. Am. J. Psychiat. 155, 774–778. Kessler, R. C., Davis, C. G., and Kendler, K. S. (1997). Childhood adversity and adult psychiatric disorder in the US National Comorbidity Survey. Psychol. Med. 27, 1101–1119. Kesterson, K. L., Lane, R. D., and Rhoades, R. W. (2002). Effects of elevated serotonin levels on patterns of GAP-43 expression during barrel development in rat somatosensory cortex. Brain Res. Dev. Brain Res. 139, 167–174. Kientz, M. A., and Dunn, W. (1997). A comparison of the performance of children with and without autism on the sensory profile. Am. J. Occup. Ther. 51, 530–537. Kim, C. K., Kalynchuk, L. E., Kornecook, T. J., Mumby, D. G., Dadgar, N. A., Pinel, J. P., and Weinberg, J. (1997). Object-recognition and spatial learning and memory in rats prenatally exposed to ethanol. Behav. Neurosci. 111, 985–995. Kim, C. S., McNamara, M. C., Lauder, J. M., and Lawson, E. E. (1994). Immunocytochemical detection of serotonin content in raphe neurons of newborn and young adult rabbits before and after acute hypoxia. Int. J. Dev. Neurosci. 12, 499–505. Kimmel, S. E., Calabrese, J. R., Woyshville, M. J., and Meltzer, H. Y. (1994). Clozapine in treatment- refractory mood disorders. J. Clin. Psychiat. 55(Suppl. B), 91–93. Klancnik, J. M., Obenaus, A., Phillips, A. G., and Baimbridge, K. G. (1991). The effects of serotonergic compounds on evoked responses in the dentate gyrus and CA1 region of the hippocampal formation of the rat. Neuropharmacology 30, 1201–1209. Koe, B. K., and Weissman, A. (1966). p-Chlorophenylalanine: A specific depletor of brain serotonin. J. Pharmacol. Exp. Ther. 154, 499–516. Kojic, L., Dyck, R. H., Gu, Q., Douglas, R. M., Matsubara, J., and Cynader, M. S. (2000). Columnar distribution of serotonin-dependent plasticity within kitten striate cortex. Proc. Natl. Acad. Sci. USA 97, 1841–1844. Kojic, L., Gu, Q., Douglas, R. M., and Cynader, M. S. (1997). Serotonin facilitates synaptic plasticity in kitten visual cortex: An in vitro study. Brain Res. Dev. Brain Res. 101, 299–304. Kojic, L., Gu, Q., Douglas, R. M., and Cynader, M. S. (2001). Laminar distribution of cholinergic- and serotonergic-dependent plasticity within kitten visual cortex. Brain Res. Dev. Brain Res. 126, 157–162. Kojima, T., Homma, S., Sako, H., Shimizu, I., Okada, A., and Okado, N. (1988). Developmental changes in density and distribution of serotoninergic fibers in the chick spinal cord. J. Comp. Neurol. 267, 580–589. 162 SODHI AND SANDERS-BUSH

Kolb, B. (1984). Functions of the frontal cortex of the rat: A comparative review. Brain Res. 320, 65–98. Kontur, P. J., Leranth, C., Redmond, D. E., Jr., Roth, R. H., and Robbins, R. J. (1993). Tyrosine hydroxylase immunoreactivity and monoamine and metabolite levels in cryopreserved human fetal ventral mesencephalon. Exp. Neurol. 121, 172–180. Koponen, H., Saari, K., Savolainen, M., and Isohanni, M. (2002). Weight gain and glucose and lipid metabolism disturbances during antipsychotic medication: A review. Eur. Arch. Psy. Clin. N. 252, 294–298. Korte, S. M., Meijer, O. C., de Kloet, E. R., Buwalda, B., Keijser, J., Sluyter, F., van Oortmerssen, G., and Bohus, B. (1996). Enhanced 5-HT1A receptor expression in forebrain regions of aggressive house mice. Brain Res. 736, 338–343. Kosofsky, B. E., and Molliver, M. E. (1987). The serotoninergic innervation of cerebral cortex: Different classes of axon terminals arise from dorsal and median raphe nuclei. Synapse 1, 153–168. Kovacs, G. G., Kloppel, S., Fischer, I., Dorner, S., Lindeck-Pozza, E., Birner, P., Botefur, I. C., Pilz, P., Volk, B., and Budka, H. (2003). Nucleus-specific alteration of raphe neurons in human neurodegenerative disorders. Neuroreport 14, 73–76. Krezel, W., Dupont, S., Krust, A., Chambon, P., and Chapman, P. F. (2001). Increased anxiety and synaptic plasticity in estrogen receptor beta-deficient mice. Proc. Natl. Acad. Sci. USA 98, 12278–12282. Lafaille, F., Welner, S. A., and Suranyi-Cadotte, B. E. (1991). Regulation of serotonin type 2 (5-HT2) and beta-adrenergic receptors in rat cerebral cortex following novel and classical antidepressant treatment. J. Psychiat. Neurosci. 16, 209–214. Lambert, H. W., and Lauder, J. M. (1999). Serotonin receptor agonists that increase cyclic AMP positively regulate IGF-I in mouse mandibular mesenchymal cells. Dev. Neurosci. 21, 105–112. Lapiz, M. D., Fulford, A., Muchimapura, S., Mason, R., Parker, T., and Marsden, C. A. (2001). Influence of postweaning social isolation in the rat on brain development, conditioned behaviour and neurotransmission. Ross Fiziol Zh Im I M Sechenova. 87, 730–751. Laruelle, M., Abi-Dargham, A., Casanova, M. F., Toti, R., Weinberger, D. R., and Kleinman, J. E. (1993). Selective abnormalities of prefrontal serotonergic receptors in schizophrenia. A postmortem study. Arch. Gen. Psychiat. 50, 810–818. Lauder, J. M. (1990). Ontogeny of the serotonergic system in the rat: Serotonin as a developmental signal. Ann. N. Y. Acad. Sci. 600, 297–313; discussion 314. Lauder, J. M., and Krebs, H. (1976). Effects of p-chlorophenylalanine on time of neuronal origin during embryogenesis in the rat. Brain Res. 107, 638–644. Lauder, J. M., and Krebs, H. (1978). Serotonin as a differentiation signal in early neurogenesis. Dev. Neurosci. 1, 15–30. Lauder, J. M., Liu, J., and Grayson, D. R. (2000). In utero exposure to serotonergic drugs alters neonatal expression of 5-HT(1A) receptor transcripts: A quantitative RT-PCR study. Int. J. Dev. Neurosci. 18, 171–176. Lauder, J. M., Towle, A. C., Patrick, K., Henderson, P., and Krebs, H. (1985). Decreased serotonin content of embryonic raphe neurons following maternal administration of p-chlorophenylalanine: A quantitative immunocytochemical study. Brain Res. 352, 107–114. Lauder, J. M., Wallace, J. A., and Krebs, H. (1981). Roles for serotonin in neuroembryogenesis. Adv. Exp. Med. Biol. 133, 477–506. Lawrie, S. M., and Abukmeil, S. S. (1998). Brain abnormality in schizophrenia: A systematic and quantitative review of volumetric magnetic resonance imaging studies. Br. J. Psychiat. 172, 110–120. Lebrand, C., Cases, O., Adelbrecht, C., Doye, A., Alvarez, C., El Mestikawy, S., Seif, I., and Gaspar, P. (1996). Transient uptake and storage of serotonin in developing thalamic neurons. Neuron 17, 823–835. SEROTONIN AND BRAIN DEVELOPMENT 163

