Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-sjtt6 Total loading time: 0 Render date: 2024-06-16T10:18:00.634Z Has data issue: false hasContentIssue false

6 - The serotonin transporter knock-out rat: a review

Published online by Cambridge University Press:  06 July 2010

Allan V. Kalueff
Affiliation:
Georgetown University Medical Center
Justin L. LaPorte
Affiliation:
National Institute of Mental Health
Get access

Summary

ABSTRACT

This chapter dicusses the most recent data on the serotonin transporter knock-out rat, a unique rat model that has been generated by target-selected N-ethyl-N-nitrosourea (ENU) driven mutagenesis. The knock-out rat is the result of a premature stopcodon in the serotonin transporter gene, and the absence of the serotonin transporter has been confirmed at mRNA, protein, and functional levels. The serotonin transporter (SERT) plays a crucial role in serotonin reuptake and its absence has a huge effect on serotonin neurotransmission – exemplified by increased extracellular serotonin levels, reduced serotonin tissue/platelet/blood levels, and reduced evoked serotonin release – yet the animals appear normal and do not differ from wildtype littermates in respect to breeding and health. Behavioral phenotypes are only apparent when the animals are exposed to certain stimuli. For instance, the serotonin transporter knock-out rat displays increased stress sensitivity in a variety of anxiety- and depression-like tests, such as the elevated plus maze test and the forced swim test. Also remarkable, while general activity is not changed, the knock-out rats show a “neurotic-like” exploratory pattern. In line with the serotonin hypothesis of impulsivity, which argues that there is an inverse relationship between the two, serotonin transporter knock-out rats show reduced motor impulsivity in the five-choice serial reaction time task, and a reduction in social interaction during play and aggressive encounters. Interestingly, abdominal fat seems to be increased in the knock-out rat, despite normal body weight. Pharmacological compounds also elicit genotype-dependent responses in the knock-out rats.

Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2010

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Abel, K, Waikar, M, Pedro, B, Hemsley, D, Geyer, M (1998). Repeated testing of prepulse inhibition and habituation of the startle reflex: a study in healthy human controls. J Psychopharmacol 12: 330–7.CrossRefGoogle ScholarPubMed
Adamec, R, Burton, P, Blundell, J, Murphy, D L, Holmes, A (2006). Vulnerability to mild predator stress in serotonin transporter knockout mice. Behav Brain Res 170: 126–40.CrossRefGoogle ScholarPubMed
Altamura, C, Dell'Acqua, M L, Moessner, R, Murphy, D L, Lesch, K P, Persico, A M (2007). Altered neocortical cell density and layer thickness in serotonin transporter knockout mice: a quantitation study. Cereb Cortex 17: 1394–401.CrossRefGoogle ScholarPubMed
Ansorge, M S, Zhou, M, Lira, A, Hen, R, Gingrich, J A (2004). Early-life blockade of the 5-HT transporter alters emotional behavior in adult mice. Science 306: 879–81.CrossRefGoogle ScholarPubMed
Asberg, M, Thoren, P, Traskman, L, Bertilsson, L, Ringberger, V (1976a). “Serotonin depression” – a biochemical subgroup within the affective disorders?Science 191: 478–80.CrossRefGoogle ScholarPubMed
Asberg, M, Träskman, L, Thorén, P (1976b). 5-HIAA in the cerebrospinal fluid. A biochemical suicide predictor?Arch Gen Psychiatry 33: 1193–97.CrossRefGoogle ScholarPubMed
Baker, K E, Parker, R (2004). Nonsense-mediated mRNA decay: terminating erroneous gene expression. Curr Opin Cell Biol 16: 293–9.CrossRefGoogle ScholarPubMed
Bengel, D, Murphy, D L, Andrews, A M, et al. (1998). Altered brain serotonin homeostasis and locomotor insensitivity to 3,4-methylenedioxymethamphetamine (“Ecstasy”) in serotonin transporter-deficient mice. Mol Pharmacol 53: 649–55.CrossRefGoogle Scholar
Benmansour, S, Cecchi, M, Morilak, D A, et al. (1999). Effects of chronic antidepressant treatments on serotonin transporter function, density and mRNA level. J Neuroscience 19: 10 494–501.CrossRefGoogle ScholarPubMed
Bentley, A, MacLennan, B, Calvo, J, Dearolf, C R (2000). Targeted recovery of mutations in Drosophila. Genetics 156: 1169–73.Google ScholarPubMed
Bill, D J, Knight, M, Forster, E A, Fletcher, A (1991). Direct evidence for an important species difference in the mechanism of 8-OH-DPAT-induced hypothermia. Br J Pharmacol 103: 1857–64.CrossRefGoogle ScholarPubMed
Birrell, J M, Brown, V J (2000). Medial frontal cortex mediates perceptual attention set shifting in the rat. J Neurosci 20: 4320–4.CrossRefGoogle ScholarPubMed
Blier, P, Montigny, C, Chaput, Y (1990). A role for the serotonin system in the mechanism of action of antidepressant treatments: preclinical evidence. J Clin Psychiatry 51: 14–20.Google ScholarPubMed
