+ All documents
Home > Documents > Multiplicity of mechanisms of serotonin receptor signal transduction

Multiplicity of mechanisms of serotonin receptor signal transduction

Date post: 09-Dec-2023
Category:
Upload: susu
View: 0 times
Download: 0 times
Share this document with a friend
34
Associate editor: B.L. Roth Multiplicity of mechanisms of serotonin receptor signal transduction John R. Raymond a,b, * , Yurii V. Mukhin b , Andrew Gelasco a,b , Justin Turner b , Georgiann Collinsworth a,b , Thomas W. Gettys c , Jasjit S. Grewal b , Maria N. Garnovskaya a,b a The Research Service of the Ralph H. Johnson Veterans Affairs Medical Center, Charleston, SC 29401, USA b Departments of Medicine (Nephrology Division), Medical University of South Carolina, 171 Ashley Avenue, Charleston, SC 29425, USA c Departments of Medicine (Gastroenterology Division) and Biochemistry and Molecular Biology, Medical University of South Carolina, 171 Ashley Avenue, Charleston, SC 29425, USA Abstract The serotonin (5-hydroxytryptamine, 5-HT) receptors have been divided into 7 subfamilies by convention, 6 of which include 13 different genes for G-protein-coupled receptors. Those subfamilies have been characterized by overlapping pharmacological properties, amino acid sequences, gene organization, and second messenger coupling pathways. Post-genomic modifications, such as alternative mRNA splicing or mRNA editing, creates at least 20 more G-protein-coupled 5-HT receptors, such that there are at least 30 distinct 5-HT receptors that signal through G-proteins. This review will focus on what is known about the signaling linkages of the G-protein-linked 5-HT receptors, and will highlight some fascinating new insights into 5-HT receptor signaling. D 2001 Elsevier Science Inc. All rights reserved. Keywords: Signal transduction; Adenylyl cyclase; Phospholipase; Kinase; Channel Abbreviations: AA, arachidonic acid; AC, adenylyl cyclase; CaM, calmodulin; cAMP, cyclic AMP; cGMP, cyclic GMP; CHO, Chinese hamster ovary; cNOS, constitutive nitric oxide synthase; iNOS, inducible nitric oxide synthase; DAG, diacylglycerol; ERK, extracellular signal-regulated kinase; GABA, g-aminobutyric acid; GIRK, G-protein-gated inwardly rectifying K + ; GPCR, G-protein-coupled receptor; HEK, human embryonic kidney; 5-HT, 5-hydroxytryptamine, serotonin; IP 3 , inositol trisphosphate; IRK, inward rectifier K + ; Jak, Janus kinase; LSD, lysergic acid diethylamide; MAP, mitogen- activated protein; MEK, mitogen and extracellular signal-regulated kinase; MUPP1, multi-PS-95 discs-large ZO-1 interaction motif-domain protein; NMDA, N-methyl-D-aspartic acid; NO, nitric oxide; NOS, nitric oxide synthase; 8-OH-DPAT, (±)-8-hydroxy-2-(di-N-propylamino)tetralin; PC-PLC, phosphatidylcholine-specific phospholipase C; PDGF, platelet-derived growth factor; PDZ, PS-95 discs-large ZO-1, which is a type of protein-protein interaction motif; PI, phosphatidylinositol; PI-PLC, phosphatidylinositol-specific phospholipase C; PKA, protein kinase A; PKC, protein kinase C; PL, phospholipase; RGS, regulators of G-protein signaling; SRL, serotonin receptor-like; STAT, signal transducers and activators of transcription. Contents 1. Introduction ............................................ 181 2. The 5-hydroxytryptamine 1 receptors ............................... 182 2.1. The 5-hydroxytryptamine 1A receptor ........................... 182 2.1.1. The 5-hydroxytryptamine 1A receptor both inhibits and activates adenylyl cyclase 184 2.1.2. The 5-hydroxytryptamine 1A receptor activates and inhibits phosphatidylinositol- specific phospholipase C ............................. 185 2.1.3. The 5-hydroxytryptamine 1A receptor activates other phospholipases ....... 185 2.1.4. The 5-hydroxytryptamine 1A receptor activates protein kinase C ......... 185 2.1.5. The 5-hydroxytryptamine 1A receptor activates the extracellular signal-regulated mitogen-activated protein kinase ......................... 185 2.1.6. The 5-hydroxytryptamine 1A receptor stimulates phosphatidylinositol-3 0 kinase. . 186 0163-7258/01/$ – see front matter D 2001 Elsevier Science Inc. All rights reserved. PII:S0163-7258(01)00169-3 * Corresponding author: Room 829 CSB, Medical University of South Carolina, 171 Ashley Avenue, Charleston, SC 29425, USA. Tel.: 843-792-4122; fax: 843-792-8399. E-mail address: [email protected] (J.R. Raymond). Pharmacology & Therapeutics 92 (2001) 179 – 212
Transcript

Associate editor: B.L. Roth

Multiplicity of mechanisms of serotonin receptor signal transduction

John R. Raymonda,b,*, Yurii V. Mukhinb, Andrew Gelascoa,b, Justin Turnerb,Georgiann Collinswortha,b, Thomas W. Gettysc, Jasjit S. Grewalb, Maria N. Garnovskayaa,b

aThe Research Service of the Ralph H. Johnson Veterans Affairs Medical Center, Charleston, SC 29401, USAbDepartments of Medicine (Nephrology Division), Medical University of South Carolina, 171 Ashley Avenue, Charleston, SC 29425, USA

cDepartments of Medicine (Gastroenterology Division) and Biochemistry and Molecular Biology,

Medical University of South Carolina, 171 Ashley Avenue, Charleston, SC 29425, USA

Abstract

The serotonin (5-hydroxytryptamine, 5-HT) receptors have been divided into 7 subfamilies by convention, 6 of which include 13 different

genes for G-protein-coupled receptors. Those subfamilies have been characterized by overlapping pharmacological properties, amino acid

sequences, gene organization, and second messenger coupling pathways. Post-genomic modifications, such as alternative mRNA splicing or

mRNA editing, creates at least 20 more G-protein-coupled 5-HT receptors, such that there are at least 30 distinct 5-HT receptors that signal

through G-proteins. This review will focus on what is known about the signaling linkages of the G-protein-linked 5-HT receptors, and will

highlight some fascinating new insights into 5-HT receptor signaling. D 2001 Elsevier Science Inc. All rights reserved.

Keywords: Signal transduction; Adenylyl cyclase; Phospholipase; Kinase; Channel

Abbreviations: AA, arachidonic acid; AC, adenylyl cyclase; CaM, calmodulin; cAMP, cyclic AMP; cGMP, cyclic GMP; CHO, Chinese hamster ovary;

cNOS, constitutive nitric oxide synthase; iNOS, inducible nitric oxide synthase; DAG, diacylglycerol; ERK, extracellular signal-regulated kinase;

GABA, g-aminobutyric acid; GIRK, G-protein-gated inwardly rectifying K+; GPCR, G-protein-coupled receptor; HEK, human embryonic kidney; 5-HT,

5-hydroxytryptamine, serotonin; IP3, inositol trisphosphate; IRK, inward rectifier K+; Jak, Janus kinase; LSD, lysergic acid diethylamide; MAP, mitogen-

activated protein; MEK, mitogen and extracellular signal-regulated kinase; MUPP1, multi-PS-95 discs-large ZO-1 interaction motif-domain protein;

NMDA, N-methyl-D-aspartic acid; NO, nitric oxide; NOS, nitric oxide synthase; 8-OH-DPAT, (±)-8-hydroxy-2-(di-N-propylamino)tetralin; PC-PLC,

phosphatidylcholine-specific phospholipase C; PDGF, platelet-derived growth factor; PDZ, PS-95 discs-large ZO-1, which is a type of protein-protein

interaction motif; PI, phosphatidylinositol; PI-PLC, phosphatidylinositol-specific phospholipase C; PKA, protein kinase A; PKC, protein kinase C; PL,

phospholipase; RGS, regulators of G-protein signaling; SRL, serotonin receptor-like; STAT, signal transducers and activators of transcription.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181

2. The 5-hydroxytryptamine1 receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182

2.1. The 5-hydroxytryptamine1A receptor . . . . . . . . . . . . . . . . . . . . . . . . . . . 182

2.1.1. The 5-hydroxytryptamine1A receptor both inhibits and activates adenylyl cyclase 184

2.1.2. The 5-hydroxytryptamine1A receptor activates and inhibits phosphatidylinositol-

specific phospholipase C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185

2.1.3. The 5-hydroxytryptamine1A receptor activates other phospholipases . . . . . . . 185

2.1.4. The 5-hydroxytryptamine1A receptor activates protein kinase C . . . . . . . . . 185

2.1.5. The 5-hydroxytryptamine1A receptor activates the extracellular signal-regulated

mitogen-activated protein kinase . . . . . . . . . . . . . . . . . . . . . . . . . 185

2.1.6. The 5-hydroxytryptamine1A receptor stimulates phosphatidylinositol-30 kinase. . 186

0163-7258/01/$ – see front matter D 2001 Elsevier Science Inc. All rights reserved.

PII: S0163 -7258 (01 )00169 -3

* Corresponding author: Room 829 CSB, Medical University of South Carolina, 171 Ashley Avenue, Charleston, SC 29425, USA. Tel.: 843-792-4122;

fax: 843-792-8399.

E-mail address: [email protected] (J.R. Raymond).

Pharmacology & Therapeutics 92 (2001) 179–212

2.1.7. The 5-hydroxytryptamine1A receptor causes production of reactive oxygen and

nitrogen species. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186

2.1.8. The 5-hydroxytryptamine1A receptor regulates K+ channels . . . . . . . . . . . 186

2.1.9. The 5-hydroxytryptamine1A receptor inhibits neuronal Ca2+ conductances . . . . 187

2.1.10. The 5-hydroxytryptamine1A receptor regulates other ion channels . . . . . . . . 187

2.1.11. The 5-hydroxytryptamine1A receptor regulates miscellaneous transport processes 187

2.1.12. The 5-hydroxytryptamine1A receptor regulates proliferation and related pathways 187

2.1.13. What is the basis for the multiplicity of 5-hydroxytryptamine1Areceptor signaling? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188

2.2. The 5-hydroxytryptamine1B receptor . . . . . . . . . . . . . . . . . . . . . . . . . . . 188

2.2.1. The 5-hydroxytryptamine1B receptor regulates adenylyl cyclase . . . . . . . . . 188

2.2.2. The 5-hydroxytryptamine1B receptor activates phospholipases . . . . . . . . . . 189

2.2.3. The 5-hydroxytryptamine1B receptor regulates the extracellular signal-regulated

kinase, phosphatidylinositol-30 kinase, p70 S6 kinase, and Akt kinase . . . . . . 189

2.2.4. The 5-hydroxytryptamine1B receptor stimulates endothelial nitric oxide production 189

2.2.5. Other signals of the 5-hydroxytryptamine1B receptor . . . . . . . . . . . . . . . 189

2.3. The 5-hydroxytryptamine1D receptor . . . . . . . . . . . . . . . . . . . . . . . . . . . 189

2.3.1. The 5-hydroxytryptamine1D receptor inhibits adenylyl cyclase . . . . . . . . . . 190

2.3.2. The 5-hydroxytryptamine1D receptor regulates ion channels . . . . . . . . . . . 190

2.3.3. The 5-hydroxytryptamine1D receptor is mitogenic . . . . . . . . . . . . . . . . 190

2.3.4. The 5-hydroxytryptamine1D receptor inhibits 5-hydroxytryptamine release . . . 190

2.4. The 5-hydroxytryptamine1E receptor. . . . . . . . . . . . . . . . . . . . . . . . . . . . 190

2.5. The 5-hydroxytryptamine1F receptor. . . . . . . . . . . . . . . . . . . . . . . . . . . . 190

3. The 5-hydroxytryptamine2 receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191

3.1. The 5-hydroxytryptamine2A receptor . . . . . . . . . . . . . . . . . . . . . . . . . . . 191

3.1.1. The 5-hydroxytryptamine2A receptor activates phospholipase C . . . . . . . . . 191

3.1.2. The 5-hydroxytryptamine2A receptor activates other phospholipases . . . . . . . 191

3.1.3. The 5-hydroxytryptamine2A receptor can regulate cyclic amp accumulation in

certain cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191

3.1.4. 5-hydroxytryptamine2A receptor activates the extracellular signal-regulated

mitogen-activated protein kinase . . . . . . . . . . . . . . . . . . . . . . . . . 192

3.1.5. The 5-hydroxytryptamine2A receptor activates the Janus kinase/signal transducers

and activators of transcription pathway . . . . . . . . . . . . . . . . . . . . . . 192

3.1.6. The 5-hydroxytryptamine2A receptor regulates apoptosis . . . . . . . . . . . . . 192

3.1.7. The 5-hydroxytryptamine2A receptor regulates channels . . . . . . . . . . . . . 192

3.1.8. The 5-hydroxytryptamine2A receptor causes production of reactive oxygen and

nitrogen species. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192

3.1.9. The 5-hydroxytryptamine2A receptor regulates calmodulin . . . . . . . . . . . . 192

3.1.10. The 5-hydroxytryptamine2A receptor regulates transport processes . . . . . . . 193

3.1.11. A unique signaling role for internalization of the 5-hydroxytryptamine2A receptor 193

3.2. The 5-hydroxytryptamine2B receptor . . . . . . . . . . . . . . . . . . . . . . . . . . . 193

3.2.1. The 5-hydroxytryptamine2B receptor activates phospholipase C . . . . . . . . . 193

3.2.2. The 5-hydroxytryptamine2B receptor can stimulate cyclic amp accumulation . . 193

3.2.3. The 5-hydroxytryptamine2B receptor regulates morphogenesis and mitogenesis . 193

3.2.4. The 5-hydroxytryptamine2B receptor activates the extracellular signal-regulated

kinase and cell cycle components . . . . . . . . . . . . . . . . . . . . . . . . 194

3.2.5. The 5-hydroxytryptamine2B receptor causes production of reactive

nitrogen species. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194

3.2.6. The 5-hydroxytryptamine2B receptor regulates channels . . . . . . . . . . . . . 194

3.2.7. The 5-hydroxytryptamine2B receptor regulates transport processes . . . . . . . . 194

3.3. The 5-hydroxytryptamine2C receptor . . . . . . . . . . . . . . . . . . . . . . . . . . . 194

3.3.1. The 5-hydroxytryptamine2C receptor activates phospholipase C . . . . . . . . . 195

3.3.2. The 5-hydroxytryptamine2C receptor can modulate cyclic amp accumulation . . 195

3.3.3. The 5-hydroxytryptamine2C receptor regulates channels . . . . . . . . . . . . . 195

3.3.4. The 5-hydroxytryptamine2C receptors regulate mitogenesis . . . . . . . . . . . 196

3.3.5. The 5-hydroxytryptamine2C receptor can regulate nitric oxide levels. . . . . . . 196

3.3.6. The 5-hydroxytryptamine2C receptor regulates transport processes . . . . . . . . 196

3.3.7. Possible signaling through PS-95 discs-large ZO-1 interaction motifs by

the 5-hydroxytryptamine2C receptor . . . . . . . . . . . . . . . . . . . . . . . 196

3.3.8. Multiple novel functional 5-hydroxytryptamine2C receptors are created by

mrna editing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212180

4. The 5-hydroxytryptamine4 receptor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197

4.1. The 5-hydroxytryptamine4 receptor activates adenylyl cyclase . . . . . . . . . . . . . . 197

4.2. The 5-hydroxytryptamine4 receptor activates protein kinase A . . . . . . . . . . . . . . 197

4.3. The 5-hydroxytryptamine4 receptor regulates channels . . . . . . . . . . . . . . . . . . 197

4.4. The 5-hydroxytryptamine4 receptor regulates transport . . . . . . . . . . . . . . . . . . 198

4.5. Functional differences among the 5-hydroxytryptamine4 receptor splice variants . . . . . 198

5. The 5-hydroxytryptamine5 receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198

5.1. Functional coupling of the 5-hydroxytryptamine5 receptors to signaling pathways . . . . 198

6. The 5-hydroxytryptamine6 receptor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199

6.1. The 5-hydroxytryptamine6 receptor activates adenylyl cyclase . . . . . . . . . . . . . . 199

6.2. The 5-hydroxytryptamine6 receptor splice variants . . . . . . . . . . . . . . . . . . . . 199

7. The 5-hydroxytryptamine7 receptor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199

7.1. The 5-hydroxytryptamine7 receptor splice variants . . . . . . . . . . . . . . . . . . . . 199

7.2. The 5-hydroxytryptamine7 receptor activates adenylyl cyclase . . . . . . . . . . . . . . 200

7.3. The 5-hydroxytryptamine7 receptor activates the extracellular signal-regulated kinase . . 200

7.4. The 5-hydroxytryptamine7 receptor stimulates vasorelaxation . . . . . . . . . . . . . . . 200

7.5. The 5-hydroxytryptamine7 receptor modulates slow afterhyperpolarization . . . . . . . . 200

8. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201

1. Introduction

Serotonin (5-HT, 5-hydroxytryptamine) was discovered

in 1948 by Rapport et al. as a potent vasotonic factor. For

many years, the physiological effects of 5-HT, including its

effects on the CNS, were attributed to only two major

subtypes of 5-HT receptors (Gaddum & Picarelli, 1957).

Because of the development of sophisticated pharmaco-

logical tools in the 1980s, it became clear that there must

be more than two subtypes of 5-HT receptors. Molecular

cloning studies over the last 14 years have confirmed the

existence of at least 14 subtypes of 5-HT receptors, each

encoded by distinct genes. Splice variants of many of the

subtypes have been identified subsequently, resulting in

the discovery of at least 30 distinct protein products

that recognize 5-HT as their physiological ligand. Most

of the receptor genes (13) belong to a large family of

receptors that encode proteins that transduce signals

through guanine nucleotide binding and regulatory pro-

teins (G-proteins).

The 5-HT receptors have been divided into 7 subfamilies

by convention. Those subfamilies have been characterized

by overlapping pharmacological properties, amino acid

sequences, gene organization, and second messenger cou-

pling pathways (Hoyer et al., 1994). The 5-HT1, 5-HT2,

5-HT4, 5-ht5, 5-HT6, and 5-HT7 receptors couple to G-pro-

teins, whereas the 5-HT3 receptors are 5-HT-gated ion

channels. The basic architecture of the G-protein-coupled

5-HT receptors is similar to that proposed for nearly all of

the G-protein-coupled receptors (GPCRs). These receptors

are integral membrane proteins with 7 putative hydrophobic

transmembrane domains connected by 3 intracellular loops

(termed i1–i3) and 3 extracellular loops (termed e1–e3).

The amino terminus is oriented toward the extracellular

space, whereas the carboxyl terminus is oriented toward the

cytoplasm. The core proteins also possess conserved or

common sites for post-translational modifications. The

extracellular domains are typically glycosylated, and pos-

sess cysteine residues that may participate in disulfide

bonds that provide structural constraints on the conforma-

tion of the receptors. The intracellular domains possess sites

for interacting with G-proteins and other regulatory pro-

teins, and sites for phosphorylation by diverse serine-

threonine kinases. Some of the 5-HT receptors contain

PS-95 discs-large ZO-1 interaction motif (PDZ) domains

within their intracellular domains. Most of the 5-HT recep-

tors also possess cysteine residues within their carboxyl

termini that may be palmitoylated, thus creating potential

membrane anchors that can form a putative fourth intra-

cellular loop.

The purpose of this review is to summarize what is

known about the second messenger and effector linkages of

5-HT receptors, as revealed in native tissues and cells and

in heterologous expression systems. This is a timely and

important topic that already has been the subject of many

excellent reviews (for example, Hoyer et al., 1994; Boess &

Martin, 1994; Zifa & Fillion, 1992; Roth et al., 1998).

Recent studies have revealed a rich diversity of coupling

mechanisms for each 5-HT receptor subtype. The multipli-

city of coupling pathways for each of the receptors suggests

that each individual 5-HT receptor subtype can regulate a

broad array of potential signals that could be affected by

variables such as cell type, receptor number, numbers and

types of G-proteins expressed in the target cells, and the

specific agonist through which the receptor is activated.

This review will focus only upon the signaling linkages of

the cloned G-protein-coupled 5-HT receptors. We will not

discuss other 5-HT receptors, such as putative peripheral

‘‘5-HT1P’’ receptors (Pan et al., 1997), which have been

reviewed elsewhere (Gershon, 1999). The discussion will

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212 181

be organized so that each major receptor subfamily will be

discussed individually.

2. The 5-hydroxytryptamine1 receptors

There are 5 members of the 5-HT1 receptor family,

termed 5-HT1A, 5-HT1B, 5-HT1D, 5-HT1E, and 5-HT1F. A

receptor formerly termed the 5-HT1C receptor is no longer

felt to be a member of the 5-HT1 receptor family, having

been reclassified as the 5-HT2C receptor (Hoyer et al.,

1994), based on similarities to other 5-HT2 receptors in

structure and second messenger systems. The 5-HT1 recep-

tors couple primarily through Gi/o-proteins to the inhibition

of adenylyl cyclase (AC) (see Fig. 1) and to a multitude of

other signaling pathways and effectors. These will be

reviewed in Sections 2.1–2.5.

2.1. The 5-hydroxytryptamine1A receptor

The 5-HT1A receptor is the best characterized of the

5-HT1 receptors (Pedigo et al., 1981; Middlemiss &

Fozard, 1983), in large part due to the wide availability

of many specific pharmacological tools and because its

cDNA and gene were cloned (Kobilka et al., 1987) and

identified (Fargin et al., 1988) over a decade ago. Like all

5-HT1 receptors, the 5-HT1A receptor is characterized

pharmacologically by its high affinity for 5-HT. It has a

uniquely high affinity for second generation, arylpipera-

zine anxiolytic agents, such as buspirone, gepirone, and

ipsapirone. The 5-HT1A receptor also has a high affinity

for ( ± )-8-hydroxy-2-(di-N-propylamino)tetralin (8-OH-

DPAT), which previously was thought to be unique.

However, it is now known that the 5-HT7 receptor shares

a relatively high affinity for 8-OH-DPAT with the 5-HT1A

receptor. The 5-HT1A receptor plays potential roles in

neuro-endocrine function and thermoregulation (Goodwin

et al., 1985, 1987; Hjorth, 1985; Gilbert et al., 1988a;

Gartside et al., 1990; Seletti et al., 1995; Balcells-Olivero

et al., 1998), vasoreactive headaches (Leone et al., 1998),

sexual behavior (Ahlenius & Larsson, 1989; Schnur et al.,

1989; Maswood et al., 1998), food intake (Dourish et al.,

1985; Gilbert & Dourish, 1987; Hutson et al., 1988;

Gilbert et al., 1988b; Yamada et al., 1998), tooth-germ

morphogenesis (Moiseiwitsch et al., 1998), memory (Eda-

gawa et al., 1998), immune function (Iken et al., 1995),

aggression (Miczek et al., 1998), depression (Blier et al.,

1997; Shiah et al., 1998), and anxiety (Parks et al., 1998;

Ramboz et al., 1998), as have been demonstrated in

various physiological, clinical, behavioral, and pharmaco-

logical studies.

The 5-HT1A receptor was one of the first GPCRs for

which the cDNA and gene were cloned (Kobilka et al.,

1987; Fargin et al., 1988; Albert et al., 1990; Fujiwara et al.,

1990; Stam et al., 1992; Chanda et al., 1993). The human

receptor gene is localized on chromosome 5q11.2-q13

(Kobilka et al., 1987), and encodes a predicted protein with

422 amino acids. The gene is intronless, and its mRNA and

protein are expressed mainly in the brain, spleen, neonatal

kidney, and gut (Kobilka et al. 1987; Fargin et al. 1988;

Kirchgessner et al., 1993; Raymond et al., 1993a) (see

Table 1). The coding block of the genes for the rat and

human 5-HT1A receptor are 88% homologous with each

other at the nucleic acid level (Kobilka et al., 1987; Fargin

et al., 1988; Albert et al., 1990; Fujiwara et al., 1990; Stam

et al., 1992). Those sequences share substantially less

homology with other G-protein-coupled 5-HT receptors,

such as the 5-HT2A (19%), 5-HT2C (18%), and 5-HT1D

(43%) receptors (Fujiwara et al., 1990). The primordial

significance of the 5-HT1A receptor is underscored by the

molecular cloning of putative homologues in non-mam-

malian species, such as Xenopus laevis (Marracci et al.,

1997) and Fugu rubripes (Yamaguchi & Brenner, 1997).

The 5-HT1A receptor arguably couples to the broadest

panel of second messengers of any of the 5-HT receptors

(see Table 2). The 5-HT1A receptor has been reported to

activate or inhibit various enzymes, channels, and kinases,

and to stimulate or inhibit production of diverse soluble

second messengers. This receptor has been reported to

Fig. 1. Prototypical signaling enzyme linkages of the G-protein-coupled

5-HT receptors. 5-HT1 receptors typically inhibit AC through pertussis toxin-

sensitive G-proteins of the Gi family, whereas 5-HT4, 5-HT6, and 5-HT7

receptors typically stimulate AC through Gs family G-proteins. Activation

of AC results in increased production of cAMP, leading to activation of PKA.

5-HT2 receptors activate PLC-b through Gq/11 family G-proteins, resulting in

accumulation of phosphatidylinositol 4,5-bisphosphate (PIP2) to IP3 and

DAG. Generation of IP3 results in elevation of intracellular Ca2 + levels,

whereas DAG activates the Ca2 + and phospholipid-dependent protein kinase

(PKC). 5-HT2 receptors also typically activate PLA2 through G-proteins to

increase the accumulation of AA.

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212182

inhibit and activate AC and phospholipase (PL)C, and to

stimulate nitric oxide synthase (NOS) and an NAD(P)H

oxidase-like enzyme. It can activate K + channels and high-

conductance anion channels, inhibit Ca2 + conductances,

and regulate a number of channels and transporters. The

5-HT1A receptor can activate protein kinase C (PKC), Src

kinase, and mitogen-activated protein (MAP) kinases. The

5-HT1A receptor can inhibit or stimulate Ca2 + mobiliza-

Table 2

Signaling characteristics of human 5-HT receptors

Receptor Common signaling linkages Other signaling linkages G-protein coupling

5-HT1A Inhibits AC Activates PLC Gia3 > Gia2 � Gia1 � Goa > Gza

Activates K + channels Activates NOS

Stimulates ERK Activates NAD(P)H oxidase

Inhibits Ca2 + Activates NHE-1

conductances

5-HT1B Inhibits AC Activates PLC Gia3 > Gia1 � Gia2 � Goa

Stimulates ERK Activates NOS

Activates AC2

Activates K + channels

5-HT1D Inhibits AC Inhibits Ca2 + conductances Gia & Goa

Activates K + channels

5-HT1E Inhibits AC Gia & Goa

5-HT1F Inhibits AC Activates PLC Gia & Goa

5-HT2A Activates PLC Activates NHE-1 Gqa & G11a � Gia

Activates PKC Activates AC

Stimulates ERK Inhibits AC

Activates PLA2 Activates Jak2/STAT3

Activates Ca2 + channels

5-HT2B Activates PLC Activates cell cycle Gqa & G11a

Activates ERK Activates iNOS

Activates PLA2 Activates cNOS

5-HT2C Activates PLC Activates Na + /Ca2 + exchanger Gqa & G11a

Activates PKC PDZ motif signals?

Activates PLA2

5-HT4 Activates AC Regulates various channels Gsa

Activates PKA

5-ht5a Unknown Unknown Unidentified

5-ht5B Unknown Unknown Unidentified

5-HT6 Activates AC Gsa

Activates PKA

5-HT7 Activates AC Activates ERK Gsa

Activates PKA

This table is not all-encompassing. See text for more details and for primary literature citations.

Table 1

Characteristics of human 5-HT receptor genes

Human receptor Archaic names

Number of

amino acids

Chromosomal

location Introns

Functional

splice variants

mRNA editing

variants

5-HT1A 422 5q11.2-q13 0 0 0

5-HT1B 5-HT1Db 390 6q13 0 0 0

5-HT1D 5-HT1Da 377 1p34.3 0 0 0

5-HT1E 365 6q14-q15 0 0 0

5-HT1F 5-HT1Eb 366 3q11 0 0 0

5-HT2A 5-HT2;

5-HT D

471 13q14-q21 2 0 0

5-HT2B SRL;

5-HT2F

481 2q36.3-q37.1 2 0 0

5-HT2C 5-HT1C 460 Xq24 3 ?1 � 14

5-HT4 388 5q31-q33 � 5 � 5 0

5-ht5a 357 7q36 1 0 0

5-ht5B ? 2q11-q13 1 0 0

5-HT6 440 1p35-p36 2 ?1 0

5-HT7 445 10q21-q24 2 � 3 0

1 Nonfunctional variants have been described. See text for primary literature citations.

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212 183

tion, activate or inhibit phosphatidyl inositol hydrolysis,

and stimulate production of reactive oxygen species (both

H2O2 and superoxide) and arachidonic acid (AA). In

nearly every case, the signals have been shown to be

almost completely sensitive to pertussis toxin, implicating

Gi/o-proteins in the signals initiated by the 5-HT1A receptor.

2.1.1. The 5-hydroxytryptamine1A receptor both inhibits and

activates adenylyl cyclase

The primary coupling of this receptor (and all 5-HT1

receptors) is to the inhibition of AC (De Vivo & Maayani,

1986; Weiss et al., 1986). In some systems, this is man-

ifested by an ability to suppress basal cyclic AMP (cAMP)

levels, whereas in others, it is primarily manifested by an

ability to inhibit forskolin-stimulated AC activity. The

number of cell types and lines in which the 5-HT1A

receptor inhibits AC is quite extensive. A partial list

includes hippocampal and cortical neurons (De Vivo &

Maayani, 1986; Weiss et al., 1986), cultured neuronal cells

(the NCB-20, F11, P-11, and HN2 lines) (Hensler et al.,

1996; Singh et al., 1996), Chinese hamster ovary (CHO)

fibroblasts (Raymond, 1991), HeLa cells (Fargin et al.,

1991), human embryonic kidney (HEK) 293 cells (Albert

et al., 1999; Kellett et al., 1999), NIH 3T3 fibroblasts

(Varrault & Bockaert, 1992; Varrault et al., 1992b), COS-7

cells (Fargin et al., 1989), GH4C1 pituitary cells and Ltk–

fibroblasts (Liu & Albert, 1991), LLC-PK1 (Langlois et al.,

1996), ventral prostate cells (Carmena et al., 1998), and even

insect (army worm) Sf9 cells (Mulheron et al., 1994). The

inhibition of AC by the 5-HT1A receptor is universally

pertussis toxin-sensitive, suggesting that this effect requires

Gi/o-proteins. Numerous studies in diverse expression sys-

tems (Escherichia coli, Spodoptera frugiperda, and mam-

malian cells) have shown remarkable consistency in the

ability of the 5-HT1A receptor to bind to G-proteins of this

class, with a consensus rank order of Gia3 > Gia2 � Gia1 �Goa > Gza (Bertin et al., 1992; Raymond et al., 1993b;

Mulheron et al., 1994; Butkerait et al., 1995; Barr et al.,

1997; Clawges et al., 1997; Garnovskaya et al., 1997). We

would like to point out that the consensus rank order of

G-protein coupling has been determined using only two

agonists (5-HT and 8-OH-DPAT). This is potentially

important in that we have previously shown using cells

transfected with the human 5-HT1A receptor, that the

relative efficacies (compared with 5-HT) of a small panel

of compounds to activate G-protein a-subunits is clearly

different, depending upon whether Gia3 or Gia2 was studied

(Gettys et al., 1994). Thus, the coupling of the 5-HT1A

receptor to various G-proteins may be influenced by the

agonist to which the receptor is exposed.