Lee, S. L., Wang, W. W., and Fanburg, B. L. (1997). Association of Tyr phosphorylation of GTPase- activating protein with mitogenic action of serotonin. Am. J. Physiol. 272, C223–C230. Leranth, C., Shanabrough, M., and Horvath, T. L. (2000). Hormonal regulation of hippocampal spine synapse density involves subcortical mediation. Neuroscience 101, 349–356. Lerer, B., Macciardi, F., Segman, R. H., Adolfsson, R., Blackwood, D., Blairy, S., Del Favero, J., Dikeos, D. G., Kaneva, R., Lilli, R., Massat, I., Milanova, V., Muir, W., Noethen, M., Oruc, L., Petrova, T., Papadimitriou, G. N., Rietschel, M., Serretti, A., Souery, D., Van Gestel, S., Van Broeckhoven, C., and Mendlewicz, J. (2001). Variability of 5-HT2C receptor cys23ser polymorphism among European populations and vulnerability to affective disorder. Mol. Psychiat. 6, 579–585. Leslie, M. J., Bennett-Clarke, C. A., and Rhoades, R. W. (1992). Serotonin 1B receptors form a transient vibrissa-related pattern in the primary somatosensory cortex of the developing rat. Brain Res. Dev. Brain Res. 69, 143–148. Levallois, C., Valence, C., Baldet, P., and Privat, A. (1997). Morphological and morphometric analysis of serotonin-containing neurons in primary dissociated cultures of human rhombenceph- alon: A study of development. Brain Res. Dev. Brain Res. 99, 243–252. Leven, R. M., Gonnella, P. A., Reeber, M. J., and Nachmias, V. T. (1983). Platelet shape change and cytoskeletal assembly: Effects of pH and monovalent cation ionophores. Thromb. Haemost. 49, 230–234. Lewis, D. A., and Levitt, P. (2002). Schizophrenia as a disorder of neurodevelopment. Annu. Rev. Neurosci. 25, 409–432. Lidov, H. G., and Molliver, M. E. (1982). Immunohistochemical study of the development of serotonergic neurons in the rat CNS. Brain Res. Bull. 9, 559–604. Lidow, M. S., and Rakic, P. (1992). Scheduling of monoaminergic neurotransmitter receptor expression in the primate neocortex during postnatal development. Cereb. Cortex 2, 401–416. Lopez, J. F., Chalmers, D. T., Little, K. Y., and Watson, S. J. (1998). A. E. Bennett Research Award. Regulation of serotonin 1A, glucocorticoid, and mineralocorticoid receptor in rat and human hippocampus: Implications for the neurobiology of depression. Biol. Psychiat. 43, 547–573. Lopez, J. F., Liberzon, I., Vazquez, D. M., Young, E. A., and Watson, S. J. (1999). Serotonin 1A receptor messenger RNA regulation in the hippocampus after acute stress. Biol. Psychiat. 45, 934–937. Lunardi, G., Mainardi, P., Rubino, V., Fracassi, M., Favale, E., and Albano, C. (1995). Serotonin and epilepsy. Epilepsia 36, S165. Luque, C. A., and Rey, J. A. (1999). Sibutramine: A serotonin- reuptake inhibitor for the treatment of obesity. Ann. Pharmacother. 33, 968–978. Manjarrez, G. G., Magdaleno, V. M., Chagoya, G., and Hernandez, J. (1996). Nutritional recovery does not reverse the activation of brain serotonin synthesis in the ontogenetically malnourished rat. Int. J. Dev. Neurosci. 14, 641–648. Mann, D. M., Yates, P. O., Marcyniuk, B., and Ravindra, C. R. (1985). Pathological evidence for neurotransmitter deficits in Down’s syndrome of middle age. J. Ment. Defic. Res. 29(Pt.2), 125–135. Mann, J. J., McBride, P. A., and Stanley, M. (1986a). Postmortem and enzyme studies in suicide. Ann. N. Y. Acad. Sci. 487, 114–121. Mann, J. J., McBride, P. A., and Stanley, M. (1986b). Postmortem serotonergic and binding to frontal cortex: Correlations with suicide. Psychopharmacol. Bull. 22, 647–649. Mann, J. J., Stanley, M., McBride, P. A., and McEwen, B. S. (1986c). Increased serotonin2 and beta- adrenergic receptor binding in the frontal cortices of suicide victims. Arch. Gen. Psychiat. 43, 954–959. 164 SODHI AND SANDERS-BUSH

Marazziti, D., Muratori, F., Cesari, A., Masala, I., Baroni, S., Giannaccini, G., Dell’Osso, L., Cosenza, A., Pfanner, P., and Cassano, G. B. (2000). Increased density of the platelet serotonin transporter in autism. Pharmacopsychiatry 33, 165–168. Marek, G. J., Li, A. A., and Seiden, L. S. (1989). Selective 5-hydroxytryptamine2 antagonists have antidepressant-like effects on differential-reinforcement-of-low-rate 72-second schedule. J. Pharmacol. Exp. Ther. 250, 52–59. Maren, S., Aharonov, G., and Fanselow, M. S. (1997). Neurotoxic lesions of the dorsal hippocampus and Pavlovian fear conditioning in rats. Behav. Brain Res. 88, 261–274. Marenco, S., and Weinberger, D. R. (2000). The neurodevelopmental hypothesis of schizophrenia: Following a trail of evidence from cradle to grave. Dev. Psychopathol. 12, 501–527. Martin, S. J., Grimwood, P. D., and Morris, R. G. (2000). Synaptic plasticity and memory: An evaluation of the hypothesis. Annu. Rev. Neurosci. 23, 649–711. Masellis, M., Basile, V., Meltzer, H. Y., Lieberman, J. A., Sevy, S., Macciardi, F. M., Cola, P., Howard, A., Badri, F., Northen, M. M., Kalow, W., and Kennedy, J. L. (1998). Serotonin subtype 2 receptor genes and clinical response to clozapine in schizophrenia patients. Neuropsychopharmacology 19, 123–132. Masellis, M., Paterson, A. D., Badri, F., Lieberman, J. A., Meltzer, H. Y., Cavazzoni, P., and Kennedy, J. L. (1995). Genetic variation of 5-HT2A receptor and response to clozapine. Lancet 346, 1108. Matilla, A., Roberson, E. D., Banfi, S., Morales, J., Armstrong, D. L., Burright, E. N., Orr, H. T., Sweatt, J. D., Zoghbi, H. Y., and Matzuk, M. M. (1998). Mice lacking ataxin-1 display learning deficits and decreased hippocampal paired-pulse facilitation. J. Neurosci. 18, 5508–5516. Matsukawa, M., Ogawa, M., Nakadate, K., Maeshima, T., Ichitani, Y., Kawai, N., and Okado, N. (1997). Serotonin and acetylcholine are crucial to maintain hippocampal synapses and memory acquisition in rats. Neurosci. Lett. 230, 13–16. Mayford, M., and Kandel, E. R. (1999). Genetic approaches to memory storage. Trends Genet. 15, 463–470. Mazer, C., Muneyyirci, J., Taheny, K., Raio, N., Borella, A., and Whitaker-Azmitia, P. (1997). Serotonin depletion during synaptogenesis leads to decreased synaptic density and learning deficits in the adult rat: A possible model of neurodevelopmental disorders with cognitive deficits. Brain Res. 760, 68–73. McDougle, C. J., Kresch, L. E., and Posey, D. J. (2000). Repetitive thoughts and behavior in pervasive developmental disorders: Treatment with serotonin reuptake inhibitors. J. Autism Dev. Disord. 30, 427–435. McDougle, C. J., Naylor, S. T., Goodman, W. K., Volkmar, F. R., Cohen, D. J., and Price, L. H. (1993). Acute tryptophan depletion in autistic disorder: A controlled case study. Biol. Psychiat. 