Blier, P, Seletti, B, Gilbert, F, Young, S N, Benkelfat, C (2002). Serotonin 1A receptor activation and hypothermia in humans: lack of evidence for a presynaptic mediation. Neuropsychopharmacology 27: 301–08.CrossRefGoogle ScholarPubMed
Bolles, R C, Woods, P J (1964). The ontogeny of behavior in the albino rat. Anim Behav 12: 427–41.CrossRefGoogle Scholar
Boulougouris, V, Dalley, J W, Robbins, T W (2007). Effects of orbitofrontal, infralimbic, and prelimbic cortical lesions on serial spatial reversal learning in the rat. Behav Brain Res 179: 219–28.CrossRefGoogle ScholarPubMed
Bouwknecht, J A, Hijzen, T H, , G J, Maes, R A, Olivier, B (2000). Stress-induced hyperthermia in mice: effects of flesinoxan on heart rate and body temperature. Eur J Pharmacol 400: 59–66.CrossRefGoogle ScholarPubMed
Boyce-Rustay, J M, Wiedholz, L M, Millstein, R A, et al. (2006). Ethanol-related behaviors in serotonin transporter knockout mice. Alcohol Clin Exp Res 30: 1957–65.CrossRefGoogle ScholarPubMed
Braff, D L, Geyer, M A (1990). Sensorimotor gating and schizophrenia. Human and animal model studies. Arch Gen Psychiatry 47: 181–8.CrossRefGoogle ScholarPubMed
Briley, M, Moret, C (1993). Neurobiological mechanisms involved in antidepressant therapies. Clin Neuropharmacol 16: 387–400.CrossRefGoogle ScholarPubMed
Cannon, T D, Zorrilla, L E, Shtasel, D, et al. (1994). Neuropsychological functioning in siblings discordant for schizophrenia and healthy volunteers. Arch Gen Psychiatry 51: 651–61.CrossRefGoogle ScholarPubMed
Capecchi, M R (1989a). Altering the genome by homologous recombination. Science 244: 1288–92.CrossRefGoogle ScholarPubMed
Capecchi, M R (1989b). The new mouse genetics: altering the genome by gene targeting. Trends Genet 5: 70–6.CrossRefGoogle ScholarPubMed
Carey, R J, DePalma, G, Damianopoulos, E, Muller, C P, Huston, J P (2004). The 5-HT1A receptor and behavioral stimulation in the rat: effects of 8-OHDPAT on spontaneous and cocaine-induced behavior. Psychopharmacology 177: 46–54.CrossRefGoogle ScholarPubMed
Carey, R J, DePalma, G, Damianopoulos, E, Shanahan, A, Muller, C P, Huston, J P (2005). Evidence that the 5-HT1A autoreceptor is an important pharmacological target for the modulation of cocaine behavioral stimulant effects. Brain Res 1034: 162–71.CrossRefGoogle ScholarPubMed
Carli, M, Robbins, T W, Evenden, J L, Everitt, B J (1983). Effects of lesions to ascending noradrenergic neurons on performance of a 5-choice serial reaction time task in rats – implications for theories of dorsal noradrenergic bundle function based on selective attention and arousal. Behav Brain Res 9: 361–80.CrossRefGoogle Scholar
Carli, M, Samanin, R (1992). Serotonin2 receptor agonists and serotonergic anorectic drugs affect rats' performance differentially in a five-choice serial reaction time task. Psychopharmacology 106: 228–34.CrossRefGoogle Scholar
Carlsson, M, Carlsson, A (1988). A regional study of sex differences in rat brain serotonin. Prog Neuropsychopharmacol Biol Psychiatry 12: 53–61.CrossRefGoogle ScholarPubMed
Carroll, J C, Boyce-Rustay, J M, Millstein, R, et al. (2007). Effects of mild early life stress on abnormal emotion-related behaviors in 5-HTT knockout mice. Behav Genet 37: 214–22.CrossRefGoogle ScholarPubMed
Carroll, M E, Lac, S T, Ascencio, M, Kragh, R (1990). Fluoxetine reduces intravenous cocaine self-administration in rats. Pharmacol Biochem Behav 35: 237–44.CrossRefGoogle ScholarPubMed
Chon, H, Gaillard, C A, Meijden, B B, et al. (2004). Broadly altered gene expression in blood leukocytes in essential hypertension is absent during treatment. Hypertension 43: 947–51.CrossRefGoogle ScholarPubMed
Chudasama, Y, Passetti, F, Rhodes, S E, Lopian, D, Desai, A, Robbins, T W (2003). Dissociable aspects of performance on the 5-choice serial reaction time task following lesions of the dorsal anterior cingulate, infralimbic and orbitofrontal cortex in the rat: differential effects on selectivity, impulsivity and compulsivity. Behav Brain Res 146: 105–19.CrossRefGoogle ScholarPubMed
Clarke, H F, Dalley, J W, Crofts, H S, Robbins, T W, Roberts, A C (2004). Cognitive inflexibility after prefrontal serotonin depletion. Science 304: 878–80.CrossRefGoogle ScholarPubMed
Clarke, H F, Walker, S C, Crofts, H S, Robbins, T W, Roberts, A C (2005). Prefrontal serotonin depletion affects reversal learning but not attentional set shifting. J Neurosci 12: 532–8.CrossRefGoogle Scholar
Connor, T J, Kelliher, P, Shen, Y, Harkin, A, Kelly, J P, Leonard, B E (2000). Effect of subchronic antidepressant treatments on behavioral, neurochemical, and endocrine changes in the forced-swim test. Pharmacol Biochem Behav 65: 591–7.CrossRefGoogle ScholarPubMed
Coppen, A, Eccleston, E, Craft, I, Bye, P (1973). Letter: Total and free plasma-tryptophan concentration and oral contraception. Lancet 2: 1498.CrossRefGoogle ScholarPubMed