Several studies strongly support a role for either Gia3 or

Gia2 in mediating the inhibition of AC activity by the 5-HT1A

receptor in diverse mammalian cells (Fargin et al., 1991;

Raymond et al., 1993b; Liu et al., 1994; Langlois et al., 1996),

although Gia1 or Goa may play limited roles in some systems

(Raymond et al., 1993b; Mulheron et al., 1994).

Surprisingly, the 5-HT1A receptor does not universally

inhibit AC in all cell types. Despite being expressed in high

density in dorsal raphe neurons, 5-HT1A receptors have not

been shown to inhibit AC in the rat dorsal raphe (Clarke et al.,

1996). Moreover, there is no evidence for functional cou-

pling to the inhibition of AC in cultured astrocytes, although

multiple mRNAs for 5-HT1 receptors (including the 5-HT1A

receptor) are expressed there (Hirst et al., 1998). In contrast

to the findings of Clarke et al. (1996) in rat dorsal raphe,

Palego et al. (1999, 2000) have shown that 5-HT1A receptors

in human dorsal raphe can inhibit forskolin-stimulated AC,

so the lack of coupling of this receptor in the raphe may be

species specific.

The 5-HT1A receptor has been reported to couple po-

sitively to AC in the hippocampus (Markstein et al., 1986;

Shenker et al., 1987; Fayolle et al., 1988; Cadogan et al.,

1994) and the prostate. Some of these responses are possibly

due to confusion with another receptor subtype (Shenker

et al., 1987), most likely the 5-HT7 receptor subtype, which

can be stimulated by the 5-HT1A receptor-selective ligand

8-OH-DPAT. The 5-HT1A receptor in the ventral prostate

has been shown to slightly activate AC under very special

conditions (Carmena et al., 1998). 8-OH-DPAT, presumably

working through the 5-HT1A receptor, was shown to aug-

ment forskolin-stimulated cAMP accumulation, but only

after pretreatment of rats or their excised prostate glands

with 8-OH-DPAT. Because the effect was small, and

because no agents more specific than 8-OH-DPAT were

used, there remains the possibility that this effect was

mediated through the 5-HT7 receptor. However, in another

report, the authors used microdialysis to detect increases in

cAMP in response to 8-OH-DPAT in rat hippocampus, an

effect that they were able to attenuate with a selective

5-HT1A receptor antagonist (N-tert-butyl-3-(4-(2-methoxy-

phenyl)-piperazin-1-yl)-2-phenylpropanamide) (Cadogan

et al., 1994). Their report provided very tantalizing evidence

that the 5-HT1A receptor can activate AC.

This evidence was perplexing in light of a relatively

detailed study that could not detect coupling of the 5-HT1A

receptor to Gs or increased AC activity in CHO cells

(Raymond et al., 1993b). A recent study confirmed that

wild-type 5-HT1A receptors in HEK 293 cells do not

stimulate cAMP accumulation; they further demonstrated

that the carboxyl terminal region of the i3 loop of the

5-HT1A receptor could be mutated to induce only a very

weak coupling to Gs (Malmberg & Strange, 2000). There-

fore, if the 5-HT1A receptor couples positively to AC, it

would seem to require a highly specific cellular milieu,

perhaps necessitating the expression of particular subtypes

of AC. The family of ACs consists of at least nine distinct

members, each possessing a diverse repertoire of regula-

tory inputs (Mons et al., 1998; Taussig & Zimmermann,

1998; Hurley, 1999). Because the 5-HT1A receptor can

activate a multitude of signaling pathways, it could regu-

late AC in a number of ways, depending upon both the

specific signals generated by receptor occupancy and the

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212184

types and amounts of ACs expressed in the cells or tissues

of interest.

Three studies have suggested that the presence of AC2 is

a key determinant for the positive coupling of the 5-HT1A

receptor to cAMP accumulation. Uezono et al. (1993) used

the Xenopus oocyte system to show that b2-adrenergicreceptor increases in cAMP were enhanced by activation

of co-expressed 5-HT1A receptors. The additional activation

by the 5-HT1A receptor was enhanced by co-expression of

AC2, but not AC3 (Uezono et al., 1993). Liu et al. (1999)

showed that depletion of Gia1 from GH4 pituitary cells

switched the effect of the 5-HT1A receptor on cAMP accu-

mulation from an inhibitory to a stimulatory one. Moreover,

the stimulatory effect was sensitive to pertussis toxin, sug-

gesting that either a Gi- or Go-protein subtype specifically

mediates the effect. The same group performed more detailed

studies in HEK 293 cells, showing a requirement for AC2

and G-protein bg-subunits for the 5-HT1A receptor to

increase cAMP in an agonist-dependent fashion. Moreover,

they demonstrated that the 5-HT1A receptor also contributed

to agonist-independent increases in basal cAMP accumula-

tion that also depended upon Gia2 (Albert et al., 1999). Thus,

the 5-HT1A receptor can stimulate cAMP accumulation in

cells that express AC2 via release of bg-subunits from

pertussis toxin-sensitive G-proteins (probably Gia2).

2.1.2. The 5-hydroxytryptamine1A receptor activates and

inhibits phosphatidylinositol-specific phospholipase C

Activation of phosphatidylinositol-specific (PI)-PLC

results in the generation of two key second messengers.

These are inositol trisphosphate (IP3), which regulates in-

tracellular Ca2 + release (Berridge et al., 1998), and diacyl-

glycerol (DAG), which binds to and activates PKC. PLC

can be activated by receptors that couple to both pertussis

toxin-sensitive and -sensitive G-proteins. Although 5-HT1A

receptors can clearly activate PLC, this effect is highly host

cell-specific. The ability of the human 5-HT1A receptor to

activate PI-PLC was first demonstrated in transfection

studies in HeLa cells (Fargin et al., 1989). The stimulation

was similar in magnitude to that induced by endogenous

histamine H1-like receptors, but was not as efficient as

coupling of the 5-HT1A receptor to the inhibition of AC

(Raymond et al., 1991). The activation of PLC by the

human 5-HT1A receptor in HeLa cells was confirmed in

several other studies (Middleton et al., 1990; Boddeke et

al., 1992; Harrington et al., 1994). Liu and Albert (1991)

subsequently demonstrated that activation of PI-PLC by the

5-HT1A receptor is cell-specific. They showed that the rat

5-HT1A receptor increased PI hydrolysis and released Ca2 +

from intracellular stores in Ltk� fibroblasts, but not in

BALB/c-3T3 cells (Abdel-Baset et al., 1992) or in GH4C1

pituitary cells (Liu & Albert, 1991). The human 5-HT1A

receptor can also activate PLC when expressed in Xen-

opus oocytes (Ni et al., 1997). This coupling may be

relevant in that endogenous 5-HT1A receptors in a human

T-cell-like line (Jurkat) activate PLC (Aune et al., 1993).

In contrast, transfected 5-HT1A receptors do not stimulate

PLC in CHO cells (Newman-Tancredi et al., 1992; Cowen

et al., 1997), COS-7 cells (Fargin et al., 1989), or NIH 3T3

cells (Varrault et al., 1992a). It bears mentioning that the

5-HT1A receptor has also been shown to inhibit PLC

activation in rat hippocampal preparations (Claustre

et al., 1991), although this effect was not sensitive to

pertussis toxin.

2.1.3. The 5-hydroxytryptamine1A receptor activates

other phospholipases

The 5-HT1A receptor probably can regulate PLs other than

PI-PLC. Cowen et al. (1997) proposed that the 5-HT1A

receptor activates a phosphatidylcholine-specific (PC)-PLC

in CHO cells, and that this activity may be important in the

regulation of nuclear factor-kB and extracellular signal-

regulated kinase (ERK). Their data showed that 8-OH-

DPAT elicited release of 3H from cells preloaded with

[3H]choline. This effect was attenuated by a PC-PLC

inhibitor (tricyclodecan-9-yl-xanthogenate). The 5-HT1A

receptor in rat hippocampus has been proposed to stimulate

PLA2 based on inhibitor studies (Claustre et al., 1991). The

recombinant 5-HT1A receptor has also been shown to

activate PLA2 in HeLa cells (Harrington et al., 1994), and

to augment Ca2 +-induced AA metabolism in CHO cells

(Raymond et al., 1992). In contrast, 5-HT1A receptor ago-

nists attenuate N-methyl-D-aspartic acid (NMDA) receptor-

evoked AA release in adult rat hippocampus, raising the

possibility that the 5-HT1A receptor might inhibit PLA2 in

some settings (Strosznajder et al., 1996).

2.1.4. The 5-hydroxytryptamine1A receptor activates protein

kinase C

The 5-HT1A receptor activates several different protein

kinases. These include serine-threonine, tyrosine, and

lipid kinases. Not surprisingly, the 5-HT1A receptor can

activate PKC in cells in which the receptor has been

shown to stimulate PLC (Raymond et al., 1989; Mid-

dleton et al., 1990; Liu & Albert, 1991). This is not

surprising, in that PKC can be activated by Ca2 + and

DAG (Tanaka & Nishizuka, 1994; Nishizuka, 1995).

Activation of PKC by the 5-HT1A receptor depends upon

PLC activation. As this effect is probably mediated by

G-protein bg-subunits, it will depend upon the expres-

sion of bg-regulated PLC in the target cell or tissue. The

coupling of the 5-HT1A receptor to PKC is as efficacious

as that induced by the endogenous histamine receptor

expressed in HeLa cells.

2.1.5. The 5-hydroxytryptamine1A receptor activates

the extracellular signal-regulated mitogen-activated

protein kinase

Several groups have documented that the 5-HT1A recep-

tor activates ERK MAP kinases in CHO cells (Cowen, D. S.

et al., 1996; Garnovskaya et al., 1996, 1998; Della Rocca

et al., 1999; Mendez et al., 1999; Mukhin et al., 2000). Like

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212 185

other GPCRs, the 5-HT1A receptor activates ERK through

an intricate signaling pathway that involves assembly of a

signaling complex that requires many of the same mole-

cules used by growth factor receptor tyrosine kinases

(Marshall, 1995; Luttrell et al., 1997). The activation of

ERK by the 5-HT1A receptor is initiated by bg-subunitsreleased from pertussis toxin-sensitive G-proteins. This

results in activation of the non-receptor tyrosine kinase

Src and tyrosine phosphorylation of the docking platform

Shc. Shc phosphorylation results in the recruitment of

PI-30 kinase, Grb2 (an adapter protein), and a Ras act-

ivator protein called Sos to the signaling platform. Ras

activation leads to sequential activation of Raf, mitogen

and extracellular signal-regulated kinase (MEK) and ERK

(Cowen, D. S. et al., 1996; Garnovskaya et al., 1996, 1998;

Mendez et al., 1999). Cowen, D. S. et al. (1996) hypothe-

sized that PC-PLC augments the activation of Raf that is

induced by Ras. Thus, it is possible that the 5-HT1A receptor

in CHO cells activates ERK through the two converging

lipid-signaling pathways. The first involves PI-30 kinase and

intersects the ERK pathway upstream of Ras, whereas the

second involves PC-PLC and intersects the ERK pathway

upstream of Raf.

Other studies have implicated a Ca2 +/calmodulin (CaM)-

dependent endocytosis step between Ras and Raf (Della

Rocca et al., 1999) and reactive oxygen species (H2O2 and

superoxide) upstream of Src in 5-HT1A receptor-mediated

ERK activation in CHO cells (Mukhin et al., 2000).

Although the coupling of the 5-HT1A receptor to ERK has

been relatively well characterized in CHO cells, studies in

other cells or tissues are still needed.

2.1.6. The 5-hydroxytryptamine1A receptor stimulates

phosphatidylinositol-30 kinase

The ability of the 5-HT1A receptor expressed in CHO

cells to activate ERK (Cowen, D. S. et al., 1996; Garnov-

skaya et al., 1996, 1998) and Na + -proton exchange (Gar-

novskaya et al., 1998) is sensitive to pharmacological

inhibitors and cDNA constructs encoding dominant inter-

fering p85 and p110 subunit of PI-30 kinase. Moreover,

Adayev et al. (1999) showed that the 5-HT1A receptor can

stimulate PI-30 kinase-g in HN2-5 cells.

2.1.7. The 5-hydroxytryptamine1A receptor causes

production of reactive oxygen and nitrogen species

The 5-HT1A receptor in CHO cells has been shown by

inhibitor, biochemical, and fluorometric methods to stimu-

late production of at least two reactive oxygen species, H2O2

and superoxide (Mukhin et al., 2000). The production of

these reactive oxygen species by the 5-HT1A receptor

appears to occur at a point upstream of Src, and involves

an NAD(P)H oxidase-like enzyme. Activation of this

enzyme and the resulting production of reactive oxygen

species is a critical step in the stimulation of ERK phospho-

rylation and kinase activity by the human 5-HT1A receptor

(Mukhin et al., 2000).

The 5-HT1A receptor can also regulate the production of

nitric oxide (NO), although there is no consensus on

whether the major effect is to stimulate or inhibit its

production. There is tantalizing evidence that the 5-HT1A

receptor can increase NO production in at least three

settings. First, 8-OH-DPAT-associated hyperphagia in rats

(routinely attributed to the 5-HT1A receptor) can be inhi-

bited by NOS inhibitors (Yamada et al., 1996; Sugimoto

et al., 1999a, 1999b). Second, putative 5-HT1A receptor-

induced renal vasodilatory effects have been ascribed to the

release of NO from endothelial cells in the renal micro-

circulation (Verbeuren et al., 1991; Verbeuren, 1993).

Third, in a single report, putative endogenous 5-HT1A

receptors in rat ventral prostate cells were shown to

stimulate NOS activity (Carmena et al., 1998). In contrast,

there is evidence that the 5-HT1A receptor can inhibit

NMDA-induced NO production in human neocortical sli-

ces (Maura et al., 2000) and in adult rat hippocampus

(Strosznajder et al., 1996), and that it can inhibit NMDA/

a-amino-3-hydroxy-5-methyl-isoxazole-4-propionate re-

ceptor-induced cyclic GMP (cGMP) accumulation (a par-

ticipant in the NO pathway) in rat cerebellar synaptosomes

(Maura & Raiteri, 1996). Clearly, more studies are needed

in this area.

2.1.8. The 5-hydroxytryptamine1A receptor regulates

K+ channels

The 5-HT1A receptor has been shown to regulate the

function of several distinct types of channels, including

inwardly rectified K + channels, high conductance anion

channels, cystic fibrosis transmembrane regulator Cl– chan-

nels, and Ca2 + channels. It has been known for at least

15 years that the 5-HT1A receptor can stimulate the opening

of G-protein-gated inwardly rectifying K + (GIRK) channels

in neurons (Andrade & Nicoll, 1987; Colino & Halliwell,

1987; Zgombick et al., 1989; Penington et al., 1993;

Doupnik et al., 1997). Those channels mediate hyperpolar-

izing postsynaptic potentials in the nervous system and in

the heart during activation of Gi/oa-coupled receptors. The

regulation of GIRK channels by receptors relies upon the

interaction of G-protein bg-subunits (Doupnik et al., 1996,

1997). The 5-HT1A receptor expressed in primary cultures

of rat atrial myocytes with a Vaccinia virus vector stimulates

an endogenous atrial inward rectifier K + current (Karschin

et al., 1991). Co-injection of rat atrial RNA, GIRK1 RNA,

or rat brain GIRK3 with 5-HT1A receptor RNA into Xen-

opus oocytes resulted in 5-HT1A receptor stimulation of

GIRK activity (Dascal et al., 1993; Dissmann et al., 1996;

Doupnik et al., 1997). An additional intriguing finding is

that various regulators of G-protein signaling (RGS) pro-

teins (RGS1, RGS3, and RGS4, but not RGS2) participated

in the functional regulation of 5-HT1A receptor-mediated

GIRK activation (Doupnik et al., 1997).

The 5-HT1A receptor, however, does not universally

stimulate inwardly rectified K + channels in all cell types.

When co-expressed by transfection into COS-7 cells, the

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212186

5-HT1A receptor inhibited the cloned rat, strongly rectify-

ing the inward rectifier K + (IRK)-type inwardly rectifying

K + channel IRK1 (Kir 2.1). The inhibition was mimicked

by elevations of internal cAMP concentrations and by

microinjection of the catalytic subunits of protein kinase

A (PKA), and was not present when mutant IRK1 chan-

nels lacking a PKA phosphorylation site were studied

(Wischmeyer & Karschin, 1996). These results suggest

that the 5-HT1A receptor can inhibit IRK channels, and

support a possible role for this receptor in stimulating

cAMP production in COS-7 cells (see Section 2.1.1).

2.1.9. The 5-hydroxytryptamine1A receptor inhibits neuronal

Ca2+ conductances

The 5-HT1A receptor has been demonstrated to inhibit a

Ca2 + current in neurons (Penington & Kelly, 1990) and

neuronal cell lines (Singh et al., 1996). These effects have

been attributed to inhibition of both N- and P/Q-type Ca2 +

channels (Penington et al., 1991; Bayliss et al., 1995; Sun

& Dale, 1998). For example, the 5-HT1A receptor in

X. laevis spinal neurons causes voltage-independent inhibi-

tion of V-agatoxin-IVA-sensitive (P/Q-type) Ca2 + channels

(Sun & Dale, 1998). Liu et al. (1994) confirmed that the

rat 5-HT1A receptor inhibited Bay K8644-mediated Ca2 +

influx when transfected in GH4C1 cells, and further dem-

onstrated that this effect required expression of Goa. The

involvement of G-proteins in the inhibition of Ca2 +

currents was confirmed by studies showing that a peptide

inhibitor of G-protein bg-subunit attenuated 5-HT1A recep-

tor-mediated inhibition of Ca2 + current in dorsal raphe

neurons (Chen & Penington, 1997).

2.1.10. The 5-hydroxytryptamine1A receptor regulates other

ion channels

The 5-HT1A receptor has been shown to regulate several

other channels in transfected cells. Ni et al. (1997) dem-

onstrated that the 5-HT1A receptor expressed in Xenopus

oocytes could stimulate an oscillatory Ca2 + -activated Cl–

current, although this effect was much less efficient than

that associated with the 5-HT2C receptor. Uezono et al.

(1993) showed that the 5-HT1A receptor expressed in

Xenopus oocytes could indirectly augment the activation

of cystic fibrosis transmembrane regulator Cl– channels

induced by b2-adrenoceptors. This effect likely proceeded

via G-protein bg-subunits, and was enhanced by co-expres-

sion of AC2 and Gsa (Uezono et al., 1993). The 5-HT1A

receptor expressed by transfection in CHO cells inhibits an

endogenous high-conductance anion channel through either

Gia2 or Gia3.

2.1.11. The 5-hydroxytryptamine1A receptor regulates

miscellaneous transport processes

The 5-HT1A receptor has been shown to regulate

several transport processes in transfected cells and native

tissues. The 5-HT1A receptor stimulates Na + /K + -ATPase

through a Ca2 + -mediated pathway (Middleton et al., 1990)

and Na + -dependent phosphate uptake through a PKC-

mediated pathway (Raymond et al., 1989, 1991) in HeLa

cells. In CHO cells, the 5-HT1A receptor stimulates an

Na + /H + exchanger (Garnovskaya et al. 1997, 1998)

through a pathway that requires Gia2 and/or Gia2, Src,

and PI-30 kinase (Garnovskaya et al., 1997, 1998). The

activation of Na + -dependent phosphate uptake suggests a

potential role for the 5-HT1A receptor in regulating cellular

energy fluxes, whereas the coupling to Na + /K + -ATPase

and Na + /H + exchange suggests potential roles in cell

volume regulation. The 5-HT1A receptor in CHO cells has

also been demonstrated to accelerate mediator-facilitated

electron fluxes across the plasma membrane, a process that

may be linked to the production of reactive oxygen species

(Mukhin et al., 2000).

The 5-HT1A receptor can also regulate 5-HT, norepi-

nephrine, and acetylcholine release from neurons. Release

of 5-HT involves presynaptic receptors (Verge et al., 1985,

1986; Sharp & Hjorth, 1990), whereas release of norepi-

nephrine (Done & Sharp, 1994; Chen & Reith, 1995) and

acetylcholine (Bianchi et al., 1990; Izumi et al., 1994)

involves postsynaptic receptors (Hall et al., 1985; Verge

et al., 1986).

2.1.12. The 5-hydroxytryptamine1A receptor regulates

proliferation and related pathways

The 5-HT1A receptor can stimulate proliferative and

mitogenic pathways in various cell systems, consistent with

its stimulation of ERK (see Section 2.1.5) and nuclear

factor-kB (Cowen et al., 1997). 5-HT1A receptors stimulate

proliferation of T lymphocytes (Iken et al., 1995), pancreatic

carcinoid tumor cells (Ishizuka et al., 1992), and human

small cell lung carcinoma cells (Cattaneo et al., 1995).

Activation of 5-HT1A receptors expressed in NIH 3T3 cells

induces focus formation and potentiates the stimulation of

DNA synthesis by epidermal growth factor (Varrault et al.,

1992a). Rat 5-HT1A receptors transfected into BALB/c-3T3

cells enhance thymidine uptake and induce focus formation

(Abdel-Baset et al., 1992). All of those studies are consistent

with a stimulatory effect of the 5-HT1A receptor on prolif-

eration. However, this is not a universal effect in that

endogenous 5-HT1A receptors in lymphocytes have been

implicated in the inhibition of mitogenic responses (Ferriere

et al., 1996).

The 5-HT1A receptor has also been implicated in morpho-

genesis and cell survival. Blockade of endogenous 5-HT1A

receptors reverses serotoninergic stimulation of tooth germ

development in mouse mandibular explant cultures (Moi-

seiwitsch et al., 1998), suggesting a developmental role for

this receptor. Transfected 5-HT1A receptors have been

shown to inhibit a caspase 3-like enzyme and to attenuate

(by � 60–70%) anoxia-induced apoptosis in the neuronal

HN2-5 cells via an ERK-dependent pathway (Adayev et al.,

1999), suggesting a potential role for the 5-HT1A receptor

in neuronal survival.

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212 187

2.1.13. What is the basis for the multiplicity of

5-hydroxytryptamine1A receptor signaling?

This question is worth addressing briefly, as it is pertinent

to all of the 5-HT receptors from the GPCR family. Many

possible explanations for the multiplicity of signaling mech-

anisms of the 5-HT receptors to diverse signaling pathways

could be proposed. These include the ability to activate

overlapping pools of G-proteins, activation of distinct sig-

nals emanating from G-protein a- and bg-subunits, differingrepertoires of G-proteins and signaling enzymes expressed in

cells, ratio of G-protein/effector/and receptor, ligand-specific

effects, differential rates of desensitization, and cellular

compartmentalization, to name just a few. This issue has

been reviewed in detail (Raymond, 1995).

In that regard, the 5-HT1A receptor has been identified in

both presynaptic and postsynaptic locations (Miquel et al.,

1992; Radja et al., 1992). The signaling properties and

pharmacology of the presynaptic and postsynaptic 5-HT1A

receptors vary somewhat (Clarke et al., 1996), leading to

speculation that there might be subtypes of 5-HT1A recep-

tors (Barnes & Sharp, 1999). Langlois et al. (1996) dem-

onstrated that 5-HT1A receptors transfected into polarized

epithelial LLC-PK1 cells were expressed on both baso-

lateral and apical membranes. Receptors on both surfaces

were able to inhibit cAMP accumulation, albeit with

different efficiencies (Langlois et al., 1996). This study

demonstrates that the variables that affect the multiplicity of

signaling from single receptor types are not likely to be

manifested as ‘‘all or none’’ effects. Thus, the specific

coupling of single types of 5-HT receptors to signaling

pathways is likely to result from the summation of effects

specific to the host cell milieu.

One other theoretical possibility is that some signals

emanating from 5-HT1A (or other 5-HT) receptors may

be regulated by effector pathways activated through non-

G-protein-activated pathways. This intriguing hypothesis

has been reviewed recently by Hall et al. (1999).

2.2. The 5-hydroxytryptamine1B receptor

The 5-HT1B receptor initially was identified as a low-

affinity [3H]spiperone-binding site (Nelson et al., 1981;

Pedigo et al., 1981), distinguishing it from the 5-HT1A

receptor. There has been considerable confusion over the

identities of the 5-HT1B and 5-HT1D receptors. Thus, a brief

history of these two closely related receptor subtypes is in

order. It originally was thought that the 5-HT1B receptor was

primarily or exclusively expressed in rodent (hamster,

mouse, and rat) tissues, whereas the closely related

5-HT1D receptor was believed to be expressed in other

species (humans, cows, dogs, and guinea pigs) (Hoyer &

Middlemiss, 1989). The two receptors are pharmacolo-

gically similar, although not identical. The major pharmaco-

logical distinction is that b-adrenergic receptor antagonists

bind with high affinity to the 5-HT1B receptor, but not to the

5-HT1D receptor. Because the distributions of the 5-HT1D

and the rodent 5-HT1B receptors showed significant overlap,

Hoyer and Middlemiss (1989) initially proposed that the

two were species homologues.

The subsequent identification of two intronless human

receptor genes encoding 5-HT receptors with pharmaco-

logical characteristics similar to the 5-HT1D receptor

resulted in their temporary designation as 5-HT1D receptors

(5-HT1Da and 5-HT1Db) (Hamblin & Metcalf, 1991; Ham-

blin et al., 1992; Hartig et al., 1992; Weinshank et al., 1992).

The cloning and identification of the rodent 5-HT1B receptor

cDNA and gene revealed a surprisingly high homology with

the human 5-HT1Db receptor (Voigt et al., 1991; Adham

et al., 1992; Jin et al., 1992; Maroteaux et al., 1992).

Whereas the human 5-HT1Da and 5-HT1Db receptors share

77% sequence homology in the transmembrane domains,

the human 5-HT1Db and rodent 5-HT1B receptors share 96%

homology (Hartig et al., 1992; Jin et al., 1992). The

uniquely high affinity of the rodent 5-HT1B receptors for

b-adrenergic receptor antagonists when compared with the

human 5-HT1Db receptors was shown to depend upon a

single amino acid residue difference (Metcalf et al., 1992;

Oksenberg et al., 1992; Parker et al., 1993). Subsequently,

the cloning by Hamblin and colleagues (1992) of a rat gene

that encodes a 5-HT receptor with pharmacological charac-

teristics more closely resembling the 5-HT1D receptor than

the 5-HT1B receptor gave further support to the idea that the

5-HT1B and 5-HT1D receptors represent separate gene

products that are not exclusively expressed in rodent or

non-rodent mammals. The preponderance of evidence

resulted in a reassessment of the nomenclature of these

receptors in which the human 5-HT1Da and rodent 5-HT1D

receptors were reclassified as 5-HT1D receptors, and the

human 5-HT1Db and all 5-HT1B receptors were classified as

5-HT1B receptors (Hartig et al., 1996).

The 5-HT1B receptor is widely expressed in brain

tissue, probably in both presynaptic and postsynaptic

locations (Middlemiss & Hutson, 1990; Buhlen et al.,

1996). The 5-HT1B receptor gene is located on chromosome

6q13 in humans and chromosome 9E in mice (Saudou &

Hen, 1994). The human gene encodes a protein with 390

amino acids.

2.2.1. The 5-hydroxytryptamine1B receptor regulates

adenylyl cyclase

In general, all of the 5-HT1 receptors can be expected to

share many of the signaling linkages of the 5-HT1A receptor

(Hoyer et al., 1994). Thus, the 5-HT1B receptor would be

expected to couple mainly to pertussis toxin-sensitive

G-proteins and to inhibit AC. The 5-HT1B receptor, in fact,

has been shown to inhibit AC when natively expressed in

tissues such as the substantia nigra (Bouhelal et al., 1988;

Schoeffter et al., 1988) and rabbit mesenteric artery (Hinton

et al., 1999), in opossum kidney cells (Ciaranello et al.,

1990; Zgombick & Branchek, 1998), human umbilical

endothelial vein cells (Schoeffter et al., 1995), bovine

basilar artery vascular smooth muscle cells (Ebersole et al.,

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212188

1993), or CHO-K1 cells (Berg et al., 1994b; Giles et al.,

1996), and when heterologously expressed in transfected

cells (Ltk-, COS-7, Sf9 cells) (Adham et al., 1992; Levy

et al., 1992b; Weinshank et al., 1992; Ng et al., 1993). Like

the 5-HT1A receptor, the 5-HT1B receptor has also been

documented to increase cAMP in HEK 293 cells through a

conditional process that involves ACII and Gi proteins

(Albert et al., 1999).

The 5-HT1B receptor has been shown to couple to

mammalian Gi/o-proteins in Sf9 cells when expressed using

a Baculovirus vector. Slightly different coupling efficiencies

were noted, such that high-affinity agonist binding to the

5-HT1B receptor was stimulated with the following rank

order: Gia3 > Gia1 � Gia2 � Goa (Clawges et al., 1997).

The same group noted that the 5-HT1A receptor coupled to

the G-proteins slightly more efficiently than the 5-HT1B

receptor. Although the G-protein coupling in Sf9 cells was

slightly more efficient for the 5-HT1A receptor than for the

5-HT1B receptor, another group showed that lower numbers

of transfected 5-HT1B receptors than 5-HT1A receptors were

needed for efficient inhibition of AC when transfected into

mammalian cells (Mendez et al., 1999).

2.2.2. The 5-hydroxytryptamine1B receptor

activates phospholipases

The 5-HT1B receptor can activate PLC in several cell

types. This effect is manifested by pertussis toxin-sensitive

increases in intracellular Ca2 + in opossum kidney cells

(Zgombick & Branchek, 1998). Although 5-HT1B receptors

natively expressed in CHO cells were not shown to increase

accumulation of inositol phosphates (Dickenson & Hill,

1995), transfected human 5-HT1B receptors were shown to

mediate pertussis toxin-sensitive increases in both intra-

cellular Ca2 + and accumulation of inositol phosphates in

CHO cells through G-proteins and PLC (Dickenson & Hill,

1998). The 5-HT1B receptor in vascular smooth muscle

cells derived from bovine basilar artery also increases

intracellular Ca2 + through a pertussis toxin-sensitive mech-

anism (Ebersole et al., 1993), and this signal may be

responsible for 5-HT1B-induced blood vessel contractions

(Ullmer et al., 1995).