33, 547–550. McGlashan, T. H., and Hoffman, R. E. (2000). Schizophrenia as a disorder of developmentally reduced synaptic connectivity. Arch. Gen. Psychiat. 57, 637–648. McKeith, I. G., Marshall, E. F., Ferrier, I. N., Armstrong, M. M., Kennedy, W. N., Perry, R. H., Perry, E. K., and Eccleston, D. (1987). 5-HT receptor binding in post-mortem brain from patients with affective disorder. J. Affect. Disord. 13, 67–74. McKittrick, C. R., Blanchard, D. C., Blanchard, R. J., McEwen, B. S., and Sakai, R. R. (1995). Serotonin receptor binding in a colony model of chronic social stress. Biol. Psychiat. 37, 383–393. McNeely, W., and Goa, K. L. (1998). Sibutramine: A review of its contribution to the management of obesity. Drugs 56, 1093–1124. Meaney, M. J., Diorio, J., Francis, D., LaRocque, S., O’Donnell, D., Smythe, J. W., Sharma, S., and Tannenbaum, B. (1994). Environmental regulation of the development of glucocorticoid receptor systems in the rat forebrain: The role of serotonin. Ann. N. Y. Acad. Sci. 746, 260–273; discussion 274, 289–293. SEROTONIN AND BRAIN DEVELOPMENT 165

Meert, T. F., Melis, W., Aerts, N., and Clincke, G. (1997). Antagonism of meta-chlorophenylpiper- azine-induced inhibition of exploratory activity in an emergence procedure, the open field test, in rats. Behav. Pharmacol. 8, 353–363. Meijer, O. C., Cole, T. J., Schmid, W., Schutz, G., Joels, M., and De Kloet, E. R. (1997). Regulation of hippocampal 5-HT1A receptor mRNA and binding in transgenic mice with a targeted disruption of the glucocorticoid receptor. Brain Res. Mol. Brain Res. 46, 290–296. Meijer, O. C., Van Oosten, R. V., and De Kloet, E. R. (1997). Elevated basal trough levels of corticosterone suppress hippocampal 5-hydroxytryptamine(1A) receptor expression in adrenally intact rats: Implication for the pathogenesis of depression. Neuroscience 80, 419–426. Meltzer, H. Y. (2002). Commentary on ‘‘Clinical studies on the mechanism of action of clozapine; the dopamine-serotonin hypothesis of schizophrenia.’’ Psychopharmacology 163, 1–3. Meltzer, H. Y., Bastani, B., Ramirez, L., and Matsubara, S. (1989a). Clozapine: New research on efficacy and mechanism of action. Eur. Arch. Psychiat. Neurol. Sci. 238, 332–339. Meltzer, H. Y., Matsubara, S., and Lee, J. C. (1989b). Classification of typical and atypical antipsychotic drugs on the basis of dopamine D-1, D-2 and serotonin2 pKi values. J. Pharmacol. Exp. Ther. 251, 238–246. Meltzer, H. Y., Matsubara, S., and Lee, J. C. (1989c). The ratios of serotonin2 and dopamine2 affinities differentiate atypical and drugs. Psychopharmacol. Bull. 25, 390–392. Menard, J., and Treit, D. (1999). Effects of centrally administered anxiolytic compounds in animal models of anxiety. Neurosci. Biobehav. Rev. 23, 591–613. Meredith, G. E., De Souza, I. E., Hyde, T. M., Tipper, G., Wong, M. L., and Egan, M. F. (2000). Persistent alterations in dendrites, spines, and dynorphinergic synapses in the nucleus accumbens shell of rats with neuroleptic-induced dyskinesias. J. Neurosci. 20, 7798–7806. Merzak, A., Koochekpour, S., Fillion, M. P., Fillion, G., and Pilkington, G. J. (1996). Expression of serotonin receptors in human fetal astrocytes and glioma cell lines: A possible role in glioma cell proliferation and migration. Brain Res. Mol. Brain Res. 41, 1–7. Micheva, K. D., and Beaulieu, C. (1996). Quantitative aspects of synaptogenesis in the rat barrel field cortex with special reference to GABA circuitry. J. Comp. Neurol. 373, 340–354. Mikuni,M.,Kagaya,A.,Takahashi,K.,andMeltzer,H.Y.(1992).Serotoninbutnot norepinephrine-induced calcium mobilization of platelets is enhanced in affective disorders. Psychopharmacology (Berl.) 106, 311–314 . Miller, J. H., and Azmitia, E. C. (1999). Growth inhibitory effects of a mu opioid on cultured cholinergic neurons from fetal rat ventral forebrain, brainstem, and spinal cord. Brain Res. Dev. Brain Res. 114, 69–77. Miyata, M., Ito, M., Sasajima, T., Ohira, H., Sato, Y., and Kasukawa, R. (2000). Development of monocrotaline-induced pulmonary hypertension is attenuated by a serotonin receptor antagon- ist. Lung 178, 63–73. Moiseiwitsch, J. R., and Lauder, J. M. (1995). Serotonin regulates mouse cranial neural crest migration. Proc. Natl. Acad. Sci. USA 92, 7182–7186. Moiseiwitsch, J. R., and Lauder, J. M. (1997). Regulation of gene expression in cultured embryonic mouse mandibular mesenchyme by serotonin antagonists. Anat. Embryol. (Berl.) 195, 71–78. Moiseiwitsch, J. R., Raymond, J. R., Tamir, H., and Lauder, J. M. (1998). Regulation by serotonin of tooth-germ morphogenesis and gene expression in mouse mandibular explant cultures. Arch. Oral Biol. 43, 789–800. Mokler, D. J., Bronzino, J. D., Galler, J. R., and Morgane, P. J. (1999). The effects of median raphe electrical stimulation on serotonin release in the dorsal hippocampal formation of prenatally protein malnourished rats. Brain Res. 838, 95–103. Monaco, F., Torta, R., Borio, R., Cicolin, A., and Ravizza, L. (1995). Epilepsy, depression, and the selective serotonin reuptake inhibitor drugs. Epilepsia 36, S175. 166 SODHI AND SANDERS-BUSH

Muneoka, K., Ogawa, T., Kamei, K., Muraoka, S., Tomiyoshi, R., Mimura, Y., Kato, H., Suzuki, M. R., and Takigawa, M. (1997). Prenatal nicotine exposure affects the development of the central serotonergic system as well as the dopaminergic system in rat offspring: Involvement of route of drug administrations. Brain Res. Dev. Brain Res. 102, 117–126. Murray, R. M., O’Callaghan, E., Castle, D. J., and Lewis, S. W. (1992). A neurodevelopmental approach to the classification of schizophrenia. Schizophr. Bull. 18, 319–332. Namerow, L. B., Thomas, P., Bostic, J. Q., Prince, J., and Monuteaux, M. C. (2003). Use of citalopram in pervasive developmental disorders. J. Dev. Behav. Pediatr. 24, 104–108. Nanson, J. L. (1992). Autism in fetal alcohol syndrome: A report of six cases. Alcohol Clin. Exp. Res. 16, 558–565. Nebigil, C. G., Choi, D. S., Dierich, A., Hickel, P., Le Meur, M., Messaddeq, N., Launay, J. M., and Maroteaux, L. (2000). Serotonin 2B receptor is required for heart development. Proc. Natl. Acad. Sci. USA 97, 9508–9513. Nelson, D. R., Thomas, D. R., and Johnson, A. M. (1989). Pharmacological effects of paroxetine after repeated administration to animals. Acta Psychiatr. Scand. Suppl. 