Cowen, P J, Parry-Billings, M, Newsholme, E A (1989). Decreased plasma tryptophan levels in major depression. J Affect Disord 16: 27–31.CrossRefGoogle ScholarPubMed
Cryan, J F, Kelliher, P, Kelly, J P, Leonard, B E (1999). Comparative effects of serotonergic agonists with varying efficacy at the 5-HT(1A) receptor on core body temperature: modification by the selective 5-HT(1A) receptor antagonist WAY 100635. J Psychopharmacol 13: 278–83.CrossRefGoogle ScholarPubMed
Dalley, J W, Cardinal, R N, Robbins, T W (2004). Prefrontal executive and cognitive functions in rodents: neural and neurochemical substrates. Neurosci Biobehav Rev 28: 771–84.CrossRefGoogle ScholarPubMed
Dalley, J W, Fryer, T D, Brichard, L, et al. (2007). Nucleus accumbens D2/3 receptors predict trait impulsivity and cocaine reinforcement. Science 315: 1267–70.CrossRefGoogle ScholarPubMed
Daws, L C, Montañez, S, Munn, J L, et al. (2006). Ethanol inhibits clearance of brain serotonin by a serotonin transporter-independent mechanism. J Neurosci 26: 6431–8.CrossRefGoogle ScholarPubMed
Dawson, L A, Nguyen, H Q, Smith, D L, Schechter, L E (2002). Effect of chronic fluoxetine and WAY-100635 treatment on serotonergic neurotransmission in the frontal cortex. J Psychopharmacol 16: 145–52.CrossRefGoogle ScholarPubMed
Boer, S F, Lesourd, M, Mocaer, E, Koolhaas, J M (2000). Somatodendritic 5-HT1A autoreceptors mediate the antiaggressive actions of 5-HT1A receptor agonists in rats: an ethopharmacological study with S-15535, Alenspirone and WAY-100635. Neuropsychopharmacology 23: 20–33.CrossRefGoogle Scholar
Bruin, J P C, Feenstra, M G P, Broersen, L M, et al. (2000). Role of the prefrontal cortex of the rat in learning and decision making: effects of transient inactivation. Prog Brain Res 126: 103–13.CrossRefGoogle ScholarPubMed
Jong, T R, Pattij, T, Veening, J G, Waldinger, M D, Cools, A R, Olivier, B (2005). Effects of chronic selective serotonin reuptake inhibitors on 8-OH-DPAT-induced facilitation of ejaculation in rats: comparison of fluvoxamine and paroxetine. Psychopharmacology 179: 509–15.CrossRefGoogle ScholarPubMed
Visser, L, Bos, R, Kuurman, W W, Kas, M J, Spruijt, B M (2006). Novel approach to the behavioral characterization of inbred mice: automated home cage observations. Genes Brain Behav 5: 458–66.CrossRefGoogle ScholarPubMed
Dekeyne, A, Brocco, M, Adhumeau, A, Gobert, A, Millan, M J (2000). The selective serotonin (5-HT)1A receptor ligand, S-15535, displays anxiolytic-like effects in the social interaction and Vogel models and suppresses dialysate levels of 5-HT in the dorsal hippocampus of freely moving rats. A comparison with other anxiolytic agents. Psychopharmacology 152: 55–66.CrossRefGoogle Scholar
Dominguez, R, Cruz-Morales, S E, Carvalho, M C, Xavier, M, Brandao, M L (2003). Sex differences in serotonergic activity in dorsal and median raphe nucleus. Physiol Behav 80: 203–10.CrossRefGoogle ScholarPubMed
Dragan, W Ł, Oniszczenko, W (2006). Association of a functional polymorphism in the serotonin transporter gene with personality traits in females in a Polish population. Neuropsychobiology 54: 45–50.CrossRefGoogle Scholar
Drevets, W C, Frank, E, Price, J C, et al. (1999). PET imaging of serotonin 1A receptor binding in depression. Biol Psychiatry 46: 1375–87.CrossRefGoogle ScholarPubMed
Edwards, D H, Kravitz, E A (1997). Serotonin, social status and aggression. Curr Opin Neurobiol 7: 812–9.CrossRefGoogle ScholarPubMed
Fabre, V, Beaufour, C, Evrard, A, et al. (2000). Altered expression and functions of serotonin 5-HT1A and 5-HT1B receptors in knock-out mice lacking the 5-HT transporter. Eur J Neurosci 12: 2299–310.CrossRefGoogle ScholarPubMed
Fairbanks, L A, Melega, W P, Jorgensen, M J, Kaplan, J R, McGuire, M T (2001). Social impulsivity inversely associated with CSF 5-HIAA and fluoxetine exposure in vervet monkeys. Neuropsychopharmacology 24: 370–8.CrossRefGoogle ScholarPubMed
Ferrari, P F, Palanza, P, Parmigiani, S, Almeida, R M, Miczek, K A (2005). Serotonin and aggressive behavior in rodents and nonhuman primates: predispositions and plasticity. Eur J Pharmacol 526: 259–73.CrossRefGoogle ScholarPubMed
Forcelli, P A, Heinrichs, S C (2007). Teratogenic effects of maternal antidepressant exposure on neural substrates of drug-seeking behavior in offspring. Addict Biol 13: 52–62.CrossRefGoogle ScholarPubMed
Fox, M A, Andrews, A M, Wendland, J R, Lesch, K P, Holmes, A, Murphy, D L (2007). A pharmacological analysis of mice with a targeted disruption of the serotonin transporter. Psychopharmacology 195: 147–66.CrossRefGoogle ScholarPubMed
Fuller, R W (1996). The influence of fluoxetine on aggressive behavior. Neuropsychopharmacology 14 (2): 77–81.CrossRefGoogle ScholarPubMed
Gerra, G, Zaimovic, A, Timpano, M, Zambelli, U, Delsignore, R, Brambilla, F (2000). Neuroendocrine correlates of temperamental traits in humans. Psychoneuroendorinology 25: 479–96.CrossRefGoogle ScholarPubMed