5-HT can also stimulate PLD via endogenous 5-HT1B

receptors in rabbit mesenteric artery through a signaling

pathway that requires extracellular Ca2 + and the activa-

tion of PKC, but is independent of PLC activation (Hinton

et al., 1999).

2.2.3. The 5-hydroxytryptamine1B receptor regulates

the extracellular signal-regulated kinase,

phosphatidylinositol-30 kinase,

p70 S6 kinase, and Akt kinase

Over a decade ago, endogenous 5-HT1B receptors in

fibroblasts were reported to stimulate DNA synthesis

through a pertussis toxin-sensitive pathway that does not

involve PLC (Seuwen et al., 1988). Subsequent studies have

revealed that the 5-HT1B receptor can activate ERK in

bovine aortic endothelial cells (McDuffie et al., 2000),

BE2-C neuroblastoma cells (Lione et al., 2000), and CHO

cells (Pullarkat et al., 1998; Mendez et al., 1999). In CHO

cells, this occurs through a pathway that involves pertussis

toxin-sensitive G-proteins, PI-30 kinase, and MEK (Pullarkat

et al., 1998; Mendez et al., 1999). Akt kinase is also

activated by 5-HT1B receptors acting through pertussis

toxin-sensitive G-proteins in BE2-C neuroblastoma cells

(Lione et al., 2000). In CHO cells, 5-HT1B receptors activate

p70 S6 kinase through a pathway that involves PI-30 kinase

and MEK, and is sensitive to rapamycin and pertussis toxin.

Interestingly, coupling to ERK of the 5-HT1B receptor in

CHO cells was shown to be relatively more efficient than

that induced by the 5-HT1A receptor (Mendez et al., 1999).

2.2.4. The 5-hydroxytryptamine1B receptor stimulates

endothelial nitric oxide production

The 5-HT1B receptor increases NO in cultured human

coronary artery rings (Auch-Schwelk et al., 2000) and

endothelial cells, probably through increasing levels of

intracellular Ca2 + (Ishida et al., 1998). McDuffie et al.

(1999) presented evidence that the 5-HT1B receptor stim-

ulates endothelial NOS through pertussis toxin-sensitive

G-proteins in cultured bovine aortic endothelial cells. This

signaling pathway (and another linked to 5-HT2B receptors)

probably accounts for most of the 5-HT-induced vasorelax-

ation in diverse vascular beds (Ullmer et al., 1995).

2.2.5. Other signals of the 5-hydroxytryptamine1B receptor

Several groups have demonstrated that presynaptic

5-HT1B receptors decrease 5-HT release (Sharp et al., 1989;

Hjorth & Tao, 1991; Lawrence & Marsden, 1992; Martin

et al., 1992; Davidson & Stamford, 1996). Human 5-HT1B

receptors stably transfected in C6 glioma cells activate

Ca2 + -dependent K + channels (Le Grand et al., 1998).

2.3. The 5-hydroxytryptamine1D receptor

The 5-HT1D receptor was discovered by homology

screening using the canine RDC4 gene. That screen

revealed two similar genes with pharmacology closely

linked to each other, which were originally termed 5-HT1Da

and 5-HT1Db. These were later reclassified respectively as

the 5-HT1D and 5-HT1B receptors (see Section 2.2). The

gene is intronless, is located on chromosome 1p34.3, and

encodes a protein of 377 amino acids. The 5-HT1D receptor

shares 43% sequence homology with the 5-HT1A receptor

and 59% with the 5-HT1B receptor (77% in the transmem-

brane regions) (Hamblin & Metcalf, 1991; Hamblin et al.,

1992; Hartig et al., 1992).

It has been difficult to establish the precise locations of

5-HT1D receptors due to the lack of specific ligands and

apparently low levels of mRNA and receptor protein in the

brain. Binding sites for the 5-HT1D receptor have localized

to the substantia nigra, globus pallidus, and caudate, and in

lower levels in the cortex and hippocampus putamen

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212 189

(Bruinvels et al., 1993), although mRNA for this receptor

was not detected in the globus pallidus or substantia nigra

(Bruinvels et al., 1994a, 1994b). These results suggest that

5-HT1D receptors may be transported along axon terminals

after their synthesis.

2.3.1. The 5-hydroxytryptamine1D receptor inhibits

adenylyl cyclase

As has been shown for all other 5-HT1 receptors, the

5-HT1D receptor can inhibit AC through pertussis toxin-

sensitive G-proteins in transfected C6 glioma and NIH 3T3

fibroblast cells (Maenhaut et al., 1991; Weinshank et al.,

1992; Zgombick et al., 1996; Lesage et al., 1998). How-

ever, when expressed at high levels in CHO or Y1 adrenal

cells, this receptor can also weakly stimulate cAMP

accumulation (Van Sande et al., 1993; Watson et al.,

1996; Wurch et al., 1997).

2.3.2. The 5-hydroxytryptamine1D receptor regulates

ion channels

Like other 5-HT1 receptors, the 5-HT1D receptor can

regulate K + and Ca2 + channels. In Xenopus spinal

neurons, the 5-HT1D receptor was reported to cause

voltage-independent inhibition of V-conotoxin-GVIA-sens-

itive (N-type) Ca2 + channels (Sun & Dale, 1998),

although the ligands used were not sufficiently discrim-

inative to rule out the involvement of 5-HT1B receptors

in mediating this effect. Like 5-HT1B receptors, recombin-

ant human 5-HT1Dreceptors stably transfected in C6

glioma cells can stimulate Ca2 + -dependent K + channels,

with the resultant outward hyperpolarizing current being

dependent upon IP3 receptor-mediated intracellular Ca2 +

release (Le Grand et al., 1998).

2.3.3. The 5-hydroxytryptamine1D receptor is mitogenic

5-HT1D receptors endogenous to two small cell lung

carcinoma cell lines (GLC8 and NCI-N-592) have been

reported to stimulate DNA synthesis, although effects of

5-HT1B receptors were not ruled out (Cattaneo et al., 1994).

However, cloned human 5-HT1D receptors transfected into

C6 glial cells promote cell growth, so there is little doubt

that the 5-HT1D receptor (like 5-HT1A and 5-HT1B recep-

tors) can be mitogenic (Pauwels et al., 1996).

2.3.4. The 5-hydroxytryptamine1D receptor inhibits

5-hydroxytryptamine release

5-HT1D receptors can inhibit 5-HT release in several

brain regions. 5-HT efflux in the ventral lateral geniculate

nucleus of the rat is modulated by presynaptic 5-HT1D

receptors (Davidson & Stamford, 1996). 5-HT1D autorecep-

tors diminish 5-HT release in guinea-pig mesencephalic

raphe, hippocampus, and frontal cortical slices through a

pathway involving G-proteins. Interestingly, 5-HT1B recep-

tors did not decrease 5-HT release in this preparation

(el Mansari & Blier, 1996). In contrast, Trillat et al.

(1997) showed that the inhibition of cortical and hippo-

campal 5-HT release was absent in 5-HT1B receptor knock-

out mice. Thus, the respective roles of the 5-HT1B and

5-HT1D receptors in the inhibition of 5-HT release remain

unresolved in many brain regions.

Inhibition of glutamate release from rat cerebellar

synaptosomes and from human neocortex is attributable

to 5-HT1D receptors (Raiteri et al., 1986; Maura & Raiteri,

1996; Maura et al., 1998). Molderings et al. (1996)

demonstrated that presynaptic 5-HT1D receptors can inhibit

norepinephrine release from isolated human right atrial

appendage sections.

2.4. The 5-hydroxytryptamine1E receptor

The 5-HT1E receptor originally was identified as a

component of [3H]5-HT binding to human cortical homo-

genates that was resistant to a cocktail of antagonists of

5-HT1A, 5-HT1B/D, and 5-HT2 receptors (Leonhardt et al.,

1989). The intronless 5-HT1E receptor gene was subse-

quently cloned from human and rat (Levy et al., 1992a;

McAllister et al., 1992; Zgombick et al., 1992; Lovenberg

et al., 1993b). The human gene encodes a protein of 365

amino acids. This protein shares 39% homology with the

5-HT1A receptor and 47% identity (64% in the transmem-

brane regions) with the 5-HT1B/D receptors. The 5-HT1E

receptor gene has been localized to human chromosome

6q14-q15 (Levy et al., 1994).

There are few details about signaling pathways linked to

the 5-HT1E receptor. Low concentrations of agonist have

been shown to inhibit AC in HeLa and BS-C-1 cells trans-

fected with the 5-HT1E receptor (Lovenberg et al., 1993b;

Adham et al., 1994). The 5-HT1E receptor can also stimulate

AC, as high concentrations of agonist also promote

increases in cAMP in BS-C-1 cells (Adham et al., 1994).

The 5-HT1E receptor has not yet been demonstrated to

stimulate PLC and/or PLA2. In BS-C-1 cells, 5-HT had no

effect on inositol phosphate release, intracellular Ca2 +

levels, or AA mobilization (Adham et al., 1994).

2.5. The 5-hydroxytryptamine1F receptor

The 5-HT1F receptor originally was cloned from

mouse, and was termed 5-HT1Eb, based on pharmaco-

logical similarities with the 5-HT1E receptor (Amlaiky

et al., 1992). The human gene is intronless like the mouse

gene, and it is located on chromosome 3q11 (Saudou &

Hen, 1994).

The 5-HT1F receptor gene encodes a protein of 366

amino acids, with 70% homology to the 5-HT1E receptor,

60% to the 5-HT1B receptor (60%), and 63% to the 5-HT1D

receptor. 5-HT1F receptor mRNA was detected in human

brain, uterus, and mesentery, but not in kidney, liver, spleen,

heart, pancreas, and testes (Adham et al., 1993b). When

stably expressed in NIH 3T3 cells, the 5-HT1F receptor

negatively couples to AC (Amlaiky et al., 1992; Adham

et al., 1993b).

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212190

The 5-HT1F receptor has been shown to couple to PI-PLC

in a cell-specific manner. When transfected into NIH 3T3

cells, no elevation of inositol phosphates or Ca2 + was

observed, whereas in LM(tk-) cells, the receptor stimulated

accumulation of inositol phosphates and induced a rapid

increase of Ca2 + . PLC activation was blocked by pertussis

toxin, supporting a role for the Gi/o-protein (Adham et al.,

1993a). The significance of the coupling to PLC is uncertain,

as the transfected cells used in this study expressed high

levels of receptors (4.4 pmol/mg of protein). The 5-HT1F

receptor has been demonstrated to decrease capsaicin-

induced c-fos expression in the rat trigeminal nucleus

caudalis, although the signaling pathway of this effect was

not defined (Mitsikostas et al., 1999).

3. The 5-hydroxytryptamine2 receptors

There are three members of the 5-HT2A receptor family,

termed 5-HT2A, 5-HT2B, and 5-HT2C (Hoyer et al., 1994).

The 5-HT2A receptor is probably the 5-HT M receptor

described by Gaddum and Picarelli (1957). The 5-HT2B

receptor was formerly referred to as the 5-HT2F receptor

[or serotonin receptor like (SRL)] (Foguet et al., 1992b),

and the 5-HT2C receptor was previously referred to as the

5-HT1C receptor. The 5-HT2 receptors couple consistently

to the PLC-b second messenger pathway (Peroutka, 1995)

in native tissues (Conn & Sanders-Bush, 1984) and hetero-

logous cells (Pritchett et al., 1988). Like the 5-HT1 recep-

tors, the 5-HT2 receptors can couple to other second

messenger pathways in a cell-specific manner. Unlike the

5-HT1 receptors, the 5-HT2 receptor genes have introns.

Although the 5-HT2 receptors are similar in structure,

pharmacology, and signaling pathways, there are a few

differences in their signaling properties (Berg et al., 1994b,

1996; Grotewiel & Sanders-Bush, 1999). These differences

will be discussed in Sections 3.1–3.3.

3.1. The 5-hydroxytryptamine2A receptor

The 5-HT2A receptor originally was identified as a

[3H]spiperone-binding site, with low affinity for 5-HT

(Leysen et al., 1978; Peroutka & Snyder, 1979). It was

classified as a 5-HT receptor based on pharmacological

similarities with other 5-HT receptors. The rat receptor

gene was identified in 1988 by Pritchett et al., and the

human receptor gene was identified in 1990 by Julius et al.

The rat and human genes are 87% homologous. The

human 5-HT2A receptor gene is located on chromosome

13q14-q21 (Hsieh et al., 1990). The 5-HT2A receptor

shares 45% homology with the 5-HT2B receptor and

49% homology with the 5-HT2C receptors (80% in the

transmembrane regions). The 5-HT2A receptor is widely

distributed in the brain (cortex, caudate nucleus, olfactory

tubercle, nucleus accumbens, and hippocampus) (Pazos et al.,

1985), skeletal (Guillet-Deniau et al., 1997; Hajduch et al.,

1999) and smooth muscle (Kuemmerle et al., 1995; Ellwood

& Curtis, 1997), kidneys (Garnovskaya et al., 1995), and

platelets (Cook et al., 1994).

3.1.1. The 5-hydroxytryptamine2A receptor activates

phospholipase C

The 5-HT2A receptor activates PLC-b in most tissues and

cells in which it is expressed, resulting in accumulation of

inositol phosphates and elevations of intracellular Ca2 +

(Conn & Sanders-Bush, 1984, 1985; Roth et al., 1984,

1986; Tamir et al., 1992; Berg et al., 1998; Briddon et al.,

1998; Grotewiel & Sanders-Bush, 1999). These effects can

result in activation of L-type Ca2 + channels (Watts, 1998)

and stimulation of PKC (Takuwa et al., 1989).

3.1.2. The 5-hydroxytryptamine2A receptor activates

other phospholipases

The 5-HT2A receptor activates other PLs, such as PLA2

and PLD. 5-HT2A receptors in CHO cells and 1C11 cells

activate PLA2-mediated AA release (Berg et al., 1996, 1998;

Tournois et al., 1998). In C6 glial cells, the 5-HT2A receptor is

coupled to themobilization of polyunsaturated fatty acids and

AA (Garcia & Kim, 1997). 5-HT2A receptor-induced pros-

taglandin release is thought to play a major role in the

acceleration of gastrointestinal transit (Matsuda et al.,

2000). The 5-HT2A receptor stimulates amyloid precursor

protein ectodomain secretion by a PLA2-dependent pathway

(Nitsch et al., 1996).

The 5-HT2A receptor has been shown to couple to

PLD in cultured rat renal mesangial cells (Kurscheid-

Reich et al., 1995). This effect is cell-specific, as the re-

ceptor does not couple to PLD in guinea pig trachea (Cox

et al., 1996).

3.1.3. The 5-hydroxytryptamine2A receptor can regulate

cyclic AMP accumulation in certain cells

The 5-HT2A receptor does not typically regulate cAMP

formation in most cells or tissues (Lucaites et al., 1996), but

it can both stimulate and diminish cAMP accumulation

in specific cell types. Berg et al. (1994a) showed that the

5-HT2A receptor can amplify cAMP formation in A1A1

cells by the intermediate actions of PKC-a and/or PKC-dand Ca2 + /CaM. In FRTL-5 thyroid cells, the 5-HT2A

receptor increases cAMP through a pertussis toxin-sensitive

mechanism (Tamir et al., 1992).

Garnovskaya et al. (1995) demonstrated that the 5-HT2A

receptor in rat renal mesangial cells could inhibit forskolin-

stimulated cAMP accumulation in intact cells and AC

activity in washed membranes. The inhibition of cAMP

accumulation did not involve PLC, Ca2 + , or PKC, but was

sensitive to pertussis toxin, suggesting that the 5-HT2A

receptor in mesangial cells might inhibit cAMP through

a membrane-delimited pathway that requires activation

of Gi/o-proteins.

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212 191

3.1.4. 5-hydroxytryptamine2A receptor activates the

extracellular signal-regulated mitogen-activated

protein kinase

Several groups have documented that the 5-HT2A

receptor activates ERK MAP kinases in cells with con-

tractile phenotypes, vascular smooth muscle cells (Florian

& Watts, 1998; Watts, 1998) and renal mesangial cells

(Grewal et al., 1999; Greene et al., 2000). In vascular

smooth muscle cells, the activation of ERK was complex,

requiring inputs from PLC, L-type Ca2 + channels, and

MEK1 (Florian & Watts, 1998; Watts, 1998). In mesangial

cells, the pathway also involves stimulation of PKC,

activation of an NAD(P)H oxidase-like enzyme, and pro-

duction of reactive oxygen species (H2O2 and/or super-

oxide) (Grewal et al., 1999; Greene et al., 2000). The

activation of ERK in vascular smooth muscle cells is in-

volved in contraction, whereas the activation in mesangial

cells results in proliferation (Takuwa et al., 1989); induction

of early response genes, such as Egr-1 and cyclo-oxygenase-

2 (Goppelt-Struebe & Stroebel, 1998); and up-regulation of

the matrix-regulatory cytokine, TGF-b (Grewal et al., 1999).In contrast, ERK does not appear to be involved in the up-

regulation of another of the matrix-regulatory cytokines

(connective tissue growth factor) by the 5-HT2A receptor

in mesangial cells. Rather, a small G-protein (Rho) and the

actin cytoskeleton appear to mediate the up-regulation

(Hahn et al., 2000).

3.1.5. The 5-hydroxytryptamine2A receptor activates the

Janus kinase/signal transducers and activators of

transcription pathway

Guillet-Deniau et al. (1997) identified a 5-HT2A receptor

in rat skeletal muscle myoblasts that is coupled to the Janus

kinase (Jak)/signal transducers and activators of transcrip-

tion (STAT) pathway. The receptor was detected on the

plasma membrane and in T-tubules in contracting myo-

tubes. It was shown to increase the expression of genes

involved in myogenic differentiation, and to trigger a

rapid tyrosine phosphorylation of Jak2 and STAT3, result-

ing in translocation into the nucleus of STAT3. The

5-HT2A receptor co-precipitated with Jak2 and STAT3,

indicating that they are physically associated (Guillet-

Deniau et al., 1997).

3.1.6. The 5-hydroxytryptamine2A receptor

regulates apoptosis

Increasing levels of 5-HT in the cerebral cortex coincide

with developmental events, such as cell proliferation, sur-

vival, and differentiation. Dooley et al. (1997) used a cell-

survival assay to show that stimulation of the 5-HT2A

receptor promoted the survival of rat cortical progenitor cells

in vitro. They further demonstrated that this effect was

specific to the glutamate-containing post-mitotic neuron

population, and not to dividing neuroepithelial cells in the

developing cortex (Dooley et al., 1997).

3.1.7. The 5-hydroxytryptamine2A receptor

regulates channels

The 5-HT2A receptor increases intracellular Ca2 + levels

by liberating intracellular stores of Ca2 + and/or by activ-

ating Ca2 + channels, depending upon the cell of interest.

The 5-HT2A receptor may activate L-type Ca2 + channels in

some cell types (Eberle-Wang et al., 1994; Jalonen et al.,

1997; Watts, 1998). Although because of significant overlap

in the pharmacology of the L-type Ca2 + channels and the

5-HT2A receptor antagonists (Okoro, 1999), one should

proceed with prudence before attributing effects of the

5-HT2A receptor to L-type Ca2 + channels. The Ca2 + channels

coupled to the 5-HT2A receptor have been characterized as

both voltage-dependent and -independent (Nakaki et al.,

1985; Eberle-Wang et al., 1994; Hagberg et al., 1998). The

increases in Ca2 + levels evoked by the 5-HT2A receptor

have been linked to subsequent opening of Ca2 + -activated

K + channels in C6 glial cells (Bartrup & Newberry, 1994)

and to an inward current mediated through Ca2 + -activated

Cl- channels in Xenopus oocytes (Montiel et al., 1997).

In rat cortical astrocytes, the 5-HT2A receptor activates

both an L-type Ca2 + channel and an apamin-sensitive

Ca2 + -activated small conductance K + channel (Jalonen

et al., 1997).

3.1.8. The 5-hydroxytryptamine2A receptor causes

production of reactive oxygen and nitrogen species

Preliminary reports suggest that the 5-HT2A receptor can

either stimulate or inhibit NO synthesis in certain tissues and

cells. For example, gastrointestinal transit is regulated by

5-HT2A receptor-induced release of NO (Matsuda et al.,

2000). On the other hand, 5-HT2A receptors also inhibit

cytokine-stimulated inducible NOS in C6 glioma cells

(Miller & Gonzalez, 1998).

The 5-HT2A receptor in renal mesangial cells has been

shown to induce the production of H2O2 and superoxide

through the action of an NAD(P)H oxidase-like enzyme

(Grewal et al., 1999). The enzyme resides downstream from

PKC, and is an important upstream activator of ERK by the

5-HT2A receptor in those cells (Greene et al., 2000).

3.1.9. The 5-hydroxytryptamine2A receptor

regulates calmodulin

It is not surprising that CaM should be a target for 5-HT2

receptor signaling, in that CaM is a major signaling target

of Ca2 + mobilization in most cells types. There are three lines

of indirect evidence that the 5-HT2A receptor signals through

CaM. First, agonist-mediated up-regulation of the 5-HT2A

receptor depends upon CaM and Ca2 +/CaM-dependent

kinase 2 (Chen et al., 1995). Second, Berg et al. (1994a)

showed that the 5-HT2A receptor increases cAMP formation

in A1A1 cells by the intermediate action of CaM. Third,

inhibition of CaM-dependent kinase 2 or calcineurin (a CaM-

dependent phosphatase) inhibits 5-HT-induced cyclo-oxy-

genase 2mRNA expression in renal mesangial cells (Stroebel

& Goppelt-Struebe, 1994; Goppelt-Struebe et al., 1999).

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212192

3.1.10. The 5-hydroxytryptamine2A receptor regulates

transport processes

Like the 5-HT1A receptor (see Section 2.1.11), the 5-HT2A

receptor has been linked to a variety of transport processes. It

activates the Type 1 Na + -proton exchanger in renal mesan-

gial cells (Saxena et al., 1993; Garnovskaya et al., 1995). The

5-HT2A receptor in airway smooth muscle stimulates the

Na + /K + -ATPase (Na + pump) through a pathway that might

involve activation of the Na + -proton exchange exchanger

(Rhoden et al., 2000). In contrast, 5-HT2A receptor trans-

fected into NIH 3T3 fibroblasts was unable to stimulate Na +

pump activity; rather, it activated Na + /K + /2Cl- cotransport

(Mayer & Sanders-Bush, 1994).

The 5-HT2A receptor expressed endogenously in rat renal

mesangial cells accelerates mediator-facilitated electron

fluxes across the plasma membrane through a process that

may be linked to the production of reactive oxygen species

(Grewal et al., 1999). The 5-HT2A receptor enhances release

of dopamine and norepinephrine, but not 5-HT, in the frontal

cortex of freely moving rats (Gobert & Millan, 1999). Both

5-HT2A and 5-HT2C receptors stimulate amyloid precursor

protein ectodomain secretion through a pathway that in-

volves PLA2 (Nitsch et al., 1996).

3.1.11. A unique signaling role for internalization of the

5-hydroxytryptamine2A receptor

The 5-HT2A receptor is internalized in response to both

agonists and antagonists, adding a very interesting twist to its

signaling properties (Bhatnagar et al., 2000). This feature of

the 5-HT2A receptor may play important roles in its signaling

and in the actions of antipsychotic medications. Bhatnagar

et al. (2000) examined the internalization process of this

receptor in detail, demonstrating that both agonist- and

antagonist-induced internalization of the 5-HT2A receptor

were dynamin-dependent. In contrast, both agonist- and

antagonist-induced internalizations of a 5-HT2A receptor-

green fluorescent protein fusion protein were insensitive to

three different dominant-negative mutants of arrestin.

Activation of the 5-HT2A receptor by agonists, but not anta-

gonists, induced greater translocation of arrestin-3 than

arrestin-2 to the plasma membrane, and resulted in differ-

ential sorting of arrestin-2, arrestin-3, and 5-HT2A receptors

into distinct plasma membrane and intracellular compart-

ments. It is likely that these differences in distribution of the

various signaling components induced by agonists and

antagonists may be important in the ‘‘ligand-directed’’ of

second messenger signals by the 5-HT2A receptor, depend-

ing upon which ligand is used to stimulate the receptor.

3.2. The 5-hydroxytryptamine2B receptor

The 5-HT2B receptor was first described as the receptor

that mediates contraction of the gastric fundus (Vane, 1959).

The receptor was cloned from rat and mouse in 1992

(Foguet et al., 1992a, 1992b) and from human in 1994

(Kursar et al., 1994). The receptor was formerly termed

either 5-HT2F (Wainscott et al., 1993) or SRL (Foguet et al.,

1992b). The human receptor protein is 481 amino acids in

length (Kursar et al., 1994), and it has 2 introns similar to

the 5-HT2A and 5-HT2C receptors. The human receptor has

45% homology with the 5-HT2A receptor and 42% homol-

ogy with the 5-HT2C receptor (Barnes & Sharp, 1999). The

human 5-HT2B receptor gene is localized to chromosome

2q36.3-2q37.1.

mRNA encoding the 5-HT2B receptor is expressed with

greatest abundance in the human liver and kidney. Lower

levels of expression have been detected in the pancreas and

spleen (Bonhaus et al., 1995), but the presence of mRNA for

the 5-HT2B receptor in the brain appears to be relatively

limited (Loric et al., 1992; Kursar et al., 1994). 5-HT2B

receptor immunoreactivity has been reported in neurons in

the cerebellum, dorsal hypothalamus, lateral septum, and

medial amygdala (Duxon et al., 1997).

3.2.1. The 5-hydroxytryptamine2B receptor activates

phospholipase C

Like the 5-HT2A and 5-HT2C receptors, the 5-HT2B

receptor couples consistently to PLC. The cloned rat and

human receptors stimulate inositol phosphate accumulation

when expressed heterologously in mammalian cells (Wain-

scott et al., 1993; Kursar et al., 1994; Schmuck et al., 1994;

Kellermann et al., 1998) or endogenously in IC11 cells (Loric

et al., 1995). The 5-HT2B receptor also has been reported to

stimulate Ca2 + mobilization in astrocytes derived from rat

cerebral cortex, hippocampus, and brain stem (Sanden et al.,

2000). The 5-HT2A and 5-HT2C receptors can activate other

PLs, but there is less evidence that the 5-HT2B receptor can

do so. The 5-HT2B receptor has been shown in one report to

activate PLA2 in IC11 cells (Tournois et al., 1998). However,

the 5-HT2B does not couple to PLD in the rat stomach fundus

(Cox et al., 1999).

3.2.2. The 5-hydroxytryptamine2B receptor can stimulate

cyclic amp accumulation

The 5-HT2B receptor does not typically regulate cAMP

production, but when heterologously expressed in AV12

cells, it induces a modest increase in cAMP levels (Lucaites

et al., 1996).

3.2.3. The 5-hydroxytryptamine2B receptor regulates

morphogenesis and mitogenesis

Choi et al. (1997) reported that 5-HT2B receptors regulate

embryonic morphogenesis, probably by preventing the dif-

ferentiation of cranial neural crest cells and myocardial

precursor cells. Their evidence was that 5-HT2B receptor

antagonists had specific effects on embryonic cultures, such

that they induced morphological defects in the cephalic

region, heart, and neural tube. 5-HT2B receptor antagonists

interfered with cranial neural crest cell migration and induced

their apoptosis. In the heart, antagonists caused abnormal

sarcomeric organization in the subepicardium and induced

the specific absence of the trabecular cell layer in the

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212 193

ventricular myocardium. The specific signaling pathways

involved in the regulation of cellular migration, apoptosis,

and morphogenesis remain to be defined, but a study by

Nebigil et al. (2000a) suggests that ErbB-2might be involved.

Nebigil et al. (2000a) generated 5-HT2B receptor knock-

out mice, which resulted in embryonic and neonatal death

caused by specific heart defects. The knockout embryos

lacked trabeculae in the heart, leading to midgestation

lethality. Surviving newborn mice displayed a severe

ventricular hypoplasia associated with impaired prolifera-

tive capacity of myocytes. In surviving adult knockout

mice, there were severe cardiac histopathological changes,

including myocyte disarray and ventricular dilation. They

also found that there was a specific reduction in the

expression levels of the receptor tyrosine kinase ErbB-2.

Their data suggest that the 5-HT2B receptor probably

requires ErbB-2 in the signaling pathway for cardiac

differentiation (Nebigil et al., 2000a).

3.2.4. The 5-hydroxytryptamine2B receptor activates

the extracellular signal-regulated kinase and cell

cycle components

Launay et al. (1996) showed that 5-HT2B receptors stably

expressed in mouse fibroblast LMTK- cells rapidly activate

the proto-oncogene Ras and ERK1/ERK2 MAP kinases.

This pathway also involves Gaq and Gbg. This pathway also

stimulated 5-HT-dependent proliferation of the transfected

cells. The 5-HT2B receptor expressed in LMTK- cells can

induce foci formation, and when injected into nude mice,

the foci cause tumors (Launay et al., 1996). This effect is

mediated through the Ras-ERK pathway.

Nebigil et al. (2000b) presented evidence that the

5-HT2B receptor can stimulate cell cycle progression

through an ERK-dependent pathway. The pathway also

leads to hyperphosphorylation of Rb through up-regulation

and activation of both cyclin D1/cdk4 and cyclin E/cdk2

kinases. The pathway leading to up-regulation and activa-

tion of ERK and cyclin D1, but not cyclin E, requires the

intermediate activation of the platelet-derived growth factor

(PDGF) receptor kinase signaling pathways. Moreover,

activation of the PDGF receptor, activation of ERK, up-

regulation and activation of both cyclin D1 and cyclin E,

and cellular proliferation all depend upon the non-receptor

tyrosine kinase Src (Nebigil et al., 2000b). Thus, the

5-HT2B receptor stimulates cellular proliferation through

two parallel pathways. These are as follows: (1) 5-HT2B

receptor ! Gq ! Src ! PDGF receptor tyrosine

kinase ! ERK ! cyclin D1/cdk4 ! Rb hyperphos-

phorylation ! proliferation and (2) 5-HT2B receptor !Gq ! Src ! ? ! cyclin E/cdk2 ! Rb hyperphos-

phorylation ! proliferation.