350, 21–23. Nichols, C. D., Garcia, E. E., and Sanders-Bush, E. (2003). Dynamic changes in prefrontal cortex gene expression following lysergic acid diethylamide administration. Brain Res. Mol. Brain Res. 111, 182–188. Nichols, C. D., and Sanders-Bush, E. (2002). A single dose of lysergic acid diethylamide influences gene expression patterns within the mammalian brain. Neuropsychopharmacology 26, 634–642. Niswender, C. M., Copeland, S. C., Herrick-Davis, K., Emeson, R. B., and Sanders-Bush, E. (1999). RNA editing of the human serotonin 5-hydroxytryptamine 2C receptor silences constitutive activity. J. Biol. Chem. 274, 9472–9478. Niswender, C. M., Herrick-Davis, K. et al. (2001). RNA editing of the human serotonin 5-HT2C receptor alterations in suicide and implications for serotonergic pharmacotherapy. Neuropsycho- pharmacology. 24, 478–491. Norrholm, S. D., and Ouimet, C. C. (2001). Altered dendritic spine density in animal models of depression and in response to antidepressant treatment. Synapse 42, 151–163. Ohashi, S., Matsumoto, M., Otani, H., Mori, K., Togashi, H., Ueno, K., Kaku, A., and Yoshioka, M. (2002). Changes in synaptic plasticity in the rat hippocampo-medial prefrontal cortex pathway induced by repeated treatments with fluvoxamine. Brain Res. 949, 131–138. Ohuoha, D. C., Hyde, T. M., and Kleinman, J. E. (1993). The role of serotonin in schizophrenia: An overview of the nomenclature, distribution and alterations of serotonin receptors in the central nervous system. Psychopharmacology (Berl.) 112, S5–15. Okado, N., Cheng, L., Tanatsugu, Y., Hamada, S., and Hamaguchi, K. (1993). Synaptic loss following removal of serotoninergic fibers in newly hatched and adult chickens. J. Neurobiol. 24, 687–698. Okado, N., Narita, M., and Narita, N. (2001). A biogenic amine-synapse mechanism for mental retardation and developmental disabilities. Brain Dev. 23(Suppl. 1), S11–15. Olivier, B., Pattij, T., Wood, S. J., Oosting, R., Sarnyai, Z., and Toth, M. (2001). The 5-HT(1A) receptor knockout mouse and anxiety. Behav. Pharmacol. 12, 439–450. O’Steen, W. K., Barnard, J. L., Jr., and Yates, R. D. (1967). Morphologic changes in skeletal muscle induced by serotonin treatment: A light- and electron-microscope study. Exp. Mol. Pathol. 7, 145–155. Osterheld-Haas, M. C., Van der Loos, H., and Hornung, J. P. (1994). Monoaminergic afferents to cortex modulate structural plasticity in the barrelfield of the mouse. Brain Res. Dev. Brain Res. 77, 189–202. Pakkenberg, B. (1993). Total nerve cell number in neocortex in chronic schizophrenics and controls estimated using optical disectors. Biol. Psychiat. 34, 768–772. SEROTONIN AND BRAIN DEVELOPMENT 167

Pakkenberg, B., and Gundersen, H. J. (1989). New stereological method for obtaining unbiased and efficient estimates of total nerve cell number in human brain areas: Exemplified by the mediodorsal thalamic nucleus in schizophrenics. Apmis 97, 677–681. Palmer, B. W., McClure, F. S., and Jeste, D. V. (2001). Schizophrenia in late life: Findings challenge traditional concepts. Harv. Rev. Psychiat. 9, 51–58. Pandey, G. N. (1997). Altered serotonin function in suicide: Evidence from platelet and neuroendocrine studies. Ann. N. Y. Acad. Sci. 836, 182–200. Pandey, G. N., Pandey, S. C., Dwivedi, Y., Sharma, R. P., Janicak, P. G., and Davis, J. M. (1995). Platelet serotonin-2A receptors: A potential biological marker for suicidal behavior. Am. J. Psychiat. 152, 850–855. Pare, W. P., and Tejani-Butt, S. M. (1996). Effect of stress on the behavior and 5-HT system in Sprague-Dawley and Wistar Kyoto rat strains. Integr. Physiol. Behav. Sci. 31, 112–121. Parks, C. L., Robinson, P. S., Sibille, E., Shenk, T., and Toth, M. (1998). Increased anxiety of mice lacking the serotonin1A receptor. Proc. Natl. Acad. Sci. USA 95, 10734–10739. Pazos, A., Hoyer, D., and Palacios, J. M. (1984). The binding of serotonergic ligands to the porcine choroid plexus: Characterization of a new type of serotonin recognition site. Eur. J. Pharmacol. 106, 539–546. Pedigo, N. W., Yamamura, H. I., and Nelson, D. L. (1981). Discrimination of multiple [3H]5- hydroxytryptamine binding sites by the neuroleptic spiperone in rat brain. J. Neurochem. 36, 220–226. Pei, Q., Lewis, L., Sprakes, M. E., Jones, E. J., Grahame-Smith, D. G., and Zetterstrom, T. S. (2000). Serotonergic regulation of mRNA expression of Arc, an immediate early gene selectively localized at neuronal dendrites. Neuropharmacology 39, 463–470. Peroutka, S. J., and Snyder, S. H. (1980). Long-term antidepressant treatment decreases spiroperidol- labeled serotonin receptor binding. Science 210, 88–90. Perrier, J. F., Alaburda, A., and Hounsgaard, J. (2002). Spinal plasticity mediated by postsynaptic L-type Ca2þ channels. Brain Res. Brain Res. Rev. 40, 223–229. Persico, A. M., Altamura, C., Calia, E., Puglisi-Allegra, S., Ventura, R., Lucchese, F., and Keller, F. (2000). Serotonin depletion and barrel cortex development: Impact of growth impairment vs. serotonin effects on thalamocortical endings. Cereb. Cortex 10, 181–191. Persico, A. M., Mengual, E., Moessner, R., Hall, F. S., Revay, R. S., Sora, I., Arellano, J., DeFelipe, J., Gimenez-Amaya, J. M., Conciatori, M., Marino, R., Baldi, A., Cabib, S., Pascucci, T., Uhl, G. R., Murphy, D. L., Lesch, K. P., Keller, F., and Hall, S. F. (2001). Barrel pattern formation requires serotonin uptake by thalamocortical afferents, and not vesicular monoamine release. J. Neurosci. 21, 6862–6873. Peters, D. A. (1988). Both prenatal and postnatal factors contribute to the effects of maternal stress on offspring behavior and central 5-hydroxytryptamine receptors in the rat. Pharmacol. Biochem. Behav. 30, 669–673. Pflieger, J. F., Clarac, F., and Vinay, L. (2002). Postural modifications and neuronal excitability changes induced by a short-term serotonin depletion during neonatal development in the rat. J. Neurosci. 22, 5108–5117. Pieri, L., Keller, H. H., Burkard, W., and Da Prada, M. (1978). Effects of and LSD on cerebral monoamine systems and hallucinosis. Nature 272, 278–280. Pilowsky, L. S., Kerwin, R. W., and Murray, R. M. (1993). Schizophrenia: A neurodevelopmental perspective. Neuropsychopharmacology 9, 83–91. Pletnikov, M. V., Rubin, S. A., Schwartz, G. J., Carbone, K. M., and Moran, T. H. (2000). Effects of neonatal rat Borna disease virus (BDV) infection on the postnatal development of the brain monoaminergic systems. Brain Res. Dev. Brain Res. 119, 179–185. Popenenkova, Z. A., Zuev, V. A., Romanovskaia, M. G., and Maslennikov, G. A. (1977). Acute and latent influenzal infection in mice with altered endogenous serotonin metabolism. Vopr Virusol 432–437. 168 SODHI AND SANDERS-BUSH