Geyer, M A, Krebs-Thomson, K, Braff, D L, Swerdlow, N R (2001). Pharmacological studies of prepulse inhibition models of sensorimotor gating deficits in schizophrenic patients: a decade in review. Psychopharmacology 156: 117–54.CrossRefGoogle Scholar
Geyer, M A, Tapson, G S (1988). Habituation of tactile startle is altered by drugs acting on serotonin-2 receptors. Neuropsychopharmacology 1: 135–47.CrossRefGoogle ScholarPubMed
Gibbs, R A, Weinstock, G M, Metzker, M L, et al. (2004). Genome sequence of the Brown Norway rat yields insight into mammalian evolution. Nature 428: 493–521.Google Scholar
Goodwin, G M, Souza, R J, Green, A R (1985). Presynaptic serotonin receptor-mediated response in mice attenuated by antidepressant drugs and electroconvulsive shock. Nature 317: 531–3.CrossRefGoogle ScholarPubMed
Goodwin, G M, Souza, R J, Green, A R, Heal, D J (1987). The pharmacology of the behavioral and hypothermic responses of rats to 8-hydroxy-2-(di-n-propylamino)tetralin (8-OH-DPAT). Psychopharmacology 91: 506–11.CrossRefGoogle Scholar
Gorman, J M (2006). Gender differences in depression and response to psychotropic medication. Gend Med 3: 93–109.CrossRefGoogle ScholarPubMed
Griffith, R R, Bigelow, G E, Henningfield, J E (1980). Similarities in animal and human drug-taking behavior. In Mello, NK, editor. Advances in substance abuse Vol. 1. Greenwich, CT: JAI Press, pp. 1–90.Google Scholar
Haberstick, B C, Smolen, A, Hewitt, J K (2006). Family-based association test of the 5HTTLPR and aggressive behavior in a general population sample of children. Biol Psychiatry 59: 36–43.CrossRefGoogle Scholar
Haleem, D J, Kennett, G A, Curzon, G (1990). Hippocampal 5-hydroxytryptamine synthesis is greater in female rats than in males and more decreased by the 5-HT1A agonist 8-OH-DPAT. J Neural Transm Gen Sect 79: 93–101.CrossRefGoogle ScholarPubMed
Harison, A A, Everitt, B J, Robbins, T W (1997). Central 5-HT depletion enhances impulsive responding without affecting the accuracy of attentional performance: interactions with dopaminergiz mechanisms. Psychopharmacology 133: 329–42.CrossRefGoogle Scholar
Heikkila, R E, Orlansky, H, Cohen, G (1975). Studies on the distinction between uptake inhibition and release of [3H]dopamine in rat brain tissue slices. Biochem Pharmacol 24: 847–52.CrossRefGoogle Scholar
Heinz, A, Jones, D W, Mazzanti, C, et al. (2000). A relationship between serotonin transporter genotype and in vivo protein expression and alcohol neurotoxicity. Biol Psychiatry 47: 643–9.CrossRefGoogle ScholarPubMed
Higley, J D, Linnoila, M (1997). Low central nervous system serotonergic activity is traitlike and correlates with impulsive behavior. A nonhuman primate model investigating genetic and environmental influences on neurotransmission. Ann NY Acad Sci USA 836: 39–56.CrossRefGoogle ScholarPubMed
Higgins, G A, Bradbury, A J, Jones, B J, Oakley, N R (1988). Behavioral and biochemical consequences following activation of 5HT1-like and GABA receptors in the dorsal raphe nucleus of the rat. Neuropharmacology 27: 993–1001.CrossRefGoogle ScholarPubMed
Hillegaart, V (1991). Effects of local application of 5-HT and 8-OH-DPAT into the dorsal and median raphe nuclei on core temperature in the rat. Psychopharmacology 103: 291–6.CrossRefGoogle ScholarPubMed
Hodos, W (1961). Progressive ratio as a measure of reward strength. Science 134: 943–4.CrossRefGoogle ScholarPubMed
Hogg, S (1996). A review of the validity and variability of the elevated plus-maze as an animal model of anxiety. Pharmacol Biochem Behav 54: 21–30.CrossRefGoogle ScholarPubMed
Hol, T, Berg, C, Ree, J M, Spruijt, B M (1999). Isolation during the play period in infancy decreases adult social interactions in rats. Behav Brain Res 100: 91–7.CrossRefGoogle ScholarPubMed
Holmes, A, Li, Q, Murphy, D L, Gold, E, Crawley, J N (2003a). Abnormal anxiety-related behavior in serotonin transporter null mutant mice: the influence of genetic background. Genes Brain Behav 2: 365–80.CrossRefGoogle ScholarPubMed
Holmes, A, Murphy, D L, Crawley, J N (2002a). Reduced aggression in mice lacking the serotonin transporter. Psychopharmacology 161: 160–7.CrossRefGoogle ScholarPubMed
Holmes, A, Murphy, D L, Crawley, J N (2003b). Abnormal behavioral phenotypes of serotonin transporter knockout mice: parallels with human anxiety and depression. Biol Psychiatry 54: 953–9.CrossRefGoogle ScholarPubMed
Holmes, A, Yang, R J, Murphy, D L, Crawley, J N (2002b). Evaluation of antidepressant-related behavioral responses in mice lacking the serotonin transporter. Neuropsychopharmacology 27: 914–23.CrossRefGoogle ScholarPubMed
Homberg, J R, Arends, B, Wardeh, G, Raasø, H S, Schoffelmeer, A N M, Vries, T J (2004). Individual differences in the effects of serotonergic anxiolytic drugs on the motivation to self-administer cocaine. Neuroscience 128: 121–30.CrossRefGoogle ScholarPubMed