3.2.5. The 5-hydroxytryptamine2B receptorcauses production

of reactive nitrogen species

The 5-HT2B receptor increases production of NO in

human coronary artery endothelial cells (Ishida et al., 1998)

and in porcine cerebral artery, where this results in vaso-

relaxation (Florian & Watts, 1998). In contrast, a role for NO

in 5-HT2B receptor-mediated contraction of the stomach

fundus has been challenged (Cox & Cohen, 1995). Manivet

et al. (2000) published an elegant study in which they

documented that 5-HT2B receptors expressed endogenously

in IC11 cells and Mastomys natalensis carcinoid cells or

expressed heterologously in LMTK- fibroblasts stimulate

intracellular cGMP production through dual activation of

constitutive NOS (cNOS) and inducible NOS (iNOS). This

coupling relied on the Group I PDZ motif in the carboxyl

terminal tail of the 5-HT2B receptor, a conclusion based on

competition with a carboxyl terminal fragment of the receptor

and upon studies in which a mutant receptor lacking the PDZ

interaction motif could not couple to cNOS or iNOS (Man-

ivet et al., 2000). Neutralizing antibodies raised against Ga13

blocked activation of iNOS, but not cNOS, activity. Thus, the

5-HT2B receptor can activate NO production through two

overlapping pathways: (1) 5-HT2B receptor ! PDZ inter-

action motif + Ga13 ! iNOS ! NO/cGMP and 5-HT2B

receptor ! PDZ interaction motif ! cNOS ! NO/cGMP

(Manivet et al., 2000). Thus, the ability of the 5-HT2B

receptor to increase NO levels depends upon interactions

with two isoforms of NOS.

3.2.6. The 5-hydroxytryptamine2B receptor

regulates channels

The 5-HT2B receptor induces release of Ca2 + from ryano-

dine-sensitive intracellular stores in human pulmonary artery

endothelial cells through a pathway that involves activation

of ryanodine receptors, but is independent of accumulation

of inositol phosphates (Ullmer et al., 1996). In contrast,

in Xenopus oocytes, the 5-HT2B receptor activates Ca2 +

oscillations through a pathway that involves IP3-dependent

release of internal Ca2 + stores, as well as Ca2 + influx. The

increased Ca2 + leads to activation of a Ca2 +-dependent Cl-

channel (Foguet et al., 1992a; Parekh et al., 1993).

3.2.7. The 5-hydroxytryptamine2B receptor regulates

transport processes

Leung et al. (1999) used the short-circuit current tech-

nique to demonstrate that the 5-HT2B receptor can stimulate

bicarbonate secretion through a process that is dependent

upon prostaglandin synthesis in cultured rat epididymal

epithelium. The 5-HT2B receptor may be able to regulate

other transport processes, in that it has been reported to

stimulate phosphorylation of the 5-HT transporter SERT and

of the Na +/K +-ATPase (Launay et al., 1998).

3.3. The 5-hydroxytryptamine2C receptor

The 5-HT2C receptor (originally called the 5-HT1C recep-

tor) was identified as a high-affinity [3H]5-HT-binding site in

the choroid plexus (Pazos et al., 1985). The cloning of mouse

(Lubbert et al., 1987; Yu et al., 1991), rat (Julius et al., 1988),

and human (Saltzman et al., 1991) 5-HT2C receptors led to the

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212194

recognition that the 5-HT2C receptor is much more closely

related to the 5-HT2 receptor family than to the 5-HT1

receptor family. The receptor gene is located on chromosome

Xq24, and contains three introns (Yu et al., 1991; Stam et al.,

1994; Xie et al., 1996). A nonfunctional truncated splice

variant of the 5-HT2C receptor has been reported (Canton

et al., 1996; Xie et al., 1996). To date, no other splice

variants of the 5-HT2C receptor have been identified. How-

ever, the 5-HT2C receptor exhibits a very unique mechanism

of generating multiple functional receptor variants through a

process called mRNA editing (Burns et al., 1997). This

unique feature will be discussed in Section 3.3.8.

The 5-HT2C receptor is expressed nearly exclusively in

the brain. High levels of 5-HT2C receptor expression have

been detected by ligand autoradiography, immunocyto-

chemistry, and in situ hybridization in the choroid plexus,

the cortex, the nucleus accumbens, the amygdala, the

hippocampus caudate nucleus, and the substantia nigra

(Pazos et al., 1985; Mengod et al., 1990; Palacios et al.,

1990; Abramowski et al., 1995).

3.3.1. The 5-hydroxytryptamine2C receptor activates

phospholipase C

The 5-HT2C receptor has long been recognized as being

capable of activating PLC in the choroid plexus (Sanders-

Bush et al., 1988) and other brain regions (Wolf & Schutz,

1997). A major signal transduction cascade linked to the

5-HT2C receptor in transfected cells is also stimulation of

PLC-mediated inositol phosphate accumulation (Julius et al.,

1988; Berg et al., 1994b, 1998; Briddon et al., 1998;

Grotewiel & Sanders-Bush, 1999), although it can activate

PLA2-mediated AA release as well (Berg et al., 1996,

1998). The activation of PLC probably is mediated through

Gq/11a-proteins, although Chen et al. (1994) presented evid-

ence that the 5-HT2C receptor could couple to Goa and Gia1 in

Xenopus oocytes. Chang et al. (2000) used an elegant method

of delivering membrane-permeable blocking peptides to

document that 5-HT2C receptors expressed in the choroid

plexus activate PLC through Gqa-proteins and not through

bg-subunits. In their study, peptide mimics of Gqa and the

Gqa interaction domain of PLCb1 prevented 5-HT2C recep-

tor-mediated accumulation of inositol phosphates, whereas

peptides derived from bg-binding sites of PLCb2 and a

phosducin-like protein C-terminus did not. The latter two

peptides were quite effective in blocking the Gbg-mediated

activation of ERK by the a2A-adrenergic receptor. Tohda

et al. (1995) also suggested that botulin toxin-sensitive, low-

molecular weight G-proteins are important in the inositol

phosphate accumulation induced by the 5-HT2C receptor in

COS-7 cells.

Like the 5-HT2A and 5-HT2B receptors, the 5-HT2C

receptor can also activate PLA2, but the story for the

5-HT2C receptor has an intriguing twist. Berg and colleagues

(1998) compared the ability of a panel of agonists to activate

both PLA2 and PLC in cells expressing transfected 5-HT2A

and 5-HT2C receptors. They showed that for the 5-HT2C

receptor, some agonists preferentially activated PLA2 or

PLC, when relative efficacies were referenced to 5-HT. This

is probably the best evidence for ‘‘agonist-directed traffick-

ing of receptor stimulus,’’ meaning that a single receptor

subtype can couple with different efficacies to several signal-

ing pathways, depending upon the nature of the agonist to

which the receptor is exposed. In contrast, the results for the

5-HT2A receptor showed that all agonists demonstrated

greater relative efficacies for PLA2 than for PLC.

3.3.2. The 5-hydroxytryptamine2C receptor can modulate

cyclic AMP accumulation

When expressed at physiological levels, the 5-HT2C

receptor has not been reported to modulate cAMP levels

in cells or tissues. When expressed at high density

(12 pmol/mg membrane protein) in stably transformed

AV12 cells, however, the 5-HT2C receptor has been shown

to inhibit forskolin-stimulated cAMP production, with an

IC50 of � 50 nM. This effect was sensitive to pertussis

toxin, suggesting the involvement of Gi/o-proteins. In

contrast, the 5-HT2A and 5-HT2B receptors expressed at

similar levels did not decrease AC activity in those cells.

Pretreatment of the cells expressing high levels of 5-HT2C

receptor with pertussis toxin also unmasked a small

stimulatory effect of the 5-HT2C receptor on cAMP accu-

mulation. When expressed at low density, the 5-HT2C

receptor could potentiate forskolin-stimulated cAMP pro-

duction by � 2-fold (Lucaites et al., 1996).

3.3.3. The 5-hydroxytryptamine2C receptor

regulates channels

The 5-HT2C receptor can regulate K + and Cl- channels.

The cloned 5-HT2C receptor has been shown to close K +

channels in the choroid plexus (Hung et al., 1993) and when

expressed in cell lines (Kelly et al., 1991; Panicker et al.,

1991). DiMagno et al. (1996) showed that the 5-HT2C

receptor, when co-expressed in Xenopus oocytes with rat

brain mRNA, could inhibit an inwardly rectified, Ba2 +-

sensitive K + conductance through a PKC-dependent path-

way. When co-expressed in Xenopus oocytes, the 5-HT2C

receptor also suppresses the activity of the Shaker-related

K + gene Kv1.3, which encodes the type n K + channel. This

effect on Kv1.3 currents proceeds via activation of a per-

tussis toxin-sensitive G-protein and a subsequent rise in

intracellular Ca2 + , but not via PKC, Ca2 + /CaM or phos-

phatases (Aiyar et al., 1993). Timpe and Fantl (1994) used a

similar system to show that the 5-HT2C receptor suppressed

the activity of the voltage-activated K + channel Kv1.5

through a PLC-dependent pathway.

The 5-HT2C receptor has also been shown to stimulate an

apical Cl- conductance in mouse choroid plexus (Hung

et al., 1993). 5-HT2C receptors have been shown to inhibit

g-aminobutyric acid (GABA)-A receptor channels by a

Ca2 + -dependent, phosphorylation-independent mechanism

in Xenopus oocytes (Huidobro-Toro et al., 1996). 5-HT2C

receptors expressed in Xenopus oocytes can stimulate Ca2 +

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212 195

release from IP3-sensitive intracellular stores, which results

in the opening of Ca2 + -gated Cl- channels (Walter et al.,

1991; DiMagno et al., 1996). This effect occurs through

endogenous Go-proteins in Xenopus oocytes. The same Cl-

current could also be stimulated by co-expression of four

isoforms of mammalian G-protein a-subunits (Goa, Gob,

Gqa, G11a), along with the 5-HT2C receptor. The effect was

mediated by a PLC-b endogenous to the Xenopus oocytes.

The same assay failed to detect any modulation of cAMP or

of three G-proteins (Gta, Golfb, Gsa, G11a) by the 5-HT2C

receptor (Quick et al., 1994). The physiological significance

of these effects is not known.

3.3.4. The 5-hydroxytryptamine2C receptors

regulate mitogenesis

The recombinant 5-HT2C receptor, when expressed in

NIH 3T3 cells, results in the generation of transformed

foci. Moreover, the injection of cells derived from trans-

formed foci into nude mice has been shown to result in the

generation of tumors (Julius et al., 1989). However, when

expressed in CCL36 fibroblasts, the 5-HT2C receptor

mediates only a very weak transforming activity (Kahan

et al., 1992).

3.3.5. The 5-hydroxytryptamine2C receptor can regulate

nitric oxide levels

The 5-HT2C receptor has been shown to both inhibit and

stimulate production of NO. It inhibits NMDA-induced

cGMP production and increases NO in rat cerebellum

(Marcoli et al., 1997) and in human neocortical slices

(Maura et al., 2000). In contrast, the 5-HT2C receptor can

stimulate cGMP and NO production in the choroid plexus

through a pathway that requires Ca2 + , PLA2, and lipox-

ygenase activity (Kaufman et al., 1995).

3.3.6. The 5-hydroxytryptamine2C receptor regulates

transport processes

The 5-HT2C receptor has been shown to modulate a

number of distinct transport processes. Recombinant

5-HT2C receptors expressed in 3T3 cells can activate pro-

cessing of amyloid precursor proteins. This process involves

both PKC and PLA2 (Nitsch et al., 1996).

The 5-HT2C receptor activates an electrogenic Na + /

Ca2 + exchanger in histaminergic neurons in the rat hypo-

thalamic tuberomammillary nucleus. This results in a doub-

ling of the firing rates of the neurons and depolarization

(Eriksson et al., 2001). The 5-HT2C receptor might also

stimulate prolactin (Cowen, P. J. et al., 1996), corticoster-

one, and ACTH secretion (Fuller, 1996), and inhibit the

neuronal release of norepinephrine and dopamine, but not

5-HT (Millan et al., 1998).

3.3.7. Possible signaling through PS-95 discs-large ZO-1

interaction motifs by the 5-hydroxytryptamine2C receptor

The 5-HT2C receptor has been shown in a yeast two-

hybrid screen to interact with a novel multivalent PDZ

protein called MUPP1. The C-terminus of the 5-HT2C

receptor selectively interacts with the tenth PDZ domain

of MUPP1. Both proteins were shown to interact by con-

focal microscopy and by co-immunoprecipitation studies in

COS-7 cells. Thus, Becamel et al. (2001) proposed that

MUPP1 might function as a multivalent scaffold protein,

which selectively assembles and targets 5-HT2C receptor-

signaling complexes. The interaction between MUPP1 and

the 5-HT2C receptor is functionally significant in two

regards. First, the interaction leads to a conformational

change in MUPP1 (Becamel et al., 2001). Second, deletion

of the 5-HT2C receptor PDZ recognition motif prevents

phosphorylation of the receptor and delays resensitization

of receptor responses (Backstrom et al., 2000). These

findings open up an exciting new area of GPCR signaling

that may function independently from, or in concert with,

G-protein-dependent pathways.

3.3.8. Multiple novel functional 5-hydroxytryptamine2Creceptors are created by mRNA editing

The 5-HT2C receptor exhibits a very unique mechanism

of generating multiple functional receptor variants through

a process called mRNA editing (Burns et al., 1997).

Niswender et al. (1998) examined the possibility that the

5-HT2A and 5-HT2B receptors might also undergo RNA

editing in the i2 loop using a polymerase chain reaction

strategy, but were unable to show any evidence that this

process occurs for those receptors. They also noted that five

genomically encoded adenosines within the putative i2 loop

of the 5-HT2C receptor were converted to guanosines at the

RNA level. These sites are referred to as sites A–E. The

editing of the mRNA probably occurs by de-amidation of

the adenosines to inosines, which are read at the RNA level

as guanosines. In the human 5-HT2C receptor, various

combinations of A ! G conversions result in at least 21

discrete mRNA species encoding 14 different editing var-

iants of the 5-HT2C receptor. The genomic (unedited)

receptor expresses amino acids INI (in the sequence IRNPI

in i2), whereas conversion of all three amino acids results in

VNV, VSV, VGV, or VDV (Niswender et al., 1998; Back-

strom et al., 1999; Fitzgerald et al., 1999). These are referred

to as 5-HT2C-INI, 5-HT2C-VNV, 5-HT2C-VSV, 5-HT2C-VGV, and

5-HT2C-VDV receptors. Other human receptor variants are

5-HT2C-VNI, 5-HT2C-VSI, 5-HT2C-ISI, 5-HT2C-INV, 5-HT2C-ISV,

5-HT2C-IGI, 5-HT2C-VGI, 5-HT2C-IDI, and 5-HT2C-IDV recep-

tors. The situation in rats is not identical, in that only 4 A

! G conversion sites (termed sites A–D) have been

identified, resulting in 11 discrete mRNAs and 7 different

receptor protein isoforms (Fitzgerald et al., 1999; Niswender

et al., 1999).

The functional significance of the various edited iso-

forms of the 5-HT2C receptor is underscored by their

differential abundances of expression in total brain and

hypothalamic mRNA populations (Fitzgerald et al., 1999).

Moreover, the variant edited receptors exhibit differential

abilities to bind various ligands, to mobilize intracellular

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212196

Ca2 + , and to stimulate accumulation of inositol phosphates

(Niswender et al., 1998; Backstrom et al., 1999; Fitzgerald

et al., 1999).

4. The 5-hydroxytryptamine4 receptor

Three types of 5-HT receptors couple primarily to the

activation of AC: the Gs-coupled 5-HT4, 5-HT6, and 5-HT7

receptors (Hamblin et al., 1998). Unlike the 5-HT6 and

5-HT7 receptors, the 5-HT4 receptor was well-characterized

pharmacologically and functionally (Dumuis et al., 1988;

Bockaert et al., 1990) prior to its cloning. The major

functional effects of the 5-HT4 receptors are prokinetic

actions in the gut and positive inotropy, chronotropy, and

lusitropy in atria, but not ventricles (Kaumann, 1991).

The human 5-HT4 gene is thought to be highly complex,

with multiple introns (Bockaert et al., 1998b). It maps to

chromosome 5q31-q33 (Claeysen et al., 1997). In 1995,

Gerald et al. reported the isolation of two rat brain 5-HT4

receptors. These were originally termed 5-HT4S and 5-HT4L

to indicate that they are short (387 amino acids) and long

versions (406 amino acids) of the 5-HT4 receptor (Gerald

et al., 1995). Similar murine cDNA homologues of these

receptors were subsequently reported (Claeysen et al.,

1996), and these were proposed to be renamed 5-HT4a

and 5-HT4b receptors, to be more in keeping with the

recommendations of the Nomenclature Committee of the

International Union of Pharmacology (Hoyer et al., 1994).

Two later reports suggested that the correct length of the

longer isoform (5-HT4b receptor) is actually 388 amino

acids (Van den Wyngaert et al., 1997; Bockaert et al.,

1998b). The putative protein products share sequence iden-

tity from amino acids 1–358, and diverge in amino acid

sequence after L358. Subsequently, four other splice variants

of this receptor have been shown to differ only in the length of

their intracellular carboxyl terminal tails. The human 5-HT4c

and 5-HT4d receptors are 380 and 360 amino acids in length,

respectively (Blondel et al., 1998; Claeysen et al., 1998). The

human 5-HT4e receptor is 378 amino acids in length (Claey-

sen et al., 1999). The corresponding rodent 5-HT4e receptor is

371 amino acids in length. The 5-HT4f receptor has been

identified in the mouse and is 363 amino acids in length; a

human 5-HT4f receptor has not been identified yet (Claeysen

et al., 1999).

There may be subtle differences in the distribution of

5-HT4a and 5-HT4b receptor splice variants (Shen et al.,

1993; Blondel et al., 1997), although this has not been a

universal finding (Claeysen et al., 1996; Vilaro et al., 1996).

5-HT4a, 5-HT4b, and 5-HT4c receptors are expressed in the

brain, the atrium, and the gut, whereas the 5-HT4d receptor

has only been detected in the gut (Blondel et al., 1998). The

human 5-HT4d receptor is localized in the atrium and the

brain, but not in the gut (Mialet et al., 2000), whereas mouse

and rat 5-HT4d receptors and mouse 5-HT4e and 5-HT4f

receptors are brain-specific (Claeysen et al., 1999).

4.1. The 5-hydroxytryptamine4 receptor activates

adenylyl cyclase

The 5-HT4 receptors prototypically stimulate AC when

expressed endogenously in tissues (Dumuis et al., 1988;

Bockaert et al., 1990; Ford et al., 1992; Albuquerque

et al., 1998; Bach et al., 2001) or heterologously in cells

(Gerald et al., 1995; Blondel et al., 1998; Bockaert et al.,

1998b; Claeysen et al., 1998; Mialet et al., 2000). When

transiently expressed in COS-7 cells, four splice variants

of the 5-HT4 receptor displayed similar abilities to

stimulate AC activity in the presence of 5-HT (Blondel

et al., 1998).

4.2. The 5-hydroxytryptamine4 receptor activates protein

kinase A

As would be expected for receptors whose primary

effect is to increase cAMP levels, most of the effects of

the 5-HT4 receptors are due to activation of PKA. These

include positive inotropic, chronotropic, and lusitropic

effects in the atria (Kaumann, 1991); inhibition of a K +

current in colliculi neurons (Fagni et al., 1992); facilitation

of striatal dopamine release (Steward et al., 1996); relaxa-

tion of the circular smooth muscle of the human colon

(McLean & Coupar, 1996); and regulation of the Ca2 + -

activated K + current in adult hippocampal neurons (Torres

et al., 1995).

4.3. The 5-hydroxytryptamine4 receptor regulates channels

5-HT4 receptors can regulate a variety of channels. For

example, activation of L-type Ca2 + channels (Kaumann,

1991; Ouadid et al., 1992) is probably responsible for the

prokinetic actions of the 5-HT4 receptors in the gut and

positive inotropy in the heart by activating cardiac L-type

Ca2 + channels. In human atrial myocytes, activation of

L-type Ca2 + channels by the 5-HT4 receptor occurs via an

elevation of intracellular cAMP levels and stimulation of

PKA (Ouadid et al., 1992). The 5-HT4 receptor increases the

If pacemaker current in atrial myocytes (Pino et al., 1998)

and stimulates a Cl- current in human jejunal mucosa and

rat distal colon (Budhoo et al., 1996a, 1996b). The 5-HT4

receptor stimulates a tetrodotoxin-insensitive Na + current

in Type II dorsal root ganglion cells through a diffusible

second messenger pathway that does not appear to involve

cAMP (Fagni et al., 1992). Thus, this receptor may modulate

channel function through pathways other than the cAMP/

PKA pathway.

The 5-HT4 receptor does not universally stimulate

channels, in that it inhibits a Ca2 + -activated K + current

responsible for the slow afterhyperpolarization in adult

hippocampal neurons (Andrade & Chaput, 1991). This

effect is mediated by increased cAMP and activation of

PKA (Torres et al., 1995). Ansanay et al. (1995) showed

that the 5-HT4 receptor induces long-lasting inhibition of

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212 197

a K + current in murine colliculi neurons through a PKA

and phosphatase-regulated pathway. 5-HT4 receptors also

inhibit (long-lasting) a delayed rectifier voltage-dependent

K + current similar to KV3.2 (Bockaert et al., 1998a).

The 5-HT4 receptor in mouse colliculi neurons inhibits a

voltage-activated K + current through cAMP and PKA

(Fagni et al., 1992).

4.4. The 5-hydroxytryptamine4 receptor regulates transport

5-HT stimulates aldosterone secretion in humans through

5-HT4 receptors positively coupled to AC (Lefebvre et al.,

1997, 2000). 5-HT4 receptors also facilitate rat striatal dop-

amine release in vitro and in vivo (Steward et al., 1996),

release of acetylcholine from rat frontal cortex (Consolo et al.,

1994), and release of 5-HT from rat hippocampus (Ge &

Barnes, 1996).

4.5. Functional differences among the 5-hydroxytryptamine4receptor splice variants

Although four splice variants of the 5-HT4 receptor, when

transiently expressed in COS-7 cells, were reported to

display similar abilities to stimulate AC activity in the

presence of 5-HT (Blondel et al., 1998), some interesting

differences in the behavior of the splice variants have been

observed (Claeysen et al., 1999). The 5-HT4e and 5-HT4f

receptors induce significantly more agonist-independent

increases in AC activity than do longer isoforms of the

5-HT4 receptor splice variants. The authors presented data

in support of a model in which a serine-threonine rich

portion of the carboxyl terminus of the receptor that is

shared by all of the splice variants (residues 346–359) holds

the receptor in an inactive conformation. Longer splice

variants, such as 5-HT4a and 5-HT4b receptors, augment

the inhibitory function of residues 346–359, whereas

shorter splice variants, such as the 5-HT4e and 5-HT4f

receptors, do not, resulting in increased basal cAMP levels

(Claeysen et al., 1999). An alternative interpretation of their

data is that the carboxyl terminus may be involved in rapid

or constitutive desensitization. Thus, removal of the carbox-

ylterminus could increase basal activity by reducing con-

stitutive desensitization. Evidence for this mechanism has

been presented recently for the B2 bradykinin receptor

(Leeb-Lundberg et al., 2001). This fascinating story is likely

to become more interesting, as the 5-HT4b receptor splice

variants differ in their expression of distinct PDZ interaction

motifs (Hamblin et al., 1998; Claeysen et al., 1999), which

may alter their signaling properties.

5. The 5-hydroxytryptamine5 receptors

Currently, little is known about 5-ht5 receptors. As func-

tional 5-ht5 receptors have not been identified in vivo yet, the

lower case designation is used. Cloning experiments have

revealed two subtypes of the 5-ht5 receptor, termed 5-ht5a and

5-ht5b. Both 5-ht5a and 5-ht5b receptors have been cloned

from rat and mouse, but only the 5-ht5a receptor has been

cloned from human (Plassat et al., 1992; Erlander et al., 1993;

Matthes et al., 1993; Wisden et al., 1993; Rees et al., 1994).

The human 5-ht5A receptor gene encodes a protein of 357

amino acid residues, has a single intron, and is located on

chromosome 7q36 (Rees et al., 1994). The human 5-ht5breceptor cDNA and its gene have not been cloned yet, but the

gene is localized to chromosome 2q11-13 (Matthes et al.,

1993). The rodent 5-ht5b receptor genes contain a single

intron in the middle of the putative i3 loops, and they encode

proteins of 370 or 371 amino acids.

The 5-ht5 receptors have high affinity for lysergic acid

diethylamide (LSD) and 5-carboxamidotryptamine. Both

receptors are expressed in multiple brain regions, but not

in peripheral tissues (Plassat et al., 1992; Rees et al., 1994).

Waeber and Moskowitz (1995) were able to identify putat-

ive 5-ht5a receptors by comparative autoradiography with

[3H]5-carboxamidotryptamine and [125I]LSD in wild-type

and 5-ht5a receptor knockout mice. Studies with [3H]5-

carboxamidotryptamine (in the presence of 8-OH-DPAT,

GR127935, and spiperone) revealed no binding in knockout

mice, but intermediate levels of binding in wild-type mice in

the olfactory bulb and neocortex. This binding probably

represents the 5-ht5a receptor. They also observed high

densities of [125I]LSD binding (in the presence of clozapine

and spiperone) in the medial habenula of wild-type and

knockout mice. However, 5-carboxamidotryptamine com-

peted for [125I]LSD binding with an affinity of 2 nM in the

olfactory bulb and neocortex of wild-type mice and of 30 nM

in the habenula of knockout mice, suggesting that binding in

the habenula might be accounted for by 5-ht5b receptors

(Waeber et al., 1998). Those results were somewhat similar

to those obtained using an in situ hybridization assay, in

which the 5-ht5a receptor mRNAwas located predominantly

in the cerebral cortex, the hippocampus, the habenula, the

olfactory bulb, and the granular layer of the cerebellum of

the mouse (Plassat et al., 1992). Reverse transcription-

polymerase chain reaction of the human 5-ht5a receptor

showed expression of mRNA in many brain areas, but none

in peripheral tissues (Rees et al., 1994).

5.1. Functional coupling of the 5-hydroxytryptamine5receptors to signaling pathways

As of yet, the 5-ht5 receptors have not been convincingly

linked to specific physiological effects or signal transduction

cascades in mammalian cells. For example, when expressed

in COS-7 cells and NIH 3T3 cells, 5-ht5 receptors were

unable to modulate AC activity (Plassat et al., 1992). Two

studies have suggested that 5-ht5a receptors might be capable

of inhibiting forskolin-stimulated cAMP accumulation. Car-

son et al. (1996) presented evidence that the 5-ht5a receptor

heterologously expressed in C6 glioma cells could inhibit

cAMP accumulation. Similarly, when the 5-ht5a receptor was

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212198

expressed at very high levels in HEK 293 cells, it was shown

to inhibit cAMP accumulation (Francken et al., 1998).

Putative 5-ht5a/5-ht5b receptor binding to [125I]LSD in the

brain (Waeber et al., 1998) and transfected cells (Plassat et al.,

1992) is GTP-sensitive, suggesting that these receptors do

couple to G-proteins. However, the specific G-proteins to

which these receptors couple, their downstream signals, and

physiological effects remain elusive.

6. The 5-hydroxytryptamine6 receptor

Two groups cloned the rat 5-HT6 receptor cDNA in 1993,

although there was disagreement over the sequence of this

receptor (Monsma et al., 1993; Ruat et al., 1993a). This was

subsequently resolved as a sequencing error after the cloning

of the human 5-HT6 receptor in 1996 (Kohen et al., 1996).

The human 5-HT6 receptor is a 440 amino acid protein. The

gene contains two introns that correspond to regions of the

putative i3 and e3 loops (Monsma et al., 1993; Ruat et al.,

1993a). The gene for the receptor maps to the human

chromosome region 1p35-p36 (Kohen et al., 1996). The

5-HT6 receptor stimulates AC and has high affinity for

typical and atypical antipsychotics, including clozapine.

The receptor is expressed in several brain regions, most

prominently in the caudate nucleus, the olfactory tubercle,

the striatum, the hippocampus, and the nucleus accum-

bens (Gerard et al., 1996, 1997; Grimaldi et al., 1998;

Sleight et al., 1998). Low levels of 5-HT6 receptor

mRNA are also expressed in the adrenal gland and the

stomach (Monsma et al., 1993; Ruat et al., 1993a).

Surprisingly, 5-HT6 receptors appear to regulate cholin-

ergic (rather than dopaminergic) neurotransmission in the

brain, implicating it as a target for the treatment of learning

and memory disorders (Bourson et al., 1998; Branchek &

Blackburn, 2000).

6.1. The 5-hydroxytryptamine6 receptor activates

adenylyl cyclase

Transfection studies have shown that the 5-HT6 receptor

activates AC (Ruat et al., 1993a; Kohen et al., 1996; Baker

et al., 1998; Grimaldi et al., 1998). 5-HT6-like receptors

have been shown to stimulate AC in striatal neurons in

culture (Sebben et al., 1994) and in pig caudate membranes

(Schoeffter & Waeber, 1994). When transfected into HEK

293 cells, the 5-HT6 receptor stimulates AC5, a Gs-sensitive

AC, but not AC1 or AC8, which are CaM-stimulated ACs.

In contrast, 5-HT7 receptors expressed in the same cells

could also activate AC1 and AC8 (Baker et al., 1998).

6.2. The 5-hydroxytryptamine6 receptor splice variants

Because of the presence of introns in the putative i3 and

e3 loops of the 5-HT6 receptor, alternative splice variants of

this receptor could exist. Thus far, only two truncation

mutants have been identified. The first is expressed in brain,

but is probably not functional as it deletes the entire seventh

transmembrane region of the receptor (Monsma et al.,

1993). Another truncation mutant of the human 5-HT6

receptor in which the protein is prematurely terminated in

the third transmembrane region has been shown to encode a

nonfunctional protein (Olsen et al., 1999).

7. The 5-hydroxytryptamine7 receptor

The 5-HT7 receptor is the third 5-HT receptor subtype

shown to couple to Gs (Bard et al., 1993; Lovenberg et al.,

1993a; Plassat et al., 1993; Ruat et al., 1993b; Shen et al.,

1993). It is highly expressed in the CNS, especially in the

hippocampus, the hypothalamus, and the neocortex. It has

been speculated to participate in the control of circadian

rhythm because it is expressed in the suprachiasmatic

nucleus (Lovenberg et al., 1993a; Stowe & Barnes, 1998).

The 5-HT7 receptor is also expressed in glial cells (Hirst

et al., 1997; Shimizu et al., 1996), the spleen (Shen et al.

1993), vascular smooth muscle (Becker et al., 1992; Bard

et al., 1993; Ullmer et al., 1995; Schoeffter et al., 1996), and

the intestine (Bard et al., 1993). It has a high affinity for

atypical antipsychotic agents, and it may mediate some of

their effects (Shen et al., 1993). It is also expressed in the

periphery, where it mediates relaxation of smooth muscle in

various organs (Eglen et al., 1997).