Popken, G. J., Bunney, W. E., Jr., Potkin, S. G., and Jones, E. G. (2000). Subnucleus-specific loss of neurons in medial thalamus of schizophrenics. Proc. Natl. Acad. Sci. USA 97, 9276–9280. Posey, D. J., and McDougle, C. J. (2000). The pharmacotherapy of target symptoms associated with autistic disorder and other pervasive developmental disorders. Harv. Rev. Psychiat. 8, 45–63. Prasad, C. (1998). Food, mood and health: A neurobiologic outlook. Braz. J. Med. Biol. Res. 31, 1517–1527. Radulovic, J., Ruhmann, A., Liepold, T., and Spiess, J. (1999). Modulation of learning and anxiety by corticotropin-releasing factor (CRF) and stress: Differential roles of CRF receptors 1 and 2. J. Neurosci. 19, 5016–5025. Rakic, P., and Goldman-Rakic, P. S. (1982). The development and modifiability of the cerebral cortex: Overview. Neurosci. Res. Program Bull. 20, 433–438. Ramos, A. J., Tagliaferro, P., Lopez-Costa, J. J., Lopez, E. M., Pecci Saavedra, J., and Brusco, A. (2002). Neuronal and inducible nitric oxide synthase immunoreactivity following serotonin depletion. Brain Res. 958, 112–121. Rapport, M. M., Green, A. A., and Page, I. H. (1948). Partial purification of the vasoconstrictor in beef serum. J. Biol. Chem. 174, 735–738. Raymond, G. V., Bauman, M. L., and Kemper, T. L. (1996). Hippocampus in autism: A Golgi analysis. Acta Neuropathol. (Berl.) 91, 117–119. Read, J., Perry, B. D., Moskowitz, A., and Connolly, J. (2001). The contribution of early traumatic events to schizophrenia in some patients: A traumagenic neurodevelopmental model. Psychiatry 64, 319–345. Remmers, J. E. (1990). Sleeping and breathing. Chest 97, 77S–80S. Rhoades, R. W., Mooney, R. D., Chiaia, N. L., and Bennett-Clarke, C. A. (1990). Development and plasticity of the serotoninergic projection to the hamster’s superior colliculus. J. Comp. Neurol. 299, 151–166. Rioux, A., Fabre, V., Lesch, K. P., Moessner, R., Murphy, D. L., Lanfumey, L., Hamon, M., and Martres, M. P. (1999). Adaptive changes of serotonin 5-HT2A receptors in mice lacking the serotonin transporter. Neurosci. Lett. 262, 113–116. Ritvo, E. R., Yuwiler, A., Geller, E., Ornitz, E. M., Saeger, K., and Plotkin, S. (1970). Increased blood serotonin and platelets in early infantile autism. Arch. Gen. Psychiat. 23, 566–572. Robinson, T. E., and Becker, J. B. (1986). Enduring changes in brain and behavior produced by chronic amphetamine administration: A review and evaluation of animal models of amphetamine psychosis. Brain Res. 396, 157–198. Rogan, M. T., and LeDoux, J. E. (1996). Emotion: Systems, cells, synaptic plasticity. Cell 85, 469–475. Roskams, A. J., Bredt, D. S., Dawson, T. M., and Ronnett, G. V. (1994). Nitric oxide mediates the formation of synaptic connections in developing and regenerating neurons. Neuron 13, 289–299. Rosoklija, G., Toomayan, G., Ellis, S. P., Keilp, J., Mann, J. J., Latov, N., Hays, A. P., and Dwork, A. J. (2000). Structural abnormalities of subicular dendrites in subjects with schizophrenia and mood disorders: Preliminary findings. Arch. Gen. Psychia. 57, 349–356. Rothman, R. B., and Baumann, M. H. (2002). Serotonin releasing agents: Neurochemical therapeutic and adverse effects. Pharmacol. Biochem. Be 71, 825–836. Rubenstein, J. L. (1998). Development of serotonergic neurons and their projections. Biol. Psychiat. 44, 145–150. Sanders-Bush, E., and Breeding, M. (1991). Choroid plexus epithelial cells in primary culture: A model of 5HT1C receptor activation by hallucinogenic drugs. Psychopharmacology (Berl.) 105, 340–346. Sanders-Bush, E., Burris, K. D., and Knoth, K. (1988). Lysergic acid diethylamide and 2,5- dimethoxy-4-methylamphetamine are partial agonists at serotonin receptors linked to phosphoinositide hydrolysis. J. Pharmacol. Exp. Ther. 246, 924–928. SEROTONIN AND BRAIN DEVELOPMENT 169

Sanders-Bush, E., Bushing, J. A., and Sulser, F. (1972a). Long-term effects of p-chloroamphetamine on tryptophan hydroxylase activity and on the levels of 5-hydroxytryptamine and 5- hydroxyindole acetic acid in brain. Eur. J. Pharmacol. 20, 385–388. Sanders-Bush, E., Bushing, J. A., and Sulser, F. (1972b). p-Chloroamphetamine: Inhibition of cerebral tryptophan hydroxylase. Biochem. Pharmacol. 21, 1501–1510. Sanders-Bush, E., and Conn, P. J. (1986). Effector systems coupled to serotonin receptors in brain: Serotonin stimulated phosphoinositide hydrolysis. Psychopharmacol. Bull. 22, 829–836. Sanders-Bush, E., Gallager, D. A., and Sulser, F. (1974). On the mechanism of brain 5- hydroxytryptamine depletion by p-chloroamphetamine and related drugs and the specificity of their action. Adv. Biochem. Psychopharmacol. 10, 185–194. Sargent, P. A., Kjaer, K. H., Bench, C. J., Rabiner, E. A., Messa, C., Meyer, J., Gunn, R. N., Grasby, P. M., and Cowen, P. J. (2000). Brain serotonin 1A receptor binding measured by positron emission tomography with [11C]WAY-100635: Effects of depression and antidepressant treatment. Arch. Gen. Psychiat. 57, 174–180. Sari, Y., Powrozek, T., and Zhou, F. C. (2001). Alcohol deters the outgrowth of serotonergic neurons at midgestation. J. Biomed. Sci. 8, 119–125. Sarihi, A., Fathollahi, Y., Motamedi, F., Naghdi, N., and Rashidy-Pour, A. (2003). Effects of lidocaine reversible inactivation of the median raphe nucleus on long-term potentiation and recurrent inhibition in the dentate gyrus of rat hippocampus. Brain Res. 962, 159–168. Sarihi, A., Motamedi, F., Naghdi, N., and Rashidy-Pour, A. (2000). Lidocaine reversible inactivation of the median raphe nucleus has no effect on reference memory but enhances working memory versions of the Morris water maze task. Behav. Brain Res. 114, 1–9. Sarihi, A., Motamedi, F., Rashidy-Pour, A., Naghdi, N., and Behzadi, G. (1999). Reversible inactivation of the median raphe nucleus enhances consolidation and retrieval but not acquisition of passive avoidance learning in rats. Brain Res. 817, 59–66. Sarnyai, Z., Sibille, E. L., Pavlides, C., Fenster, R. J., McEwen, B. S., and Toth, M. (2000). Impaired hippocampal-dependent learning and functional abnormalities in the hippocampus in mice lacking serotonin(1A) receptors. Proc. Natl. Acad. Sci. USA 97, 14731–14736. Satoh, J., Gallyas, F., Jr., Endoh, M., Yamamura, T., Kunishita, T., Kobayashi, T., and Tabira, T. (1992). Establishment of mouse-immortalized hybrid clones expressing characteristics of differ- entiated neurons derived from the cerebellar and brain stem regions. J. Neurobiol. 