Homberg, J R, Braam, B, Ellenbroek, B, Cuppen, E, Joles, J A (2006). Blood pressure in mutant rats lacking the 5-hydroxytryptamine (5-HT) transporter. Hypertension 48: e115–6.CrossRefGoogle Scholar
Homberg, J R, Olivier, J D A, Smits, B M G, et al. (2007a). Characterization of the serotonin transporter knockout rat: a selective change in the functioning of the serotonergic system. Neuroscience 146: 1662–76.CrossRefGoogle ScholarPubMed
Homberg, J R, Pattij, T, Janssen, M C W, et al. (2007c). Serotonin transporter deficiency in rats improves inhibitory control but not behavioral flexibility. Eur J Neurosci 26: 2066–73.CrossRefGoogle ScholarPubMed
Homberg, J R, Schiepers, O J G, Schoffelmeer, A N M, Cuppen, E, Vanderschuren, L J M J (2007b). Acute and constitutive increases in central serotonin levels reduce social play behavior in periadolescent rats. Psychopharmacology 195: 175–82.CrossRefGoogle Scholar
Jans, L A, Riedel, W J, Markus, C R, Blokland, A (2007). Serotonergic vulnerability and depression: assumptions, experimental evidence and implications. Mol Psychiatry 12: 522–43.CrossRefGoogle ScholarPubMed
Jansen, G, Hazendonk, E, Thijssen, K L, Plasterk, R H (1997). Reverse genetics by chemical mutagenesis in Caenorhabditis elegans. Nat Genet 17: 119–21.CrossRefGoogle ScholarPubMed
Kalueff, A V, Fox, M A, Gallagher, P S, Murphy, D L (2007a). Hypolocomotion, anxiety and serotonin syndrome-like behavior contribute to the complex phenotype of serotonin transporter knockout mice. Genes Brain Behav 6: 389–400.CrossRefGoogle ScholarPubMed
Kalueff, A V, Gallagher, P S, Murphy, D L (2006). Are serotonin transporter knockout mice ‘depressed’?: hypoactivity but no anhedonia. Neuroreport 17: 1347–51.CrossRefGoogle ScholarPubMed
Kalueff, A V, Jensen, C L, Murphy, D L (2007b). Locomotory patterns, spatiotemporal organization of exploration and spatial memory in serotonin transporter knockout mice. Brain Res 1169: 87–97.CrossRefGoogle ScholarPubMed
Kelaï, S, Aïssi, F, Lesch, K P, Cohen-Salmon, C, Hamon, M, Lanfumey, L (2003). Alcohol intake after serotonin transporter inactivation in mice. Alcohol Alcohol 38: 386–9.CrossRefGoogle ScholarPubMed
Kim, D K, Tolliver, T J, Huang, S J, et al. (2005). Altered serotonin synthesis, turnover and dynamic regulation in multiple brain regions of mice lacking the serotonin transporter. Neuropharmacology 49: 798–810.CrossRefGoogle ScholarPubMed
Koob, G F (1992). Drugs of abuse: anatomy, pharmacology and function of reward pathways. Trends Pharmacol Sci 13: 177–84.CrossRefGoogle ScholarPubMed
Kranzler, H, Lappalainen, J, Nellissery, M, Gelernter, J (2002). Association study of alcoholism subtypes with a functional promoter polymorphism in the serotonin transporter protein gene. Alcohol Clin Exp Res 26: 1330–5.CrossRefGoogle ScholarPubMed
Larson, E T, Summers, C H (2001). Serotonin reverses dominant social status. Behav Brain Res 121: 95–102.CrossRefGoogle ScholarPubMed
Lauder, J M (1990). Ontogeny of the serotonin system in the rat: serotonin as a developmental signal. Ann NY Acad Sci 600: 297–313.CrossRefGoogle ScholarPubMed
Leibowitz, S F, Alexander, J T (1998). Hypothalamic serotonin in control of eating behavior, meal size, and body weight. Biol Psychiatry 44: 851–64.CrossRefGoogle ScholarPubMed
LeMarquand, D, Pihl, R O, Benkelfat, C (1994). Serotonin and alcohol intake, abuse, and dependence: findings of animal studies. Biol Psychiatry 36: 395–421.CrossRefGoogle ScholarPubMed
Linnoila, M, Virkkunen, M, Scheinin, M, Nuutile, A, Rimon, R, Goodwin, F K (1983). Low cerebrospinal-fluid 5-hydroxyindoleacetic acid concentration differentiates impulsive from non-impulsive violent behavior. Life Sci 33: 2609–14.CrossRefGoogle Scholar
Lira, A, Zhou, M, Castanon, N, et al. (2003). Altered depression-related behaviors and functional changes in the dorsal raphe nucleus of serotonin transporter-deficient mice. Biol Psychiatry 54: 960–71.CrossRefGoogle ScholarPubMed
Mann, J J, Huang, Y Y, Underwood, M D, et al. (2000). A serotonin transporter gene promoter polymorphism. (5-HTTLPR) and prefrontal cortical binding in major depression and suicide. Arch Gen Psychiatry 57: 729–38.CrossRefGoogle ScholarPubMed
Masaki, D, Yokoyama, C, Kinoshita, S, et al. (2006). Relationship between limbic and cortical 5-HT neurotransmission and acquisition and reversal learning in a go/no-go task in rats. Psychopharmacology 189: 249–58.CrossRefGoogle Scholar
Maurel, S, Vry, J, Schreiber, R (1999). Comparison of the effects of the selective serotonin-reuptake inhibitors fluoxetine, paroxetine, citalopram and fluvoxamine in alcohol-preferring cAA rats. Alcohol 17: 195–201.CrossRefGoogle ScholarPubMed
Martin, K F, Phillips, I, Hearson, M, Prow, M R, Heal, D J (1992). Characterization of 8-OH-DPAT-induced hypothermia in mice as a 5-HT1A autoreceptor response and its evaluation as a model to selectively identify antidepressants. Br J Pharmacol 107: 15–21.CrossRefGoogle ScholarPubMed