5-HT7 receptors have been isolated from a number of

mammalian species (Bard et al., 1993; Lovenberg et al.,

1993a; Meyerhof et al., 1993; Plassat et al., 1993; Ruat

et al., 1993b; Shen et al., 1993; Tsou et al., 1994) and

also from X. laevis (Nelson et al., 1995). These receptors

appear to be homologues to the fruit fly 5-HTdro1 receptor

(Witz et al., 1990). The human 5-HT7 receptor initially was

reported to be a protein of 445 amino acids, with 57%

homologywith theDrosophila melanogaster 5-HTdro1 recep-

tor in the transmembrane regions, but only 39–53% homo-

logy with cloned human 5-HT1, 5-HT2, 5-HT5, and 5-HT6

receptors. Surprisingly, there is only 39% homology with

the 5-HT6 receptor within the transmembrane regions (Bard

et al., 1993). The human 5-HT7 receptor gene is located on

chromosome 10 (10q21-q24) (Gelernter et al., 1995), and

contains two introns (Ruat et al., 1993b; Erdmann et al.,

1996; Heidmann et al., 1997).

7.1. The 5-hydroxytryptamine7 receptor splice variants

The presence of introns in the 5-HT7 receptor gene is

significant in that a number of functional splice variants of

this receptor have been identified. No splice variants have

been reported for the first intron, which is in the putative i2

loop of the receptor. The second intron corresponds to the

carboxyl terminus of the 5-HT7 receptor, from which at least

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212 199

four functional splice variants have been documented. The

first clue that functional splice variants existed came from

the observation that the sequence identified by Lovenberg

and colleagues was truncated by 13 amino acids, compared

with other cloned sequences, and that the truncation

occurred at the second intron (Bard et al., 1993; Lovenberg

et al., 1993a; Shen et al., 1993). Subsequently, four splice

variants, termed 5-HT7a, 5-HT7b, 5-HT7c, and 5-HT7d, were

identified by several groups (Heidmann et al., 1997, 1998;

Jasper et al., 1997; Stam et al., 1997). The 5-HT7a, 5-HT7b,

and 5-HT7c receptor isoforms have been shown to be

expressed in rats, whereas the 5-HT7a, 5-HT7b, and 5-HT7d

receptor isoforms have been shown to be expressed in

humans (Heidmann et al., 1998). The three human splice

variants encode proteins of 448 (5-HT7a), 435 (5-HT7b), and

479 (5-HT7d) amino acids.

The functional consequences of the individual splice

variants are not clear, although all of them seem to be able

to stimulate AC (Jasper et al., 1997; Stam et al., 1997;

Heidmann et al., 1998). However, because the different

variants may have differing numbers of phosphorylation

sites, their regulation by kinases may be different. An

additional difference in the splice variants is the presence

or absence of PDZ motifs in the distal carboxyl terminus.

The 5-HT7a and 5-HT7d receptors do not have a PDZ

motif, whereas the 5-HT7b and 5-HT7c receptors have

group II and group I PDZ motifs, respectively (Hamblin

et al., 1998).

7.2. The 5-hydroxytryptamine7 receptor activates

adenylyl cyclase

Transfection studies have shown that the 5-HT7 receptor

stimulates AC in mammalian (Bard et al., 1993; Lovenberg

et al., 1993a; Plassat et al., 1993; Ruat et al., 1993b; Shen

et al., 1993; Tsou et al., 1994; Heidmann et al., 1997, 1998;

Stam et al., 1997) and insect Sf9 cells (Hirst et al., 1997).

Presumably, most of the effects of the 5-HT7 receptor are

mediated through coupling to Gs. In that regard, the i3 loop

of the 5-HT7 receptor has been shown to be critical in

coupling to G-proteins (Obosi et al., 1997). The 5-HT7a

receptor stimulates several forms of AC, including Gs-sen-

sitive AC5 and AC1 and AC8 (Baker et al., 1998). The ability

of the 5-HT7a receptor to activate those latter two isoforms of

AC raises the issue of whether this receptor can also couple to

G-proteins other than Gs. Baker et al. (1998) partially inves-

tigated the question by demonstrating that activation of the

receptor resulted in increases in intracellular Ca2 + , which

is consistent with the known Ca2 + /CaM sensitivity of AC1

and AC8 (Mons et al., 1998; Taussig & Zimmermann,

1998; Hurley, 1999). However, this effect was independent

of PKC, phosphoinositide turnover, and Gi-proteins (Baker

et al., 1998). These results are very intriguing in light of

studies showing that 5-HT can activate Ca2 + -stimulated

AC in the rat cerebral cortex and hippocampus (Mork &

Geisler, 1990).

7.3. The 5-hydroxytryptamine7 receptor activates the

extracellular signal-regulated kinase

Errico et al. (2001) recently reported that 5-HT7

receptors endogenously expressed by cultured rat hippo-

campal neurons stimulate the MAP kinase ERKs ERK1

and ERK2. Details of the signaling pathway used by the

receptor to activate ERK remain to be elucidated, but the

stimulation of ERK was not sensitive to pertussis toxin

and, therefore, is unlikely to involve Gi/o-proteins (Errico

et al., 2001).

7.4. The 5-hydroxytryptamine7 receptor

stimulates vasorelaxation

The 5-HT7 receptor stimulates vasorelaxation (Ullmer

et al., 1995; Terron, 1996; Villalon et al., 1997). The signaling

pathway involved in vasorelaxation by the 5-HT7 receptor is

not known, but it has been hypothesized to involve release of

NO (Morecroft & MacLean, 1998) or cAMP-induced inhibi-

tion of myosin light chain kinase in vascular smooth muscle

cells (Ullmer et al., 1995).

7.5. The 5-hydroxytryptamine7 receptor modulates

slow afterhyperpolarization

The 5-HT7 receptor has been implicated in the inhibition

of Ca2 + -spike-induced slow afterhyperpolarization ampli-

tude in CA3 hippocampal neurons, although the signaling

pathway used to modulate this conductance is not known

(Bacon & Beck, 2000). In contrast, Chapin and Andrade

(2000, 2001a, 2001b) have shown that the 5-HT7 receptor

can actually stimulate afterdepolarization. They performed

detailed studies of the regulation of afterhyperpolarization

by the 5-HT7 receptor in neurons from the anterodorsal

nucleus of the thalamus (Chapin & Andrade, 2001a), and

showed that the signal was mediated by cAMP and by the

hyperpolarization-activated nonselective cation current

(Chapin & Andrade, 2001b). Furthermore, the afterdepola-

rization was independent of PKA (Chapin & Andrade,

2001b) or alterations in cellular Ca2 + levels (Chapin &

Andrade, 2000), suggesting that novel signal transduction

pathways may mediate this effect.

8. Conclusion

The rich diversity of coupling of the 5-HT receptors to

distinct (and sometimes opposing) signaling pathways is

becoming more broadly appreciated as a mechanism of

‘fine-tuning’’ receptor response to the requirements of the

specific cells and tissues in which they are expressed. The

diversity of signaling is mediated by a large number of 5-HT

receptors (14 genomically encoded) and their variants (over

20 splice variants or variants resulting from mRNA editing),

and also by cell-specific and ligand-specific variables. The

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212200

diversity of signaling is so staggering that one naturally

might ask: ‘‘What is the purpose of this immense diversity of

signaling mechanisms coupled to 5-HT receptors?’’ Unfortu-

nately, there currently are no clear-cut answers to this

fundamental question. One of us reviewed this topic several

years ago, and many of the points that were raised in that

review remain pertinent today (Raymond, 1995). We will

speculate somewhat on this issue in the next two paragraphs.

The diversity in 5-HT receptor signals may provide

different cells with the capacity to determine the specific

effectors that are modulated by 5-HT receptors, based on

the intrinsic or variable needs of the cells or the organs in

which they reside. Thus, different cell types could respond

differently to the same stimulus, based on their anatomical

location, physiological state, and functions. These diffe-

rential responses in distinct cell types need not be haphaz-

ard; they would most likely be coordinated into some type

of ‘‘response network.’’ For example, glial cells and

neurons could intrinsically coordinate their responsiveness

to 5-HT (or other 5-HT receptor ligands), based on their

complements of 5-HT receptors, G-proteins, and effectors.

This makes sense, as these cell types should be intrins-

ically able to function in a coordinated manner. Their

coordinated responses should also be plastic, in that they

can be shaped by physiological alterations, such as hypoxia

or hormonal influences.

Both the coordination and plasticity of cellular res-

ponses to 5-HT receptors could be influenced by variables

such as (1) the types and numbers of receptors present in

each cell type; (2) the subcellular localizations of the

receptors and their targets into macrodomains or organ-

elles; (3) physical or functional compartmentalization of

the signaling components into microdomains; (4) differ-

ential and/or overlapping sensitivities to various ligands;

(5) differential rates of receptor synthesis, internalization,

degradation, and desensitization; (6) differential signaling

from different ligands acting through single receptor sub-

types due to ‘‘agonist-directed trafficking;’’ (7) the pres-

ence or absence of receptor reserve; (8) different levels of

spontaneous (non-agonist-mediated) receptor activity; (9)

co-stimulation with other receptor types; (10) non-G-pro-

tein-mediated signals; and (11) the presence or absence of

regulatory factors. These same differences could also be

applied to individual G-proteins and effectors. Moreover, all

of those variables need to be recognized as dynamic. Thus,

cells can modify their capacity for signaling from 5-HT

receptors to effectors in order to produce distinct (or even

opposing) responses. Thus, the specific response of a cell to

5-HT will be determined as the sum of many variables. That

is why signal transduction studies performed in transfected

cells ultimately must be validated contextually in cells,

organs, or animals in which the 5-HT receptors are endoge-

nously expressed.

This review has highlighted some fascinating new

insights into 5-HT receptor signaling that have resulted

from multiple lines of investigation over the last decade.

Substantial recent progress has been made in defining the

potential diversity of signaling from 5-HT receptors. The

discovery process has been accelerated by the cloning of

cDNA and genes for 5-HT receptors, by transfection

studies, by the development of new ligands, by the

creation of transgenic and receptor knockout animals,

and by the generation of receptor-specific tools, such as

antibodies and selective mRNA probes. Despite the accel-

erating pace of exciting discoveries in this field, one is

struck by a sense that the most exciting new discoveries

are yet to come.

Acknowledgements

The authors were supported by grants from the Depart-

ment of Veterans Affairs (Merit Awards to M.N.G. and

J.R.R. and a REAP Award to Y.V.M., A.G., J.R.R., G.C.,

and M.N.G.), the National Institutes of Health (DK52448

and DK54720 to J.R.R., DK053981 to T.W.G., and

DK02694 to Y.V.M.), and a laboratory endowment jointly

supported by the M.U.S.C. Division of Nephrology and

Dialysis Clinics, Incorporated (J.R.R.). A.G. and M.N.G.

were Associate Investigators of the Department of Veterans

Affairs during the composition of this manuscript.

References

Abdel-Baset, H., Bozovic, V., Szyf, M., & Albert, P. R. (1992). Conditional

transformation mediated via a pertussis toxin-sensitive receptor signal-

ling pathway. Mol Endocrinol 6, 730–740.

Abramowski, D., Rigo, M., Duc, D., Hoyer, D., & Staufenbiel, M.

(1995). Localization of the 5-hydroxytryptamine2C receptor protein

in human and rat brain using specific antisera. Neuropharmacology

34, 1635–1645.

Adayev, T., El-Sherif, Y., Barua, M., Penington, N. J., & Banerjee, P.

(1999). Agonist stimulation of the serotonin1A receptor causes suppres-

sion of anoxia-induced apoptosis via mitogen-activated protein kinase

in neuronal HN2-5 cells. J Neurochem 72, 1489–1496.

Adham, N., Romanienko, P., Hartig, P., Weinshank, R. L., & Branchek, T.

(1992). The rat 5-hydroxytryptamine1B receptor is the species homo-

logue of the human 5-hydroxytryptamine1D beta receptor. Mol Pharma-

col 41, 1–7.

Adham, N., Borden, L. A., Schechter, L. E., Gustafson, E. L., Cochran, T. L.,

Vaysse, P. J., Weinshank, R. L., & Branchek, T. A. (1993a). Cell-specific

coupling of the cloned human 5-HT1F receptor to multiple signal trans-

duction pathways. Naunyn Schmiedebergs Arch Pharmacol 348,

566–575.

Adham, N., Kao, H. T., Schecter, L. E., Bard, J., Olsen, M., Urquhart, D.,

Durkin, M., Hartig, P. R., Weinshank, R. L., & Branchek, T. A. (1993b).

Cloning of another human serotonin receptor (5-HT1F): a fifth 5-HT1

receptor subtype coupled to the inhibition of adenylate cyclase. Proc

Natl Acad Sci USA 90, 408–412.

Adham, N., Vaysse, P. J., Weinshank, R. L., & Branchek, T. A. (1994). The

cloned human 5-HT1E receptor couples to inhibition and activation of

adenylyl cyclase via two distinct pathways in transfected BS-C-1 cells.

Neuropharmacology 33, 403–410.

Ahlenius, S., & Larsson, K. (1989). Antagonism by pindolol, but not

betaxolol, of 8-OH-DPAT-induced facilitation of male rat sexual beha-

vior. J Neural Transm 77, 163–170.

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212 201

Aiyar, J., Grissmer, S., & Chandy, K. G. (1993). Full-length and truncated

Kv1.3 K + channels are modulated by 5-HT1c receptor activation and

independently by PKC. Am J Physiol 265, C1571–C1578.

Albert, P. R., Zhou, Q. Y., Van Tol, H. H., Bunzow, J. R., & Civelli, O.

(1990). Cloning, functional expression, and mRNA tissue distribution

of the rat 5-hydroxytryptamine1A receptor gene. J Biol Chem 265,

5825–5832.

Albert, P. R., Sajedi, N., Lemonde, S., & Ghahremani, M. H. (1999).

Constitutive Gi2-dependent activation of adenylyl cyclase type II by

the 5-HT1A receptor. Inhibition by anxiolytic partial agonists. J Biol

Chem 274, 35469–35474.

Albuquerque, F. C., Smith, E. H., & Kellum, J. M. (1998). 5-HT induces

cAMP production in crypt colonocytes at a 5-HT4 receptor. J Surg Res

77, 137–140.

Amlaiky, N., Ramboz, S., Boschert, U., Plassat, J. L., & Hen, R. (1992).

Isolation of a mouse ‘‘5HT1E-like’’ serotonin receptor expressed pre-

dominantly in hippocampus. J Biol Chem 267, 19761–19764.

Andrade, R., & Chaput, Y. (1991). 5-Hydroxytryptamine4-like receptors

mediate the slow excitatory response to serotonin in the rat hippo-

campus. J Pharmacol Exp Ther 257, 930–937.

Andrade, R., & Nicoll, R. A. (1987). Pharmacologically distinct actions of

serotonin on single pyramidal neurones of the rat hippocampus recorded

in vitro. J Physiol 394, 99–124.

Ansanay, H., Dumuis, A., Sebben, M., Bockaert, J., & Fagni, L. (1995).

cAMP-dependent, long-lasting inhibition of a K + current in mamma-

lian neurons. Proc Natl Acad Sci USA 92, 6635–6639.

Auch-Schwelk, W., Paetsch, I., Krackhardt, F., Grafe, M., Hetzer, R., &

Fleck, E. (2000). Modulation of contractions to ergonovine and meth-

ylergonovine by nitric oxide and thromboxane A2 in the human coro-

nary artery. J Cardiovasc Pharmacol 36, 631–639.

Aune, T. M., McGrath, K. M., Sarr, T., Bombara, M. P., & Kelley, K. A.

(1993). Expression of 5HT1a receptors on activated human T cells.

Regulation of cyclic AMP levels and T cell proliferation by 5-hydroxy-

tryptamine. J Immunol 151, 1175–1183.

Bach, T., Syversveen, T., Kvingedal, A. M., Krobert, K. A., Brattelid, T.,

Kaumann, A. J., & Levy, F. O. (2001). 5HT4(a) and 5-HT4(b) receptors

have nearly identical pharmacology and are both expressed in human

atrium and ventricle. Naunyn Schmiedebergs Arch Pharmacol 363,

146–160.

Backstrom, J. R., Chang, M. S., Chu, H., Niswender, C. M., & Sanders-

Bush, E. (1999). Agonist-directed signaling of serotonin 5-HT2C recep-

tors: differences between serotonin and lysergic acid diethylamide

(LSD). Neuropsychopharmacology 21, 77S–81S.

Backstrom, J. R., Price, R. D., Reasoner, D. T., & Sanders-Bush, E. (2000).

Deletion of the serotonin 5-HT2C receptor PDZ recognition motif pre-

vents receptor phosphorylation and delays resensitization of receptor

responses. J Biol Chem 275, 23620–23626.

Bacon, W. L., & Beck, S. G. (2000). 5-Hydroxytryptamine7 receptor acti-

vation decreases slow afterhyperpolarization amplitude in CA3 hippo-

campal pyramidal cells. J Pharmacol Exp Ther 294, 672–679.

Baker, L. P., Nielsen, M. D., Impey, S., Metcalf, M. A., Poser, S. W.,

Chan, G., Obrietan, K., Hamblin, M. W., & Storm, D. R. (1998).

Stimulation of type 1 and type 8 Ca2 + /calmodulin-sensitive adenylyl

cyclases by the Gs-coupled 5-hydroxytryptamine subtype 5-HT7A

receptor. J Biol Chem 273, 17469–17476.

Balcells-Olivero, M., Cousins, M. S., & Seiden, L. S. (1998). Holtzman and

Harlan Sprague-Dawley rats: differences in DRL 72-sec performance

and 8-hydroxy-di-propylamino tetralin-induced hypothermia. J Phar-

macol Exp Ther 286, 742–752.

Bard, J. A., Zgombick, J., Adham, N., Vaysse, P., Branchek, T. A., &

Weinshank, R. L. (1993). Cloning of a novel human serotonin receptor

(5-HT7) positively linked to adenylate cyclase. J Biol Chem 268,

23422–23426.

Barnes, N. M., & Sharp, T. (1999). A review of central 5-HT receptors and

their function. Neuropharmacology 38, 1083–1152.

Barr, A. J., Brass, L. F., & Manning, D. R. (1997). Reconstitution of

receptors and GTP-binding regulatory proteins (G proteins) in Sf9 cells.

A direct evaluation of selectivity in receptor-G protein coupling. J Biol

Chem 272, 2223–2229.

Bartrup, J. T., & Newberry, N. R. (1994). 5-HT2A receptor-mediated out-

ward current in C6 glioma cells is mimicked by intracellular IP3 release.

Neuroreport 5, 1245–1248.

Bayliss, D. A., Umemiya, M., & Berger, A. J. (1995). Inhibition of N- and

P-type calcium currents and the after-hyperpolarization in rat moto-

neurones by serotonin. J Physiol 485, 635–647.

Becamel, C., Figge, A., Poliak, S., Dumuis, A., Peles, E., Bockaert, J.,

Lubbert, H., & Ullmer, C. (2001). Interaction of serotonin 5-HT2C

receptors with PDZ10 of the multi PDZ protein MUPP1. J Biol Chem

275, 12974–12982.

Becker, B. N., Gettys, T. W., Middleton, J. P., Olsen, C. L., Albers, F. J.,

Lee, S. L., Fanburg, B. L., & Raymond, J. R. (1992). 8-Hydroxy-

2-(di-n-propylamino)tetralin-responsive 5-hydroxytryptamine4-like re-

ceptor expressed in bovine pulmonary artery smooth muscle cells.

Mol Pharmacol 42, 817–825.

Berg, K. A., Clarke, W. P., Chen, Y., Ebersole, B. J., McKay, R. D., &

Maayani, S. (1994a). 5-Hydroxytryptamine type 2A receptors regulate

cyclic AMP accumulation in a neuronal cell line by protein kinase

C-dependent and calcium/calmodulin-dependent mechanisms. Mol

Pharmacol 45, 826–836.

Berg, K. A., Clarke, W. P., Sailstad, C., Saltzman, A., & Maayani, S.

(1994b). Signal transduction differences between 5-hydroxytryptamine

type 2A and type 2C receptor systems. Mol Pharmacol 46, 477–484.

Berg, K. A., Maayani, S., & Clarke, W. P. (1996). 5-Hydroxytryptamine2C

receptor activation inhibits 5-hydroxytryptamine1B-like receptor func-

tion via arachidonic acid metabolism. Mol Pharmacol 50, 1017–1023.

Berg, K. A., Maayani, S., Goldfarb, J., & Clarke, W. P. (1998). Pleiotropic

behavior of 5-HT2A and 5-HT2C receptor agonists. Ann N Y Acad Sci

861, 104–110.

Berridge, M. J., Bootman, M. D., & Lipp, P. (1998). Calcium— a life and

death signal. Nature 395, 645–648.

Bertin, B., Freissmuth, M., Breyer, R. M., Schutz, W., Strosberg, A. D., &

Marullo, S. (1992). Functional expression of the human serotonin

5-HT1A receptor in Escherichia coli. Ligand binding properties and in-

teraction with recombinant G protein alpha-subunits. J Biol Chem 267,

8200–8206.

Bhatnagar, A., Willins, D. L., Gray, J. A., Woods, J., Benovic, J. L., &

Roth, B. L. (2000). The dynamin-dependent, arrestin-independent in-

ternalization of 5-HT2A serotonin receptors reveals differential sorting

of arrestins and 5-HT2A receptors during endocytosis. J Biol Chem 276,

8269–8277.

Bianchi, C., Siniscalchi, A., & Beani, L. (1990). 5-HT1A agonists increase

and 5-HT3 agonists decrease acetylcholine efflux from the cerebral

cortex of freely-moving guinea-pigs. Br J Pharmacol 101, 448–452.

Blier, P., Bergeron, R., & de Montigny, C. (1997). Selective activation of

postsynaptic 5-HT1A receptors induces rapid antidepressant response.

Neuropsychopharmacology 16, 333–338.

Blondel, O., Vandecasteele, G., Gastineau, M., Leclerc, S., Dahmoune, Y.,

Langlois, M., & Fischmeister, R. (1997). Molecular and functional

characterization of a 5-HT4 receptor cloned from human atrium. FEBS

Lett 412, 465–474.

Blondel, O., Gastineau, M., Dahmoune, Y., Langlois, M., & Fischmeister,

R. (1998). Cloning, expression, and pharmacology of four human

5-hydroxytryptamine4 receptor isoforms produced by alternative splicing

in the carboxyl terminus. J Neurochem 70, 2252–2261.

Bockaert, J., Sebben, M., & Dumuis, A. (1990). Pharmacological cha-

racterization of 5-hydroxytryptamine4 (5-HT4) receptors positively

coupled to adenylate cyclase in adult guinea pig hippocampal mem-

branes: effect of substituted benzamide derivatives. Mol Pharmacol

37, 408–411.

Bockaert, J., Ansanay, H., Letty, S., Marchetti-Gauthier, E., Roman, F.,

Rondouin, G., Fagni, L., Soumireu-Mourat, B., & Dumuis, A.

(1998a). 5-HT4 receptors: long-term blockade of K + channels and

effects on olfactory memory. C R Acad Sci III 321, 217–221.

Bockaert, J., Claeysen, S., Sebben, M., & Dumuis, A. (1998b). 5-HT4

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212202

receptors: gene, transduction and effects on olfactory memory. Ann

N Y Acad Sci 861, 1–15.

Boddeke, H. W., Fargin, A., Raymond, J. R., Schoeffter, P., & Hoyer, D.

(1992). Agonist/antagonist interactions with cloned human 5-HT1A

receptors: variations in intrinsic activity studied in transfected HeLa

cells. Naunyn Schmiedebergs Arch Pharmacol 345, 257–263.

Boess, F. G., & Martin, I. L. (1994). Molecular biology of 5-HT receptors.

Neuropharmacology 33, 275–317.

Bonhaus, D. W., Bach, C., DeSouza, A., Salazar, F. H., Matsuoka, B. D.,

Zuppan, P., Chan, H. W., & Eglen, R. M. (1995). The pharmacology

and distribution of human 5-hydroxytryptamine2B (5-HT2B) receptor

gene products: comparison with 5-HT2A and 5-HT2C receptors. Br J

Pharmacol 115, 622–628.

Bouhelal, R., Smounya, L., & Bockaert, J. (1988). 5-HT1B receptors are

negatively coupled with adenylate cyclase in rat substantia nigra. Eur J

Pharmacol 151, 189–196.

Bourson, A., Boess, F. G., Bos, M., & Sleight, A. J. (1998). Involvement of

5-HT6 receptors in nigro-striatal function in rodents. Br J Pharmacol

125, 1562–1566.

Branchek, T. A., & Blackburn, T. P. (2000). 5-ht6 receptors as emerging

targets for drug discovery. Annu Rev Pharmacol Toxicol 40, 319–334.

Briddon, S. J., Leslie, R. A., & Elliott, J. M. (1998). Comparative

desensitization of the human 5-HT2A and 5-HT2C receptors expressed

in the human neuroblastoma cell line SH-SY5Y. Br J Pharmacol

125, 727–734.

Bruinvels, A. T., Palacios, J. M., & Hoyer, D. (1993). Autoradiographic

characterisation and localisation of 5-HT1D compared to 5-HT1B binding

sites in rat brain.Naunyn Schmiedebergs Arch Pharmacol 347, 569–582.

Bruinvels, A. T., Landwehrmeyer, B., Gustafson, E. L., Durkin, M.M., Men-

god, G., Branchek, T. A., Hoyer, D., & Palacios, J. M. (1994a). Local-

ization of 5-HT1B, 5-HT1D alpha, 5-HT1E and 5-HT1F receptor messenger

RNA in rodent and primate brain. Neuropharmacology 33, 367–386.

Bruinvels, A. T., Landwehrmeyer, B., Probst, A., Palacios, J. M., & Hoyer,

D. (1994b). A comparative autoradiographic study of 5-HT1D binding

sites in human and guinea-pig brain using different radioligands. Brain

Res Mol Brain Res 21, 19–29.

Budhoo, M. R., Harris, R. P., & Kellum, J. M. (1996a). 5-Hydroxytrypt-

amine-induced Cl- transport is mediated by 5-HT3 and 5-HT4 receptors

in the rat distal colon. Eur J Pharmacol 298, 137–144.

Budhoo, M. R., Harris, R. P., & Kellum, J. M. (1996b). The role of the

5-HT4 receptor in Cl- secretion in human jejunal mucosa. Eur J

Pharmacol 314, 109–114.

Buhlen, M., Fink, K., Boing, C., & Gothert, M. (1996). Evidence for

presynaptic location of inhibitory 5-HT1D beta-like autoreceptors in

the guinea-pig brain cortex. Naunyn Schmiedebergs Arch Pharmacol

353, 281–289.

Burns, C. M., Chu, H., Rueter, S. M., Hutchinson, L. K., Canton, H.,

Sanders-Bush, E., & Emeson, R. B. (1997). Regulation of serotonin-

2C receptor G-protein coupling by RNA editing. Nature 387, 303–308.

Butkerait, P., Zheng, Y., Hallak, H., Graham, T. E., Miller, H. A., Burris,

K. D., Molinoff, P. B., & Manning, D. R. (1995). Expression of the

human 5-hydroxytryptamine1A receptor in Sf9 cells. Reconstitution of a

coupled phenotype by co-expression of mammalian G protein subunits.

J Biol Chem 270, 18691–18699.

Cadogan, A. K., Kendall, D. A., & Marsden, C. A. (1994). Serotonin

5-HT1A receptor activation increases cyclic AMP formation in the rat

hippocampus in vivo. J Neurochem 62, 1816–1821.

Canton, H., Emeson, R. B., Barker, E. L., Backstrom, J. R., Lu, J. T., Chang,

M. S., & Sanders-Bush, E. (1996). Identification, molecular cloning, and

distribution of a short variant of the 5-hydroxytryptamine2C receptor

produced by alternative splicing. Mol Pharmacol 50, 799–807.

Carmena, M. J., Camacho, A., Solano, R. M., Montalvo, L., Garcia-

Lopez, E., Arias, A., & Prieto, J. C. (1998). 5-Hydroxytryptamine1A

receptor-mediated effects on adenylate cyclase and nitric oxide synthase

activities in rat ventral prostate. Cell Signal 10, 583–587.

Carson, M. J., Thomas, E. A., Danielson, P. E., & Sutcliffe, J. G. (1996).

The serotonin 5-HT5A receptor is expressed predominantly by astro-

cytes in which it inhibits cAMP accumulation: a mechanism for neuro-

nal suppression of reactive astrocytes. Glia 17, 317–326.

Cattaneo, M. G., Palazzi, E., Bondiolotti, G., & Vicentini, L. M. (1994).

5-HT1D receptor type is involved in stimulation of cell proliferation

by serotonin in human small cell lung carcinoma. Eur J Pharmacol

268, 425–430.

Cattaneo, M. G., Fesce, R., & Vicentini, L. M. (1995). Mitogenic effect of

serotonin in human small cell lung carcinoma cells via both 5-HT1A and

5-HT1D receptors. Eur J Pharmacol 291, 209–211.

Chanda, P. K., Minchin, M. C., Davis, A. R., Greenberg, L., Reilly, Y.,

McGregor, W. H., Bhat, R., Lubeck, M. D., Mizutani, S., & Hung, P. P.

(1993). Identification of residues important for ligand binding to the

human 5-hydroxytryptamine1A serotonin receptor. Mol Pharmacol 43,

516–520.

Chang, M., Zhang, L., Tam, J. P., & Sanders-Bush, E. (2000). Dissecting G

protein-coupled receptor signaling pathways with membrane-permeable

blocking peptides. Endogenous 5-HT2C receptors in choroid plexus

epithelial cells. J Biol Chem 275, 7021–7029.

Chapin, E. M., & Andrade, R. (2000). Calcium-independent afterdepolari-

zation regulated by serotonin in anterior thalamus. J Neurophysiol 83,

3173–3176.

Chapin, E. M., & Andrade, R. (2001a). A 5-HT7 receptor-mediated depo-

larization in the anterodorsal thalamus. I. Pharmacological character-

ization. J Pharmacol Exp Ther 297, 395–402.

Chapin, E. M., & Andrade, R. (2001b). A 5-HT7 receptor-mediated depola-

rization in the anterodorsal thalamus. II. Involvement of the hyperpola-

rization-activated current I(h). J Pharmacol Exp Ther 297, 403–409.

Chen, H., Li, H., & Chuang, D. M. (1995). Role of second messengers in

agonist up-regulation of 5-HT2A receptor binding sites in cerebellar

granule neurons: involvement of calcium influx and a calmodulin-

dependent pathway. J Pharmacol Exp Ther 275, 674–680.

Chen, N. H., & Reith, M. E. (1995). Monoamine interactions measured by

microdialysis in the ventral tegmental area of rats treated systemically

with (+/-)-8-hydroxy-2-(di-n-propylamino)tetralin. J Neurochem 64,

1585–1597.

Chen, Y., & Penington, N. J. (1997). QEHA27, a peptide that binds to

G-protein beta gamma-subunits, reduces the inhibitory effect of 5-HTon

the Ca2 + current of rat dorsal raphe neurons. Neurosci Lett 224, 87–90.