23, 905–919. Savic, I., Gulyas, B., Lindstrom, P., Halldin, C., and Farde, L. (2001). Mesial temporal lobe epilepsy is associated with limbic reductions in serotonin 5HT1A receptor binding. Neuroimage 13, S833. Schildkraut, J. J. (1965). The catecholamine hypothesis of affective disorders: A review of supporting evidence. Am. J. Psychiat. 122, 509–522. Scott, B. S., Becker, L. E., and Petit, T. L. (1983). Neurobiology of Down’s syndrome. Prog. Neurobiol. 21, 199–237. Seidl,R., Kaehler, S. T.,Prast, H., Singewald, N., Cairns,N., Gratzer, M., and Lubec, G.(1999). Serotonin (5-HT) in brains of adult patients with Down syndrome. J. Neural Transm. Suppl. 57, 221–232. Selemon, L. D., and Goldman-Rakic, P. S. (1999). The reduced neuropil hypothesis: A circuit based model of schizophrenia. Biol. Psychiat. 45, 17–25. Selemon, L. D., Rajkowska, G., and Goldman-Rakic, P. S. (1995). Abnormally high neuronal density in the schizophrenic cortex: A morphometric analysis of prefrontal area 9 and occipital area 17. Arch. Gen. Psychiat. 52, 805–818; discussion 819–820. Selemon, L. D., Rajkowska, G., and Goldman-Rakic, P. S. (1998). Elevated neuronal density in prefrontal area 46 in brains from schizophrenic patients: Application of a three-dimensional, stereologic counting method. J. Comp. Neurol. 392, 402–412. Shen, W. Z., Luo, Z. B., Zheng, D. R., and Yew, D. T. (1989). Immunohistochemical studies on the development of 5-HT (serotonin) neurons in the nuclei of the reticular formations of human fetuses. Pediatr. Neurosci. 15, 291–295. 170 SODHI AND SANDERS-BUSH

Shuey, D. L., Sadler, T. W., and Lauder, J. M. (1992). Serotonin as a regulator of craniofacial morphogenesis: Site specific malformations following exposure to serotonin uptake inhibitors. Teratology 46, 367–378. Shuey, D. L., Sadler, T. W., Tamir, H., and Lauder, J. M. (1993). Serotonin and morphogenesis: Transient expression of serotonin uptake and binding protein during craniofacial morphogenesis in the mouse. Anat. Embryol. (Berl.) 187, 75–85. Shutoh, F., Hamada, S., Shibata, M., Narita, M., Shiga, T., Azmitia, E. C., and Okado, N. (2000). Long term depletion of serotonin leads to selective changes in subunits. Neurosci. Res. 38, 365–371. Sibille, E., Pavlides, C., Benke, D., and Toth, M. (2000). Genetic inactivation of the Serotonin(1A) receptor in mice results in downregulation of major GABA(A) receptor alpha subunits, reduction of GABA(A) receptor binding, and -resistant anxiety. J. Neurosci. 20, 2758–2765. Silva, A. J. (2003). Molecular and cellular cognitive studies of the role of synaptic plasticity in memory. J. Neurobiol. 54, 224–237. Simpson, M. D., Lubman, D. I., Slater, P., and Deakin, J. F. (1996). Autoradiography with [3H]8- OH-DPAT reveals increases in 5-HT(1A) receptors in ventral prefrontal cortex in schizophrenia. Biol. Psychiat. 39, 919–928. Slade, P. D. (1976). An investigation of psychological factors involved in the predisposition to auditory hallucinations. Psychol. Med. 6, 123–132. Smith, K. A., Fairburn, C. G., and Cowen, P. J. (1997). Relapse of depression after rapid depletion of tryptophan. Lancet 349, 915–919. Soares, C. N., Almeida, O. P., Joffe, H., and Cohen, L. S. (2001). Efficacy of estradiol for the treatment of depressive disorders in perimenopausal women: A double-blind, randomized, placebo-controlled trial. Arch. Gen. Psychiat. 58, 529–534. Soares, C. N., and Cohen, L. S. (2001). The perimenopause, depressive disorders, and hormonal variability. Sao Paulo Med. J. 119, 78–83. Sodhi, M. S., Aitchison, K. J., Kerwin, R. W., and Sham, P. (1999). The role of the 5-HT2A receptor gene in psychosis. Am. J. Hum. Genet. 65, A469. Sodhi, M. S., Arranz, M. J., Curtis, D., Ball, D. M., Sham, P., Roberts, G. W., Price, J., Collier, D. A., and Kerwin, R. W. (1995). Association between clozapine response and allelic variation in the 5-HT2C receptor gene. Neuroreport 7, 169–172. Sodhi, M. S., Burnet, P. W. J., Makoff, A. J., Kerwin, R. W., and Harrison, P. J. (2001). RNA editing of the 5-HT2C receptor is reduced in schizophrenia. Mol. Psychiat. 6, 373–379. Sodhi, M. S., and Murray, R. M. (1997). Future therapies for schizophrenia. Expert. Opin. Ther. Pat. 7, 151–165. Sodhi, M. S., Sham, P. C., Makoff, A. J., Collier, D. A., Arranz, M. J., Munro, J., and Kerwin, R. W. (1999b). Replication and meta-analysis of allelic association of a 5-HT2C receptor polymorph- ism with good response to clozapine. Mol. Psychiat. 4, S94–S95. Southam, E., and Garthwaite, J. (1993). The nitric oxide-cyclic GMP signalling pathway in rat brain. Neuropharmacology 32, 1267–1277. Spurlock, G., Heils, A., Holmans, P., Williams, J., D’Souza, U. M., Cardno, A., Murphy, K. C., Jones, L., Buckland, P. R., McGuffin, P., Lesch, K. P., and Owen, M. J. (1998). A family based association study of T102C polymorphism in 5HT2A and schizophrenia plus identification of new polymorphisms in the promoter. Mol. Psychiat. 3, 42–49. Staner, L., Kempenaers, C., Simonnet, M. P., Fransolet, L., and Mendlewicz, J. (1992). 5-HT2 receptor antagonism and slow-wave sleep in major depression. Acta Psychiatr Scand 86, 133–137. Stanley, M., Mann, J. J., and Cohen, L. S. (1986a). Role of the serotonergic system in the postmortem analysis of suicide. Psychopharmacol. Bull. 22, 735–740. SEROTONIN AND BRAIN DEVELOPMENT 171

Stanley, M., Mann, J. J., and Cohen, L. S. (1986b). Serotonin and serotonergic receptors in suicide. Ann. N. Y. Acad. Sci. 487, 122–127. Statnick, M. A., Dailey, J. W., Jobe, P. C., and Browning, R. A. (1996). Abnormalities in brain serotonin concentration, high-affinity uptake, and tryptophan hydroxylase activity in severe- seizure genetically epilepsy-prone rats. Epilepsia 37, 311–321. Stiedl, O., Misane, I., Spiess, J., and Ogren, S. O. (2000). Involvement of the 5-HT1A receptors in classical fear conditioning in C57BL/6J mice. J. Neurosci. 20, 8515–8527. Stober, G., Franzek, E., Haubitz, I., Pfuhlmann, B., and Beckmann, H. (1998). Gender differences and age of onset in the catatonic subtypes of schizophrenia. Psychopathology 31, 307–312. Stockmeier, C. A., Dilley, G. E., Shapiro, L. A., Overholser, J. C., Thompson, P. A., and Meltzer, H. Y. (1997). Serotonin receptors in suicide victims with major depression. Neuropsychopharmacology. 