Mathews, T A, Fedele, D E, Coppelli, F M, Avila, A M, Murphy, D L, Andrews, A M (2004). Gene dose-dependent alterations in extraneuronal serotonin but not dopamine in mice with reduced serotonin transporter expression. J Neurosci Methods 140: 169–81.CrossRefGoogle Scholar
McBride, W J, Murphy, J M, Gatto, G J, et al. (1993). CNS mechanisms of alcohol administration. Alcohol Alcohol 2: 463–7.Google Scholar
McCallum, C M, Comai, L, Greene, E A, Henikoff, S (2000). Targeted screening for induced mutations. Nat Biotechnol 18: 455–7.CrossRefGoogle ScholarPubMed
Millan, M J, Rivet, J M, Canton, H, et al. (1993a). S 15535: a highly selective benzodioxopiperazine 5-HT1A receptor ligand which acts as an agonist and an antagonist at presynaptic and postsynaptic sites respectively. Eur J Pharmacol 230: 99–102.CrossRefGoogle ScholarPubMed
Millan, M J, Rivet, J M, Canton, H, Marmouille-Girardon, S, Gobert, A (1993b). Induction of hypothermia as a model of 5-hydroxytryptamine1A receptor-mediated activity in the rat: a pharmacological characterization of the actions of novel agonists and antagonists. J Pharmacol Exp Ther 264: 1364–76.Google ScholarPubMed
Montañez, S, Owens, W A, Gould, G G, Murphy, D L, Daws, L C (2003). Exaggerated effect of fluvoxamine in heterozygote serotonin transporter knockout mice. J Neurochem 86: 210–9.CrossRefGoogle ScholarPubMed
Muir, J L, Everitt, B J, Robbins, T W (1996). The cerebral cortex of the rat and visual attentional function: dissociable effects of mediofrontal, cingulate, anterior dorsolateral, and parietal cortex lesions on a five-choice serial reaction time task. Cereb Cortex 6: 470–81.CrossRefGoogle ScholarPubMed
Murphy, F C, Smith, K A, Cowen, P J, Robbins, T W, Sahakian, B J (2002). The effects of tryptophan depletion on cognitive and affective processing in healthy volunteers. Psychopharmacology 163: 42–53.CrossRefGoogle ScholarPubMed
Muscat, R, Papp, M, Willner, P (1992). Reversal of stress-induced anhedonia by the atypical antidepressants, fluoxetine and maprotiline. Psychopharmacology 109: 433–8.CrossRefGoogle ScholarPubMed
Myers, D P, Renaut, S D, Collier, M J (2004). The Morris maze as a test of learning and memory in rats – a case study demonstrating the value of re-testing the same set of animals. Toxicology 202: 132–3.Google Scholar
Naranjo, C A, Kadlec, K E, Sanhueza, P, Woodley-Remus, D, Sellers, E M (1990). Fluoxetine differentially alters alcohol intake and other consummatory behaviors in problem drinkers. Clin Pharmacol Ther 47: 490–8.CrossRefGoogle ScholarPubMed
New, A S, Buchsbaum, M S, Hazlett, E A, et al. (2004). Fluoxetine increases relative metabolic rate in prefrontal cortex in impulsive aggression. Psychopharmacology 176: 451–8.CrossRefGoogle ScholarPubMed
Newman-Tancredi, A, Rivet, J, Chaput, C, Touzard, M, Verriele, L, Millan, M J (1999). The 5HT(1A) receptor ligand, S-15535, antagonises G-protein activation: a [35S] GTPgammaS and [3H]S-15535 autoradiography study. Eur J Pharmacol 384: 111–21.CrossRefGoogle Scholar
Ni, W, Lookingland, K, Watts, S W (2006). Arterial 5-hydroxytryptamine transporter function is impaired in deoxycorticosterone acetate and N{omega}-nitro-L-arginine but not spontaneously hypertensive rats. Hypertension 48: 134–40.CrossRefGoogle Scholar
Nishizawa, S, Benkelfat, C, Young, S N, et al. (1997). Differences between males and females in rates of serotonin synthesis in human brain. Proc Natl Acad Sci USA 94: 5308–13.CrossRefGoogle ScholarPubMed
Olivier, B (2004). Serotonin and aggression. Ann NY Acad Sci 1036: 382–92.CrossRefGoogle ScholarPubMed
Olivier, B, Chan, J S, Pattij, T, et al. (2006). Psychopharmacology of male rat sexual behavior: modeling human sexual dysfunctions?Int J Impot Res 18 (suppl. 1): S14–23.CrossRefGoogle ScholarPubMed
Olivier, J D A, Hart, M G C, Swelm, R P L, et al. (2008). A study in male and female serotonin transporter knockout rats: an animal model for anxiety and depression disorders. Neuroscience 152: 573–84.CrossRefGoogle ScholarPubMed
Olivier, B, Zethof, T J, Ronken, E, Heyden, J A (1998). Anxiolytic effects of flesinoxan in the stress-induced hyperthermia paradigm in singly-housed mice are 5-HT1A receptor mediated. Eur J Pharmacol 342: 177–82.CrossRefGoogle ScholarPubMed
Owens, M J, Nemeroff, C B (1998). The serotonin transporter and depression. Depression Anxiety 8: 5–12.3.0.CO;2-I>CrossRefGoogle ScholarPubMed
Paine, T A, Jackman, S L, Olmstead, M C (2002). Cocaine-induced anxiety: alleviation by diazepam, but not buspirone, dimenhydrinate or diphenhydramine. Behav Pharmacol 13: 511–23.CrossRefGoogle ScholarPubMed
Passetti, F, Dalley, J W, Robbins, T W (2003). Double dissociation of serotonergic and dopaminergic mechanisms on attentional performance using a rodent five-choice reaction time task. Psychopharmacology 165: 136–45.CrossRefGoogle ScholarPubMed