Chen, Y., Baez, M., & Yu, L. (1994). Functional coupling of the 5-HT2C

serotonin receptor to G proteins in Xenopus oocytes. Neurosci Lett 179,

100–102.

Choi, D. S., Ward, S. J., Messaddeq, N., Launay, J. M., & Maroteaux, L.

(1997). 5-HT2B receptor-mediated serotonin morphogenetic functions in

mouse cranial neural crest and myocardiac cells. Development 124,

1745–1755.

Ciaranello, R. D., Tan, G. L., & Dean, R. (1990). G-protein-linked seroto-

nin receptors in mouse kidney exhibit identical properties to 5-HT1b

receptors in brain. J Pharmacol Exp Ther 252, 1347–1354.

Claeysen, S., Sebben, M., Journot, L., Bockaert, J., & Dumuis, A. (1996).

Cloning, expression and pharmacology of the mouse 5-HT4L receptor.

FEBS Lett 398, 19–25.

Claeysen, S., Faye, P., Sebben, M., Lemaire, S., Bockaert, J., Dumuis, A.,

& Taviaux, S. (1997). Assignment of 5-hydroxytryptamine receptor

(HTR4) to human chromosome 5 bands q31–>q33 by in situ hybrid-

ization. Cytogenet Cell Genet 78, 133–134.

Claeysen, S., Faye, P., Sebben, M., Taviaux, S., Bockaert, J., & Dumuis, A.

(1998). 5-HT4 receptors: cloning and expression of new splice variants.

Ann N Y Acad Sci 861, 49–56.

Claeysen, S., Sebben, M., Becamel, C., Bockaert, J., & Dumuis, A. (1999).

Novel brain-specific 5-HT4 receptor splice variants show marked con-

stitutive activity: role of the C-terminal intracellular domain. Mol Phar-

macol 55, 910–920.

Clarke, W. P., Yocca, F. D., & Maayani, S. (1996). Lack of 5-hydroxytryp-

tamine1A-mediated inhibition of adenylyl cyclase in dorsal raphe of

male and female rats. J Pharmacol Exp Ther 277, 1259–1266.

Claustre, Y., Benavides, J., & Scatton, B. (1991). Potential mechanisms

involved in the negative coupling between serotonin 5-HT1A receptors

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212 203

and carbachol-stimulated phosphoinositide turnover in the rat hippo-

campus. J Neurochem 56, 1276–1285.

Clawges, H. M., Depree, K. M., Parker, E. M., & Graber, S. G. (1997).

Human 5-HT1 receptor subtypes exhibit distinct G protein coupling be-

haviors in membranes from Sf9 cells. Biochemistry 36, 12930–12938.

Colino, A., & Halliwell, J. V. (1987). Differential modulation of three

separate K-conductances in hippocampal CA1 neurons by serotonin.

Nature 328, 73–77.

Conn, P. J., & Sanders-Bush, E. (1984). Selective 5HT-2 antagonists inhibit

serotonin stimulated phosphatidylinositol metabolism in cerebral cortex.

Neuropharmacology 23, 993–996.

Conn, P. J., & Sanders-Bush, E. (1985). Serotonin-stimulated phosphoino-

sitide turnover: mediation by the S2 binding site in rat cerebral cortex

but not in subcortical regions. J Pharmacol Exp Ther 234, 195–203.

Consolo, S., Arnaboldi, S., Giorgi, S., Russi, G., & Ladinsky, H. (1994).

5-HT4 receptor stimulation facilitates acetylcholine release in rat frontal

cortex. Neuroreport 5, 1230–1232.

Cook, E. H., Fletcher, K. E., Wainwright, M., Marks, N., Yan, S. Y., &

Leventhal, B. L. (1994). Primary structure of the human platelet sero-

tonin 5-HT2A receptor: identify with frontal cortex serotonin 5-HT2A

receptor. J Neurochem 63, 465–469.

Cowen, D. S., Sowers, R. S., & Manning, D. R. (1996). Activation of a

mitogen-activated protein kinase (ERK2) by the 5-hydroxytryptamine1Areceptor is sensitive not only to inhibitors of phosphatidylinositol

3-kinase, but to an inhibitor of phosphatidylcholine hydrolysis. J Biol

Chem 271, 22297–22300.

Cowen, D. S., Molinoff, P. B., & Manning, D. R. (1997). 5-Hydroxytryp-

tamine1A receptor-mediated increases in receptor expression and acti-

vation of nuclear factor-kappaB in transfected Chinese hamster ovary

cells. Mol Pharmacol 52, 221–226.

Cowen, P. J., Clifford, E. M., Walsh, A. E., Williams, C., & Fairburn,

C. G. (1996). Moderate dieting causes 5-HT2C receptor supersensitivity.

Psychol Med 26, 1155–1159.

Cox, D. A., & Cohen, M. L. (1995). Is nitric oxide involved in 5-HT2B

receptor-mediated contraction in the rat stomach fundus? Life Sci 56,

L333–L338.

Cox, D. A., Watts, S. W., & Cohen, M. L. (1996). Neomycin selectively

inhibits 5-hydroxytryptamine-induced contraction in the guinea pig

trachea. J Pharmacol Exp Ther 277, 954–959.

Cox, D. A., Blase, D. K., & Cohen, M. L. (1999). Bradykinin and phorbol

ester but not 5-HT2B receptor activation stimulate phospholipase D

activity in the rat stomach fundus. Prog Neuropsychopharmacol Biol

Psychiatry 23, 697–704.

Dascal, N., Lim, N. F., Schreibmayer, W., Wang,W., Davidson, N., & Lester,

H. A. (1993). Expression of an atrial G-protein-activated potassium chan-

nel in Xenopus oocytes. Proc Natl Acad Sci USA 90, 6596–6600.

Davidson, C., & Stamford, J. A. (1996). Serotonin efflux in the rat ventral

lateral geniculate nucleus assessed by fast cyclic voltammetry is modu-

lated by 5-HT1B and 5-HT1D autoreceptors. Neuropharmacology 35,

1627–1634.

De Vivo, M., & Maayani, S. (1986). Characterization of the 5-hydroxytryp-

tamine1a receptor-mediated inhibition of forskolin-stimulated adenylate

cyclase activity in guinea pig and rat hippocampal membranes. J Phar-

macol Exp Ther 238, 248–253.

Della Rocca, G. J., Mukhin, Y. V., Garnovskaya, M. N., Daaka, Y., Clark,

G. J., Luttrell, L.M., Lefkowitz,R. J.,&Raymond, J.R. (1999). Serotonin

5-HT1A receptor-mediated Erk activation requires calcium/calmodulin-

dependent receptor endocytosis. J Biol Chem 274, 4749–4753.

Dickenson, J. M., & Hill, S. J. (1995). Coupling of an endogenous 5-HT1B-

like receptor to increases in intracellular calcium through a pertussis

toxin-sensitive mechanism in CHO-K1 cells. Br J Pharmacol 116,

2889–2896.

Dickenson, J. M., & Hill, S. J. (1998). Human 5-HT1B receptor stimulated

inositol phospholipid hydrolysis in CHO cells: synergy with Gq-coupled

receptors. Eur J Pharmacol 348, 279–285.

DiMagno, L., Dascal, N., Davidson, N., Lester, H. A., & Schreibmayer, W.

(1996). Serotonin and protein kinase C modulation of a rat brain in-

wardly rectifying K + channel expressed in xenopus oocytes. Pflugers

Arch 431, 335–340.

Dissmann, E., Wischmeyer, E., Spauschus, A., Pfeil, D. V., Karschin, C., &

Karschin, A. (1996). Functional expression and cellular mRNA local-

ization of a G protein-activated K + inward rectifier isolated from rat

brain. Biochem Biophys Res Commun 223, 474–479.

Done, C. J., & Sharp, T. (1994). Biochemical evidence for the regulation of

central noradrenergic activity by 5-HT1A and 5-HT2 receptors: micro-

dialysis studies in the awake and anaesthetized rat. Neuropharmacology

33, 411–421.

Dooley, A. E., Pappas, I. S., & Parnavelas, J. G. (1997). Serotonin promotes

the survival of cortical glutamatergic neurons in vitro. Exp Neurol 148,

205–214.

Doupnik, C. A., Dessauer, C.W., Slepak, V. Z., Gilman, A. G., Davidson, N.,

& Lester, H. A. (1996). Time resolved kinetics of direct G beta 1 gamma 2

interactions with the carboxyl terminus of Kir3.4 inward rectifier K +

channel subunits. Neuropharmacology 35, 923–931.

Doupnik, C. A., Davidson, N., Lester, H. A., & Kofuji, P. (1997). RGS

proteins reconstitute the rapid gating kinetics of Gbg-activated inwardlyrectifying K+ channels. Proc Natl Acad Sci USA 94, 10461–10466.

Dourish, C. T., Hutson, P. H., & Curzon, G. (1985). Characteristics of

feeding induced by the serotonin agonist 8-hydroxy-2-(di-n-propylami-

no) tetralin (8-OH-DPAT). Brain Res Bull 15, 377–384.

Dumuis, A., Bouhelal, R., Sebben, M., Cory, R., & Bockaert, J. (1988). A

nonclassical 5-hydroxytryptamine receptor positively coupled with

adenylate cyclase in the central nervous system. Mol Pharmacol 34,

880–887.

Duxon, M. S., Flanigan, T. P., Reavley, A. C., Baxter, G. S., Blackburn, T. P.,

& Fone, K. C. (1997). Evidence for expression of the 5-hydroxytrypt-

amine-2B receptor protein in the rat central nervous system. Neuro-

science 76, 323–329.

Eberle-Wang, K., Braun, B. T., & Simansky, K. J. (1994). Serotonin contracts

the isolated rat pylorus via a 5-HT2-like receptor. Am J Physiol 266,

R284–R291.

Ebersole, B. J., Diglio, C. A., Kaufman, D. W., & Berg, K. A. (1993).

5-Hydroxytryptamine1-like receptors linked to increases in intracellu-

lar calcium concentration and inhibition of cyclic AMP accumulation

in cultured vascular smooth muscle cells derived from bovine basilar

artery. J Pharmacol Exp Ther 266, 692–699.

Edagawa, Y., Saito, H., & Abe, K. (1998). 5-HT1A receptor-mediated in-

hibition of long-term potentiation in rat visual cortex. Eur J Pharmacol

349, 221–224.

Eglen, R.M., Alvarez, R., Carter, D., Leung, E., Jakeman, L., To, Z., & Tsou,

A. P. (1997). Cloned and native guinea pig 5-ht7 receptors. Character-

ization using an integrated approach. Ann N Y Acad Sci 812, 216–217.

Ellwood, A. J., & Curtis, M. J. (1997). Involvement of 5-HT1B/1D and

5-HT2A receptors in 5-HT-induced contraction of endothelium-denuded

rabbit epicardial coronary arteries. Br J Pharmacol 122, 875–884.

el Mansari, M., & Blier, P. (1996). Functional characterization of 5-HT1D

autoreceptors on the modulation of 5-HT release in guinea-pig mesen-

cephalic raphe, hippocampus and frontal cortex. Br J Pharmacol 118,

681–689.

Erdmann, J., Nothen, M. M., Shimron-Abarbanell, D., Rietschel, M., Albus,

M., Borrmann, M., Maier, W., Franzek, E., Korner, J., Weigelt, B., Fim-

mers, R., & Propping, P. (1996). The human serotonin 7 (5-HT7) receptor

gene: genomic organization and systematic mutation screening in schiz-

ophrenia and bipolar affective disorder. Mol Psychiatry 1, 392–397.

Eriksson, K. S., Stevens, D. R., & Haas, H. L. (2001). Serotonin excites

tuberomammillary neurons by activation of Na + /Ca2 + -exchange.

Neuropharmacology 40, 345–351.

Erlander, M. G., Lovenberg, T. W., Baron, B. M., de Lecea, L., Danielson,

P. E., Racke, M., Slone, A. L., Siegel, B. W., Foye, P. E., Cannon, K.,

Burns, J. E., & Sutcliffe, J. G. (1993). Two members of a distinct

subfamily of 5-hydroxytryptamine receptors differentially expressed

in rat brain. Proc Natl Acad Sci USA 90, 3452–3456.

Errico, M., Crozier, R. A., Plummer, M. R., & Cowen, D. S. (2001). 5-HT7

receptors activate the mitogen activated protein kinase extracellular

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212204

signal related kinase in cultured rat hippocampal neurons. Neuroscience

102, 361–367.

Fagni, L., Dumuis, A., Sebben, M., & Bockaert, J. (1992). The 5-HT4

receptor subtype inhibits K + current in colliculi neurones via activa-

tion of a cyclic AMP-dependent protein kinase. Br J Pharmacol 105,

973–979.

Fargin, A., Raymond, J. R., Lohse, M. J., Kobilka, B. K., Caron, M. G., &

Lefkowitz, R. J. (1988). The genomic clone G-21 which resembles a

beta-adrenergic receptor sequence encodes the 5-HT1A receptor. Nature

335, 358–360.

Fargin, A., Raymond, J. R., Regan, J. W., Cotecchia, S., Lefkowitz, R. J.,

& Caron, M. G. (1989). Effector coupling mechanisms of the cloned

5-HT1A receptor. J Biol Chem 264, 14848–14852.

Fargin, A., Yamamoto, K., Cotecchia, S., Goldsmith, P. K., Spiegel, A. M.,

Lapetina, E. G., Caron, M. G., & Lefkowitz, R. J. (1991). Dual cou-

pling of the cloned 5-HT1A receptor to both adenylyl cyclase and

phospholipase C is mediated via the same Gi protein. Cell Signal 3,

547–557.

Fayolle, C., Fillion, M. P., Barone, P., Oudar, P., Rousselle, J. C., & Fillion,

G. (1988). 5-Hydroxytryptamine stimulates two distinct adenylate cy-

clase activities in rat brain: high-affinity activation is related to a 5-HT1

subtype different from 5-HT1A, 5-HT1B, and 5-HT1C. Fundam Clin

Pharmacol 2, 195–214.

Ferriere, F., Khan, N. A., Troutaud, D., & Deschaux, P. (1996). Serotonin

modulation of lymphocyte proliferation via 5-HT1A receptors in rain-

bow trout (Oncorhynchus mykiss). Dev Comp Immunol 20, 273–283.

Fitzgerald, L. W., Iyer, G., Conklin, D. S., Krause, C. M., Marshall, A.,

Patterson, J. P., Tran, D. P., Jonak, G. J., & Hartig, P. R. (1999).

Messenger RNA editing of the human serotonin 5-HT2C receptor. Neu-

ropsychopharmacology 21, 82S–90S.

Florian, J. A., & Watts, S. W. (1998). Integration of mitogen-activated

protein kinase kinase activation in vascular 5-hydroxytryptamine2Areceptor signal transduction. J Pharmacol Exp Ther 284, 346–355.

Foguet, M., Hoyer, D., Pardo, L. A., Parekh, A., Kluxen, F. W., Kalkman,

H. O., Stuhmer, W., & Lubbert, H. (1992a). Cloning and functional

characterization of the rat stomach fundus serotonin receptor. EMBO J

11, 3481–3487.

Foguet, M., Nguyen, H., Le, H., & Lubbert, H. (1992b). Structure of the

mouse 5-HT1C, 5-HT2 and stomach fundus serotonin receptor genes.

Neuroreport 3, 345–348.

Ford, A. P., Baxter, G. S., Eglen, R. M., & Clarke, D. E. (1992). 5-Hydrox-

ytryptamine stimulates cyclic AMP formation in the tunica muscularis

mucosae of the rat oesophagus via 5-HT4 receptors. Eur J Pharmacol

211, 117–120.

Francken, B. J. B., Jurzak, M., Luyten, W. M. H., & Leysen, J. E.

(1998). h5-HT5A receptor in stably transfected HEK293 cells couples

to G-proteins and receptor activation inhibits adenylate cyclase. Se-

rotonin Club. In P. R. Saxena, R. A. Green, & M. D. Ferrari (Eds.),

4th IUPHAR Satellite Meeting on Serotonin, Rotterdam, Netherlands,

July 23-25 (p. 188). Rotterdam: Erasmus University Press.

Fujiwara, Y., Nelson, D. L., Kashihara, K., Varga, E., Roeske, W. R., &

Yamamura, H. I. (1990). The cloning and sequence analysis of the rat

serotonin-1A receptor gene. Life Sci 47, L127–L132.

Fuller, R. W. (1996). Serotonin receptors involved in regulation of pituitary-

adrenocortical function in rats. Behav Brain Res 73, 215–219.

Gaddum, J. H., & Picarelli, Z. P. (1957). Two kinds of tryptamine receptor.

Br J Pharmacol 12, 323–328.

Garcia, M. C., & Kim, H. Y. (1997). Mobilization of arachidonate and

docosahexaenoate by stimulation of the 5-HT2A receptor in rat C6

glioma cells. Brain Res 768, 43–48.

Garnovskaya, M. N., Nebigil, C. G., Arthur, J. M., Spurney, R. F., &

Raymond, J. R. (1995). 5-Hydroxytryptamine2A receptors expressed

in rat renal mesangial cells inhibit cyclic AMP accumulation. Mol

Pharmacol 48, 230–237.

Garnovskaya, M. N., van Biesen, T., Hawe, B., Casanas Ramos, S.,

Lef kowitz, R. J., & Raymond, J. R. (1996). Ras-dependent activa-

tion of fibroblast mitogen-activated protein kinase by 5-HT1A recep-

tor via a G protein bg-subunit-initiated pathway. Biochemistry 35,

13716–13722.

Garnovskaya, M. N., Gettys, T. W., van Biesen, T., Prpic, V., Chuprun, J. K.,

& Raymond, J. R. (1997). 5-HT1A receptor activates Na + /H + ex-

change in CHO-K1 cells through Gia2 and Gia3. J Biol Chem 272,

7770–7776.

Garnovskaya, M. N., Mukhin, Y., & Raymond, J. R. (1998). Rapid activa-

tion of sodium-proton exchange and extracellular signal-regulated pro-

tein kinase in fibroblasts by G protein-coupled 5-HT1A receptor

involves distinct signalling cascades. Biochem J 330, 489–495.

Gartside, S. E., Cowen, P. J., & Hjorth, S. (1990). Effects of MDL 73005EF

on central pre- and postsynaptic 5-HT1A receptor function in the rat in

vivo. Eur J Pharmacol 191, 391–400.

Ge, J., &Barnes, N.M. (1996). 5-HT4 receptor-mediatedmodulation of 5-HT

release in the rat hippocampus in vivo. Br J Pharmacol 117, 1475–1480.

Gelernter, J., Rao, P. A., Pauls, D. L., Hamblin, M. W., Sibley, D. R., &

Kidd, K. K. (1995). Assignment of the 5HT7 receptor gene (HTR7) to

chromosome 10q and exclusion of genetic linkage with Tourette syn-

drome. Genomics 26, 207–209.

Gerald, C., Adham, N., Kao, H. T., Olsen, M. A., Laz, T. M., Schechter,

L. E., Bard, J. A., Vaysse, P. J., Hartig, P. R., & Branchek, T. A.

(1995). The 5-HT4 receptor: molecular cloning and pharmacological

characterization of two splice variants. EMBO J 14, 2806–2815.

Gerard, C., el Mestikawy, S., Lebrand, C., Adrien, J., Ruat, M., Traiffort, E.,

Hamon, M., & Martres, M. P. (1996). Quantitative RT-PCR distribution

of serotonin 5-HT6 receptor mRNA in the central nervous system of

control or 5,7-dihydroxytryptamine-treated rats. Synapse 23, 164–173.

Gerard, C., Martres, M. P., Lefevre, K., Miquel, M. C., Verge, D., Lanfu-

mey, L., Doucet, E., Hamon, M., & el Mestikawy, S. (1997). Immuno-

localization of serotonin 5-HT6 receptor-like material in the rat central

nervous system. Brain Res 746, 207–219.

Gershon, M. D. (1999). Review article: roles played by 5-hydroxytrypt-

amine in the physiology of the bowel. Aliment Pharmacol Ther 13

(suppl. 2), 15–30.

Gettys, T. W., Fields, T. A., & Raymond, J. R. (1994). Selective activation

of inhibitory G-protein alpha-subunits by partial agonists of the human

5-HT1A receptor. Biochemistry 33, 4283–4290.

Gilbert, F., & Dourish, C. T. (1987). Effects of the novel anxiolytics gepir-

one, buspirone and ipsapirone on free feeding and on feeding induced

by 8-OH-DPAT. Psychopharmacology 93, 349–352.

Gilbert, F., Brazell, C., Tricklebank, M. D., & Stahl, S. M. (1988a). Acti-

vation of the 5-HT1A receptor subtype increases rat plasma ACTH

concentration. Eur J Pharmacol 147, 431–439.

Gilbert, F., Dourish, C. T., Brazell, C., McClue, S., & Stahl, S. M. (1988b).

Relationship of increased food intake and plasmaACTH levels to 5-HT1A

receptor activation in rats. Psychoneuroendocrinology 13, 471–478.

Giles, H., Lansdell, S. J., Bolofo, M. L., Wilson, H. L., & Martin, G. R.

(1996). Characterization of a 5-HT1B receptor on CHO cells: functional

responses in the absence of radioligand binding. Br J Pharmacol 117,

1119–1126.

Gobert, A., & Millan, M. J. (1999). Serotonin (5-HT)2A receptor activation

enhances dialysate levels of dopamine and noradrenaline, but not 5-HT,

in the frontal cortex of freely-moving rats. Neuropharmacology 38,

315–317.

Goodwin, G. M., De Souza, R. J., & Green, A. R. (1985). The pharmacol-

ogy of the hypothermic response in mice to 8-hydroxy-2-(di-n-propy-

lamino)tetralin (8-OH-DPAT). A model of presynaptic 5-HT1 function.

Neuropharmacology 24, 1187–1194.

Goodwin, G. M., De Souza, R. J., Green, A. R., & Heal, D. J. (1987). The

pharmacology of the behavioural and hypothermic responses of rats to

8-hydroxy-2-(di-n-propylamino)tetralin (8-OH-DPAT). Psychopharma-

cology 91, 506–511.

Goppelt-Struebe, M., & Stroebel, M. (1998). Signaling pathways mediating

induction of the early response genes prostaglandin G/H synthase-2 and

egr-1 by serotonin via 5-HT2A receptors. J Cell Physiol 175, 341–347.

Goppelt-Struebe, M., Hahn, A., Stroebel, M., & Reiser, C. O. (1999).

Independent regulation of cyclo-oxygenase 2 expression by p42/44

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212 205

mitogen-activated protein kinases and Ca2 + /calmodulin-dependent kin-

ase. Biochem J 339, 329–334.

Greene, E. L., Houghton, O., Collinsworth, G., Garnovskaya, M. N., Nagai,

T., Sajjad, T., Bheemanathini, V., Grewal, J. S., Paul, R. V., & Ray-

mond, J. R. (2000). 5-HT2A receptors stimulate mitogen-activated pro-

tein kinase via H2O2 generation in rat renal mesangial cells. Am J

Physiol Renal Physiol 278, F650–F658.

Grewal, J. S., Mukhin, Y. V., Garnovskaya, M. N., Raymond, J. R., &

Greene, E. L. (1999). Serotonin 5-HT2A receptor induces TGF-b1 ex-

pression in mesangial cells via ERK: proliferative and fibrotic signals.

Am J Physiol 276, F922–F930.

Grimaldi, B., Bonnin, A., Fillion, M. P., Ruat, M., Traiffort, E., & Fillion,

G. (1998). Characterization of 5-ht6 receptor and expression of 5-ht6mRNA in the rat brain during ontogenetic development. Naunyn

Schmiedebergs Arch Pharmacol 357, 393–400.

Grotewiel, M. S., & Sanders-Bush, E. (1999). Differences in agonist-inde-

pendent activity of 5-HT2A and 5-HT2C receptors revealed by heterolo-

gous expression. Naunyn Schmiedebergs Arch Pharmacol 359, 21–27.

Guillet-Deniau, I., Burnol, A. F., & Girard, J. (1997). Identification and

localization of a skeletal muscle secrotonin 5-HT2A receptor coupled to

the Jak/STAT pathway. J Biol Chem 272, 14825–14829.

Hagberg, G. B., Blomstrand, F., Nilsson, M., Tamir, H., & Hansson, E.

(1998). Stimulation of 5-HT2A receptors on astrocytes in primary culture

opens voltage-independent Ca2 + channels.Neurochem Int 32, 153–162.

Hahn, A., Heusinger-Ribeiro, J., Lanz, T., Zenkel, S., & Goppelt-Struebe,

M. (2000). Induction of connective tissue growth factor by activation of

heptahelical receptors. Modulation by Rho proteins and the actin cytos-

keleton. J Biol Chem 275, 37429–37435.

Hajduch, E., Dombrowski, L., Darakhshan, F., Rencurel, F., Marette, A., &

Hundal, H. S. (1999). Biochemical localisation of the 5-HT2A (seroto-

nin) receptor in rat skeletal muscle. Biochem Biophys Res Commun 257,

369–372.

Hall, M. D., el Mestikawy, S., Emerit, M. B., Pichat, L., Hamon, M., &

Gozlan, H. (1985). [3H]8-hydroxy-2-(di-n-propylamino)tetralin binding

to pre- and postsynaptic 5-hydroxytryptamine sites in various regions of

the rat brain. J Neurochem 44, 1685–1696.

Hall, R. A., Premont, R. T., & Lefkowitz, R. J. (1999). Heptahelical receptor

signaling: beyond the G protein paradigm. J Cell Biol 145, 927–932.

Hamblin, M. W., & Metcalf, M. A. (1991). Primary structure and functional

characterization of a human 5-HT1D-type serotonin receptor. Mol Phar-

macol 40, 143–148.

Hamblin, M. W., Metcalf, M. A., McGuffin, R. W., & Karpells, S.

(1992). Molecular cloning and functional characterization of a human

5-HT1B serotonin receptor: a homologue of the rat 5-HT1B receptor

with 5-HT1D-like pharmacological specificity. Biochem Biophys Res

Commun 184, 752–759.

Hamblin, M. W., Guthrie, C. R., Kohen, R., & Heidmann, D. E. (1998). Gs

protein-coupled serotonin receptors: receptor isoforms and functional

differences. Ann N Y Acad Sci 861, 31–37.

Harrington, M. A., Shaw, K., Zhong, P., & Ciaranello, R. D. (1994). Ago-

nist-induced desensitization and loss of high-affinity binding sites of

stably expressed human 5-HT1A receptors. J Pharmacol Exp Ther 268,

1098–1106.

Hartig, P. R., Branchek, T. A., & Weinshank, R. L. (1992). A subfamily of

5-HT1D receptor genes. Trends Pharmacol Sci 13, 152–159.

Hartig, P. R., Hoyer, D., Humphrey, P. P., & Martin, G. R. (1996). Align-

ment of receptor nomenclature with the human genome: classification

of 5-HT1B and 5-HT1D receptor subtypes. Trends Pharmacol Sci 17,

103–105.

Heidmann, D. E., Metcalf, M. A., Kohen, R., & Hamblin, M. W. (1997).

Four 5-hydroxytryptamine7 (5-HT7) receptor isoforms in human and

rat produced by alternative splicing: species differences due to altered

intron-exon organization. J Neurochem 68, 1372–1381.

Heidmann, D. E., Szot, P., Kohen, R., & Hamblin, M. W. (1998). Function

and distribution of three rat 5-hydroxytryptamine7 (5-HT7) receptor

isoforms produced by alternative splicing. Neuropharmacology 37,

1621–1632.

Hensler, J. G., Cervera, L. S., Miller, H. A., & Corbitt, J. (1996). Expres-

sion and modulation of 5-hydroxytryptamine1A receptors in P11 cells.

J Pharmacol Exp Ther 278, 1138–1145.

Hinton, J. M., Adams, D., & Garland, C. J. (1999). 5-Hydroxytryptamine

stimulation of phospholipase D activity in the rabbit isolated mesenteric

artery. Br J Pharmacol 126, 1601–1608.

Hirst, W. D., Price, G. W., Rattray, M., & Wilkin, G. P. (1997). Identifica-

tion of 5-hydroxytryptamine receptors positively coupled to adenylyl

cyclase in rat cultured astrocytes. Br J Pharmacol 120, 509–515.

Hirst, W. D., Cheung, N. Y., Rattray, M., Price, G. W., & Wilkin, G. P.

(1998). Cultured astrocytes express messenger RNA for multiple sero-

tonin receptor subtypes, without functional coupling of 5-HT1 receptor

subtypes to adenylyl cyclase. Brain Res Mol Brain Res 61, 90–99.

Hjorth, S. (1985). Hypothermia in the rat induced by the potent serotoni-

nergic agent 8-OH-DPAT. J Neural Transm 61, 131–135.

Hjorth, S., & Tao, R. (1991). The putative 5-HT1B receptor agonist CP-

93,129 suppresses rat hippocampal 5-HT release in vivo: comparison

with RU 24969. Eur J Pharmacol 209, 249–252.

Hoyer, D., & Middlemiss, D. N. (1989). Species differences in the phar-

macology of terminal 5-HT autoreceptors in mammalian brain. Trends

Pharmacol Sci 10, 130–132.

Hoyer, D., Clarke, D. E., Fozard, J. R., Hartig, P. R., Martin, G. R.,

Mylecharane, E. J., Saxena, P. R., & Humphrey, P. P. (1994). Interna-

tional Union of Pharmacology classification of receptors for 5-hydroxy-

tryptamine (serotonin). Pharmacol Rev 46, 157–203.

Hsieh, C. L., Bowcock, A. M., Farrer, L. A., Hebert, J. M., Huang, K. N.,

Cavalli-Sforza, L. L., Julius, D., & Francke, U. (1990). The serotonin

receptor subtype 2 locus HTR2 is on human chromosome 13 near genes

for esterase D and retinoblastoma-1 and on mouse chromosome 14.

Somat Cell Mol Genet 16, 567–574.

Huidobro-Toro, J. P., Valenzuela, C. F., & Harris, R. A. (1996). Modulation

of GABAA receptor function by G protein-coupled 5-HT2C receptors.

Neuropharmacology 35, 1355–1363.

Hung, B. C., Loo, D. D., & Wright, E. M. (1993). Regulation of mouse

choroid plexus apical Cl- and K + channels by serotonin. Brain Res 617,

285–295.

Hurley, J. H. (1999). Structure, mechanism, and regulation of mammalian

adenylyl cyclase. J Biol Chem 274, 7599–7602.

Hutson, P. H., Dourish, C. T., & Curzon, G. (1988). Evidence that the

hyperphagic response to 8-OH-DPAT is mediated by 5-HT1A receptors.

Eur J Pharmacol 150, 361–366.

Iken, K., Chheng, S., Fargin, A., Goulet, A. C., & Kouassi, E. (1995).

Serotonin upregulates mitogen-stimulated B lymphocyte proliferation

through 5-HT1A receptors. Cell Immunol 163, 1–9.