16, 162–173. Stolz, J. F., Marsden, C. A., and Middlemiss, D. N. (1983). Effect of chronic antidepressant treatment and subsequent withdrawal on [3H]-5-hydroxytryptamine and [3H]-spiperone binding in rat frontal cortex and serotonin receptor mediated behaviour. Psychopharmacology (Berl.) 80, 150–155. Stunkard, A. J., and Allison, K. C. (2003). Two forms of disordered eating in obesity: Binge eating and night eating. Int. J. Obesity 27, 1–12. Sumiyoshi, T., Matsui, M., Yamashita, I., Nohara, S., Kurachi, M., Uehara, T., Sumiyoshi, S., Sumiyoshi, C., and Meltzer, H. Y. (2001). The effect of , a serotonin(1A) agonist, on memory function in schizophrenia. Biol. Psychiat. 49, 861–868. Sumiyoshi, T., Stockmeier, C. A., Overholser, J. C., Dilley, G. E., and Meltzer, H. Y. (1996). Serotonin(1A) receptors are increased in postmortem prefrontal cortex in schizophrenia. Brain Res. 708, 209–214. Sundstrom, E., Kolare, S., Souverbie, F., Samuelsson, E. B., Pschera, H., Lunell, N. O., and Seiger, A. (1993). Neurochemical differentiation of human bulbospinal monoaminergic neurons during the first trimester. Brain Res. Dev. Brain Res. 75, 1–12. Szara, S. (1967). The hallucinogenic drugs: Curse or blessing? Am. J. Psychiat. 123, 1513–1518. Tajuddin, N. F., and Druse, M. J. (1993). Treatment of pregnant alcohol-consuming rats with buspirone: effects on serotonin and 5-hydroxyindoleacetic acid content in offspring. Alcohol Clin. Exp. Res. 17, 110–114. Takashima, S., lida, K., Mito, T., and Arima, M. (1994). Dendritic and histochemical development and ageing in patients with Down’s syndrome. J. Intellect. Disabil. Res. 38(Pt.3), 265–273. Tecott, L. H., Sun, L. M., Akana, S. F., Strack, A. M., Lowenstein, D. H., Dallman, M. F., and Julius, D. (1995). Eating disorder and epilepsy in mice lacking 5-Ht2c serotonin receptors. Nature 374, 542–546. Teyler, T. J., and Discenna, P. (1984). Long-term potentiation as a candidate mnemonic device. Brain Res. 319, 15–28. Thorin, E., Capdeville, C., Trockle, G., Wiernsperger, N., and Atkinson, J. (1990). Chronic treatment with attenuates the development of vascular hypersensitivity to serotonin in the spontaneously hypertensive rat. J. Cardiovasc. Pharm. 16, S54–S57. Titeler, M., Lyon, R. A., and Glennon, R. A. (1988). Radioligand binding evidence implicates the brain 5-HT2 receptor as a site of action for LSD and phenylisopropylamine hallucinogens. Psychopharmacology (Berl.) 94, 213–216. Toth, G., and Fekete, M. (1986). 5-Hydroxyindole acetic in newborns, infants and children. Acta Paediatr. Hung. 27, 221–226. Treiser, S., and Kellar, K. J. (1980). Lithium: Effects on serotonin receptors in rat brain. Eur. J. Pharmacol. 64, 183–185. Tu, J. B., and Zellweger, H. (1965). Blood-serotonin deficiency in Down’s syndrome. Lancet 2, 715–716. 172 SODHI AND SANDERS-BUSH

Turlejski, K. (1996). Evolutionary ancient roles of serotonin: Long-lasting regulation of activity and development. Acta Neurobiol. Exp. 56, 619–636. Twarog, B. M., and Page, I. H. (1953). Serotonin content of some mammalian tissues and urine and a method for its determination. Am. J. Physiol. 175, 157–161. Ueda, S., Aikawa, M., Kawata, M., Naruse, I., Whitaker-Azmitia, P. M., and Azmitia, E. C. (1996). Neuro-glial neurotrophic interaction in the S-100 beta retarded mutant mouse (Polydactyly Nagoya). III. Transplantation study. Brain Res. 738, 15–23. Ueda, S., Gu, X. F., Whitaker-Azmitia, P. M., Naruse, I., and Azmitia, E. C. (1994a). Neuroglial neurotrophic interaction in the S-100 beta retarded mutant mouse (Polydactyly Nagoya). I. Immunocytochemical and neurochemical studies. Brain Res. 633, 275–283. Ueda, S., Hou, X. P., Whitaker-Azmitia, P. M., and Azmitia, E. C. (1994b). Neuro-glial neurotrophic interaction in the S-100 beta retarded mutant mouse (Polydactyly Nagoya). II. Co-cultures study. Brain Res. 633, 284–288. Ueda, S., Kokotos Leonardi, E. T., Bell, J., III, and Azmitia, E. C. (1995). Serotonergic sprouting into transplanted C-6 gliomas is blocked by S-100 beta antisense gene. Brain Res. Mol. Brain Res. 29, 365–368. Vaidya, V. A., Terwilliger, R. M., and Duman, R. S. (1999). Role of 5-HT2A receptors in the stress- induced down-regulation of brain-derived neurotrophic factor expression in rat hippocampus. Neurosci. Lett. 262, 1–4. van Os, J., Takei, N., Castle, D. J., Wessely, S., Der, G., and Murray, R. M. (1995). Premorbid abnormalities in mania, schizomania, acute schizophrenia and chronic schizophrenia. Soc. Psychiat. Psychiat. Epidemiol. 30, 274–278. van Riel, E., Meijer, O. C., Veenema, A. H., and Joels, M. (2002). Hippocampal serotonin responses in short and long attack latency mice. J. Neuroendocrinol. 14, 234–239. Vitalis, T., Cases, O., Callebert, J., Launay, J. M., Price, D. J., Seif, I., and Gaspar, P. (1998). Effects of monoamine oxidase A inhibition on barrel formation in the mouse somatosensory cortex: Determination of a sensitive developmental period. J. Comp. Neurol. 393, 169–184. Vollenweider, F. X., and Geyer, M. A. (2001). A systems model of altered consciousness: Integrating natural and drug-induced psychoses. Brain Res. Bull. 56, 495–507. Vollenweider, F. X., Vollenweider-Scherpenhuyzen, M. F., Babler, A., Vogel, H., and Hell, D. (1998). induces schizophrenia-like psychosis in humans via a serotonin-2 agonist action. Neuroreport 9, 3897–3902. Vu, D. H., and Tork, I. (1992). Differential development of the dual serotoninergic fiber system in the cerebral cortex of the cat. J. Comp. Neurol. 317, 156–174. Wajda, I. J., Banay-Schwartz, M., Manigault, I., and Lajtha, A. (1986). Modulation of the serotonin S2-receptor in brain after chronic lithium. Neurochem. Res. 11, 949–957. Wallace, J. A., and Lauder, J. M. (1983). Development of the serotonergic system in the rat embryo: An immunocytochemical study. Brain Res. Bull. 10, 459–479. Wang, S. J., Cheng, L. L., and Gean, P. W. (1999). Cross-modulation of synaptic plasticity by beta-adrenergic and 5-HT1A receptors in the rat basolateral amygdala. J. Neurosci. 19, 570–577. Wang, Y., Gu, Q., and Cynader, M. S. (1997). Blockade of serotonin-2C receptors by mesulergine reduces ocular dominance plasticity in kitten visual cortex. Exp. Brain Res. 114, 321–328. Watanabe, Y., Sakai, R. R., McEwen, B. S., and Mendelson, S. (1993). Stress and antidepressant effects on hippocampal and cortical 5-HT1A and 5-HT2 receptors and transport sites for serotonin. Brain Res. 615, 87–94. Waxham, M. N., Grotta, J. C., Silva, A. J., Strong, R., and Aronowski, J. (1996). Ischemia-induced neuronal damage: A role for calcium/calmodulin-dependent protein kinase II. J. Cereb Blood Flow Metab. 16, 1–6. Wechsler, J. G. (1998). Drug treatment of obesity. Acta Med. Aust. 25, 138–141. SEROTONIN AND BRAIN DEVELOPMENT 173