Pellis, S M, Pasztor, T J (1999). The developmental onset of a rudimentary form of play fighting in C57 mice. Dev Psychobiol 34: 175–82.3.0.CO;2-#>CrossRefGoogle ScholarPubMed
Peltier, R, Schenk, S (1993). Effects of serotonergic manipulations on cocaine self-administration in rats. Psychopharmacology 110: 390–4.CrossRefGoogle ScholarPubMed
Perry, J A, Wang, T L, Welham, T J, et al. (2003). A TILLING reverse genetics tool and a web-accessible collection of mutants of the legume Lotus japonicus. Plant Physiol 131: 866–71.CrossRefGoogle Scholar
Persico, A M, Mengual, E, Moessner, R, et al. (2001). Barrel pattern formation requires serotonin uptake by thalamocortical afferents, and not vesicular monoamine release. J Neurosci 21: 6862–73.CrossRefGoogle Scholar
Pierce, R C, Kumaresan, V (2006). The mesolimbic dopamine system: the final common pathway for the reinforcing effect of drugs of abuse?Neurosci Biobehav Rev 30: 215–38.CrossRefGoogle ScholarPubMed
Poole, T B, Fish, J (1975). An investigation of playful behavior in Rattus norvegicus and Mus musculus (Mammalia). J Zool 175: 61–71.CrossRefGoogle Scholar
Popova, L K (2006). From genes to aggressive behavior: the role of serotonergic system. Bioessays 28: 495–503.CrossRefGoogle ScholarPubMed
Prickaerts, J, Raaijmakers, W, Blokland, A (1996). Effects of myocardial infarction and captopril therapy on anxiety-related behaviors in the rat. Physiol Behav 60: 43–50.CrossRefGoogle ScholarPubMed
Prut, L, Belzung, C (2003). The open field as a paradigm to measure the effects of drugs on anxiety-like behaviors: a review. Eur J Pharmacol 463: 3–33.CrossRefGoogle Scholar
Reith, M E, Sershen, H, Lajtha, A (1980). Saturable [3H]cocaine binding in central nervous system of mouse. Life Sci 27: 1055–62.CrossRefGoogle ScholarPubMed
Richardson, N R, Roberts, D C S (1996). Fluoxetine pretreatment reduces breaking points on a progressive ratio schedule reinforced by intravenous cocaine self-administration in the rat. Life Sci 49: 833–840.CrossRefGoogle Scholar
Robbins, T W (2005). Chemistry of the mind: neurochemical modulation of prefrontal cortical function. J Comp Neurol 493: 140–6.CrossRefGoogle ScholarPubMed
Rogers, R D, Blackshaw, A J, Middleton, H C, et al. (1999). Tryptophan depletion impairs stimulus–reward learning while methylphenidate disrupts attentional control in healthy young adults: implications for the monaminergic basis of impulsive behavior. Psychopharmacology 146: 482–91.CrossRefGoogle Scholar
Rogers, D C, Fisher, E M, Brown, S D, Peters, J, Hunter, A J, Martin, J E (1997). Behavioral and functional analysis of mouse phenotype: SHIRPA, a proposed protocol for comprehensive phenotype assessment. Mamm Genome 8: 711–3.CrossRefGoogle ScholarPubMed
Salichon, N, Gaspar, P, Upton, A L (2001). Excessive activation of serotonin (5-HT) 1B receptors disrupts the formation of sensory maps in monoamine oxidase a and 5-HT transporter knock-out mice. J Neurosci 21: 884–96.CrossRefGoogle ScholarPubMed
Sargent, P A, Kjaer, K H, Bench, C J, et al. (2000). Brain serotonin1A receptor binding measured by positron emission tomography with [11C]WAY-100635: effects of depression and antidepressant treatment. Arch Gen Psychiatry 57: 174–80.CrossRefGoogle Scholar
Saykin, A J, Shtasel, D, Gur, R E, et al. (1994). Neuropsychological deficits in neuroleptic naive patients with first-episode schizophrenia. Arch Gen Psychiatry 51: 124–31.CrossRefGoogle ScholarPubMed
Seibell, P J, Demarest, J, Rhoads, D E (2003). 5-HT1A receptor activity disrupts spontaneous alternation behavior in rats. Pharmacol Biochem Behav 74: 559–64.CrossRefGoogle ScholarPubMed
Serretti, A, Mandelli, L, Lorenzi, C, et al. (2006). Temperament and character in mood disorders: influence of DRD4, SERTPR, TPH and MAO-A polymorphisms. Neuropsychobiology 53: 9–16.CrossRefGoogle ScholarPubMed
Shen, H W, Hagino, Y, Kobayashi, H, et al. (2004). Regional differences in extracellular dopamine and serotonin assessed by in vivo microdialysis in mice lacking dopamine and/or serotonin transporters. Neuropsychopharmacology 29: 1790–9.CrossRefGoogle ScholarPubMed
Shephard, R A, Broadhurst, P L (1982). Hyponeophagia and arousal in rats: effects of diazepam, 5-methoxy-N,N-dimethyltryptamine, d-amphetamine and food deprivation. Psychopharmacology 78: 368–72.CrossRefGoogle ScholarPubMed
Simpson, K L, Fisher, T M, Waterhouse, B D, Lin, R C (1998). Projection patterns from the raphe nuclear complex to the ependymal wall of the ventricular system in the rat. J Comp Neurol 399: 61–72.3.0.CO;2-8>CrossRefGoogle ScholarPubMed
Smits, B M G, Cuppen, E (2005). Rat genetics: the next episode. Trends Genetics 22: 232–40.CrossRefGoogle Scholar
Smits, B M G, Mudde, J, Plasterk, R H A, Cuppen, E (2004). Target-selected mutagenesis of the rat. Genomics 83: 332–4.CrossRefGoogle ScholarPubMed