Ishida, T., Kawashima, S., Hirata, K., & Yokoyama, M. (1998). Nitric oxide

is produced via 5-HT1B and 5-HT2B receptor activation in human coro-

nary artery endothelial cells. Kobe J Med Sci 44, 51–63.

Ishizuka, J., Beauchamp, R. D., Townsend, C. M., Greeley, G. H., &

Thompson, J. C. (1992). Receptor-mediated autocrine growth-stimula-

tory effect of 5-hydroxytryptamine on cultured human pancreatic carci-

noid cells. J Cell Physiol 150, 1–7.

Izumi, J., Washizuka, M., Miura, N., Hiraga, Y., & Ikeda, Y. (1994). Hippo-

campal serotonin 5-HT1A receptor enhances acetylcholine release in

conscious rats. J Neurochem 62, 1804–1808.

Jalonen, T. O., Margraf, R. R., Wielt, D. B., Charniga, C. J., Linne, M. L.,

& Kimelberg, H. K. (1997). Serotonin induces inward potassium and

calcium currents in rat cortical astrocytes. Brain Res 758, 69–82.

Jasper, J. R., Kosaka, A., To, Z. P., Chang, D. J., & Eglen, R. M. (1997).

Cloning, expression and pharmacology of a truncated splice variant of

the human 5-HT7 receptor (h5-HT7b). Br J Pharmacol 122, 126–132.

Jin, H., Oksenberg, D., Ashkenazi, A., Peroutka, S. J., Duncan, A. M.,

Rozmahel, R., Yang, Y., Mengod, G., Palacios, J. M., & O’Dowd, B.

F. (1992). Characterization of the human 5-hydroxytryptamine1B recep-

tor. J Biol Chem 267, 5735–5738.

Julius, D., MacDermott, A. B., Axel, R., & Jessell, T. M. (1988). Molecular

characterization of a functional cDNA encoding the serotonin 1c recep-

tor. Science 241, 558–564.

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212206

Julius, D., Livelli, T. J., Jessell, T. M., & Axel, R. (1989). Ectopic expres-

sion of the serotonin 1c receptor and the triggering of malignant trans-

formation. Science 244, 1057–1062.

Julius, D., Huang, K. N., Livelli, T. J., Axel, R., & 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.

Kahan, C., Julius, D., Pouyssegur, J., & Seuwen, K. (1992). Effects of

5-HT1C-receptor expression on cell proliferation control in hamster

fibroblasts: serotonin fails to induce a transformed phenotype. Exp Cell

Res 200, 523–527.

Karschin, A., Ho, B. Y., Labarca, C., Elroy-Stein, O., Moss, B., Davidson,

N., & Lester, H. A. (1991). Heterologously expressed serotonin 1A

receptors couple to muscarinic K + channels in heart. Proc Natl Acad

Sci USA 88, 5694–5698.

Kaufman, M. J., Hartig, P. R., & Hoffman, B. J. (1995). Serotonin 5-HT2C

receptor stimulates cyclic GMP formation in choroid plexus. J Neuro-

chem 64, 199–205.

Kaumann, A. J. (1991). 5-HT4-like receptors in mammalian atria. J Neural

Transm Suppl 34, 195–201.

Kellermann, O., Tournois, C., Richard, S., Manivet, P., Maroteaux, L., &

Launay, J. M. (1998). Signaling pathways and targets of the 5-HT2B

receptor in the 1C11 serotonergic cell line. Ann N Y Acad Sci 861, 248.

Kellett, E., Carr, I. C., & Milligan, G. (1999). Regulation of G protein

activation and effector modulation by fusion proteins between the hu-

man 5-hydroxytryptamine1A receptor and the alpha subunit of Gi1:

differences in receptor-constitutive activity imparted by single amino

acid substitutions in Gi1a. Mol Pharmacol 56, 684–692.

Kelly, J. S., Larkman, P., Penington, N. J., Rainnie, D. G., McAllister-

Williams, H., & Hodgkiss, J. (1991). Serotonin receptor heterogeneity

and the role of potassium channels in neuronal excitability. Adv Exp

Med Biol 287, 177–191.

Kirchgessner, A. L., Liu, M. T., Howard, M. J., & Gershon, M. D. (1993).

Detection of the 5-HT1A receptor and 5-HT1A receptor mRNA in the rat

bowel and pancreas: comparison with 5-HT1P receptors. J Comp Neurol

327, 233–250.

Kobilka, B. K., Frielle, T., Collins, S., Yang-Feng, T., Kobilka, T. S.,

Francke, U., Lefkowitz, R. J., & Caron, M. G. (1987). An intronless

gene encoding a potential member of the family of receptors coupled to

guanine nucleotide regulatory proteins. Nature 329, 75–79.

Kohen, R., Metcalf, M. A., Khan, N., Druck, T., Huebner, K., Lachowicz,

J. E., Meltzer, H. Y., Sibley, D. R., Roth, B. L., & Hamblin, M. W.

(1996). Cloning, characterization, and chromosomal localization of a

human 5-HT6 serotonin receptor. J Neurochem 66, 47–56.

Kuemmerle, J. F., Murthy, K. S., Grider, J. R., Martin, D. C., & Makhlouf,

G. M. (1995). Coexpression of 5-HT2A and 5-HT4 receptors coupled to

distinct signaling pathways in human intestinal muscle cells. Gastro-

enterology 109, 1791–1800.

Kursar, J. D., Nelson, D. L., Wainscott, D. B., & Baez, M. (1994).

Molecular cloning, functional expression, and mRNA tissue distribution

of the human 5-hydroxytryptamine2B receptor. Mol Pharmacol 46,

227–234.

Kurscheid-Reich, D., Throckmorton, D. C., & Rasmussen, H. (1995).

Serotonin activates phospholipase D in rat mesangial cells. Am J Physiol

268, F997–F1003.

Langlois, X., el Mestikawy, S., Arpin, M., Triller, A., Hamon, M., &

Darmon, M. (1996). Differential addressing of 5-HT1A and 5-HT1B

receptors in transfected LLC-PK1 epithelial cells: a model of receptor

targeting in neurons. Neuroscience 74, 297–302.

Launay, J. M., Birraux, G., Bondoux, D., Callebert, J., Choi, D. S., Loric, S.,

& Maroteaux, L. (1996). Ras involvement in signal transduction by the

serotonin 5-HT2B receptor. J Biol Chem 271, 3141–3147.

Launay, J. M., Loric, S., Mutel, V., & Kellermann, O. (1998). The 5-HT2B

receptor controls the overall 5-HT transport system in the 1C11 sero-

tonergic cell line. Ann N Y Acad Sci 861, 247.

Lawrence, A. J., & Marsden, C. A. (1992). Terminal autoreceptor control of

5-hydroxytryptamine release as measured by in vivo microdialysis in

the conscious guinea-pig. J Neurochem 58, 142–146.

Le Grand, B., Panissie, A., Pauwels, P. J., & John, G. W. (1998). Activation

of recombinant h5-HT1B and h5-HT1D receptors stably expressed in C6

glioma cells produces increases in Ca2 + -dependent K + current. Nau-

nyn Schmiedebergs Arch Pharmacol 358, 608–615.

Leeb-Lundberg, L. M. F., Kang, D. S., Lamb, M. E., & Fathy, D. B. (2001).

The human B1 bradykinin receptor exhibits high ligand-independent,

constitutive activity. J Biol Chem 276, 8785–8792.

Lefebvre, H., Dhib, M., Godin, M., Contesse, V., Delarue, C., Rieu, M.,

Wolf, L. M., Vaudry, H., & Kuhn, J. M. (1997). Effect of the serotonin

5-HT4 receptor agonist cisapride on aldosterone secretion in cortico-

tropic insufficiency and primary hyperaldosteronism. Neuroendocrinol-

ogy 66, 229–233.

Lefebvre, H., Cartier, D., Duparc, C., Contesse, V., Lihrmann, I., Delarue, C.,

Vaudry, H., Fischmeister, R., & Kuhn, J. M. (2000). Effect of serotonin4(5-HT4) receptor agonists on aldosterone secretion in idiopathic hyper-

aldosteronism. Endocr Res 26, 583–587.

Leone, M., Attanasio, A., Croci, D., Ferraris, A., D’Amico, D., Grazzi, L.,

Nespolo, A., & Bussone, G. (1998). 5-HT1A receptor hypersensitivity in

migraine is suggested by the m-chlorophenylpiperazine test. Neurore-

port 9, 2605–2608.

Leonhardt, S., Herrick-Davis, K., & Titeler, M. (1989). Detection of a novel

serotonin receptor subtype (5-HT1E) in human brain: interaction with a

GTP-binding protein. J Neurochem 53, 465–471.

Lesage, A. S., Wouters, R., Van Gompel, P., Heylen, L., Vanhoenacker, P.,

Haegeman, G., Luyten, W. H., & Leysen, J. E. (1998). Agonistic proper-

ties of alniditan, sumatriptan and dihydroergotamine on human 5-HT1B

and 5-HT1D receptors expressed in various mammalian cell lines. Br J

Pharmacol 123, 1655–1665.

Leung, G. P., Dun, S. L., Dun, N. J., & Wong, P. Y. (1999). Serotonin via

5-HT1B and 5-HT2B receptors stimulates anion secretion in the rat

epididymal epithelium. J Physiol (Lond) 51, 657–667.

Levy, F. O., Gudermann, T., Birnbaumer, M., Kaumann, A. J., & Birn-

baumer, L. (1992a). Molecular cloning of a human gene (S31) encoding

a novel serotonin receptor mediating inhibition of adenylyl cyclase.

FEBS Lett 296, 201–206.

Levy, F. O., Gudermann, T., Perez-Reyes, E., Birnbaumer, M., Kaumann,

A. J., & Birnbaumer, L. (1992b). Molecular cloning of a human sero-

tonin receptor (S12) with a pharmacological profile resembling that of

the 5-HT1D subtype. J Biol Chem 267, 7553–7562.

Levy, F. O., Holtgreve-Grez, H., Tasken, K., Solberg, R., Ried, T., & Gu-

dermann, T. (1994). Assignment of the gene encoding the 5-HT1E sero-

tonin receptor (S31) (locus HTR1E) to human chromosome 6q14-q15.

Genomics 22, 637–640.

Leysen, J. E., Niemegeers, C. J., Tollenaere, J. P., & Laduron, P. M. (1978).

Serotonergic component of neuroleptic receptors. Nature 272, 168–171.

Lione, A. M., Errico, M., Lin, S. L., & Cowen, D. S. (2000). Activation of

extracellular signal-regulated kinase (ERK) and Akt by human seroto-

nin 5-HT1B receptors in transfected BE(2)-C neuroblastoma cells is

inhibited by RGS4. J Neurochem 75, 934–938.

Liu, Y. F., & Albert, P. R. (1991). Cell-specific signaling of the 5-HT1A

receptor. Modulation by protein kinases C and A. J Biol Chem 266,

23689–23697.

Liu, Y. F., Jakobs, K. H., Rasenick, M. M., & Albert, P. R. (1994). G protein

specificity in receptor-effector coupling. Analysis of the roles of Go and

Gi2 in GH4C1 pituitary cells. J Biol Chem 269, 13880–13886.

Liu, Y. F., Ghahremani, M. H., Rasenick, M. M., Jakobs, K. H., & Albert, P.

R. (1999). Stimulation of cAMP synthesis by Gi-coupled receptors

upon ablation of distinct Gai protein expression. Gi subtype specificity

of the 5-HT1A receptor. J Biol Chem 274, 16444–16450.

Loric, S., Launay, J. M., Colas, J. F., & Maroteaux, L. (1992). New mouse

5-HT2-like receptor. Expression in brain, heart and intestine. FEBS Lett

312, 203–207.

Loric, S., Maroteaux, L., Kellermann, O., & Launay, J. M. (1995). Func-

tional serotonin-2B receptors are expressed by a teratocarcinoma-

derived cell line during serotoninergic differentiation. Mol Pharmacol

47, 458–466.

Lovenberg, T. W., Baron, B. M., de Lecea, L., Miller, J. D., Prosser, R. A.,

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212 207

Rea,M. A., Foye, P. E., Racke,M., Slone, A. L., & Siegel, B.W. (1993a).

A novel adenylyl cyclase-activating serotonin receptor (5-HT7) impli-

cated in the regulation of mammalian circadian rhythms. Neuron 11,

449–458.

Lovenberg, T. W., Erlander, M. G., Baron, B. M., Racke, M., Slone, A. L.,

Siegel, B. W., Craft, C. M., Burns, J. E., Danielson, P. E., & Sutcliffe, J.

G. (1993b). Molecular cloning and functional expression of 5-HT1E-like

rat and human 5-hydroxytryptamine receptor genes. Proc Natl Acad Sci

USA 90, 2184–2188.

Lubbert, H., Hoffman, B. J., Snutch, T. P., van Dyke, T., Levine, A. J.,

Hartig, P. R., Lester, H. A., & Davidson, N. (1987). cDNA cloning

of a serotonin 5-HT1C receptor by electrophysiological assays of

mRNA-injected Xenopus oocytes. Proc Natl Acad Sci USA 84,

4332–4336.

Lucaites, V. L., Nelson, D. L., Wainscott, D. B., & Baez, M. (1996).

Receptor subtype and density determine the coupling repertoire of the

5-HT2 receptor subfamily. Life Sci 59, 1081–1095.

Luttrell, L. M., van Biesen, T., Hawes, B. E., Koch, W. J., Krueger, K. M.,

Touhara, K., & Lefkowitz, R. J. (1997). G-protein-coupled receptors

and their regulation: activation of the MAP kinase signaling pathway by

G-protein-coupled receptors. Adv Second Messenger Phosphoprotein

Res 31, 263–277.

Maenhaut, C., Van Sande, J., Massart, C., Dinsart, C., Libert, F., Monferini,

E., Giraldo, E., Ladinsky, H., Vassart, G., & Dumont, J. E. (1991). The

orphan receptor cDNA RDC4 encodes a 5-HT1D serotonin receptor.

Biochem Biophys Res Commun 180, 1460–1468.

Malmberg, A., & Strange, P. G. (2000). Site-directed mutations in the third

intracellular loop of the serotonin 5-HT1A receptor alter G protein cou-

pling from Gi to Gs in a ligand-dependent manner. J Neurochem 75,

1283–1293.

Manivet, P., Mouillet-Richard, S., Callebert, J., Nebigil, C. G., Maro-

teaux, L., Hosoda, S., Kellermann, O., & Launay, J. M. (2000).

PDZ-dependent activation of nitric-oxide synthases by the serotonin

2B receptor. J Biol Chem 275, 9324–9331.

Marcoli, M., Maura, G., Tortarolo, M., & Raiteri, M. (1997). Serotonin

inhibition of the NMDA receptor/nitric oxide/cyclic GMP pathway

in rat cerebellum: involvement of 5-hydroxytryptamine2C receptors.

J Neurochem 69, 427–430.

Markstein, R., Hoyer, D., & Engel, G. (1986). 5-HT1A-receptors mediate

stimulation of adenylate cyclase in rat hippocampus. Naunyn Schmiede-

bergs Arch Pharmacol 333, 335–341.

Maroteaux, L., Saudou, F., Amlaiky, N., Boschert, U., Plassat, J. L., & Hen,

R. (1992). Mouse 5HT1B serotonin receptor: cloning, functional expres-

sion, and localization in motor control centers. Proc Natl Acad Sci USA

89, 3020–3024.

Marracci, S., Cini, D., & Nardi, I. (1997). Cloning and developmental

expression of 5-HT1A receptor gene in Xenopus laevis. Brain Res

Mol Brain Res 47, 67–77.

Marshall, C. J. (1995). Specificity of receptor tyrosine kinase signaling:

transient versus sustained extracellular signal-regulated kinase activa-

tion. Cell 80, 179–185.

Martin, K. F., Hannon, S., Phillips, I., & Heal, D. J. (1992). Opposing roles

for 5-HT1B and 5-HT3 receptors in the control of 5-HT release in rat

hippocampus in vivo. Br J Pharmacol 106, 139–142.

Maswood, N., Caldarola-Pastuszka, M., & Uphouse, L. (1998). Functional

integration among 5-hydroxytryptamine receptor families in the control

of female rat sexual behavior. Brain Res 802, 98–103.

Matsuda, H., Li, Y., & Yoshikawa, M. (2000). Possible involvement of

5-HT and 5-HT2 receptors in acceleration of gastrointestinal transit by

escin Ib in mice. Life Sci 66, 2233–2238.

Matthes, H., Boschert, U., Amlaiky, N., Grailhe, R., Plassat, J. L., Musca-

telli, F., Mattei, M. G., & Hen, R. (1993). Mouse 5-hydroxytryptami-

ne5A and 5-hydroxytryptamine5B receptors define a new family of

serotonin receptors: cloning, functional expression, and chromosomal

localization. Mol Pharmacol 43, 313–319.

Maura, G., & Raiteri, M. (1996). Serotonin 5-HT1D and 5-HT1A receptors

respectively mediate inhibition of glutamate release and inhibition of

cyclic GMP production in rat cerebellum in vitro. J Neurochem 66,

203–209.

Maura, G., Marcoli, M., Tortarolo, M., Andrioli, G. C., & Raiteri, M.

(1998). Glutamate release in human cerebral cortex and its modulation

by 5-hydroxytryptamine acting at h 5-HT1D receptors. Br J Pharmacol

123, 45–50.

Maura, G., Marcoli, M., Pepicelli, O., Rosu, C., Viola, C., & Raiteri, M.

(2000). Serotonin inhibition of the NMDA receptor/nitric oxide/cyclic

GMP pathway in human neocortex slices: involvement of 5-HT2C and

5-HT1A receptors. Br J Pharmacol 130, 1853–1858.

Mayer, S. E., & Sanders-Bush, E. (1994). 5-Hydroxytryptamine type 2A

and 2C receptors linked to Na + /K + /Cl- cotransport. Mol Pharmacol

45, 991–996.

McAllister, G., Charlesworth, A., Snodin, C., Beer, M. S., Noble, A. J.,

Middlemiss, D. N., Iversen, L. L., & Whiting, P. (1992). Molecular

cloning of a serotonin receptor from human brain (5HT1E): a fifth

5HT1-like subtype. Proc Natl Acad Sci USA 89, 5517–5521.

McDuffie, J. E., Coaxum, S. D., & Maleque, M. A. (1999). 5-Hydroxy-

tryptamine evokes endothelial nitric oxide synthase activation in bovine

aortic endothelial cell cultures. Proc Soc Exp Biol Med 221, 386–390.

McDuffie, J. E., Motley, E. D., Limbird, L. E., & Maleque, M. A. (2000).

5-Hydroxytryptamine stimulates phosphorylation of p44/p42 mitogen-

activated protein kinase activation in bovine aortic endothelial cell

cultures. J Cardiovasc Pharmacol 35, 398–402.

McLean, P. G., & Coupar, I. M. (1996). Further investigation into the signal

transduction mechanism of the 5-HT4-like receptor in the circular

smooth muscle of human colon. Br J Pharmacol 118, 1058–1064.

Mendez, J., Kadia, T. M., Somayazula, R. K., El-Badawi, K. I., & Cowen,

D. S. (1999). Differential coupling of serotonin 5-HT1A and 5-HT1B

receptors to activation of ERK2 and inhibition of adenylyl cyclase in

transfected CHO cells. J Neurochem 73, 162–168.

Mengod, G., Nguyen, H., Le, H., Waeber, C., Lubbert, H., & Palacios, J. M.

(1990). The distribution and cellular localization of the serotonin 1C

receptor mRNA in the rodent brain examined by in situ hybridization

histochemistry. Comparison with receptor binding distribution. Neuro-

science 35, 577–591.

Metcalf, M. A., McGuffin, R. W., & Hamblin, M. W. (1992). Conversion of

the human 5-HT1D beta serotonin receptor to the rat 5-HT1B ligand-

binding phenotype by Thr355Asn site directed mutagenesis. Biochem

Pharmacol 44, 1917–1920.

Meyerhof, W., Obermuller, F., Fehr, S., & Richter, D. (1993). A novel rat

serotonin receptor: primary structure, pharmacology, and expression

pattern in distinct brain regions. DNA Cell Biol 12, 401–409.

Mialet, J., Berque-Bestel, I., Eftekhari, P., Gastineau, M., Giner, M., Dah-

moune, Y., Donzeau-Gouge, P., Hoebeke, J., Langlois, M., Sicsic, S.,

Fischmeister, R., & Lezoualc’h, F. (2000). Isolation of the serotoniner-

gic 5-HT4e receptor from human heart and comparative analysis of its

pharmacological profile in C6-glial and CHO cell lines. Br J Pharmacol

129, 771–781.

Miczek, K. A., Hussain, S., & Faccidomo, S. (1998). Alcohol-heightened

aggression in mice: attenuation by 5-HT1A receptor agonists. Psycho-

pharmacology (Berl) 139, 160–168.

Middlemiss, D. N., & Fozard, J. R. (1983). 8-Hydroxy-2-(di-n-propylami-

no)-tetralin discriminates between subtypes of the 5-HT1 recognition

site. Eur J Pharmacol 90, 151–153.

Middlemiss, D. N., & Hutson, P. H. (1990). The 5-HT1B receptors. Ann N Y

Acad Sci 600, 132–147.

Middleton, J. P., Albers, F. J., Dennis, V. W., & Raymond, J. R. (1990).

Thapsigargin demonstrates calcium-dependent regulation of phosphate

uptake in HeLa cells. Am J Physiol 259, F727–F731.

Millan, M. J., Dekeyne, A., & Gobert, A. (1998). Serotonin (5-HT)2Creceptors tonically inhibit dopamine (DA) and noradrenaline (NA),

but not 5-HT, release in the frontal cortex in vivo. Neuropharmacology

37, 953–955.

Miller, K. J., & Gonzalez, H. A. (1998). Serotonin 5-HT2A receptor acti-

vation inhibits cytokine-stimulated inducible nitric oxide synthase in C6

glioma cells. Ann N Y Acad Sci 861, 169–173.

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212208

Miquel, M. C., Doucet, E., Riad, M., Adrien, J., Verge, D., & Hamon, M.

(1992). Effect of the selective lesion of serotoninergic neurons on the

regional distribution of 5-HT1A receptor mRNA in the rat brain. Brain

Res Mol Brain Res 14, 357–362.

Mitsikostas, D. D., Sanchez del Rio, M., Moskowitz, M. A., & Waeber, C.

(1999). Both 5-HT1B and 5-HT1F receptors modulate c-fos expression

within rat trigeminal nucleus caudalis. Eur J Pharmacol 369, 271–277.

Moiseiwitsch, J. R., Raymond, J. R., Tamir, H., & Lauder, J. M. (1998).

Regulation by serotonin of tooth-germ morphogenesis and gene expres-

sion in mouse mandibular explant cultures. Arch Oral Biol 43, 789–800.

Molderings, G. J., Frolich, D., Likungu, J., & Gothert, M. (1996). Inhibition

of noradrenaline release via presynaptic 5-HT1Da receptors in human

atrium. Naunyn Schmiedebergs Arch Pharmacol 353, 272–280.

Mons, N., Decorte, L., Jaffard, R., & Cooper, D. M. (1998). Ca2 + -sensitive

adenylyl cyclases, key integrators of cellular signalling. Life Sci 62,

1647–1652.

Monsma, F. J., Shen, Y., Ward, R. P., Hamblin, M. W., & Sibley, D. R.

(1993). Cloning and expression of a novel serotonin receptor with high

affinity for tricyclic psychotropic drugs. Mol Pharmacol 43, 320–327.

Montiel, C., Herrero, C. J., Garcia-Palomero, E., Renart, J., Garcia, A. G.,

& Lomax, R. B. (1997). Serotonergic effects of dotarizine in coronary

artery and in oocytes expressing 5-HT2 receptors. Eur J Pharmacol

332, 183–193.

Morecroft, I., & MacLean, M. R. (1998). 5-Hydroxytryptamine receptors

mediating vasoconstriction and vasodilation in perinatal and adult rabbit

small pulmonary arteries. Br J Pharmacol 125, 69–78.

Mork, A., & Geisler, A. (1990). 5-Hydroxytryptamine receptor agonists

influence calcium-stimulated adenylate cyclase activity in the cerebral

cortex and hippocampus of the rat. Eur J Pharmacol 175, 237–244.

Mukhin, Y. V., Garnovskaya, M. N., Collinsworth, G., Grewal, J. S.,

Pendergrass, D., Nagai, T., Pinckney, S., Greene, E. L., & Raymond,

J. R. (2000). 5-Hydroxytryptamine1A receptor/Gibg stimulates mitogen-

activated protein kinase via NAD(P)H oxidase and reactive oxygen

species upstream of src in Chinese hamster ovary fibroblasts. Biochem

J 347, 61–67.

Mulheron, J. G., Casanas, S. J., Arthur, J. M., Garnovskaya, M. N., Gettys,

T. W., & Raymond, J. R. (1994). Human 5-HT1A receptor expressed in

insect cells activates endogenous Go-like G protein(s). J Biol Chem 269,

12954–12962.

Nakaki, T., Roth, B. L., Chuang, D. M., & Costa, E. (1985). Phasic and

tonic components in 5-HT2 receptor-mediated rat aorta contraction:

participation of Ca++ channels and phospholipase C. J Pharmacol

Exp Ther 234, 442–446.

Nebigil, C. G., Choi, D. S., Dierich, A., Hickel, P., Le Meur, M., Mes-

saddeq, N., Launay, J. M., & Maroteaux, L. (2000a). Serotonin 2B

receptor is required for heart development. Proc Natl Acad Sci USA

97, 9508–9513.

Nebigil, C. G., Launay, J. M., Hickel, P., Tournois, C., & Maroteaux, L.

(2000b). 5-Hydroxytryptamine 2B receptor regulates cell-cycle progres-

sion: cross-talk with tyrosine kinase pathways. Proc Natl Acad Sci USA

97, 2591–2596.

Nelson, C. S., Cone, R. D., Robbins, L. S., Allen, C. N., & Adelman, J. P.

(1995). Cloning and expression of a 5HT7 receptor from Xenopus

laevis. Recept Channels 3, 61–70.

Nelson, D. L., Pedigo, N. W., & Yamamura, H. I. (1981). Multiple3H-5-hydroxytryptamine binding sites in rat brain. J Physiol 77,

369–372.

Newman-Tancredi, A., Wootton, R., & Strange, P. G. (1992). High-level

stable expression of recombinant 5-HT1A 5-hydroxytryptamine receptors

in Chinese hamster ovary cells. Biochem J 285, 933–938.

Ng,G.Y.,George, S.R., Zastawny,R.L.,Caron,M.,Bouvier,M.,Dennis,M.,

& O’Dowd, B. F. (1993). Human serotonin1B receptor expression in Sf 9

cells: phosphorylation, palmitoylation, and adenylyl cyclase inhibition.

Biochemistry 32, 11727–11733.

Ni, Y. G., Panicker, M. M., & Miledi, R. (1997). Efficient coupling of

5-HT1a receptors to the phospholipase C pathway in Xenopus oocytes.

Brain Res Mol Brain Res 51, 115–122.

Nishizuka, Y. (1995). Protein kinase C and lipid signaling for sustained

cellular responses. FASEB J 9, 484–496.

Niswender, C. M., Sanders-Bush, E., & Emeson, R. B. (1998). Identifica-

tion and characterization of RNA editing events within the 5-HT2C

receptor. Ann N Y Acad Sci 861, 38–48.

Niswender, C. M., Copeland, S. C., Herrick-Davis, K., Emeson, R. B., &

Sanders-Bush, E. (1999). RNA editing of the human serotonin 5-hy-

droxytryptamine2C receptor silences constitutive activity. J Biol Chem

274, 9472–9478.

Nitsch, R. M., Deng, M., Growdon, J. H., & Wurtman, R. J. (1996).

Serotonin 5-HT2a and 5-HT2c receptors stimulate amyloid precursor

protein ectodomain secretion. J Biol Chem 271, 4188–4194.

Obosi, L. A., Hen, R., Beadle, D. J., Bermudez, I., & King, L. A. (1997).

Mutational analysis of the mouse 5-HT7 receptor: importance of the

third intracellular loop for receptor-G-protein interaction. FEBS Lett

412, 321–324.

Okoro, E. O. (1999). Overlap in the pharmacology of L-type Ca2 + -channel

blockers and 5-HT2 receptor antagonists in rat aorta. J Pharm Pharma-

col 51, 953–957.

Oksenberg, D., Marsters, S. A., O’Dowd, B. F., Jin, H., Havlik, S., Per-

outka, S. J., & Ashkenazi, A. (1992). A single amino-acid difference

confers major pharmacological variation between human and rodent

5-HT1B receptors. Nature 360, 161–163.

Olsen, M. A., Nawoschik, S. P., Schurman, B. R., Schmitt, H. L., Burno, M.,

Smith, D. L., & Schechter, L. E. (1999). Identification of a human 5-HT6

receptor variant produced by alternative splicing. Brain Res Mol Brain

Res 64, 255–263.

Ouadid, H., Seguin, J., Dumuis, A., Bockaert, J., & Nargeot, J. (1992).

Serotonin increases calcium current in human atrial myocytes via the

newly described 5-hydroxytryptamine4 receptors. Mol Pharmacol 41,

346–351.

Palacios, J. M., Waeber, C., Hoyer, D., & Mengod, G. (1990). Distribution

of serotonin receptors. Ann N Y Acad Sci 600, 36–52.

Palego, L., Giromella, A., Marazziti, D., Borsini, F., Naccarato, A. G.,

Giannaccini, G., Lucacchini, A., Cassano, G. B., & Mazzoni, M. R.

(1999). Effects of postmortem delay on serotonin and (+)8-OH-DPAT-

mediated inhibition of adenylyl cyclase activity in rat and human brain

tissues. Brain Res 816, 165–174.

Palego, L., Giromella, A., Marazziti, D., Giannaccini, G., Borsini, F., Big-

azzi, F., Naccarato, A. G., Lucacchini, A., Cassano, G. B., & Mazzoni,

M. R. (2000). Lack of stereoselectivity of 8-hydroxy-2(di-N-propyla-

mino)tetralin-mediated inhibition of forskolin-stimulated adenylyl cy-

clase activity in human pre- and post-synaptic brain regions. Neurochem

Int 36, 225–232.

Pan, H., Wang, H. Y., Friedman, E., & Gershon, M. D. (1997). Mediation

by protein kinases C and A of Go-linked slow responses of enteric

neurons to 5-HT. J Neurosci 17, 1011–1024.

Panicker, M. M., Parker, I., & Miledi, R. (1991). Receptors of the serotonin

1C subtype expressed from cloned DNA mediate the closing of K +

membrane channels encoded by brain mRNA. Proc Natl Acad Sci

USA 88, 2560–2562.

Parekh, A. B., Foguet, M., Lubbert, H., & Stuhmer, W. (1993). Ca2 +

oscillations and Ca2 + influx in Xenopus oocytes expressing a novel

5-hydroxytryptamine receptor. J Physiol 469, 653–671.