Weinberger, D. R. (1987). Implications of normal brain development for the pathogenesis of schizophrenia. Arch. Gen. Psychiat. 44, 660–669. Weinberger, D. R. (1995). From neuropathology to neurodevelopment. Lancet 346, 552–557. Weinberger, D. R. (1999). Cell biology of the hippocampal formation in schizophrenia. Biol. Psychiat. 45, 395–402. Weinberger, D. R., and Berman, K. F. (1996). Prefrontal function in schizophrenia: Confounds and controversies. Phil. Trans. R. Soc. Lond. B Biol. Sci. 351, 1495–1503. Weissman, N. J. (2001). Appetite suppressants and valvular heart disease. Am. J. Med. Sci. 321, 285–291. Whitaker-Azmitia, P., Zhou, F., Hobin, J., and Borella, A. (2000). Isolation-rearing of rats produces deficits as adults in the serotonergic innervation of hippocampus. Peptides 21, 1755–1759. Whitaker-Azmitia, P. M. (1998). Role of the neurotrophic properties of serotonin in the delay of brain maturation induced by cocaine. Ann. N. Y. Acad. Sci. 846, 158–164. Whitaker-Azmitia, P. M. (2001). Serotonin and brain development: Role in human developmental diseases. Brain Res. Bull. 56, 479–485. Whitaker-Azmitia, P. M., Borella, A., and Raio, N. (1995). Serotonin depletion in the adult rat causes loss of the dendritic marker MAP-2: A new animal model of schizophrenia? Neuropsychopharma- cology 12, 269–272. Whitaker-Azmitia, P. M., Druse, M., Walker, P., and Lauder, J. M. (1996). Serotonin as a developmental signal. Behav. Brain Res. 73, 19–29. Whitaker-Azmitia, P. M., Lauder, J. M., Shemmer, A., and Azmitia, E. C. (1987). Postnatal changes in serotonin receptors following prenatal alterations in serotonin levels: Further evidence for functional fetal serotonin receptors. Brain Res. 430, 285–289. Whitaker-Azmitia, P. M., Molino, L. J., Caruso, J., and Shemer, A. V. (1990). Serotonergic agents restore appropriate decision-making in neonatal rats displaying dopamine D1 receptor-mediated vacillatory behavior. Eur. J. Pharmacol. 180, 305–309. Williams, J., McGuffin, P., Nothen, M., and Owen, M. J. (1997). Meta-analysis of association between the 5-HT2a receptor T102C polymorphism and schizophrenia. EMASS Collaborative Group. European Multicentre Association Study of Schizophrenia. Lancet 349, 1221. Williams, J., Spurlock, G., McGuffin, P., Mallet, J., Nothen, M. M., Gill, M., Aschauer, H., Nylander, P. O., Macciardi, F., and Owen, M. J. (1996). Association between schizophrenia and T102C polymorphism of the 5-hydroxytryptamine type 2a-receptor gene. European Multicentre Association Study of Schizophrenia (EMASS) Group. Lancet 347, 1294–1296. Wilson, C. C., Faber, K. M., and Haring, J. H. (1998). Serotonin regulates synaptic connections in the dentate molecular layer of adult rats via 5-HT1a receptors: Evidence for a glial mechanism. Brain Res. 782, 235–239. Wissink, S., Meijer, O., Pearce, D., van Der Burg, B., and van Der Saag, P. T. (2000). Regulation of the rat serotonin-1A receptor gene by corticosteroids. J. Biol. Chem. 275, 1321–1326. Woods, B. T. (1998). Is schizophrenia a progressive neurodevelopmental disorder? Toward a unitary pathogenetic mechanism Am. J. Psychiatr. 155, 1661–1670. Wright, I. C., Rabe-Hesketh, S., Woodruff, P. W., David, A. S., Murray, R. M., and Bullmore, E. T. (2000). Meta-analysis of regional brain volumes in schizophrenia. Am. J. Psychiat. 157, 16–25. Wu, C., Singh, S. K., Dias, P., Kumar, S., and Mann, D. M. (1999). Activated astrocytes display increased 5-HT2a receptor expression in pathological states. Exp. Neurol. 158, 529–533. Yamashita, Y., Ito, T., Hashimoto, I., Ohyama, A., and Kuriyama, K. (1989). Changes in cerebral level of monoamines by Japanese encephalitis virus infection. Uirusu 39, 47–54. Yan, Q. S., Jobe, P. C., Cheong, J. H., Ko, K. H., and Dailey, J. W. (1994). Role of serotonin in the anticonvulsant effect of fluoxetine in genetically epilepsy-prone rats. N-S Arch. Pharmacol. 350, 149–152. 174 SODHI AND SANDERS-BUSH

Yan, W., Wilson, C. C., and Haring, J. H. (1997a). 5-HT1a receptors mediate the neurotrophic effect of serotonin on developing dentate granule cells. Brain Res. Dev. Brain Res. 98, 185–190. Yan, W., Wilson, C. C., and Haring, J. H. (1997b). Effects of neonatal serotonin depletion on the development of rat dentate granule cells. Brain Res. Dev. Brain Res. 98, 177–184. Yang, Y., and Schmitt, H. P. (2001). Frontotemporal dementia: Evidence for impairment of ascending serotoninergic but not noradrenergic innervation. Immunocytochemical and quantitative study using a graph method. Acta Neuropathol. (Berl.) 101, 256–270. Yates, M., Leake, A., Candy, J. M., Fairbairn, A. F., McKeith, I. G., and Ferrier, I. N. (1990). 5HT2 receptor changes in major depression. Biol. Psychiat. 27, 489–496. Yavarone, M. S., Shuey, D. L., Tamir, H., Sadler, T. W., and Lauder, J. M. (1993). Serotonin and cardiac morphogenesis in the mouse embryo. Teratology 47, 573–584. Yocca, F. D., Eison, A. S., Hyslop, D. K., Ryan, E., Taylor, D. P., and Gianutsos, G. (1991). Unique modulation of central 5-HT2 receptor binding sites and 5-HT2 receptor-mediated behavior by continuous treatment. Life Sci. 49, 1777–1785. Yocca, F. D., Wright, R. N., Margraf, R. R., and Eison, A. S. (1990). 8-OH-DPAT and buspirone analogs inhibit the ketanserin-sensitive -induced head shake response in rats. Pharmacol. Biochem. Behav. 35, 251–254. Young, K. A., Manaye, K. F., Liang, C., Hicks, P. B., and German, D. C. (2000). Reduced number of mediodorsal and anterior thalamic neurons in schizophrenia. Biol. Psychiat. 47, 944–953. Zhang, Z. W. (2003). Serotonin induces tonic firing in layer V pyramidal neurons of rat prefrontal cortex during postnatal development. J. Neurosci. 23, 3373–3384. Zhou, F. C., Sari, Y., Zhang, J. K., Goodlett, C. R., and Li, T. (2001). Prenatal alcohol exposure retards the migration and development of serotonin neurons in fetal C57BL mice. Brain Res. Dev Brain Res. 126, 147–155. Zilles, K., Wu, J., Crusio, W. E., and Schwegler, H. (2000). Water maze and radial maze learning and the density of binding sites of glutamate, GABA, and serotonin receptors in the hippocampus of inbred mouse strains. Hippocampus 10, 213–225.