Smits, B M, Mudde, J B, Belt, J, et al. (2006). Generation of gene knockouts and mutant models in the laboratory rat by ENU-driven target-selected mutagenesis. Pharmacogenet Genomics 16: 159–69.Google Scholar
Sora, I, Hall, F S, Andrews, A M, et al. (2001). Molecular mechanisms of cocaine reward: combined dopamine and serotonin transporter knockouts eliminate cocaine place preference. Proc Natl Acad Sci USA 98: 5300–05.CrossRefGoogle ScholarPubMed
Sora, I, Wichems, C, Takahashi, N, et al. (1998). Cocaine reward models: conditioned place preference can be established in dopamine- and in serotonin-transporter knockout mice. Proc Natl Acad Sci USA 13: 7699–704.CrossRefGoogle Scholar
Taravosh-Lahn, K, Bastida, C, Delville, Y (2006). Differential responsiveness to fluoxetine during puberty. Behav Neurosci 120: 1084–92.CrossRefGoogle ScholarPubMed
Till, B J, Reynolds, S H, Weil, C, et al. (2004). Discovery of induced point mutations in maize genes by TILLING. BMC Plant Biol 28: 4–12.Google Scholar
Berg, C L, Hol, T, Ree, J M, Spruijt, B M, Everts, H, Koolhaas, J M (1999). Play is indispensable for an adequate development of coping with social challenges in the rat. Dev Psychobiol 34: 129–38.3.0.CO;2-L>CrossRefGoogle ScholarPubMed
Vanderschuren, L J M J, Niesink, R J M, Ree, J M (1997). The neurobiology of social play behavior in rats. Neurosci Biobehav Rev 21: 309–26.CrossRefGoogle ScholarPubMed
Vizi, E S, Zsilla, G, Caron, M G, Kiss, J P (2004). Uptake and release of norepinephrine by serotonergic terminals in norepinephrine transporter knock-out mice: implications for the action of selective serotonin reuptake inhibitors. J Neurosci 24: 7888–94.CrossRefGoogle ScholarPubMed
Walsh, S L, Cunningham, K A (1997). Serotonergic mechanisms involved in the discriminative stimulus, reinforcing and subjective effects of cocaine. Psychopharmacology 130: 41–58.CrossRefGoogle ScholarPubMed
Watts, S W (2005). 5-HT in systemic hypertension: foe, friend or fantasy?Clin Sci (Lond.) 108: 399–412. Review. PMID: 15831089.CrossRefGoogle ScholarPubMed
Watts, A G, Stanley, H F (1984). Indoleamines in the hypothalamus and area of the midbrain raphe nuclei of male and female rats throughout postnatal development. Neuroendocrinology 38: 461–6.CrossRefGoogle ScholarPubMed
Wellman, C L, Izquierdo, A, Garrett, J E, et al. (2007). Impaired stress-coping and fear extinction and abnormal corticolimbic morphology in serotonin transporter knock-out mice. J Neurosci 27: 684–91.CrossRefGoogle ScholarPubMed
Whitaker-Azmitia, P M (2005). Behavioral and cellular consequences of increasing serotonergic activity during brain development: a role in autism?Int J Dev Neurosci 23: 75–83.CrossRefGoogle ScholarPubMed
Wichems, C H, Andrews, A M, Heils, A, Li, Q, Lesch, K P, Murphy, D L (1998). Spontaneous behavior differences and altered responses to psychomotor stimulants in mice lacking the serotonin transporter. 28th Annual Meeting of the Society for Neuroscience.Google Scholar
Willner, P, Towell, A, Sampson, D, Sophokleous, S, Muscat, R (1987). Reduction of sucrose preference by chronic unpredictable mild stress, and its restoration by a tricyclic antidepressant. Psychopharmacology 93: 358–64.CrossRefGoogle ScholarPubMed
Winstanley, C A, Chadusama, Y, Dalley, J W, Theobald, D E, Glennon, J C, Robbins, T W (2003). Intra-prefrontal 8-OHDPAT and M100907 improve visuospatial attention and decrease impulsivity on the five-choice serial reaction time task in rats. Psychopharmacology 167: 304–14.CrossRefGoogle Scholar
Winstanley, C A, Dalley, J W, Theobald, D E, Robbins, T W (2004). Fractionating impulsivity: contrasting effects of central 5-HT depletion on different measures of impulsive behavior. Neuropsychopharmacology 7: 1331–43.CrossRefGoogle Scholar
Winstanley, C A, Eagle, D M, Robbins, T W (2006). Behavioral models of impulsivity in relation to ADHD: translation between clinical and preclinical studies. Clin Psychol Rev 26: 379–95.CrossRefGoogle Scholar
Wise, R A (2002). Brain reward circuitry: insights from unsensed incentives. Neuron 36: 229–40.CrossRefGoogle ScholarPubMed
Yamada, J, Sugimoto, Y, Horisaka, K (1988). The behavioral effects of 8-hydroxy-2-(di-N-propylamino)tetralin (8-OH-DPAT) in mice. Eur J Pharmacol 154: 299–304.CrossRefGoogle ScholarPubMed
Zan, Y, Haag, J D, Chen, K S, et al. (2003). Production of knockout rats using ENU mutagenesis and a yeast-based screening assay. Nat Biotech 21: 645–51.CrossRefGoogle Scholar
Zhao, S, Edwards, J, Carroll, J, et al. (2006). Insertion mutation at the C-terminus of the serotonin transporter disrupts brain serotonin function and emotion-related behaviors in mice. Neuroscience 140: 321–34.CrossRefGoogle ScholarPubMed
Zhou, F C, Lesch, K P, Murphy, D L (2002). Serotonin uptake into dopamine neurons via dopamine transporters: a compensatory alternative. Brain Res 942: 109–19.CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×