Parker, E. M., Grisel, D. A., Iben, L. G., & Shapiro, R. A. (1993). A single

amino acid difference accounts for the pharmacological distinctions

between the rat and human 5-hydroxytryptamine1B receptors. J Neuro-

chem 60, 380–383.

Parks, C. L., Robinson, P. S., Sibille, E., Shenk, T., & Toth, M. (1998).

Increased anxiety of mice lacking the serotonin1A receptor. Proc Natl

Acad Sci USA 95, 10734–10739.

Pauwels, P. J., Palmier, C., Wurch, T., & Colpaert, F. C. (1996). Pharmacol-

ogy of cloned human 5-HT1D receptor-mediated functional responses in

stably transfected rat C6-glial cell lines: further evidence differentiating

human 5-HT1D and 5-HT1B receptors. Naunyn Schmiedebergs Arch

Pharmacol 353, 144–156.

Pazos, A., Cortes, R., & Palacios, J. M. (1985). Quantitative autoradio-

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212 209

graphic mapping of serotonin receptors in the rat brain. II. Serotonin-2

receptors. Brain Res 346, 231–249.

Pedigo, N. W., Yamamura, H. I., & Nelson, D. L. (1981). Discrimination

of multiple [3H]5-hydroxytryptamine binding sites by the neuroleptic

spiperone in rat brain. J Neurochem 36, 220–226.

Penington, N. J., & Kelly, J. S. (1990). Serotonin receptor activation reduces

calcium current in an acutely dissociated adult central neuron. Neuron 4,

751–758.

Penington, N. J., Kelly, J. S., & Fox, A. P. (1991). A study of the mech-

anism of Ca2 + current inhibition produced by serotonin in rat dorsal

raphe neurons. J Neurosci 11, 3594–3609.

Penington, N. J., Kelly, J. S., & Fox, A. P. (1993). Whole-cell recordings of

inwardly rectifying K + currents activated by 5-HT1A receptors on dor-

sal raphe neurones of the adult rat. J Physiol 469, 387–405.

Peroutka, S. J. (1995). 5-HT receptors: past, present and future. Trends

Neurosci 18, 68–69.

Peroutka, S. J., & Snyder, S. H. (1979). Multiple serotonin receptors: differ-

ential binding of [3H]5-hydroxytryptamine, [3H]lysergic acid diethyla-

mide and [3H]spiroperidol. Mol Pharmacol 16, 687–699.

Pino, R., Cerbai, E., Calamai, G., Alajmo, F., Borgioli, A., Braconi, L.,

Cassai, M., Montesi, G. F., & Mugelli, A. (1998). Effect of 5-HT4

receptor stimulation on the pacemaker current I(f ) in human isolated

atrial myocytes. Cardiovasc Res 40, 516–522.

Plassat, J. L., Boschert, U., Amlaiky, N., & Hen, R. (1992). The mouse

5HT5 receptor reveals a remarkable heterogeneity within the 5HT1D

receptor family. EMBO J 11, 4779–4786.

Plassat, J. L., Amlaiky, N., & Hen, R. (1993). Molecular cloning of a

mammalian serotonin receptor that activates adenylate cyclase. Mol

Pharmacol 44, 229–236.

Pritchett, D. B., Bach, A.W., Wozny, M., Taleb, O., Dal Toso, R., Shih, J. C.,

& Seeburg, P. H. (1988). Structure and functional expression of cloned

rat serotonin 5HT-2 receptor. EMBO J 7, 4135–4140.

Pullarkat, S. R., Mysels, D. J., Tan, M., & Cowen, D. S. (1998). Coupling

of serotonin 5-HT1B receptors to activation of mitogen-activated protein

kinase (ERK-2) and p70 S6 kinase signaling systems. J Neurochem 71,

1059–1067.

Quick, M. W., Simon, M. I., Davidson, N., Lester, H. A., & Aragay, A.

M. (1994). Differential coupling of G protein alpha subunits to sev-

en-helix receptors expressed in Xenopus oocytes. J Biol Chem 269,

30164–30172.

Radja, F., Daval, G., Hamon, M., & Verge, D. (1992). Pharmacological and

physicochemical properties of pre-versus postsynaptic 5-hydroxytrypta-

mine1A receptor binding sites in the rat brain: a quantitative autoradio-

graphic study. J Neurochem 58, 1338–1346.

Raiteri, M., Maura, G., Bonanno, G., & Pittaluga, A. (1986). Differential

pharmacology and function of two 5-HT1 receptors modulating trans-

mitter release in rat cerebellum. J Pharmacol Exp Ther 237, 644–648.

Ramboz, S., Oosting, R., Amara, D. A., Kung, H. F., Blier, P., Mendel-

sohn, M., Mann, J. J., Brunner, D., & Hen, R. (1998). Serotonin

receptor 1A knockout: an animal model of anxiety-related disorder.

Proc Natl Acad Sci USA 95, 14476–14481.

Rapport, M. M., Green, A. A., & Page, I. H. (1948). Crystalline serotonin.

Science 108, 329–330.

Raymond, J. R. (1991). Protein kinase C induces phosphorylation and

desensitization of the human 5-HT1A receptor. J Biol Chem 266,

14747–14753.

Raymond, J. R. (1995). Multiple mechanisms of receptor-G protein signal-

ing specificity. Am J Physiol 269, F141–F158.

Raymond, J. R., Fargin, A., Middleton, J. P., Graff, J. M., Haupt, D. M.,

Caron, M. G., Lefkowitz, R. J., & Dennis, V. W. (1989). The human

5-HT1A receptor expressed in HeLa cells stimulates sodium-depend-

ent phosphate uptake via protein kinase C. J Biol Chem 264,

21943–21950.

Raymond, J. R., Albers, F. J., Middleton, J. P., Lefkowitz, R. J., Caron,M. G.,

Obeid, L. M., & Dennis, V. W. (1991). 5-HT1A and histamine H1 recep-

tors in HeLa cells stimulate phosphoinositide hydrolysis and phosphate

uptake via distinct G protein pools. J Biol Chem 266, 372–379.

Raymond, J. R., Albers, F. J., & Middleton, J. P. (1992). Functional expres-

sion of human 5-HT1A receptors and differential coupling to second

messengers in CHO cells. Naunyn Schmiedebergs Arch Pharmacol

346, 127–137.

Raymond, J. R., Kim, J., Beach, R. E., & Tisher, C. C. (1993a). Immunohis-

tochemical mapping of cellular and subcellular distribution of 5-HT1A

receptors in rat and human kidneys. Am J Physiol 264, F9–F19.

Raymond, J. R., Olsen, C. L., & Gettys, T. W. (1993b). Cell-specific

physical and functional coupling of human 5-HT1A receptors to inhib-

itory G protein alpha-subunits and lack of coupling to Gs alpha. Bio-

chemistry 32, 11064–11073.

Rees, S., den Daas, I., Foord, S., Goodson, S., Bull, D., Kilpatrick, G., &

Lee, M. (1994). Cloning and characterisation of the human 5-HT5A

serotonin receptor. FEBS Lett 355, 242–246.

Rhoden, K. J., Dodson, A. M., & Ky, B. (2000). Stimulation of the Na + -K +

pump in cultured guinea pig airway smooth muscle cells by serotonin.

J Pharmacol Exp Ther 293, 107–112.

Roth, B. L., Nakaki, T., Chuang, D. M., & Costa, E. (1984). Aortic recog-

nition sites for serotonin (5HT) are coupled to phospholipase C

and modulate phosphatidylinositol turnover. Neuropharmacology 23,

1223–1225.

Roth, B. L., Nakaki, T., Chuang, D. M., & Costa, E. (1986). 5-Hydroxy-

tryptamine2 receptors coupled to phospholipase C in rat aorta: modu-

lation of phosphoinositide turnover by phorbol ester. J Pharmacol Exp

Ther 238, 480–485.

Roth, B. L., Willins, D. L., Kristiansen, K., & Kroeze, W. K. (1998).

5-Hydroxytryptamine2-family receptors (5-hydroxytryptamine2A, 5-hy-

droxytryptamine2B, 5-hydroxytryptamine2C): where structure meets

function. Pharmacol Ther 79, 231–257.

Ruat, M., Traiffort, E., Arrang, J. M., Tardivel-Lacombe, J., Diaz, J.,

Leurs, R., & Schwartz, J. C. (1993a). A novel rat serotonin (5-HT6)

receptor: molecular cloning, localization and stimulation of cAMP

accumulation. Biochem Biophys Res Commun 193, 268–276.

Ruat, M., Traiffort, E., Leurs, R., Tardivel-Lacombe, J., Diaz, J., Arrang,

J. M., & Schwartz, J. C. (1993b). Molecular cloning, characterization,

and localization of a high-affinity serotonin receptor (5-HT7) activating

cAMP formation. Proc Natl Acad Sci USA 90, 8547–8551.

Saltzman, A. G., Morse, B., Whitman, M. M., Ivanshchenko, Y., Jaye, M.,

& Felder, S. (1991). Cloning of the human serotonin 5-HT2 and 5-HT1C

receptor subtypes. Biochem Biophys Res Commun 181, 1469–1478.

Sanden, N., Thorlin, T., Blomstrand, F., Persson, P. A. I., & Hansson, E.

(2000). 5-Hydroxytryptamine2B receptors stimulate Ca2 + increases in

cultured astrocytes from three different brain regions. Neurochem Int

36, 427–434.

Sanders-Bush, E., Burris, K. D., & Knoth, K. (1988). Lysergic acid dieth-

ylamide and 2,5-dimethoxy-4-methylamphetamine are partial agonists

at serotonin receptors linked to phosphoinositide hydrolysis. J Pharma-

col Exp Ther 246, 924–928.

Saudou, F., & Hen, R. (1994). 5-Hydroxytryptamine receptor subtypes in

vertebrates and invertebrates. Neurochem Int 25, 503–532.

Saxena, R., Saksa, B. A., Fields, A. P., & Ganz, M. B. (1993). Activation of

Na/H exchanger in mesangial cells is associated with translocation of

PKC isoforms. Am J Physiol 265, F53–F60.

Schmuck, K., Ullmer, C., Engels, P., & Lubbert, H. (1994). Cloning and

functional characterization of the human 5-HT2B serotonin receptor.

FEBS Lett 342, 85–90.

Schnur, S. L., Smith, E. R., Lee, R. L., Mas, M., & Davidson, J. M. (1989).

A component analysis of the effects of DPAT on male rat sexual

behavior. Physiol Behav 45, 897–901.

Schoeffter, P., & Waeber, C. (1994). 5-Hydroxytryptamine receptors with a

5-HT6 receptor-like profile stimulating adenylyl cyclase activity in pig

caudate membranes. Naunyn Schmiedebergs Arch Pharmacol 350,

356–360.

Schoeffter, P., Waeber, C., Palacios, J. M., & Hoyer, D. (1988). The

5-hydroxytryptamine 5-HT1D receptor subtype is negatively coupled to

adenylate cyclase in calf substantia nigra. Naunyn Schmiedebergs Arch

Pharmacol 337, 602–608.

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212210

Schoeffter, P., Ullmer, C., Gutierrez, M., Weitz-Schmidt, G., & Lubbert, H.

(1995). Functional serotonin 5-HT1D receptors and 5-HT1D beta receptor

mRNA expression in human umbilical vein endothelial cells. Naunyn

Schmiedebergs Arch Pharmacol 352, 580–582.

Schoeffter, P., Ullmer, C., Bobirnac, I., Gabbiani, G., & Lubbert, H. (1996).

Functional, endogenously expressed 5-hydroxytryptamine 5-ht7 recep-

tors in human vascular smooth muscle cells. Br J Pharmacol 117,

993–994.

Sebben, M., Ansanay, H., Bockaert, J., & Dumuis, A. (1994). 5-HT6 recep-

tors positively coupled to adenylyl cyclase in striatal neurones in culture.

Neuroreport 5, 2553–2557.

Seletti, B., Benkelfat, C., Blier, P., Annable, L., Gilbert, F., & de Montigny,

C. (1995). Serotonin1A receptor activation by flesinoxan in humans.

Body temperature and neuroendocrine responses. Neuropsychopharma-

cology 13, 93–104.

Seuwen, K., Magnaldo, I., & Pouyssegur, J. (1988). Serotonin stimulates

DNA synthesis in fibroblasts acting through 5-HT1B receptors coupled

to a Gi-protein. Nature 335, 254–256.

Sharp, T., & Hjorth, S. (1990). Application of brain microdialysis to study

the pharmacology of the 5-HT1A autoreceptor. J Neurosci Methods 34,

83–90.

Sharp, T., Bramwell, S. R., & Grahame-Smith, D. G. (1989). 5-HT1 ago-

nists reduce 5-hydroxytryptamine release in rat hippocampus in vivo as

determined by brain microdialysis. Br J Pharmacol 96, 283–290.

Shen, Y., Monsma, F. J. Jr., Metcalf, M. A., Jose, P. A., Hamblin, M. W., &

Sibley, D. R. (1993). Molecular cloning and expression of a 5-hydroxy-

tryptamine7 serotonin receptor subtype. J Biol Chem 268, 18200–18204.

Shenker, A., Maayani, S., Weinstein, H., & Green, J. P. (1987). Pharmaco-

logical characterization of two 5-hydroxytryptamine receptors coupled

to adenylate cyclase in guinea pig hippocampal membranes. Mol Phar-

macol 31, 357–367. Erratum: Mol Pharmacol 32, 564 (1987).

Shiah, I. S., Yatham, L. N., Lam, R. W., Tam, E. M., & Zis, A. P. (1998).

Cortisol, hypothermic, and behavioral responses to ipsapirone in pa-

tients with bipolar depression and normal controls. Neuropsychobiology

38, 6–12.

Shimizu, M., Nishida, A., Zensho, H., & Yamawaki, S. (1996). Chronic

antidepressant exposure enhances 5-hydroxytryptamine7 receptor-medi-

ated cyclic adenosine monophosphate accumulation in rat frontocortical

astrocytes. J Pharmacol Exp Ther 279, 1551–1558.

Singh, J. K., Yan, Q., Dawson, G., & Banerjee, P. (1996). Cell-specific

regulation of the stably expressed serotonin 5-HT1A receptor and altered

ganglioside synthesis. Biochim Biophys Acta 1310, 201–211.

Sleight, A. J., Boess, F. G., Bos, M., & Bourson, A. (1998). The putative

5-ht6 receptor: localization and function. Ann N YAcad Sci 861, 91–96.

Stam, N. J., Van Huizen, F., Van Alebeek, C., Brands, J., Dijkema, R.,

Tonnaer, J. A., & Olijve, W. (1992). Genomic organization, coding

sequence and functional expression of human 5-HT2 and 5-HT1A

receptor genes. Eur J Pharmacol 227, 153–162.

Stam, N. J., Vanderheyden, P., van Alebeek, C., Klomp, J., de Boer, T., van

Delft, A. M., & Olijve, W. (1994). Genomic organisation and functional

expression of the gene encoding the human serotonin 5-HT2C receptor.

Eur J Pharmacol 269, 339–348.

Stam, N. J., Roesink, C., Dijcks, F., Garritsen, A., van Herpen, A., &

Olijve, W. (1997). Human serotonin 5-HT7 receptor: cloning and phar-

macological characterisation of two receptor variants. FEBS Lett 413,

489–494.

Steward, L. J., Ge, J., Stowe, R. L., Brown, D. C., Bruton, R. K., Stokes,

P. R., & Barnes, N. M. (1996). Ability of 5-HT4 receptor ligands to

modulate rat striatal dopamine release in vitro and in vivo. Br J

Pharmacol 117, 55–62.

Stowe, R. L., & Barnes, N. M. (1998). Selective labelling of 5-HT7 receptor

recognition sites in rat brain using [3H]5-carboxamidotryptamine.

Neuropharmacology 37, 1611–1619.

Stroebel, M., & Goppelt-Struebe, M. (1994). Signal transduction pathways

responsible for serotonin-mediated prostaglandin G/H synthase expres-

sion in rat mesangial cells. J Biol Chem 269, 22952–22957.

Strosznajder, J., Chalimoniuk, M., & Samochocki, M. (1996). Activation of

serotonergic 5-HT1A receptor reduces Ca2 + and glutamatergic receptor-

evoked arachidonic acid and NO/cGMP release in adult hippocampus.

Neurochem Int 28, 439–444.

Sugimoto, Y., Yamada, J., & Yoshikawa, T. (1999a). A neuronal nitric oxide

synthase inhibitor 7-nitroindazole reduces the 5-HT1A receptor against

8-OH-DPAT-elicited hyperphagia in rats. Eur J Pharmacol 376, 1–5.

Sugimoto, Y., Yoshikawa, T., & Yamada, J. (1999b). Involvement of nitric

oxide in the 5-HT1A autoreceptor-mediated hyperphagia in rats. Adv

Exp Med Biol 467, 109–111.

Sun, Q. Q., & Dale, N. (1998). Differential inhibition of N and P/Q Ca2 +

currents by 5-HT1A and 5-HT1D receptors in spinal neurons of Xenopus

larvae. J Physiol (Lond) 510, 103–120.

Takuwa, N., Ganz, M., Takuwa, Y., Sterzel, R. B., & Rasmussen, H. (1989).

Studies of the mitogenic effect of serotonin in rat renal mesangial cells.

Am J Physiol 257, F431–F439.

Tamir, H., Hsiung, S. C., Yu, P. Y., Liu, K. P., Adlersberg, M., Nunez, E. A.,

& Gershon, M. D. (1992). Serotonergic signalling between thyroid

cells: protein kinase C and 5-HT2 receptors in the secretion and action

of serotonin. Synapse 12, 155–168.

Tanaka, C., & Nishizuka, Y. (1994). The protein kinase C family for neuro-

nal signaling. Annu Rev Neurosci 17, 551–567.

Taussig, R., & Zimmermann, G. (1998). Type-specific regulation of mam-

malian adenylyl cyclases by G protein pathways. Adv Second Messen-

ger Phosphoprotein Res 32, 81–98.

Terron, J. A. (1996). The relaxant 5-HT receptor in the dog coronary artery

smooth muscle: pharmacological resemblance to the cloned 5-ht7

receptor subtype. Br J Pharmacol 118, 1421–1428.

Timpe, L. C., & Fantl, W. J. (1994). Modulation of a voltage-activated

potassium channel by peptide growth factor receptors. J Neurosci 14,

1195–1201.

Tohda, M., Tohda, C., Oda, H., & Nomura, Y. (1995). Possible involvement

of botulinum ADP-ribosyltransferase sensitive low molecular G-protein

on 5-hydroxytryptamine (5-HT)-induced inositol phosphates formation

in 5-HT2c cDNA transfected cells. Neurosci Lett 190, 33–36.

Torres, G. E., Chaput, Y., & Andrade, R. (1995). Cyclic AMP and protein

kinase A mediate 5-hydroxytryptamine type 4 receptor regulation of

calcium-activated potassium current in adult hippocampal neurons.

Mol Pharmacol 47, 191–197.

Tournois, C., Mutel, V., Manivet, P., Launay, J. M., & Kellermann, O.

(1998). Cross-talk between 5-hydroxytryptamine receptors in a seroto-

nergic cell line. Involvement of arachidonic acid metabolism. J Biol

Chem 273, 17498–17503.

Trillat, A. C., Malagie, I., Scearce, K., Pons, D., Anmella, M. C., Jacquot, C.,

Hen, R., & Gardier, A. M. (1997). Regulation of serotonin release in

the frontal cortex and ventral hippocampus of homozygous mice lack-

ing 5-HT1B receptors: in vivo microdialysis studies. J Neurochem 69,

2019–2025.

Tsou, A. P., Kosaka, A., Bach, C., Zuppan, P., Yee, C., Tom, L., Alvarez, R.,

Ramsey, S., Bonhaus, D. W., & Stefanich, E. (1994). Cloning and ex-

pression of a 5-hydroxytryptamine7 receptor positively coupled to ad-

enylyl cyclase. J Neurochem 63, 456–464.

Uezono, Y., Bradley, J., Min, C., McCarty, N. A., Quick, M., Riordan, J. R.,

Chavkin, C., Zinn, K., Lester, H. A., & Davidson, N. (1993). Receptors

that couple to 2 classes of G proteins increase cAMP and activate CFTR

expressed in Xenopus oocytes. Recept Channels 1, 233–241.

Ullmer, C., Schmuck, K., Kalkman, H. O., & Lubbert, H. (1995). Expression

of serotonin receptor mRNAs in blood vessels. FEBS Lett 370, 215–221.

Ullmer, C., Boddeke, H. G., Schmuck, K., & Lubbert, H. (1996). 5-HT2B

receptor-mediated calcium release from ryanodine-sensitive intracellu-

lar stores in human pulmonary artery endothelial cells. Br J Pharmacol

117, 1081–1088.

Van den Wyngaert, I., Gommeren, W., Verhasselt, P., Jurzak, M., Leysen, J.,

Luyten, W., & Bender, E. (1997). Cloning and expression of a human

serotonin 5-HT4 receptor cDNA. J Neurochem 69, 1810–1819.

Vane, J. R. (1959). The relative activities of some tryptamine analogues on

the isolated rat stomach strip preparation. Br J Pharmacol Chemother

14, 87–98.

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212 211

Van Sande, J., Allgeier, A., Massart, C., Czernilofsky, A., Vassart, G.,

Dumont, J. E., & Maenhaut, C. (1993). The human and dog 5-HT1D

receptors can both activate and inhibit adenylate cyclase in transfected

cells. Eur J Pharmacol 247, 177–184.

Varrault, A., & Bockaert, J. (1992). Differential coupling of 5-HT1A recep-

tors occupied by 5-HT or 8-OH-DPAT to adenylyl cyclase. Naunyn

Schmiedebergs Arch Pharmacol 346, 367–374.

Varrault, A., Bockaert, J., & Waeber, C. (1992a). Activation of 5-HT1A

receptors expressed in NIH-3T3 cells induces focus formation and

potentiates EGF effect on DNA synthesis. Mol Biol Cell 3, 961–969.

Varrault, A., Journot, L., Audigier, Y., & Bockaert, J. (1992b). Transfection

of human 5-hydroxytryptamine1A receptors in NIH-3T3 fibroblasts:

effects of increasing receptor density on the coupling of 5-hydroxytryp-

tamine1A receptors to adenylyl cyclase. Mol Pharmacol 41, 999–1007.

Verbeuren, T. J. (1993). Vasodilator effect of tertatolol in isolated perfused

rat kidneys: involvement of endothelial 5-HT1A receptors. Cardiology

83, 5–9.

Verbeuren, T. J., Mennecier, P., & Laubie, M. (1991). 5-Hydroxytrypt-

amine-induced vasodilatation in the isolated perfused rat kidney: are

endothelial 5-HT1A receptors involved? Eur J Pharmacol 201, 17–27.

Verge, D., Daval, G., Patey, A., Gozlan, H., el Mestikawy, S., & Hamon, M.

(1985). Presynaptic 5-HT autoreceptors on serotonergic cell bodies and/

or dendrites but not terminals are of the 5-HT1A subtype. Eur J Phar-

macol 113, 463–464.

Verge, D., Daval, G., Marcinkiewicz, M., Patey, A., el Mestikawy, S.,

Gozlan, H., & Hamon, M. (1986). Quantitative autoradiography of

multiple 5-HT1 receptor subtypes in the brain of control or 5,7-dihy-

droxytryptamine-treated rats. J Neurosci 6, 3474–3482.

Vilaro, M. T., Cortes, R., Gerald, C., Branchek, T. A., Palacios, J. M., &

Mengod, G. (1996). Localization of 5-HT4 receptor mRNA in rat brain

by in situ hybridization histochemistry. Brain Res Mol Brain Res 43,

356–360.

Villalon, C. M., Centurion, D., Lujan-Estrada, M., Terron, J. A., & San-

chez-Lopez, A. (1997). Mediation of 5-HT-induced external carotid

vasodilatation in GR 127935-pretreated vagosympathectomized dogs

by the putative 5-HT7 receptor. Br J Pharmacol 120, 1319–1327.

Voigt, M. M., Laurie, D. J., Seeburg, P. H., & Bach, A. (1991). Molecular

cloning and characterization of a rat brain cDNA encoding a 5-hydrox-

ytryptamine1B receptor. EMBO J 10, 4017–4023.

Waeber, C., & Moskowitz, M. A. (1995). Autoradiographic visualisation of

[3H]5-carboxamidotryptamine binding sites in the guinea pig and rat

brain. Eur J Pharmacol 283, 31–46.

Waeber, C., Grailhe, R., Yu, X. J., Hen, R., & Moskowitz, M. A. (1998).

Putative 5-ht5 receptors: localization in the mouse CNS and lack of

effect in the inhibition of dural protein extravasation. Ann N Y Acad

Sci 861, 85–90.

Wainscott, D. B., Cohen, M. L., Schenck, K. W., Audia, J. E., Nissen, J. S.,

Baez, M., Kursar, J. D., Lucaites, V. L., & Nelson, D. L. (1993).

Pharmacological characteristics of the newly cloned rat 5-hydroxytryp-

tamine2F receptor. Mol Pharmacol 43, 419–426.

Walter, A. E., Hoger, J. H., Labarca, C., Yu, L., Davidson, N., & Lester, H.

A. (1991). Low molecular weight mRNA encodes a protein that con-

trols serotonin 5-HT1c and acetylcholine M1 receptor sensitivity in

Xenopus oocytes. J Gen Physiol 98, 399–417.

Watson, J. M., Burton, M. J., Price, G. W., Jones, B. J., & Middlemiss, D.

N. (1996). GR127935 acts as a partial agonist at recombinant human

5-HT1Da and 5-HT1Db receptors. Eur J Pharmacol 314, 365–372.

Watts, S. W. (1998). Activation of the mitogen-activated protein kinase

pathway via the 5-HT2A receptor. Ann N Y Acad Sci 861, 162–168.

Weinshank, R. L., Zgombick, J. M., Macchi, M. J., Branchek, T. A., &

Hartig, P. R. (1992). Human serotonin 1D receptor is encoded by a

subfamily of two distinct genes: 5-HT1Da and 5-HT1Db. Proc Natl Acad

Sci USA 89, 3630–3634.

Weiss, S., Sebben, M., Kemp, D. E., & Bockaert, J. (1986). Serotonin 5-HT1

receptors mediate inhibition of cyclic AMP production in neurons. Eur J

Pharmacol 120, 227–230.

Wischmeyer, E., & Karschin, A. (1996). Receptor stimulation causes slow

inhibition of IRK1 inwardly rectifying K + channels by direct protein

kinase A-mediated phosphorylation. Proc Natl Acad Sci USA 93,

5819–5823.

Wisden, W., Parker, E. M., Mahle, C. D., Grisel, D. A., Nowak, H. P.,

Yocca, F. D., Felder, C. C., Seeburg, P. H., & Voigt, M. M. (1993).

Cloning and characterization of the rat 5-HT5B receptor. Evidence that

the 5-HT5B receptor couples to a G protein in mammalian cell mem-

branes. FEBS Lett 333, 25–31.

Witz, P., Amlaiky, N., Plassat, J. L., Maroteaux, L., Borrelli, E., & Hen, R.

(1990). Cloning and characterization of a Drosophila serotonin recep-

tor that activates adenylate cyclase. Proc Natl Acad Sci USA 87,

8940–8944.

Wolf, W. A., & Schutz, L. J. (1997). The serotonin 5-HT2C receptor is a

prominent serotonin receptor in basal ganglia: evidence from functional

studies on serotonin-mediated phosphoinositide hydrolysis. J Neuro-

chem 69, 1449–1458.

Wurch, T., Palmier, C., Colpaert, F. C., & Pauwels, P. J. (1997). Sequence

and functional analysis of cloned guinea pig and rat serotonin 5-HT1D

receptors: common pharmacological features within the 5-HT1D recep-

tor subfamily. J Neurochem 68, 410–418.

Xie, E., Zhu, L., Zhao, L., & Chang, L. S. (1996). The human serotonin

5-HT2C receptor: complete cDNA, genomic structure, and alternatively

spliced variant. Genomics 35, 551–561.

Yamada, J., Sugimoto, Y., Yoshikawa, T., & Horisaka, K. (1996). Effects of a

nitric oxide synthase inhibitor on 5-HT1A receptor agonist 8-OH-DPAT-

induced hyperphagia in rats. Eur J Pharmacol 316, 23–26.

Yamada, J., Sugimoto, Y., & Yoshikawa, T. (1998). Effects of adrenalec-

tomy on hyperphagia induced by the 5-HT1A receptor agonist 8-OH-

DPAT and 2-deoxy-D-glucose in rats. Neuroreport 9, 1831–1833.

Yamaguchi, F., & Brenner, S. (1997). Molecular cloning of 5-hydroxytrypt-

amine (5-HT) type 1 receptor genes from the Japanese puffer fish, Fugu

rubripes. Gene 191, 219–223.

Yu, L., Nguyen, H., Le, H., Bloem, L. J., Kozak, C. A., Hoffman, B. J.,

Snutch, T. P., Lester, H. A., Davidson, N., & Lubbert, H. (1991). The

mouse 5-HT1C receptor contains eight hydrophobic domains and is

X-linked. Brain Res Mol Brain Res 11, 143–149.

Zgombick, J. M., & Branchek, T. A. (1998). Native 5-HT1B receptors

expressed in OK cells display dual coupling to elevation of intracellular

calcium concentrations and inhibition of adenylate cyclase. Naunyn

Schmiedebergs Arch Pharmacol 358, 503–508.

Zgombick, J. M., Beck, S. G., Mahle, C. D., Craddock-Royal, B., &

Maayani, S. (1989). Pertussis toxin-sensitive guanine nucleotide-bind-

ing protein(s) couple adenosine A1 and 5-hydroxytryptamine1A recep-

tors to the same effector systems in rat hippocampus: biochemical and

electrophysiological studies. Mol Pharmacol 35, 484–494.

Zgombick, J. M., Schechter, L. E., Macchi, M., Hartig, P. R., Branchek,

T. A., & Weinshank, R. L. (1992). Human gene S31 encodes the

pharmacologically defined serotonin 5-hydroxytryptamine1E receptor.

Mol Pharmacol 42, 180–185.

Zgombick, J. M., Schechter, L. E., Adham, N., Kucharewicz, S. A., Wein-

shank, R. L., & Branchek, T. A. (1996). Pharmacological characteriza-

tions of recombinant human 5-HT1Da and 5-HT1Db receptor subtypes

coupled to adenylate cyclase inhibition in clonal cell lines: apparent

differences in drug intrinsic efficacies between human 5-HT1D subtypes.

Naunyn Schmiedebergs Arch Pharmacol 354, 226–236.

Zifa, E., & Fillion, G. (1992). 5-Hydroxytryptamine receptors. Pharmacol

Rev 44, 401–458.

J.R. Raymond et al. / Pharmacology & Therapeutics 92 (2001) 179–212212


Recommended